You are on page 1of 805

Mathematical and Analytical Techniques with

Applications to Engineering

Petre P. Teodorescu

Treatise on
Classical Elasticity
Theory and Related Problems
Mathematical and Analytical Techniques
with Applications to Engineering

Series Editor
Alan Jeffrey
Newcastle upon Tyne, UK
Deceased 6 June 2010

For further volumes:


http://www.springer.com/series/7311
The importance of mathematics in the study of problems arising from the real world, and the
increasing success with which it has been used to model situations ranging from the purely
deterministic to the stochastic, in all areas of today’s Physical Sciences and Engineering, is
well established. The progress in applicable mathematics has been brought about by the
extension and development of many important analytical approaches and techniques, in
areas both old and new, frequently aided by the use of computers without which the solution
of realistic problems in modern Physical Sciences and Engineering would otherwise have
been impossible. The purpose of the series is to make available authoritative, up to date, and
self-contained accounts of some of the most important and useful of these analytical
approaches and techniques. Each volume in the series will provide a detailed introduction to
a specific subject area of current importance, and then will go beyond this by reviewing
recent contributions, thereby serving as a valuable reference source.
Petre P. Teodorescu

Treatise on Classical
Elasticity
Theory and Related Problems

123
Petre P. Teodorescu
Faculty of Mathematics
University of Bucharest
Bucharest
Romania

ISSN 1559-7458 ISSN 1559-7466 (electronic)


ISBN 978-94-007-2615-4 ISBN 978-94-007-2616-1 (eBook)
DOI 10.1007/978-94-007-2616-1
Springer Dordrecht Heidelberg New York London

Library of Congress Control Number: 2012952139

Ó Springer Science?Business Media Dordrecht 2013


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of
the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or
information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed. Exempted from this legal reservation are brief
excerpts in connection with reviews or scholarly analysis or material supplied specifically for the
purpose of being entered and executed on a computer system, for exclusive use by the purchaser of the
work. Duplication of this publication or parts thereof is permitted only under the provisions of
the Copyright Law of the Publisher’s location, in its current version, and permission for use must always
be obtained from Springer. Permissions for use may be obtained through RightsLink at the Copyright
Clearance Center. Violations are liable to prosecution under the respective Copyright Law.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt
from the relevant protective laws and regulations and therefore free for general use.
While the advice and information in this book are believed to be true and accurate at the date of
publication, neither the authors nor the editors nor the publisher can accept any legal responsibility for
any errors or omissions that may be made. The publisher makes no warranty, express or implied, with
respect to the material contained herein.

Printed on acid-free paper

Springer is part of Springer Science?Business Media (www.springer.com)


Preface

The mechanics of deformable solids brings its contribution both to the knowledge
of the phenomena of the surrounding physical reality, from a theoretical point of
view, and to the determination of the state of strain and stress in various elements
of construction, from practical considerations. The deformable solids have a
particularly complex character; a mathematical modelling of them is not simple
and often leads to inextricable difficulties of computation. One of the most simple
mathematical models and, at the same time, the most used one, is the model of the
elastic body—especially the linear elastic model; despite its simplicity, even this
model of real body may lead to great difficulties of calculation.
In general, the engineering constructions have been based, during the centuries,
on empirical methods; beginning with the seventeenth century, one has obtained a
lot of results, which form what now is called ‘‘Strength of Materials’’, where
simplifying supplementary hypotheses have been introduced. As a matter of fact,
this denomination is not a proper one, because it corresponds only to a mechanical
phenomenon modelled by the so-called ‘‘strength theory’’; we maintained this
denomination, being still currently used. The theory of elasticity, chapter of the
mechanics of deformable solids with a theoretical character, succeeds to express
better the physical phenomenon, giving results closer to the reality, in certain
limits; it became a science only at the middle of ninth century, being in continuous
development even today.
The practical importance of a book on the theory of elasticity, which is—at the
same time—an introduction to the mechanics of deformable solids, consists in
putting in evidence points of view and scientific methods of computation in a
domain in which simplified methods or with a non-accurate limit of validity are
still used. The actual technical progress and the necessity to use a minimum of
materials in various constructions ask for a better determination of the state of
strain and stress which takes place in a civil or mechanical construction; the
engineering design may be thus improved.
The first eight chapters deal with the construction of the mathematical model of
a deformable solid, giving special attention to the linear elastic bodies; the for-
mulation of the fundamental problems is followed by their solution in

v
vi Preface

displacements end stresses. The importance of the concentrated loads is put into
evidence, as well in the case of Cosserat-type bodies. Another group of four
chapters contains static and dynamic spatial problems, treated systematically by
the same method of potential functions.
The following two chapters deal with some special problems: particular cases,
treated in the same systematical manner and the case of anisotropic and non-
homogeneous bodies.
The last two chapters contain introductions to thermoelasticity and linear vis-
coelasticity. Special accent is put on the solving methodology as well as on the
mathematical tool used: vectors, tensors and notions of the field theory. Contin-
uous and discontinuous phenomena and various mechanical quantities are pre-
sented in unitary form by means of the theory of distributions. Some appendices
give the book an autonomy with respect to other works, a special mathematical
knowledge being not necessary.
Concerning the first six chapters, I must mention the kind co-operation of
Professor Vasile Ille, Technical University of Cluj-Napoca, who unfortunately has
passed away. I am grateful to Mariana Gheorghitßă for her valuable help in the
presentation of this book. The excellent cooperation of the team of Springer,
Dordrecht, is gratefully acknowledged.
The book covers a wide number of problems (classical or new ones) as one can
see from its contents. It used the known literature, as well as the original results of
the author and his more than 50 years’ experience as Professor of Mechanics and
Elasticity at the University of Bucharest. It is addressed to a large circle of readers:
mathematicians (especially those involved in applied mathematics), physicists
(particularly those interested in mechanics and its connections), engineers of
various specialities (civil, mechanical engineers, etc., who are scientific
researchers or designers), students in various domains etc.

Bucharest, Romania, January 2013 P. P. Teodorescu


Contents

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Aim of Mechanics of Deformable Solids. . . . . . . . . . . . . . . . 1
1.1.1 General Considerations . . . . . . . . . . . . . . . . . . . . . . 1
1.1.2 Models in the Mechanics of Deformable Solids . . . . . 4
1.2 Fundamental Computation Hypotheses.
Short Historical Account . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.1 Fundamental Computation Hypotheses . . . . . . . . . . . 9
1.2.2 Short Historical Account. Development Trends . . . . . 15
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

2 Geometry and Kinematics of Deformation. . . . . . . . . . . . . . . . . . 33


2.1 Finite Deformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.1.1 General Considerations . . . . . . . . . . . . . . . . . . . . . . 33
2.1.2 Material and Space Co-ordinates. Strains. . . . . . . . . . 35
2.2 Infinitesimal Deformations. . . . . . . . . . . . . . . . . . . . . . . . . . 43
2.2.1 Displacement Gradient . . . . . . . . . . . . . . . . . . . . . . 44
2.2.2 Continuity Equations. Computation of Displacements
and Rotations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
2.2.3 Elementary States of Deformation . . . . . . . . . . . . . . 60
2.2.4 Displacements and Strains in Curvilinear
Co-ordinates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
2.2.5 Kinematics of Deformation . . . . . . . . . . . . . . . . . . . 68
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

3 Mechanics of Stresses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
3.1 Stress Vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
3.1.1 General Considerations . . . . . . . . . . . . . . . . . . . . . . 73
3.1.2 Variation of Stresses Around a Point . . . . . . . . . . . . 78
3.2 Stress Tensor. Equations of Equilibrium and Motion . . . . . . . 84
3.2.1 Stress Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

vii
viii Contents

3.2.2 Equations of Equilibrium and Motion . . . . . . . . . . . . 95


3.2.3 Elementary States of Stress . . . . . . . . . . . . . . . . . . . 98
3.2.4 Finite Deformations . . . . . . . . . . . . . . . . . . . . . . . . 101
3.2.5 Stresses in Curvilinear Co-ordinates . . . . . . . . . . . . . 108
References . ............................. . . . . . . . . . . . 112

4 Mathematical Models in Mechanics of Deformable Solids . . . . . . 115


4.1 Elastic Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
4.1.1 Construction of the Elastic Model . . . . . . . . . . . . . . 116
4.1.2 Elastic Potential. Green’s Theory . . . . . . . . . . . . . . . 122
4.1.3 Hooke’s Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
4.2 Inelastic Bodies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
4.2.1 Bodies with Plastic Properties . . . . . . . . . . . . . . . . . 155
4.2.2 Bodies with Viscous Properties . . . . . . . . . . . . . . . . 169
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188

5 General Equations of the Theory of Elasticity. Formulation


of Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
5.1 General Equations of the Theory of Elasticity . . . . . . . . . . . . 191
5.1.1 Statical Problems . . . . . . . . . . . . . . . . . . . . . . . . . . 192
5.1.2 Dynamical Problems . . . . . . . . . . . . . . . . . . . . . . . . 196
5.2 Statical Problem. Potential Functions . . . . . . . . . . . . . . . . . . 199
5.2.1 Formulations in Displacements. . . . . . . . . . . . . . . . . 199
5.2.2 Formulations in Stresses . . . . . . . . . . . . . . . . . . . . . 205
5.3 Dynamical Problem. Potential Functions . . . . . . . . . . . . . . . . 219
5.3.1 Formulations in Displacements. . . . . . . . . . . . . . . . . 219
5.3.2 Formulations in Stresses . . . . . . . . . . . . . . . . . . . . . 227
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240

6 Principles and General Theorems of the Theory of Elasticity.


Computation Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ... 243
6.1 Principles and General Theorems of the Theory
of Elasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
6.1.1 Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
6.1.2 General Principles . . . . . . . . . . . . . . . . . . . . . . . . . 248
6.1.3 Other Considerations. . . . . . . . . . . . . . . . . . . . . . . . 256
6.1.4 Simply Connected Domains.
Multiply Connected Domains. . . . . . . . . . . . . . . . . . 262
6.2 Computation Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266
6.2.1 Particular Integrals . . . . . . . . . . . . . . . . . . . . . . . . . 266
6.2.2 General Methods of Computation . . . . . . . . . . . . . . . 279
6.2.3 Variational Methods . . . . . . . . . . . . . . . . . . . . . . . . 281
6.2.4 Method of Fundamental Solutions . . . . . . . . . . . . . . 288
6.2.5 Other Computation Methods . . . . . . . . . . . . . . . . . . 301
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 304
Contents ix

7 Introduction to the Theory of Cosserat Type Bodies . . . . . . . . . . 307


7.1 General Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307
7.1.1 Introduction to Cosserat Type Bodies . . . . . . . . . . . . 307
7.1.2 State of Deformation. . . . . . . . . . . . . . . . . . . . . . . . 312
7.1.3 State of Stress and Couple-Stress . . . . . . . . . . . . . . . 322
7.1.4 Constitutive Laws . . . . . . . . . . . . . . . . . . . . . . . . . . 329
7.2 Formulations of the Static and Dynamic Problems.
General Theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331
7.2.1 Formulation of the Static Problem . . . . . . . . . . . . . . 332
7.2.2 Formulation of the Dynamic Problem . . . . . . . . . . . . 336
7.2.3 General Theorems . . . . . . . . . . . . . . . . . . . . . . . . . 345
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 347

8 Theory of Concentrated Loads . . . . . . . . . . . . . . . . . . . . . . . . . . 357


8.1 Case of Linearly Elastic Bodies . . . . . . . . . . . . . . . . . . . . . . 357
8.1.1 Construction of Concentrated Loads . . . . . . . . . . . . . 357
8.1.2 Tensor Properties of Concentrated Loads. . . . . . . . . . 372
8.1.3 Solutions for Concentrated Loads . . . . . . . . . . . . . . . 376
8.2 Case of Linearly Elastic Cosserat Type Bodies . . . . . . . . . . . 383
8.2.1 Solutions for Concentrated Loads . . . . . . . . . . . . . . . 384
8.2.2 Centres of Dilatation. Centre of Rotation. . . . . . . . . . 386
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 390

9 Elastic Space. Elastic Half-Space. . . . . . . . . . . . . . . . . . . . . . . . . 393


9.1 Elastic Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393
9.1.1 Volume Loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . 394
9.1.2 Concentrated Loads . . . . . . . . . . . . . . . . . . . . . . . . 397
9.2 Elastic Half-Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 399
9.2.1 Action of a Periodic Load . . . . . . . . . . . . . . . . . . . . 400
9.2.2 Action of a Local Load . . . . . . . . . . . . . . . . . . . . . . 404
9.2.3 Applications. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 408
9.2.4 Methods of the Theory of Distributions. . . . . . . . . . . 412
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 424

10 Elastic Eighth-Space. Elastic Quarter-Space . . . . . . . . . . . . . . . . 427


10.1 Elastic Eighth-Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 427
10.1.1 Action of a Periodic Normal Load . . . . . . . . . . . . . . 428
10.1.2 Action of a Local Normal Load . . . . . . . . . . . . . . . . 430
10.1.3 Action of a Local Tangential Load . . . . . . . . . . . . . . 446
10.1.4 Particular Cases. Application . . . . . . . . . . . . . . . . . . 455
10.2 Elastic Quarter-Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 462
10.2.1 Action of a Local Normal Load . . . . . . . . . . . . . . . . 462
10.2.2 Action of a Local Tangential Load . . . . . . . . . . . . . . 467
10.2.3 Particular Cases. Applications . . . . . . . . . . . . . . . . . 473
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 478
x Contents

11 Elastic Parallelepiped. Elastic Strip. Elastic Layer.


Thick Plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ........... 481
11.1 Elastic Parallelepiped . . . . . . . . . . . . . . . . . . ........... 481
11.1.1 Stress Functions. Boundary Conditions ........... 482
11.1.2 Infinite System of Linear Equations.
State of Strain and Stress. . . . . . . . . . . . . . . . . . . . . 489
11.2 Elastic Strip. Elastic Layer . . . . . . . . . . . . . . . . . . . . . . . . . 493
11.2.1 Elastic Strip . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 493
11.2.2 Elastic Layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 496
11.3 Thick Plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 505
11.3.1 General Considerations . . . . . . . . . . . . . . . . . . . . . . 507
11.3.2 State of Strain and Stress. . . . . . . . . . . . . . . . . . . . . 510
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 513

12 Dynamical Problems of Elastic Bodies. . . . . . . . . . . . . . . ...... 517


12.1 Axisymmetrical Problems . . . . . . . . . . . . . . . . . . . . ...... 517
12.1.1 Formulation in Displacements
of the Limit Problem . . . . . . . . . . . . . . . . . . . . . . . 517
12.1.2 Solutions by Potential Functions. . . . . . . . . . . . . . . . 519
12.2 Progressive Waves. Free and Characteristic Vibrations . . . . . . 522
12.2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 522
12.2.2 Plane Progressive Waves . . . . . . . . . . . . . . . . . . . . . 524
12.2.3 Free and Characteristic Waves . . . . . . . . . . . . . . . . . 527
12.3 Forced and Free Vibrations . . . . . . . . . . . . . . . . . . . . . . . . . 531
12.3.1 Forced Vibrations . . . . . . . . . . . . . . . . . . . . . . . . . . 532
12.3.2 Free Vibrations . . . . . . . . . . . . . . . . . . . . . . . . . . . 540
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 545

13 Particular Cases of States of Strain and Stress . . . . . . . . . . . . . . 547


13.1 Conditions for Stresses . . . . . . . . . . . . . . . . . . . . . . . . . . . . 548
13.1.1 Case of a Zero Normal Stress . . . . . . . . . . . . . . . . . 548
13.1.2 Particular Cases . . . . . . . . . . . . . . . . . . . . . . . . . . . 562
13.1.3 Case of Two Zero Tangential Stresses . . . . . . . . . . . 577
13.1.4 Particular Cases . . . . . . . . . . . . . . . . . . . . . . . . . . . 582
13.1.5 Problems of the Straight Cylinder. Discussion . . . . . . 587
13.2 Conditions for Strains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 592
13.2.1 Case of a Zero Linear Strain . . . . . . . . . . . . . . . . . . 592
13.2.2 Case of Two Zero Angular Strains . . . . . . . . . . . . . . 602
13.3 Plane and Antiplane Problems . . . . . . . . . . . . . . . . . . . . . . . 603
13.3.1 Plane and Antiplane States of Stress . . . . . . . . . . . . . 603
13.3.2 Plane and Antiplane States of Strain . . . . . . . . . . . . . 605
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 612
Contents xi

14 Anisotropic and Non-homogeneous Bodies . . . . . . . . . . . . . . . . . . 615


14.1 Anisotropic Elastic Bodies. . . . . . . . . . . . . . . . . . . . . . . . . . 615
14.1.1 Various Cases of Anisotropy . . . . . . . . . . . . . . . . . . 616
14.1.2 Elements of Crystallography . . . . . . . . . . . . . . . . . . 638
14.2 Non-homogeneous Elastic Bodies . . . . . . . . . . . . . . . . . . . . . 652
14.2.1 Three-Dimensional Problems . . . . . . . . . . . . . . . . . . 652
14.2.2 Two-Dimensional Problems . . . . . . . . . . . . . . . . . . . 656
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 667

15 Introduction to Thermoelasticity . . . . . . . . . . . . . . . . . . . . . . . . . 671


15.1 Basic Relations and Equations . . . . . . . . . . . . . . . . . . . . . . . 671
15.1.1 Heat Conduction. Equations of Thermodynamics . . . . 671
15.1.2 Stationary and Quasi-Static Problems . . . . . . . . . . . . 674
15.1.3 Dynamic Problems . . . . . . . . . . . . . . . . . . . . . . . . . 678
15.1.4 General Considerations . . . . . . . . . . . . . . . . . . . . . . 680
15.2 Applications. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 685
15.2.1 Elementary Examples . . . . . . . . . . . . . . . . . . . . . . . 685
15.2.2 Problems with Axial Symmetry . . . . . . . . . . . . . . . . 688
15.2.3 Plane Problems of Thermoelasticity . . . . . . . . . . . . . 693
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 698

16 Introduction to Linear Viscoelasticity . . . . . . . . . . . . . . . . . . . . . 699


16.1 Linear Viscoelastic Solids . . . . . . . . . . . . . . . . . . . . . . . . . . 699
16.1.1 Constitutive Laws . . . . . . . . . . . . . . . . . . . . . . . . . . 700
16.1.2 The Complex Moduli of Relaxation and Creep . . . . . 709
16.2 Limit Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 714
16.2.1 Formulation of the Problems of the Linear Theory
of Viscoelasticity . . . . . . . . . . . . . . . . . . . . . . . . .. 715
16.2.2 Applications. . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 720
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 727

Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 729

Subject Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 793

Author Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 797


Chapter 1
Introduction

The mathematical model of mechanics of deformable bodies is shortly presented


in this chapter; stress is put on the notion of mechanical system. Using these
introductory notions, one can pass to a mathematical study of these mechanical
systems.

1.1 Aim of Mechanics of Deformable Solids

The scope of mechanics of deformable solids is to determine the state of strain and
stress of a solid body subjected to the action of external loads in static or dynamic
equilibrium.
In what follows we make some general considerations and put in evidence the
basic computational hypotheses, which specify the mathematical model of the
considered solids.

1.1.1 General Considerations

To specify the position of mechanics of deformable solids with respect to various


technical disciplines, especially strength of materials, one must discuss its math-
ematical model. To do this, it is necessary to put in evidence the state of strain and
stress of a solid body.

P. P. Teodorescu, Treatise on Classical Elasticity, Mathematical and Analytical 1


Techniques with Applications to Engineering, DOI: 10.1007/978-94-007-2616-1_1,
Ó Springer Science+Business Media Dordrecht 2013
2 1 Introduction

1.1.1.1 State of Strain and Stress

Under the action of external loads (either concentrated or distributed forces,


concentrated or distributed moments, other concentrated loads (applied on the
external surface of the body or within it), volume forces or moments, inertial
forces, loads due to a thermal field, loads due to an electromagnetic field, loads due
to radioactive radiations, deformations induced by various means, imposed
displacements etc.), the particles (infinitesimal elements) constituting a solid body
change (eventually in time) the positions (with respect to a fixed frame of refer-
ence) they had before the action of these forces. If, by a translation and a rotation,
we can impose on the particles of the body undergoing the action of the external
forces to regain the positions they had before the application of these loads, then
we have to deal with a rigid body motion. Otherwise, the body is subjected to a
deformation. The totality of the deformations (which we shall define later)
undergone by a particle of the body constitutes the state of deformation at a point
(the centre of mass of the particle). The totality of the states of deformation
corresponding to all the points of the solid body constitutes the state of defor-
mation of the body.
At the same time, the notion of displacement may be emphasized. The totality
of the displacements corresponding to a point of the solid body constitutes the state
of displacement at a point. The totality of the states of displacement corresponding
to all the points of the solid body constitutes the state of displacement of the body.
Often, by the state of deformation of the body its state of displacement is equally
intended, the final scope being to determine the components of the displacement
vector. In general, we use the notion of state of deformation (state of strain).
The bodies undergoing only rigid body displacements are called rigid solids;
the other solid bodies are called deformable solids.
Owing to the deformations, the balance (static or dynamic) of the forces linking
the various particles of the body is broken and additional internal forces arise; the
totality of these internal forces (called efforts, if one deals with an arbitrary section
in the body, or stresses, in case of efforts which act upon a unit area), corre-
sponding to a particle, constitutes the state of stress at a point (the centre of mass
of the above considered particle). The totality of the states of stress corresponding
to all the points (particles) of the solid body constitutes the state of stress of the
body.
The state of deformation (or the state of displacement) and the state of stress of
a body form the state of strain and stress of it.

1.1.1.2 Mechanics of Deformable Solids and Strength of Materials

The purpose of mechanics of deformable solids is to determine the state of strain


and stress of a solid body subjected to the action of external loads that are in static
or dynamic equilibrium. One has to deal with a theoretical discipline, the results of
which are of great importance for many technical disciplines.
1.1 Aim of Mechanics of Deformable Solids 3

In general, the strength of materials deals with the same problems as the
mechanics of deformable solids. But this discipline involves a technical solving
method that for each problem or group of problems formulates separate hypoth-
eses, reflecting more or less the physical reality in isolated cases. On the basis of
these working hypotheses, one obtains the equations which govern the phenome-
non and endeavour thus to simplify the computation as much as possible.
By contrast, the mechanics of deformable solids is a chapter of mechanics,
fundamental science of nature, which is based on a mathematical theory, built on
unitary hypotheses; it is applicable whatever the form of the body, the kind of the
material or the manner of action of the external loads may be. Each problem is
specified by its own limiting conditions (boundary conditions and initial condi-
tions (for the movable bodies)), since the fundamental hypotheses always remain
the same. The computational methods applied are general methods leading to an
accurate solution of the problem (more properly, as accurate as necessary and
possible).
In the strength of materials, the construction elements are simplified, by
reducing them to the medium fiber or to the middle surface, so as to diminish the
number of variables involved in the computation. On the other hand, additional
simplifying hypotheses concerning the distribution of the deformations on a cross
section of the body, e.g., the hypothesis of plane cross section [96], as it has been
given by Jacob Bernoulli [4] for the straight bars (equally applied later in the case
of curved bars). If to this hypothesis is added a linear relation between strain and
stress (Hooke’s law) and the mechanical equilibrium is written for a portion of the
bar, then we shall find the known formulae that give the so-called simple stresses
for a straight bar. Assuming that the strains are small with respect to unity and can
be thus disregarded, it follows that the principle of superposition of effects can be
applied, hence the results obtained can be extended to the case of combined
stresses. Thus, computation methods valid for ordinary beams are obtained.
Let us also consider the hypothesis of the straight line element, due to Kirchhoff
[38] for thin plates and to A.-E.-H. Love [43] for thin shells, which leads to the
usual computational methods for these construction elements.
In case in which the state of strain and stress is to be determined for wall beams,
thick (or moderately thick) plates, blocks or in case of stress concentrations etc.,
more accurate computational methods should be applied. In this case, it is no
longer permitted to consider the equilibrium of a finite element (section), but that
of an infinitesimal element cut out of the body. A local study in the neighbourhood
of every point of the body is completed by a global study, referring to the whole
body. The computational method (exact, approximate and even experimental) is
general and is only restricted by difficulties which are liable to arise in practical
applications.
The methods of the mechanics of deformable solids allow one to verify the
limits of applicability of the results obtained by the methods of the strength of
materials. Chiefly, it permits one to solve problems that cannot be tackled by more
elementary methods, such as all three-dimensional problems.
Hereafter, we shall apply the methods of the mechanics of deformable solids.
4 1 Introduction

1.1.2 Models in the Mechanics of Deformable Solids

To make a study of a deformable solid from the mechanical point of view, it is


necessary, firstly, to set up a mathematical model of it.
Usually, by model one means an object or a device artificially created by man,
resembling to a certain extent to another one, the latter one being an object of
scientific research or of practical interest. The scientific notion of model refers to a
certain technique to know the reality representing the phenomenon under study by
an artificially-built system. The most general property of a model is therefore its
capacity to reflect and to reproduce properties and phenomena of the objective
world, as well as their necessary order and their structure.

1.1.2.1 Technical Models

From the very start, the models can be divided into two large classes: technical (or
material) models and ideal (or imagined) models; this division is made according
to the manner of construction the models and according to the means by which the
objects under study may be reproduced.
The technical models are created by man, but they exist objectively, indepen-
dently of his awareness, being materialized in metal, wood, electromagnetic fields
etc. Their purpose is to reproduce for a cognitive goal the object under study, so as
to put in evidence its structure or some of its properties. The model can or cannot
retain the physical nature of the object under study or its geometrical likeness. If
the similitude is maintained, but the model differs in its physical nature, we have to
deal with analogic systems. For instance, electrical models can reproduce pro-
cesses analogue to those taking place in the mechanics of deformable solids, which
are qualitatively different, but are described by similar equations. These models, as
well as other of the same kind, are classed as mathematical models.
One can construct such models, e.g., to study the torsion of a cylindrical straight
bar of arbitrary simply or multiply connected cross section. If the bar is homo-
geneous, isotropic and linearly elastic (subjected to infinitesimal deformations),
then the phenomenon is governed by a Poisson type equation in B. de Saint-
Venant’s [12] theory. L. Prandtl showed that the same partial differential equation
is valid in the case of a membrane supported on a given contour and undergoing a
constant internal pressure; if the contour is similar to the frontier of the plane
domain corresponding to the cross section of the straight bar, then we have a
correspondence of the boundary conditions, hence the classical membrane analogy
(or of the soap-bubble film) is obtained. Other analogies for these problems are
used too, i.e.: electrical modelling, optical interference modelling, hydrodynamical
modelling etc.
It is interesting to remark that the above analogies can moreover be extended to
more complex phenomena. Thus, in case of a multiply connected domain (espe-
cially that of a tubular bar with intermediate thin walls), it is possible to use a
1.1 Aim of Mechanics of Deformable Solids 5

membrane supported along internal contours. As well, in case of a perfect plastic


straight bar, the analogy with small sand heaps of A. Nadai is used. In the case of
elastic plastic bodies, the analogy to be used is obtained be combining the Prandtl
analogy with the Nadai one.
A slightly different mathematical analogy is that between the classical plane
problem of the theory of elasticity and the bending problem of a thin plate. In this
case, the model (a thin plate subjected to a bending in the absence of loads normal
to its middle plane) has the same physical nature as the object under study (thin
plate loaded in its middle plane, therefore undergoing a plane state of stress). By
bending the thin plate, which is thus undergoing certain displacements and slopes
of the middle surface, deformed along the contour, one obtains an image of the
G.-B. Airy stress function of the plane problem, for which the values as well as
those of its normal derivative along the boundary are given. This analogy, due to
K. Wieghardt [99], can be extended to multiply connected domains or to nonho-
mogeneous and anisotropic bodies. Let us also remark that the contour values of
the Airy function and of its normal derivative are obtained as the bending moment
and the axial force in a fictitious bar along the boundary of the considered plane
domain, subjected to the action of given external loads, the contour being travelled
through counterclockwise, from an arbitrarily chosen point; by so doing, an
interesting analogy appears, which is particularly useful in computation. Recently,
this analogy has been partially extended to the space problem. It is also to be
mentioned that, by a judicious choice of the similitude, the model can be made of a
material with more accentuated bending properties, which is very important from a
practical standpoint.
Another type of technical model, used in the mechanics of deformable solids, is
that which corresponds to the intuitive notion of model. Various construction
elements are performed partially or wholly at a reduced scale; this allows to obtain
results concerning the maximal deformations and stresses liable to occur. These
models can be built up of the same material as the object to be studied or of other
materials, which may raise rather difficult similitude problems.
Technical models of a slightly different type are used in photoelastic studies; in
this case, although the geometrical likeness is preserved, the physical nature of the
model is different from that of the object (a thin plate, made of an optically
sensible material, which acquires birefringence properties under the action of
external loads imposed in its middle plane). In this manner, an additional phe-
nomenon appears that enables one to specify the various properties related to the
object which is studied.

1.1.2.2 Character of the Ideal Models

Generally, the ideal models are not materialized and—sometimes—they neither


can be. From the viewpoint of their form, they can be of two types.
The models of first order are built up of intuitive elements having a certain
likeness to the corresponding elements of the real modelled phenomenon; we
6 1 Introduction

observe that this similitude must not be limited only to space relationships but can
be extended also to other aspects of the model and of the object (e.g., the character
of the motion). The intuitiveness of these models is put into evidence on one hand
by the fact that the models themselves, formed by elements sensorial perceptible
(planes, levers, tubes, fluids, vortices etc.) and on the other hand by the fact that
they are intuitive images of the objects themselves. Sometimes, these models are
fixed in the form of schemata.
The models of second order are systems of signs the elements of which are
special signs; the logical relations between them form—at the same time—a
system and are expressed by special signs. In this case, there is no likeness
between the elements of the models and the corresponding elements of the object.
These models have not intuitiveness with respect to geometrical likeness or
physical analogy; they have, by their physical nature, nothing in common with the
nature of the modelled objects. The models of second order reflect the reality on a
gnosiological plane, by virtue of their isomorphism with the reality; a one to one
correspondence is thus assumed between each element and each relation of the
model. These models reproduce the objects under study in a simplified form,
constituting thus (as, in fact, all models do) a certain idealization of the reality.
The types of ideal models mentioned above can be considered as limit cases.
Indeed, there are ideal models combining common features of both model types
which have been described; they contain systems of notions and axioms which
characterize quantitatively and qualitatively the phenomena of nature, for instance
representing mathematical models. Such models are particularly important and
their systematic use has permitted—among other things—the large development of
the mechanics of deformable solids in the last time.
The basic dialectic contradiction of modelling (the model serves to the
knowledge of the object just because it is not identical with the latter one) is
useful, e.g., to put into evidence the properties of continuous deformable solids. In
fact, a model contains the more information concerning the object, when it is more
like this one. Nevertheless, the physical reality is rather intricate; the contradiction
is solved by using a sequence of models which are more and more complete, where
each one brings new contributions to the knowledge of real deformable solids. We
shall endeavour to emphasize the very process of continuously improving methods
of the mechanics of deformable solids, a process constituting—as a matter of
fact—the main tenor of the development of this branch of mechanics.
In general, after a certain ideal model is adopted, it is absolutely necessary to
compare the results obtained by theoretical reasoning with physical reality. If these
results are not satisfactory (sometime this can occur between some limits, which
can be sufficiently narrow), then it is necessary to make corrections or to improve
the chosen model. In fact, this is the manner in which the mechanics of deformable
solids was developed, the word ‘‘model’’ being more and more used by researchers
dealing with this branch of mechanics.
1.1 Aim of Mechanics of Deformable Solids 7

1.1.2.3 Ideal Models in General Mechanics

Both the general (classical) mechanics and the mechanics of deformable contin-
uous media (in particular, the mechanics of deformable solids and the mechanics of
fluids) are studying the objective laws of the simplest form of motion, namely the
mechanical motion.
First of all, for the study of the mechanical motion, a representation of space
and time is necessary; thus, in classical mechanics, the physical space is the three-
dimensional Euclidean space E3 , while the time (considered as universal and
characterizing the duration, the succession and the simultaneity of the material
processes) is still assimilable to the one-dimensional Euclidean space E1 . Thus, the
geometric models of space and time, as used in general mechanics, are reflecting
properties of the real space and time as forms of existence of the matter.
A real movable body is generally thought to be rigid and is often reduced to a
particle. In the same manner, the systems of particles studied can be deformable.
For various values of the time t in E1 , we are able to find the position of the body
or of the system of particles in E3 .
Another intervening element is the cause of the mechanical motion (which, for
the sake of simplicity, will be called ‘‘motion’’). The bodies interact mechanically
with one another and in many cases it is difficult to establish the physical nature of
this action; generically, it was called force. This idea, which originates in the
action of a human organism upon the external world, acquires a precise meaning in
classical mechanics. It is the abstract expression of the measure of the transmission
of motion. Without investigating the nature of the respective force, it was math-
ematically modelled with the help of vectors (sliding vectors in case of rigid solids
and bound vectors in case of deformable systems of particles). We must mention
that a mechanical motion could exist even in the absence of any force, namely the
inertial motion (rectilinear and uniform).
The notion of mass of the particle must be introduced; this is a fundamental
property of matter, exists objectively and is independent of the place where it is
measured. I. Newton [56] conceived the mass as the measure of a quantity of
matter. The notions of gravitational mass and inertial mass should be introduced;
numerically, they are equal to each other, as it has been proved by L. Eötvös. In
this manner, we dispose of various possibilities of measuring the mass.
In the classical Newtonian conception, space, time and mass are considered as
independent from one another.
After these elements were introduced, the Newtonian model was born, by
adopting the principles of Newton [56] (the principle of inertia, the principle of
action of forces, the principle of action and reaction, the principle of the paral-
lelogram of forces and the principle of initial conditions, in a modern view, to can
put in evidence its deterministic aspect). The model was verified by direct practice
in the case of bodies moving at relatively low velocities (negligible with respect to
the velocity of light in vacuum).
8 1 Introduction

1.1.2.4 Introduction of the Ideal Models in Mechanics


or Deformable Solids

General mechanics endeavour in general to solve the problems of motion of solids


on the assumption that they are rigid, although certain established general results
are equally valid for deformable solids (particularly when applied to discrete
deformable systems of particles).
In the study of deformable continuous media, the Newtonian model has to be
completed; instead of a rigid solid, we introduce various models of deformable
media. By examining the historic process of the development of these models,
from a knowledge of the manner in which bodies undergoing certain actions are
deformed and begin to flow, we shall have implicitly the process of the appearance
of the theory of elasticity, of the theory of plasticity, of the perfect and viscous
fluids, of the rheology etc. Hereafter, we shall deal with deformable solids in the
frame of classical mechanics.
The most general problem which has arisen can be formulated as follows: Let
be a given solid of volume V and boundary S in E3 . On S, the action of other
bodies is known; on the other hand, one assumes that the action of other external
loads (e.g., within the volume V) is known equally. Owing to these external loads,
the given solid is deformed, the boundary S becoming S . We wish to find the new
boundary S and to show how it varies in time; we look for the ratio between the
dimensions of the body and the intensity of the external loads as to avoid the
fracture of the body.
To this end, in view of a space-time representation, the same E3 þ E1 geo-
metric model is adopted. As new elements, a study is made of the geometry and
kinematics of deformations and of the mechanics of stresses.

1.2 Fundamental Computation Hypotheses.


Short Historical Account

As any other discipline, the mechanics of deformable solids (and especially the
theory of elasticity) has a few fundamental hypotheses which allow us to simplify
and generalize the phenomena, retaining the essential ones and rending them
accessible to the mathematical computation. It is very important to know these
hypotheses if we want to realize the possibility of practical application of the
results thus obtained.
After reviewing such hypotheses, we shall supply some general data concerning
the development of the mechanics of deformable solids and then insist on the
development of the theory of elasticity.
1.2 Fundamental Computation Hypotheses. Short Historical Account 9

1.2.1 Fundamental Computation Hypotheses

First of all, we present the bodies with which one deals in mechanics of
deformable solids, as well as the fundamental computation hypotheses in the
theory of elasticity. Then we specify the position of the theory of elasticity in the
field of the mechanics of deformable solids.

1.2.1.1 Bodies in the Mechanics of Deformable Solids

Hereafter, we shall give a classification of the solid bodies according to the ratio of
their dimensions.
Bodies with a much greater dimension (length) than the other two ones (cor-
responding to their cross section) shall be called bars. To make the notion clearer,
we shall give a constructive definition.
Let there be a curve C of finite length l, namely the bar axis. In the plane normal
to the curve C, at one of its points P, we shall consider a closed curve C bounding
a plane domain D (the bar cross section); the centre of gravity of this domain is
assumed to be on the curve at the point P. If the point P travels through the curve
C, the curve C, which can also be deformable, will generate a surface bounding a
three-dimensional domain which will be called bar (Fig. 1.1). According to the
form of the axis, the bars can be straight or curved. Moreover, the curved bars can
be classified as skew curved bars and plane curved bars.
Both the (average) dimensions a and b of the cross section shall be considered
as being of the same order, provided the condition a; b  l is fulfilled. If all three
dimensions are of different order (a  b  l), we shall have to consider thin wall
bars (Fig. 1.2a). Lastly, if the cross section dimensions are negligible with regard
to the bar length, so that the bar should be perfectly flexible (unable to retain a
bending), then we have a string (Fig. 1.2b).
Bodies having one dimension (the thickness) much smaller than the other two-
dimension (corresponding to the middle surface) shall be called plates.
Let be a surface S of finite area, the middle surface of the plate. On the normal
to the surface S at a point P of the latter one, we shall consider a segment of a line
of length h (the thickness), the middle of which is assumed to be on the surface S,
at the point P. If the point P travels through the surface S, the extremities of the
segment of a line, which can be of variable length, will generate two surfaces

Fig. 1.1 Bar

l
C P
D
b

a
10 1 Introduction

Fig. 1.2 Thin wall bar (a). (a) (b)


String (b)

a
l

b
Fig. 1.3 Plate

hh
22
P

a
b

bounding a three-dimensional domain, which we shall call plate (Fig. 1.3).


According to the form of the middle surface, the plates may be plane or curved;
the plane plates are called simply plates.
According to the ratio between the (average) dimensions a and b of the plate,
contained in the middle surface, and its thickness h, we shall consider thin plates
(for h  a; b) (Fig. 1.4a), moderately thick plates (Fig. 1.4b) and thick plates (for
h\a; b) (Fig. 1.4c); the curved thin plates are called shells too. Let us remark that
accurate delimitations between these plate categories cannot be made; they depend
on the computation possibilities of the states of strain and stress and can vary from
one case to another. Finally, if the thickness h is quite negligible with regard to the
other two dimensions, so that the plate should be perfectly flexible, we shall obtain
a membrane (Fig. 1.4d).
Bodies with all three dimensions a; b; c of the same order of magnitude shall be
called blocks (Fig. 1.5).
Generally, the real bodies occupy finite three-dimensional domains. We shall,
however, consider also infinite domains, the study of which is of particular
interest; as a matter of fact, these domains idealize real cases, often occurring in
practice, or else they can be used as intermediate stages in solving problems
corresponding to other domains.

1.2.1.2 Fundamental Computation Hypotheses in the Theory


of Elasticity

The fundamental computation hypotheses that can be made and that—some-


times—particularize the body to be studied are the following:
1.2 Fundamental Computation Hypotheses. Short Historical Account 11

Fig. 1.4 Plate: thin (a), (a) (b)


moderately thick (b), thick
(c), Membrane (d)

a
h
h
b b

(c) (d)

a
h

Fig. 1.5 Block

c
a
b

(i) The solid body we are studying (considered to be at rest with respect to an
inertial (fixed) frame of reference) is subjected to the action of balanced
external loads. If the body is in motion, then we introduce moreover the inertial
forces; therefore, the external loads are in dynamic equilibrium. On the other
hand, each part of the body and each particle detached from it will be subjected
to loads statically or dynamically balanced (load system equivalent to zero),
This hypothesis allows to write the partial differential equations verified by the
stresses within the body and to express the boundary and initial conditions.
(ii) The solid body is considered as a continuous medium (without holes, internal
microscopic cracks etc.). This encourages us to assume that the body defor-
mation will be equally continuous and therefore the strains and the stresses
will be mathematically expressed by continuous functions.
For some points of the body (singular points), where the strains and the stresses
tend to infinity, an additional study must be made; it is important to use the
methods of the theory of distributions in this case. Likewise, the case of the bodies
12 1 Introduction

with internal holes (continuous bodies to which correspond multiply connected


domains) and the case of the bodies with internal cuts (particular cases of multiple
connection) must be considered separately. In the case of the disperse media (that
do not verify the continuity hypotheses) it is necessary to have recourse to aleatory
variables.
(iii) The solid body is isotropic, i.e. it has the same mechanical (and physical)
properties in all directions in the neighbourhood of each of its points. This
property is expressed by the relations between strains and stresses. When this
hypothesis is not observed, then we have to deal with anisotropic (aeolo-
tropic) bodies.
(iv) The solid body is homogeneous, i.e. it has the same mechanical (and physical)
properties at each of its points. On the basis of this property, the mechanical
coefficients of the material, occurring in the constitutive law, are constant
with respect to the space variables. When these coefficients are variable, the
body is non-homogeneous.
The properties of isotropy and homogeneity do not condition one another. A
solid body can be homogeneous and isotropic (as we shall generally admit it) or
homogeneous and anisotropic (i.e. having the same properties at each point for a
given direction) or isotropic and nonhomogeneous (i.e. having the same properties
for any direction, but different from a point to another one) or even non-homo-
geneous and anisotropic.
We mention that, by changing the system of co-ordinates in the case of an
anisotropic body, we can obtain, formally, a body with non-homogeneous prop-
erties. We shall admit, in this case, that an anisotropic body is non-homogeneous if
it remains non-homogeneous in any system of curvilinear coordinates (in other
cases, it can be homogeneous, with a curvilinear anisotropy).
We also mention that the composite bodies can be modelled as anisotropic ones.
(v) The body under study is perfectly elastic. Under the action of external loads,
the body is deformed, but when the action of these loads ceases to operate, it
resumes its initial position (in the case of static loads) and its initial form
(without any hysteresis phenomenon taking place); the deformation phe-
nomenon is reversible. Therefore, the deformed form of a solid body will only
be influenced by the external loads that act on it at that moment. A one-to-one
relationship exists between strains and stresses; as already mentioned, the state
of stress at a point of the solid body will only depend on the state of strain in
the neighbourhood of this point. In this manner, we happen to be within the
framework of the theory of elasticity.
In the case of many bodies currently used, if the stresses go beyond the elastic
limit, after unloading residual deformations remain. Beyond the elastic limit, the
body presents elastic plastic properties. When this is the case, whether the elastic
deformation can or cannot be neglected, the computation of the state of strain and
stress must be made with the help of the theory of plasticity.
1.2 Fundamental Computation Hypotheses. Short Historical Account 13

Generally, the one-to-one relation between strains and stresses is expressed


mathematically by a linear relation (Hooke’s law). This hypothesis corresponds—
often—well enough to the physical phenomenon and, on the other hand, leads to
considerable simplifications of the mathematical computation. Sometimes, a more
intricate mathematical relation is considered; this leads to a theory of elasticity,
non-linear from a physical standpoint.
Moreover, special relations can be considered, e.g., as in the case of the hypo-
elasticity theory, created by C. Truesdell, or in the case of hyperelastic bodies.
If, besides the stresses, we also take into consideration the couple-stresses
(micromoments), we shall find a theory of asymmetrical elasticity for bodies of
Cosserat type (after the name of the brothers E. and F. Cosserat [13] who were the
first to study, in 1907, such a problem), i.e. bodies with microstructure. In such a
case, a volume has six degrees of freedom and the volume moments can also act as
external loads; the constitutive law—even in the linear case—necessitates the
introduction of several mechanical constants of the material.
(vi) The deformations and displacements of the solid bodies, submitted to the
action of external loads, are very small with regard to the general dimensions
of the body; it follows that the strains are negligible with regard to unity. If
the rigid body local rotations are equally negligible with regard to unity, then
we have to discuss a linear theory from a geometrical standpoint (the terms of
second order are negligible with respect to the terms of first order), as well as
from a mechanical standpoint (the equations of static and dynamic equilib-
rium can be written for the undeformed form of the body). Therefore, con-
cerning the strains, we can apply the principle of superposition of effects.
There are, however, cases where it is necessary to apply non-linear methods
from a mechanical standpoint or a non-linear theory from both geometrical and
mechanical standpoint (the case of finite deformations). These considerations are
absolutely necessary in the case of stability problems (a solid body loses its
stability when, for a certain intensity of the external loads (critical values), at least
two distinct states of strain and stress result). Phenomena of bifurcation and chaos
can also occur.
Generally, the deformation of the bodies is concomitant with a change of volume
(the solid bodies are compressible). Particularly, the case of incompressible solids
and of bodies subjected to isochoric deformations are considered equally; this can
also constitute a first computing approximation.
Let us mention also that we shall not take into account the rate of deformation
of the body; thus, we shall not consider bodies with rheological properties
(for which one takes into account the influence of viscosity), therefore bodies of
Maxwell or Voigt type, mixed or of a more complex nature, for which the con-
stitutive law (generally, integro-differential) depends moreover on a time variable.
Consequently, we shall not consider creep and relaxation phenomena.
14 1 Introduction

(vii) When the deformations are propagated, within the solid body, at a very high
velocity (a case recently appeared in a unified theory of elementary parti-
cles), the phenomenon should be studied from the standpoint of the theory of
relativity.
(viii) The influence of the temperature variation is not taken into consideration
neither in the constitutive laws nor in the computation of the deformations.
But it will be introduced in the considerations concerning thermodynamical
principles.
(ix) The body under study does not have initial stresses, which could be due to
initial deformations of the material. They could result from machinery
operations (rolling, drawing etc.), from assembling operations or could be
due to phenomena occurring during the working, before the action of the
external loads, contraction phenomena of concrete etc.
Although these phenomena do exist and cannot be eliminated, to simplify the
computation, one assumes that the initial stresses are missing. Thus, to an
unloaded body (case of vanishing external loads) corresponds a null state of stress.
If the initial stresses cannot be neglected, then one makes a supplementary study to
take into account their influence.
We will assume—in general—that all these hypotheses are taken into account;
if, in particular cases, one of these hypotheses is not respected and if we are placed
in a more general situation (a less restrictive hypothesis), then we make a special
mention about it.

1.2.1.3 Position of the Theory of Elasticity in the Frame of Mechanics


of Deformable Solids

As we have seen, in the mechanics of deformable solids, the geometrical and


mechanical aspect of the considered problems must be completed with an aspect of
physical, experimental nature, which specifies the nature of the solid body. One
introduces a certain constitutive law of the deformable solid, a certain relation
between strains and stresses.
In general, one can establish (especially on an experimental way) relations
between strains and stresses in a one-dimensional case. The experiments in two- or
three-dimensional cases are much more difficult (e.g., in mechanics of earth, the
experiments with the three-axial apparatuses are of great interest); the most time—on
the basis of various hypotheses—one extends the results obtained in the linear case.
As well, the fundamental law of the deformable solids can be established on the
basis of various considerations and hypotheses of energetical and thermodynamical
nature. In particular, the relations which are obtained (and in which intervene—the
most time—the invariants of the strain and stress tensors) depend on certain
coefficients (constant or variable), which are determined on an experimental way.
We have seen that an elastic body is that one for which the deformation
phenomenon is reversible. As a chapter of the mechanics of deformable solids,
1.2 Fundamental Computation Hypotheses. Short Historical Account 15

the theory of elasticity has as object of study the determination of the state of strain
and stress of an elastic solid body subjected to the action of certain external loads
in static or dynamic equilibrium. To the fundamental equations of the mechanics
of deformable solids one must add a constitutive law, which characterizes the
respective elastic body.
We mention that, the most times, a solid body may be considered as being
elastic only till a certain intensity of the stresses or till a certain magnitude of the
strains (till the limit of elasticity); after this, it becomes elastic plastic properties.
Thus, our study will be valid for the so-called elastic zone of the respective body.

1.2.2 Short Historical Account. Development Trends

Certainly, since most ancient times, since men began to build, some problems of
mechanics arose—in general—and of strength of materials—in particular; various
problems about deformable solid bodies (although this property was not quite clear
in their minds) were imposed by daily practice.
We shall supply some general data concerning the development of the
mechanics of deformable solids and then we shall insist on the development of the
theory of elasticity. We shall mention also the researches made in Romania.
As well, we will put in evidence some development trends and the new
problems which are put.

1.2.2.1 Mechanics of Deformable Solids

Strength of materials problems occured and were solved—in a certain form—by


famous builders, such as Archytas of Tarentum (approx. 440–360 A.C.) or
architects such as Vitruvius (the second half of the first century A.C.). The
Egyptians had, probably, empirical rules that they took into account. The Greeks,
more advanced—it is sufficient to mention Archimedes (287–212 A.C.)—devel-
oped statics, that is the very basis of the study of real bodies. The Romans were
equally great builders, many of their structures being even now in use; we mention
thus the famous ‘‘Pont du Gard’’ in the south of France.
Much of the empirical knowledge acquired A.C. or at the beginning A.D. was
lost in the course of the feudal epoch, being rediscovered only during the
Renaissance. Leonardo da Vinci (1452–1519) was the first to make fracture tests
on a steel string. So, at the end of a suspended string, a vessel is hung into which a
quantity of sand is slowly flowing. The string extends until it breaks. Leonardo da
Vinci also mentioned that it is good practice to make a great many experiments
and to take the average of these results.
Later on, Galileo Galilei (1564–1642) showed moreover how tensile strength
experiments can be made on a bar; he was the first to admit a certain stress
distribution in a cantilever bar, the free end of which is subjected to a concentrated
16 1 Introduction

force. According to Galileo, in the built-in cross section should appear normal
stresses uniformly distributed on the whole cross section, which later on proved to
be only a rather vague approximation of the physical reality. Architect Fontana
(1543–1607) applied these results to hoist an obelisk at the Vatican.
Robert Hooke (1635–1703) [32], on the basis of various experiments, stated in
1678 his famous law ‘‘ut tensio sic vis’’ (such is the force, as the extension is).
Edmé Mariotte (1620–1684) verified this law on wood tensile test samples.
Valuable ideas about the notion of elasticity were set forth by Mikhail Vasilievich
Lomonosov (1711–1765).
In the seventeenth and eighteenth centuries, progress was realized, especially in
the field of the mechanics of structures and of the strength of materials. Many of
the acquired results have applications corresponding to the new problems met
within civil engineering and machine construction. We mention moreover some
theoretical results, obtained by Jacob (1654–1705) [4] and Jean (1667–1748)
Bernoulli brothers (to the former is due the hypothesis of the plane cross sections
in straight bars). The latter’s son Daniel Bernoulli (1700–1782), Daniel’s nephew
Jacques Bernoulli (1759–1789) and chiefly Leonhard Euler (1707–1783) and
Joseph-Louis Lagrange (1736–1813) dealt with problems of deformation of thin
elastic bars.
Other theoretical and experimental studies are due to Charles-Augustin de
Coulomb (1736–1806), Jean-Victor Poncelet (1788–1867) and Franz Grashof
(1826–1893). Jean le Rond d’Alembert (1717–1783) and Peter Gustav Lejeune-
Dirichlet (1803–1859) contributed by their studies in the field of mechanics to the
birth of the theory of elasticity.
The first studies about sliding lines were made in the soil mechanics by Ch.-A.
de Coulomb, William-John Macquorn Rankine (1820–1872) and Maurice Lévy
(1839–1910). Their importance was grasped by Henri-Édouard Tresca
(1814–1885) who, in 1864 and 1872, published several notes, concerning a con-
dition of plasticity in which the maximal shear stress is constant at each point of
the elastic-plastic zone, and by Christian Otto Mohr (1835–1918) who, on this
basis, formulated in 1882 a theory of the strength of materials. Adhémar-Jean-
Claude Barré de Saint-Venant (1797–1886) reviewed H.-É. Tresca’s works at the
Academy of Science (Paris) and, on this occasion, developed the fundamental
equations of the plasticity theory, by admitting that the cubical dilatation vanishes
during the plastic deformation (which is thus incompressible), that the principal
directions of the tensor of the rate of strain coincide with the principal directions of
the stress tensor and that the maximal shear stress is constant at each point of the
solid body; and we can assert that it is with this development that the theory of
plasticity was born in 1870.
And, still in the second half of the ninteenth century, the problem of the study of
the solid bodies for which the deformations rate has been taken into account began
to be formulated. Thus, James Clerk Maxwell (1831–1879), William Thomson
(Lord Kelvin) (1842–1907) and Woldemar Voigt (1850–1919) define particular
viscoelastic bodies (bodies of Maxwell or of Kelvin (or Voigt) type; the last two
arrived at analogous results by independent studies), corresponding to certain real
1.2 Fundamental Computation Hypotheses. Short Historical Account 17

bodies. Nevertheless, from a formal standpoint, the rheology was born only at an
international congress in 1929 (it is on this occasion that the name ‘‘rheology’’
(from ‘‘panta rei’’) was introduced); especially, the theory of viscoelasticity has
been developed.
Thus, the bases are set of a new branch of mechanics that deals with the study
of deformable solids.

1.2.2.2 Theory of Elasticity

The bases of the mathematical theory of elasticity were set forth, at the beginning
of nineteenth century, by the works of Louis-Marie-Henri Navier (1785–1836)
[54] and Mikhail Vasilievich Ostrogradskı (1801–1861); but it was Augustin-
Louis Cauchy (1789–1857) who, making use of these results, established the
fundamental equations of elastic bodies (the equations of static and dynamic
equilibrium and the relations between displacements and strains, valid for any
deformable solid bodies) in the form still used today. Siméon-Denis Poisson
(1781–1841) contributed to elucidate these problems, by introducing also the
notion of transverse contraction, while Benoît-Paul-Émile Clapeyron (1799–1869)
establishes a theorem of equivalence between the internal work (of deformation)
and the external work. The formulation in displacements of the fundamental
problems of the theory of elasticity as well as an important treatise on the theory of
elasticity, where is equally formulated the famous problem of the elastic paral-
lelepipedon, which only in the last years obtained solution, are chiefly due to
Gabriel Lamé (1795–1870) [39]; on the other hand, starting from the studies of
Karl Friedrich Gauss (1777–1855), he introduced [40] the systematic use of cur-
vilinear co-ordinates in solving problems of the theory of elasticity.
During the nineteenth century and the beginning of twentieth century, many
researches in the domain of the theory of elasticity have been performed, results
which are very difficult to pass, even summarily, in review.
At the same time, the strength of materials acquires equally a large develop-
ment concerning the study of plates related to the design of railway bridges. As to
the problem of plates, we may recall, besides the name of J.-L. Lagrange, those of
Sophie Germain (1776–1831) and Gustav Robert Kirchhoff (1824–1887) [36],
who continued—theoretically—the experimental research begun by Ernst
Friederich Chladni (1756–1827).
A controvercy (which lasted several decades) about the elastic constants of the
material confronted the supporters of the conception according to which a single
elastic constant is sufficient to characterize the mechanical properties of the material
(L.-M.-H. Navier [54], A.-L. Cauchy and their pupils) against those who held that
two elastic constants are necessary to achieve this end. George Green (1793–1841),
starting from considerations of an energetic nature, showed that, in the general case
of anisotropy, 21 constants are necessary, which in case of isotropic bodies are
reduced to two distinct constants. Adolf Yakovlevich Kupffer (1799–1865), at the
Central Laboratory of Weights and Measures in Sankt Petersburg, Wilhelm
18 1 Introduction

Wertheim (1815–1861), in Paris, and—later—Franz Ernst Neumann (1798–1895)


[55] and his pupils (G. R. Kirchhoff and W. Voigt) made experimental research by
which they verified the theoretical considerations of G. Green. The two elastic
constants, usually used, are the modulus of longitudinal elasticity, named after
Thomas Young (1773–1829) but introduced by L. Euler, and the coefficient of
transverse contraction of S.-D. Poisson. Among the elastic constants characterizing
the anisotropic bodies, we mention the coefficients of N. G. Chentsov.
In England, George-Bidell Airy (1801–1892) established in 1862–1863 the
bases of the plane theory of elasticity, by introducing the stress function (the first
important stress potential function in the theory of elasticity), that bears his name
today. George Gabriel Stokes (1819–1900) found many important results, chiefly
in the field of vibrations of elastic bodies and introduced, moreover, in 1849, some
displacement potential functions.
In France, we must chiefly mention the studies of B. de Saint-Venant, whose
ideas—as it is interesting to mention—did not appear in a volume, but were
published as articles, papers, notes or appendices to the successive editions of
L.-M.-H. Navier’s [54] course and to the French translation of the treatise of
Rudolf Friederich Alfred Clebsch (1833–1872) [12]; even the equations of con-
tinuity, although bearing his name, cannot be found except in these works. Among
the many and various studies of B. de Saint-Venant, we mention especially those
referring to the problem of torsion of a straight bar of any cross section however,
which are at the basis of modern research in this field; here the semi-inverse
computation method was applied first.
The first formulation of the thermoelasticity problem, in the case of uncoupled
equations, was stated in the field of displacements by Jean-Marie-Constant Duha-
mel (1797–1872) and Fr. E. Neumann. A great many results of Neumann [55] are to
be found in his treatise on the problems of the theory of elasticity and optics; let us
mention his studies in the field of double refraction by which he continues and
specifies the results acquired by David Brewster (1781–1868) and Ludwig
Friederich Seebeck (1805–1849). These studies constitute the physical basis of
photoelasticity. The coupled equations of thermoelasticity and the formulation of
its problems in its most general form are due to other researchers who subsequently
completed the equation of heat propagation of François-Marie-Charles Fourier
(1772–1837) with the terms corresponding to the elastic deformation of the body.
G. R. Kirchhoff demonstrated, under certain conditions, the theorem of
uniqueness of the solution of the elasticity theory problem. In what concerns the
existence of the solution of the elasticity equations, the first studies were made by
Lord Kelvin; the study of the problem was resumed by Arthur Korn, using the
method of successive approximations, and by Erik Ivar Fredholm (1866–1927),
using the method of integral equations.
R. F. A. Clebsch brought important contributions to the plane problem of the
theory of elasticity (conditions on the boundary, representation of the displace-
ments etc.). On the other hand, he deals with the vibrations of the elastic sphere;
particularly, in the static case, he obtained results which later were thoroughly
discussed by Lord Kelvin. The latter brought many contributions in the theory of
1.2 Fundamental Computation Hypotheses. Short Historical Account 19

elasticity, such as the use of conjugate functions, introduced by R. F. A. Clebsch to


solve the torsion problem of B. de Saint-Venant. We think however, that his most
important contribution consists in the linking elasticity problems and thermody-
namic problems; on this occasion, Lord Kelvin showed that the volume density of
strain energy does not depend in any way on the manner by which the deformation
was obtained, but only on the value of this deformation. Together with Peter-
Guthrie Tait (1831–1901), he published [74] the first volume (containing many
problems of mechanics and theory of elasticity) of a vast treatise of theoretical
physics, which, unfortunately, could not be achieved.
To J. C. Maxwell, besides the important equations of the electromagnetic field,
are also due the first complete photoelastic studies about some plane domains
(among which are certain domains that possessed theoretical solutions); on the
other hand, starting from Airy’s representation of the plane problem, he estab-
lished similar results permitting the solution, in stresses, of the space problem.
A similar representation is due to Giacinto Morera (1856–1909); we also mention
that both these representations do not verify, unfortunately, the continuity equa-
tions in stresses, established by Eugenio Beltrami (1835–1900) and achieved by
John Henri Michell (1863–1940) in the case of any volume force whatsoever. Only
later, representations were proposed which take into account these equations too.
J. C. Maxwell established a reciprocity relation of deformations in the particular
case of the action of any two concentrated loads of equal intensity; this result was
extended by Enrico Betti (1823–1892) to systems of any static loads (the theorem
of work reciprocity) and by John William Strutt (Lord Rayleigh) (1842–1919) to
the dynamic case. Still another generalization of these results is due to Carlo
Somigliana.
The notion of strain energy [92, 93], first used by G. Green, began to play a
more and more important part. Thus, Alberto Castigliano (1847–1884) [10],
continuing these studies and achieving the undemonstrated results stated by
Ludovico Frederico Menabrea (1809–1896), established the results concerning the
strain of stress determination at a point of an elastic solid. His results are valid if a
constitutive linear law (Hooke’s law) does exist; if this law is non-linear, then the
complementary energy, a notion due to Friedrich Engesser (1848–1931), should be
introduced. His results complete the ones due to G. Green.
An important contribution was achieved by the German school concerning the
strength of materials. Emil Winkler (1835–1888) [83] contributed to the study of a
straight bar on an elastic medium. To C. O. Mohr [48] is due the graphical repre-
sentation of the strain and of the stress states around a point with the help o circle
(two-dimensional case) or of three circles (three-dimensional case); on the other
hand, this representation is equally valid for any symmetrical tensor of second
order. August Föppl (1854–1924) wrote a six volume treatise on technical
mechanics, the third one [16] of which is devoted to the strength of materials with a
supplement giving the principal results of the theory of elasticity, while the fifth
volume [17], which appeared later, deals only with the theory of elasticity. As his
two-volumes book [18], concerning the strains and stresses (‘‘Drag und Zwang’’),
a book written subsequently together with his son Ludwig Föppl (1887–1976)
20 1 Introduction

(a third volume was published later only by L. Föppl), these works played an
important part, revealing to very large, especially engineering, circles the realiza-
tions of the theory of elasticity; in their time, the first works were translated into
French and Russian.
The pupils of A.-J.-C. B. de Saint-Venant successfully continued the studies he
had begun. The most important among pupils was Joseph Valentin Boussinesq
(1842–1929); to him is equally due the application of the theory of the potential to
both the static and the dynamic problems of elasticity. As an example,
J. V. Boussinesq [7] establishes the state of stress within the elastic half-space,
undergoing the action of a concentrated normal force or of a distributed load on the
separation plane. Alfred Aimé Flamant (1839–1914) studied the corresponding
two-dimensional problem (the elastic half-plane case). J. H. Michell applied these
results to the study of a whole series of plane problems of the elasticity theory in
polar co-ordinates, chiefly to the influence of the concentrated internal forces.
Another pupil of B. de Saint-Venant, Maurice Lévy (1838–1910), dealt with the
plane problem of the theory of elasticity, especially in the case of an elastic wedge.
Both to him and to J. H. Michell, an important theorem is due, according to which
the state of stress, in the plane case, in a simply connected domain, being given the
loads on the boundary and in the absence of the volume forces, does not depend on
the elastic constants of the material.
Lord Rayleigh brought important contributions to the theory of vibrations
(elastodynamics), his most important results being contained in a treatise on the
theory of sound [70]. The computation method pointed out by him and resumed
later by Walter Ritz (1878–1909) is known today as the Rayleigh-Ritz variational
method. Another variational method, often used in the computation practice, was
later elaborated by Erich Immanuel Trefftz (1888–1937). Starting from the results
supplied by Lord Rayleigh, Horace Lamb (1849–1934) and Augustus-Edward-
Hough Love (1863–1940) entered upon various studies concerning the vibrations
of plates. To the latter is also due an important treatise on the theory of elasticity,
which was published in several ulterior editions and played an important rôle in
the scientific information of several generations and is still successfully used [43].
In the same period, one of the most developed histories of the theory of elas-
ticity until the present time was published by Isaac Todhunter (1820–1884) and
Karl Pearson (1857–1936) [78].
W. Voigt was the first to introduce the notion of tensor in the theory of elas-
ticity. Hermann Ludwig Ferdinand von Helmholtz (1821–1894) was known by his
studies in the domain of kinematics of motion. Heinrich Rudolf Hertz (1857–1894)
researched in the field of hardness of materials and of local stresses (contact
problems).
In the plane problems of the theory of elasticity, a distinction is to be made
between the state of plane stress (short cylinder problem) and the state of plane
strain (long cylinder problem, theoretically of infinite length). To the studies of
G.-B. Airy, J. C. Maxwell, and M. Lévy and A. A. Flamant mentioned earlier,
other studies followed, which raised a particular interest both from theoretical and
practical standpoint. Let us discuss the research of Aksel Wihelmovich Gadolin
1.2 Fundamental Computation Hypotheses. Short Historical Account 21

(1828–1904) and Harlampii Sergeevich Golovin (1844–1904), who dealt with a


state of plane strain by studying the thick cylinder problem. G. G. Stokes applied
A. A. Flamant’s results to the approximate study of the straight beam. Among the
plane problems studied by J. H. Michell, we mention chiefly the elastic wedge and
the circular disc. The first detailed study of the plane problem, by means of
biharmonic polynomials (what we now call elementary computation methods),
was made by Augustin-Charles-Marie Mesnager (1862–1933).
In Russia, studies in this direction were at first related to the nature of materials
and to problems imposed by the technique (research on the strength of materials).
We must however mention the results found by Nikolai Mikhailovich Belyaev
(1890–1944) in the theory of elasticity. Ivan Grigorievich Bubnov (1872–1919)
was renowned for his studies in the field of plates; he offered also an idea con-
cerning a variational computing method. This idea was later resurrected by Boris
Grigorievich Galerkin (1871–1945), who used it successfully in various problems
of the theory of elasticity (it is known today as the Bubnov-Galerkin method). To
Boris Borisovich Golitsyn (1862–1916) are due the equations of damped small
motion in an elastic medium, as a result of his studies on seismic problems.
Aleksandr Nikolaevich Dinnik (1876–1950) made various researches into a few
direct computation methods (variational methods), chiefly related to problems of
elastic stability; he was also noticeable in the study of torsion problems after B. de
Saint-Venant. Important approximate computation methods, applied to various
problems of the theory of elasticity, were elaborated by Aleksei Nikolaevich
Krylov (1863–1945), to whom are also due interesting studies related to the theory
of ship building.
Work concerning the approach between the mathematical theory of elasticity
and the mechanics of structures was due to Viktor Lvovich Kirpichev
(1845–1913). Some remarks made by the famous Russian inventor Ivan Petrovich
Kulibin (1735–1818) were used and developed by Dmitrii Ivanovich Juravskı
(1821–1891), who completed by his formulae for the shear stresses the results
supplied by L.-M.-H. Navier for the straight beam problem. The studies of
Juravskı were later applied by Jacques-Antoine-Charles Bresse (1822–1888) to a
general expression of the stresses occurring in a straight beam.
An important contribution to solving in displacements the problems of the
theory of elasticity was made by B. G. Galerkin, who expressed the components of
the displacement vector with the help of three biharmonic functions; B.
G. Galerkin applied these results to many elasticity problems, such as the problem
of thick plates; his formulae were generalized later in various ways. Later on, Piotr
Fedorovich Papkovich (1887–1946) [60] expressed the solution in displacements
of the general problem of the theory of elasticity with the help of four (contin-
gently three) harmonic functions, solution of the equation of Pierre Simon de
Laplace (1749–1813); this representation was previously presented in a lecture by
G. D. Grodski. It was found again, independently, by Heinz Neuber (1906–1989)
and is now at the basis of a great many elasticity studies, thus contributing to solve
many actual boundary problems. Also to P. F. Papkovich are due many studies in
the field of ship building, e.g., an idea related to the elastic rectangle problem.
22 1 Introduction

His treatise [60] on the theory of elasticity was a particularly valuable book for
many generations of Russian elasticians.
In the climate impressed at Göttingen by Felix Klein (1849–1925) and David
Hilbert (1862–1943), many researchers asserted themselves into the field of
mechanics of deformable solids. We mention Ludwig Prandtl (1875–1953) who, in
the field of the theory of elasticity, was especially renowned for his membrane
analogy in the torsion problem of a straight bar. Another remarkable representative
of this school was Theodor von Kárman (1881–1963).
In the theory of elasticity, problems of multiply connected bodies were studied
and established on a mathematical basis—introducing the distorsions—by Vito
Volterra (1860–1940), whose results dating from the beginning of twentieth
century were gathered up in a volume, published subsequently by his son Enrico
Volterra [79].
In the domain of thin wall bars and of thin shells, interesting computational
methods were proposed by Vasilii Zakharovich Vlasov (1906–1958).
The application of the methods of integral equations to the problems of the
theory of elasticity is due to Giuseppe Lauricella (1869–1913). The theories of
strength of materials marked a new upsurge by the work of E. Beltrami, Maksy-
milian Tytus Huber (1872–1950) [34], Heinrich Hencky (1885–1951) and Richard
von Mises (1883–1953). The spatial representation of the principal stresses, due to
Harald Malcolm Westergaard (1888–1950), is much used.
The plane problem of the theory of elasticity acquires a remarkable develop-
ment after the studies of Charles Henri Ribière and Louis-Napoleon George Filon,
who use harmonic functions in the form of Fourier series or Fourier integrals,
obtained by means of certain terms which are product of trigonometric lines and
hyperbolic or exponential lines. Other similar solutions were given at the begin-
ning of twentieth century in the doctoral thesis of Aloys Timpe. These ideas have
acquired a modern form among the operational computation methods. Jacques-
Salomon Hadamard (1865–1963) gave a general method of solution of the two-
dimensional biharmonic problem (the mathematical aspect of the plane problem of
the elasticity theory). G. V. Kolosov proposed another formulation of this problem
with the help of the complex variable functions; this method was subsequently
developed at length by Nikolai Ivanovich Muskhelishvili (1891–1976) and his
pupils, and constitutes one of the finest realizations of the Russian school of
elasticity theory. The above-mentioned results were expounded at length by
N. I. Muskhelishvili [51] in his treatise translated into several languages.
Among the treatises on the theory of elasticity which played an important part
in its development, we mention those of Henri Poincaré (1854–1912) [63] and of
Roberto Marcolongo (1862–1943) [46, 47]; Eugène-Maurice-Pierre Cosserat
(1864–1931) and his brother François Cosserat [13] published a book an the theory
of asymmetric elasticity (a particle of the body has six degrees of freedom), at the
beginning of twentieth century. From 1914, when the first edition appeared in
Russian, up to now, the treatises of Stepan Prokofievich Timoshenko (1878–1972)
had a great influence on the development of many generations; they were trans-
lated into several languages. We particularly mention his book on the theory of
1.2 Fundamental Computation Hypotheses. Short Historical Account 23

elasticity, the last edition of which was published together with J. N. Goodier [77].
His history on the strength of materials, containing also a history of the theory of
elasticity, is equally interesting [75]. In Russia, we point out the treatises of
Mikhail Mitrofanovich Filonenko-Borodich (1885–1962) and of Leonid Samu-
ilovich Leıbenzon (1879–1951) [42]. In Poland, the treatise of M. T. Huber [34] on
the theory of elasticity was wide spread. Among the treatises with a general
character, including also volumes devoted to the theory of elasticity, we mention
the treatise on general mechanics of Paul-Émile Appell (1855–1939) [1], the
cycles of volumes on theoretical physics of Arnold Sommerfeld (1868–1951) [68]
and of Lev Davidovich Landau and Evgenii M. Lifshitz [41].
Today, the theory of elasticity is well developed in multiple directions
including new computational methods, both analytical and numerical; more and
more disciplines are joined with the theory of elasticity in research of a common
character. For instance, to the theory of thermoelasticity, dating from nineteenth
century, are added magnetoelasticity, viscoelasticity, the theory of finite elastic
deformations or magneto-thermoelasticity. Because of this great development, it is
very difficult to mention, even briefly, the most famous men of science who dealt
or deal with these problems. That is why we preferred to mention only the research
which has become classical and to present only the research made until the first
decades of twentieth century.

1.2.2.3 Researches into the Mechanics of Deformable Solids


in Romania

At its beginnings, the research into the mechanics of deformable solids had a
chiefly technical and practical character; engineers of wide experience both in
design and execution, professors at the Polytechnical School in Bucharest, such as
Elie Radu (1853–1931), Anghel Saligny (1854–1937), Ion Ionescu (1870–1946),
Gheorghe Emanoil Filipescu (1882–1937) or such as Nicolae Profiri (1886–1967),
Aurel A. Belesß (1891–1976), Constantin C. Teodorescu (1892–1972), Cristea
Mateescu (1894–1979) and Mihail Hangan (1897–1964) studied problems of
calculus of structures by method of strength of materials and of statics of structure.
The research on this line continued successfully for the last years, supplying
interesting results in mechanics of structures (statics, stability and dynamics), in
machine construction mechanics etc.
As to the mathematical study of deformable solids, the first research dating
from the beginning of twentieth century, had—chiefly—a characteristic of equa-
tions of mathematical physics due to a few professors of the faculties of science in
the Bucharest and Jassy Universities, but who obtained their doctor titles in France
or in Germany. For instance, the first work in this domain was the doctoral thesis
of Anton Davidoglu (1876–1958) [84, 85], defended at Paris in 1900, concerning
the application of the successive approximations method of Émile Picard
(1856–1941) to the study of some differential equations of fourth order, corre-
sponding to the transverse vibrations of nonhomogeneous elastic bars.
24 1 Introduction

In connection with the same ideas, we bring into relief the doctoral thesis of
Alexandru Myller (1879–1965) [53], defended at Göttingen in 1906; it treats
ordinary differential equations in connection with integral equations and it offers
examples concerning the elastic bar equations, in various supporting cases.
Besides, Simion Sanielevici (1870–1963) [65, 94], defended at Paris, in 1908, a
doctoral thesis on the differential equations of the vibrating string and membranes.
Other mathematical researches into the deformable solids appear beginning with
1929, but are rather sporadic until 1948. In this time interval, the doctoral thesis of
Grigore C. Moisil (1906–1973) [49, 87, 88] appeared, concerning the application of
the non-linear functional analysis to the study of dynamics of strings; we mention
also his work in which some dynamic problems of the theory of elasticity are
studied [50, 89, 90]. On the other hand, we must emphasize the doctoral thesis of
Nicolae Teodorescu (1908–2000), as well as his work concerning the application
of the areal derivative and of the generalized potentials to problems of mechanics of
deformable continuous media [71, 95]. We mention also a paper on the vibrations
of a rectangular membrane published by Mircea Drăganu (1911–1984) [86].
Since this period, the problems of the theory of elasticity were commenced
within the framework of the study of systems of partial differential equations. First
of all, to be able to study the general properties of the systems of partial differential
equations with constant coefficients, the method of the associated matrices to these
systems off equations, elaborated by Gr. C. Moisil [50], was used; by so doing,
Gr. C. Moisil initiated, in 1949, at Bucharest, at the Institute of Mathematics of the
Romanian Academy, the first systematic research into the mathematical study of
the mechanics of deformable solids, thus creating—by the specialists formed
there—a true school in this field. These researches were developed later particu-
larly within the frame of the Mathematics and Mechanics Faculties of Bucharest
and Jassy Universities of the Bucharest and Jassy Mathematics Institutes and of the
Solid Mechanics Institute of the Romanian Academy, as well as—nowadays—in
many Technical Universities and in various research institutes. As in the previous
subsection, we do not go into details as to the present researches in this domain,
but we confine ourselves to the information mentioned above.

1.2.2.4 Development Trends of the Theory of Elasticity

Hereafter we shall try to bring into relief, quite succinctly, some of the develop-
ment trends of the theory of elasticity.
We mention, in the first plane, the study of other types of bodies (non-classi-
cal), offering also—in general—an elastic character, just as the study of certain
nonclassical problems of the theory of elasticity.
Thus, most of the studies made on isotropic bodies were extended to aniso-
tropic ones, in general, and to various particular cases of anisotropy. In most of the
cases, these generalizations do not present essential difficulties; obviously, the
form of the results is more intricate, since they depend on the 21 elastic constants
involved. In the static case, the harmonic or the biharmonic equations governing
1.2 Fundamental Computation Hypotheses. Short Historical Account 25

the problem become elliptic differential equations of a general form. In the


dynamic case, however, the problem becomes more intricate, the elastic waves
being—in general—no longer decomposable into primary and secondary waves;
on the other hand, it is interesting to see the form assumed by the surface waves of
Rayleigh type, as well as by other type of waves.
In case of non-homogeneous bodies, the problem is complicated because the
elastic coefficients of the material are function of point. We shall distinguish
between discontinuous non-homogeneity (e.g., stratified bodies) and continuous
homogeneity. In the first case, conditions of continuity are imposed to the dis-
placements and to the stresses on the separation surfaces, the respective equations
being written for each subdomain. In the case of a continuous non-homogeneity,
the elastic coefficients are continuous functions of point, while the equations of the
problems have variable coefficients; in the case and only in the case in which the
rigidity of the solid body is constant in parallel planes, the system of equations of
the theory of elasticity is with constant coefficients. To such bodies there have
been given solutions in various particular cases of non-homogeneity.
But it is important to remark that the difference between the non-homogeneous
and the anisotropic bodies is not yet sufficiently specified. Indeed, let us suppose
that we have to do with an anisotropic and homogeneous body the properties of
which are expressed in Cartesian orthogonal coordinates. By passing into any
curvilinear co-ordinates whatever, a body is obtained with curvilinear anisotropic
properties; but the elastic coefficients of such a body will be functions on the
curvilinear co-ordinates of the point. Therefore, we have to deal with a formal non-
homogeneity. In order to have a non-homogeneous body from the physical
standpoint, this body should be non-homogeneous in any system of curvilinear
coordinates; otherwise, we deal with a homogeneous body with a curvilinear
anisotropy. The mathematical conditions that could bring relief to these properties
are not yet clearly established. We mention moreover the application of the ale-
atory variables to certain nonhomogeneous bodies, statistically homogeneous.
It is interesting to remark that the aleatory variables can also be used in
studying other problems, for instance some vibration problems with application to
the vibration of automobiles.
Coming to a non-linear constitutive law, we find problems offering such
characteristics, while remaining in the elastic domain; we mention also the hyp-
oelastic bodies and the hyperelastic bodies.
A step forward was made by considering an asymmetric stress tensor; this led to
the introduction of couple-stresses (micro moments) and of six elastic constants of
the material for a linear constitutive law. In this case, the local rotation will play
an important part too. We shall distinguish between bodies with free rotations
(a particle of which has six degrees of freedom) and bodies with constrained
rotations; in the latter case (which also contains the case of classical elastic
bodies), the number of elastic constants is reduced to four. Such bodies are called
bodies of Cosserat type, after the name of Cosserat brothers, who began to study
them at the beginning of twentieth century.
26 1 Introduction

In connection with these bodies, we must equally recall the bodies with
microstructure. It is to be remarked that the structure of certain constructions can
be approximated with the help of such bodies, from the standpoint of static
computation.
Beside the above mentioned physical non-linearity, we deem it equally
important to emphasize the geometric non-linearity, due to nonlinear relations
between strains and displacements. Tensor calculus methods play an increasingly
important part, where the use of curvilinear co-ordinates is essential. Owing to this
non-linearity, yet another non-linearity of a mechanical order appears, concerning
the equations of equilibrium and motion.
Because of the great computational difficulties, various approximations are
used, as a function of the nature and form of the body as well as of the problems
arising. We must distinguish, for instance, between finite and infinitesimal
deformations and finite and infinitesimal rotations; on the other hand, any of the
combinations liable to be made is possible. The classical case is that of infini-
tesimal deformations and of infinitesimal rotations, the last hypothesis being
essential for a correct computation.
Let us remark that the non-linear constitutive laws can also occur by considering
an elastic potential depending not only on the first-order gradient of the displace-
ment (like in the classical case), but on higher-order gradients too; if we confine
ourselves to second-order gradients, then we deal with bodies of second order
(bipolar bodies); more general studies lead to multipolar bodies. We mention also
studies of the elastic potential of these bodies from the dynamical standpoint. The
case of bodies with initial stresses has been brought into relief as well.
Let us mention that the bodies of Cosserat type cannot be bodies of second
order since we have to do with two distinct directions of generalization. The
identity only occurs in the case of bodies with constraint rotations.
In connection with the various generalization trends of the theory of elasticity
and with the introduction of its specific methods and forms in the more embracing
frame of the mechanics of deformable solids, we must mention studies about
dislocations (linear defects in crystals).
The problem possesses a particular importance for the physics of solids; its
theoretical aspects are increasingly of interest in mechanics too. Indeed, in the
frame of the so-called continuous theory of dislocations we endeavour to present,
in a unitary form, the theory of elasticity, the theory of plasticity and the theory of
creep. In Japan, RAAG (Research Association of Applied Geometry) dealt, under
the guidance of Kazuo Kondo, with the unification of various engineering theories,
through the agency of geometry; one of its objectives was mentioned above.
Another current line of research is that of its connection with the thermal,
magnetic, electric etc. fields. Thus, were born thermoelasticity, magnetoelasticity,
electroelasticity etc. On the other hand, there are cases when all these effects have
to be considered simultaneously.
It is to be remarked, for instance, that the thermal effect influences the deformation
of the body, but it is reciprocal; heat propagation can be influenced by the defor-
mation of the body, while Fourier’s equation, e.g., must be completed accordingly.
1.2 Fundamental Computation Hypotheses. Short Historical Account 27

So appear the so-called coupled problems of thermoelasticity, magnetoelasticity etc.,


which are non-linear, but are—ordinarily—linearized. In general, the coupling
influence is very small (at least as concerns the problems studied up to now), so that
the problems can be studied as non-coupled ones, which is simpler.
The reciprocity theorems established in this direction are quite important, since
they can lead to useful integration methods.
A great many experimental facts, among which we mention: the dependence on
the wave number of the group velocity of the propagation of a short wave per-
turbation, the amplitude attenuation of a vibration propagated in a solid with a
crystalline structure, the influence of the deformation gradient upon the coefficient
of stress concentration around a crack, the appearance of surface stresses in a body
upon which no surface loads are acting, the influence of the micro-non-homoge-
neities of a body on its behaviour as observed at a macroscopic scale cannot be
satisfactorily explained by means of the continuous model of linear or non-linear
elasticity.
Deviations of the theoretical forecasts, obtained on the basis of the classical
model, with regard to the experimentally obtained results, imposed the necessity of
broadening the classical frame. The new generalized mechanics, few of which
were already mentioned, endeavour to describe more subtly, more accurately the
behaviour of solids, but using in this endeavour the language of continuous
mathematics too. The new models are based on the generalization of the classical
model from both the kinematic and kinetic standpoints and succeed in presenting
both the influence of micro motions on the behaviour of the deformable solid and
the role of the distant interactions in the observable phenomenological response of
a medium submitted to complex external actions.
Also, more and complex practical problems necessitated the study of the
interactions between a variable electromagnetic field and a deformable, polarizable
and magnetizable solid body. New models liable to reflect better this body-field
interaction can only be constructed in the frame of a relativistic theory. The
attempts made in the world literature follow mostly this line. With a view to
construct relativistic models, it is necessary to elaborate a corresponding theory of
deformation, so as to make a link with the Maxwellian theory of electromagnetism.
The relativistic models give an excellent description of the phenomena which
occur owing to body field interaction, but they are extremely intricate and the
solution, on their basis, of specific problems of a practical importance leads to
great mathematical difficulties. That is why it is necessary to elaborate non-linear
electro-mechanical models that, in many important cases, give a good approxi-
mation of the exact relativistic model. Comparing the relativistic models with the
non-relativistic ones allows us to estimate the errors occurring in the cases when
the electro-mechanical interaction is studied on the basis of a simpler non-rela-
tivistic model which is satisfied with the solution of less intricate mathematical
problems. This fact emphasizes the importance of both model types, as well as the
necessity of constructing and studying them.
The introduction of additional conditions about the stresses (e.g., a null normal
stress or two null tangential stresses), the strains or the displacements leads to
28 1 Introduction

simpler problems by diminishing the number of variables. Beside the classical


problems (plane or anti-plane), a great number of other two-dimensional problems
can be considered (where the third variable appears explicitly), the unknown
functions depending only on two variables.
Other additional conditions lead to the study of thin plates (hypothesis of the
straight line element of G. R. Kirchhoff), of thin shells (hypothesis of the straight
line element of A.-E.-H. Love) and of straight bars (hypothesis of the plane cross
section of Jacob Bernoulli).
In regard to such types of bodies, some more special problems begin to be
considered, for instance the case of the oblique plates and the introduction of the
couple-stresses in the study of plates and shells. The current trend is chiefly to
study the validity of the corresponding approximations. We must mention, par-
ticularly, the study of the plates, in the frame of the three-dimensional problems of
the theory of elasticity without making use of the Kirchhoff-Love hypotheses.
Some special problems formed and still form the object of study of many
researchers. Lately, the respective computation methods have been improved
greatly by using a stronger mathematical apparatus; we mention, e.g., the appli-
cation of the variational methods to non-linear problems with the help of the
nonlinear functional analysis.
Also, we shall bring into relief the study of the stress concentrations, liable to be
due to various causes: holes, cross section variations, contact between two bodies,
cracks etc. Most of the researches deal with the plane case, where the method of
the functions of complex variable plays an important part. We must also take into
account the additional conditions imposed in the case of multiply connected
domains.
Another case is the problem of concentrated loads, where we emphasized a
certain important tensor character and showed that a classification of these loads is
possible. It is also to be remarked that all the concentrated loads can be acquired
by starting from a single one of them, e.g., a concentrated force. In case of bodies
of Cosserat type with free rotations it was pointed out that there are two funda-
mental concentrated loads (singularities): a volume (concentrated) force and a
volume (concentrated) moment. A single fundamental concentrated load exists
only in the case of bodies with constrained rotations.
We should like to throw into relief the fact that the theory of distributions plays
an important role in the study of the influence of the concentrated loads; moreover,
the theory of distributions leads to a general methodology for solving the problems
of the theory of elasticity (unifying discontinuous problems with continuous ones).
The possibility of connecting the theory of elasticity with certain problems of
the mechanics of fluids is also to be stated; such a connection is illustrated by the
theory of aeroelasticity, which is closely related to the theory of stability of elastic
systems.
In case of anelastic bodies, some trends of development are to be observed. A lot
of research deals with problems of loading and unloading of materials, to put in
evidence various theories of plasticity. Another direction of study is that of the
behaviour of bodies by great pressures (e.g., the case of explosions). The apparition
1.2 Fundamental Computation Hypotheses. Short Historical Account 29

of elastic-plastic waves must be also mentioned; an important role is played by the


shock waves too. The apparition of a viscous behaviour of the material leads to
interesting researches; we mention the studies concerning the elastic-viscoplastic
bodies.
We mention also the research on optimum design of constructions, where one
takes into account that the element of construction can have plastic or viscous
properties.
As it could be seen, the spline functions lead to good approximations of the
solutions for one-dimensional problems. Using various analytical methods, finite
element methods or boundary element methods, one can build unitary programs of
computation for the most important types of elements of construction; these
programs of computation must allow their assembling for spatial systems (con-
sidering them as subprograms of a complex program), so as to can use electronic
computers.
Another method of calculation, very convenient for non-linear ordinary differ-
ential equations is the linear equivalence method (LEM), which leads to convenient
solutions from a quantitative point of view as well as from a qualitative one.
In the last few years, new directions of research appeared, e.g., problems of
nanomechanics (chiefly nanoelasticity) or problems of auxetic bodies, for which
Poisson’s coefficient can be also negative.
It is very difficult to present even succinctly the development trends in such a
vast field as the one which we propose to examine; that is why we have confined
ourselves to some of its aspects, perhaps not always the most important ones,
mentioning only a few of the researches.

References

A. Books

1. Appell, P.: Traité de mécanique rationnelle, vol. IV. Gauthier-Villars, Paris (1926)
2. Bach, C.: Elasticität und Festigkeit. Die für die Technik wichtigsten Sätze und deren
erfahrungmässige Grundlage, 2nd edn. J. Springer, Berlin (1894)
3. Belluzi, O.: Scienza delle costruzioni, vol. III. N. Zanichelli, Bologna (1956)
4. Bernoulli, J., Œuvres choisies, vol. II. Genève (1744)
5. Bezukhov, N.I.: Teoriya uprugosti (Theory of Elasticity). Gostekhizdat, Moskva (1953)
6. Biezeno, C.B., Grammel R.: Technische Dynamik, vols. I, II. 2nd edn. Berlin (1953)
7. Boussinesq, J.: Applications des potentiels à l’étude de l’équilibre et du mouvement des
solides élastiques, New ed., Al. Blanchard, Paris (1969)
8. Brdička, M.: Mechanika kontinua (Mechanics of Continuum). Nakladat. Českoplov. Akad.
Věd, Praha (1959)
9. Butty, H.: Tratado de Elasticidad TeoricoTehnica, vol. I. Buenos Aires (1946)
10. Castigliano, A.: Théorie de l’équilibre des systèmes élastiques. Fr. Bocca. Ed., Torino (1879)
11. Cesàro, E.: Introdutione alla teoria matematica della Elasticità. Fr. Bocca Ed., Torino (1894)
30 1 Introduction

12. Clebsch, A.: Théorie de l’élasticité des corps solides (traduite par MM. B. de Saint-Venant et
Flamant avec de notes étendues de M. de Saint-Venant). Paris (1883)
13. Cosserat, E., Cosserat, F.: Théorie des corps déformables. E. Hermaun et fils, Paris (1909)
14. Eringen, A.C.: Nonlinear Theory of Continuous Media. McGraw-Hill Book Co., Inc., New
York (1962)
15. Fl}uge, S. (ed.): Handbuch der Physik. VI. Elasticity and Plasticity. Springer, Berlin (1958)
16. F}oppl, A.: Vorlesungen } uber technische Mechanik. III. Festigkeitslehre. B. G. Teubner
Verlag, Leipzig (1898)
17. F}oppl, A.: Vorlesungen } uber technische Mechanik. V. Die wichtigsten Lehren der h} oheren
Elastizitätstheorie. B. G. Teubner Verlag, Leipzig (1907)
18. F}oppl, A. and L., Drang und Zwang. Eine h} ohere Festigkeitslehre f}ur Ingenieure. I, II.
R. Oldenbourg Verlag, München (1920)
19. Franciosi, V.: Scienza delle costruzioni, vol. I. Liguori, Napoli (1965)
20. Fung, V.C.: Foundations of Solid Mechanics. Prentice Hall, Inc., Englewood Cliffs (1965)
21. Galerkin, B.G.: Sobronie sochineniı (Complete Works), vols. I, II. Izd. Akad. Nauk SSSR,
Moskva (1952–1953)
22. Germain, P.: Mécanique des milieux continus. Masson Éd., Paris (1962)
23. Gol’denblat, I.I.: Nelineınye problemy teoriı uprugosti (Non-Linear Problems of the Theory
of Elasticity). Izd. ‘‘Nauka’’, Moskva (1969)
24. Goodier, J.N., Hodge Jr, P.G.: Elasticity and Plasticity. Wiley, New York (1958)
25. Green, A.E., Adkins, J.E.: Large Elastic Deformations and Non Linear Continuum
Mechanics. Clarendon Press, Oxford (1960)
26. Green, A.E., Zerna, W.: Theoretical Elasticity. Clarendon Press, Oxford (1954)
27. Grioli, G.: Mathematical Theory of Elastic Equilibrium (Recent Result). Springer, Berlin
(1962)
28. Gurtin, M.E.: The Linear Theory of Elasticity (in Handbuch der Physik. VIa/2). Springer,
New York (1972)
29. Haimovici, M.: Teoria elasticitătßii (Theory of Elasticity). Ed. did. ped, Bucuresßti (1969)
30. Hermite, R.L’.: Résistance des matériaux, théorique et expérimentale. I. Dunod, Paris (1954)
31. Hertz, H.: Gesammelte Werke, vol. I. Leipzig (1895)
32. Hooke, R.: Lectures de Potentia Restitutive or of Springs Explaining the Power of Springing
Bodies. John Martin, London (1678)
33. Howink, R.: Elastizität, Plastizität und Struktur der Materie. Dresden-Leipzig (1957)
34. Huber, M.T.: Teoria spre_zistości (Theory of Elasticity). I, II. Państ. Wydawn. Naukove,
Warszawa (1954)
35. Kecs, W., Teodorescu, P.P.: Applications of the Theory of Distributions in Mechanics. Ed.
Acad., Bucuresßti, Abacus Press, Tunbridge Wells (1974)
36. Kirchhoff, G.R.: Gesammelte Abhandlungen. Leipzig (1882)
37. Knops, R.J., Payne, L.E.: Uniqueness Theorems in Linear Elasticity. Springer, New York
(1971)
38. Kr}oner, E.: Kontinuumstheorie der Versetzungen und Eigenspannungen. Springer, Berlin
(1958)
39. Lamé, G.: Leçons sur la théorie mathématique de l’élasticité des corps solides. Paris (1852)
40. Lamé, G.: Leçons sur les coordonnées curvilignes. Paris (1859)
41. Landau, L., Lifchitz, E.: Théorie de l’élasticité. Ed. Mir, Moskva (1967)
42. Leıbezon, L.S.: Kurs teoriı uprugosti (Course of Theory of Elasticity). Ogiz, Moskva-
Leningrad (1947)
43. Love, A.E.H.: A Treatise of the Mathematical Theory of Elasticity, 4th edn. Cambridge
University Press, Cambridge (1934)
44. Lurje, A.I.: Räumliche Probleme der Elastizitätstheorie. Akad Verlag, Berlin (1963)
45. Lur’e, A.I.: Teoríya uprugosti (Theory of Elasticity). Izd. ‘‘Nauka’’, Moskva (1970)
46. Marcolongo, R.: Teoria matematica della elasticità. Milano (1902)
47. Marcolongo, R.: Teoria matematica dell’equilibrio dei corpi elastici. Milano (1904)
48. Mohr, O.: Technische Mechanik, 2nd edn. Verlag von W. Ernst & Sohn, Berlin (1914)
References 31

49. Moisil, Gr.C.: La mécanique analytique des systèmes continus. Thèse, Gauthier-Villars, Paris
(1929)
50. Moisil, Gr.C.: Matricele asociate sistemelor de ecuatßii cu derivate partßiale. Introducere în
studiul cercetărilor lui I.N.Lopatinschi (Matrices associated to systems of partial differential
equations. Introduction to the study of Lopatinski’s researches) Ed. Academiei, Bucuresßti
(1950)
51. Muskhelishvili, N.I.: Some Basic Problems of the Mathematical Theory of Elasticity.
P. Noordhoff Ltd., Groningen (1963)
52. M}uller, W.: Theorie der elastichen Verformung. Akad. Verlagsgesellschaft Geest & Portig
K.G, Leipzig (1959)
53. Myller, Al.: Gew}ohnliche Differenzialgleichungen h} ocherer Ordnung in ihrer Beziehung zu
den Integralgleichungen. Inaugural Dissertation zur Erlangung der Doktorw} urde, G}ottingen
(1906)
54. Navier, L.M.H.: Résumé des leçons sur l’application de la mécanique a l’établissement des
constructions et des machines. 3me (édn.) avec de notes et des appendices de B. de Saint-
Venant. Paris (1864)
55. Neumann, F.: Vorlesungen } uber die Theorie der Elasticität der festen Körper und des
Lichtäthers. Leipzig (1885)
56. Newton, I.: Philosophiae naturalis principia mathematica. S. Pepys, Reg. Soc. Praeses,
Londini (1687)
57. Novozhilov, V.V.: Osnovy nelineınoı teoriı uprugosti (Fundamentals of the non-linear theory
of elasticity). Gostekhizdat, Moskva-Leningrad (1948)
58. Novozhilov, V.V.: Teoriya uprugosti (Theory of Elasticity). Sudpromgiz, Moskva (1958)
59. Nowacki, W.: Teoria sprepystości (Theory of Elasticity). Państ. Wydawn. Naukowe,
Warszawa (1970)
60. Papkovich, P.F.: Teoriya uprugosti (Theory of Elasticity). Gostekhizdat, Moskva (1939)
61. Pearson, C.E.: Theoretical Elasticity. Harvard University Press, Cambridge (1959)
62. Planck, M.: Mechanik deformierbarer K} orper, 3rd edn. S. Hirzel, Leipzig (1931)
63. Poincaré, H.: Leçons sur la théorie de l’élasticité. Paris (1892)
64. Prager, W.: Introduction to Mechanics of Continua. Ginn & Co., Boston (1961)
65. Sanielevici, S.: Sur les équations différentielles des cordes et des membranes vibrantes.
Thèse, Gauthier-Villars, Paris (1908)
66. Sokolnikoff, I.S.: Mathematical Theory of Elasticity, 2nd edn. McGraw-Hill Book, Co., Inc.,
New York (1956)
67. Sokolnikoff, I.S.: Tensor Analysis. Theory and Applications to Geometry and Mechanics of
Continua, 2nd edn. Wiley, New York (1964)
68. Sommerfeld, A.: Mechanik der deformierbaren Medien, 6th edn. Akad. Verlags-gesellschaft
Geest & Portig K.-G., Leipzig (1949)
69. Southwell, R.V.: An Introduction to the Theory of Elasticity for Engineers and Physicists,
2nd edn. Oxford University Press, London (1953)
70. Strutt, J.W. (Lord Rayleigh): The Theory of Sound. London (1877)
71. Teodorescu, N.: La dérivée aréolaire et des applications à la physique mathématique. Thèse,
Gauthier-Villars, Paris (1931)
72. Teodorescu, P.P.: Probleme spatßiale în teoria elasticitătßii (Space problems in the theory of
elasticity). Ed. Academiei, Bucuresßti (1970)
73. Teodorescu, P.P.: Dynamics of Linear Elastic. Bodies. Ed. Acad. Bucuresßti, Abacus Press,
Tunbridge Wells, Kent (1975)
74. Thomson, W., Tait, P.G.: Treatise on Natural Phylosophy, vol. I. Cambridge University
Press, Cambridge (1861)
75. Timoshenko, S.P.: History of Strength of Materials. McGraw-Hill Book Co., Inc., New York
(1953)
76. Timoshenko, S.P.: Collected Papers. McGraw-Hill Book Co., Inc., New York (1953)
77. Timoshenko, S.P., Goodier, J.N.: Theory of Elasticity, 2nd edn. McGraw-Hill Book Co., Inc.,
New York (1951)
32 1 Introduction

78. Todhunter, I., Pearson, K.: History of the theory of elasticity, vols. I–III. Cambridge
(1886–1893)
79. Volterra, V., Volterra, R.: Sur les distorsions des corps élastiques (théorie et applications).
Mém. Sci. Math., CXLVII. Gauthier-Villars, Paris (1960)
80. Wang, C.C., Truesdell, C.: Introduction to Rational Elasticity. Noordhoff International
Publishing, Leyden (1973)
81. Wang, Chi.Teh.: Applied Elasticity. McGraw-Hill Publishing Company, New York (1953)
82. Westergaard, H.M.: Theory of Elasticity and Plasticity. Harvard University Press, Cambridge
(1952)
83. Winkler, E.: Die Lehre von der Elastizität und Plastizität. Praha (1867)

B. Papers

84. Davidoglu, A.: Sur l’équation des vibrations transversales des verges élastiques. Thèse. Ann.
Sci. de l’École norm. sup. Paris, 3rd ser., 17, 359 (1900)
85. Davidoglu, A.: Sur une application de la méthode des approximations successives. C. Rend.
hebd. des séauces de l’Acad. Sci. 130, 692; 1341 (1900)
86. Drăganu, M.: Sur une résolution de l’équation différentielle des mouvements vibratoires
d’une membrane rectangulaire par la méthode de la transformation multiple de Laplace.
Mathematica 22, 206 (1946)
87. Moisil, Gr.C.: Sur la dynamique du fil. Bull. Math. de la Soc. Roum. Sci., 31, 2; 170 (1929)
88. Moisil, Gr.C.: Sur le mouvement d’un fil sur une surface. Mathematica 3, 144 (1930)
89. Moisil, Gr.C.: Asupra sistemelor de ecuatßii cu derivate partßiale liniare cu coeficientßi
constantßi (On the systems of linear partial differential equations with constant coefficients).
Bul. Sßt. Acad., ser A, 1, 4; 341 (1949)
90. Moisil, Gr.C.: Teoria preliminară a sistemelor de ecuatßii cu derivate partßiale liniare cu
coeficientßi constantßi (Preliminary theory of systems of linear partial differential equations
with constant coefficients). Bul. Sßt. Acad., S
ßt. Mat.-Fiz., 4, 2; 319 (1952)
91. Myller, Al.: Asupra ecuatßiilor coardelor vibrante (On the equations of vibrating strings). Gaz.
Mat. 14, 6 (1909)
92. Oravas, G.Æ., Lean, L.Mc: Historical development of energetical principles in
elastomechanics. I. From Heraclites to Maxwell. Appl. Mech. Rev. 19, 8; 647 (1966)
93. Oravas, G.Æ., Lean, L.Mc: Historical development of energetical principles in
elastomechanics. II. From Cotterill to Prange. Appl. Mech. Rev. 19, 11; 919 (1966)
94. Sanielevici, S.: Sur l’équation aux dérivées partielles des membranes vibrantes. C. Rend.
hebd. des séances de l’Acad. Sci. 146, 1249; 1387 (1908)
95. Teodorescu, N.: Sur une formulation généralisant l’intégrale de Cauchy et sur les équations
de l’élasticité plane. C. Rend hebd. des séances de l’Acad. Sci. 189, 565 (1930)
96. Teodorescu, P.P.: Asupra ipotezei sectßiunilor plane în rezistentßa materialelor (On the
hypothesis of the plane cross-sections in strength of materials). An. Univ. Bucuresßti, ser. ßst.
nat. 6, 41 (1957)
97. Teodorescu, P.P.: One hundred years of investigations of the plane problem of the theory of
elasticity. Appl. Mech. Rev. 17, 175 (1964) (also in ‘‘Applied Mechanics Surveys’’, Spartan
Books, Washigton, 245 (1966))
98. Teodorescu, P.P.: Entwicklungstendenzen in der Elastizitätstheorie. Mitteilungen der Math.
Gesellschaft der D. D. R. 1, 5 (1970)
}
99. Wieghardt, K.: Uber ein neues Verfahren, verwickelte Spannungsverteilungen in elastischen
K}orpern auf experimentellen Wege zu finden. Mitt. } uber Forsch. Geb. Ing. 49, 15 (1908)
Chapter 2
Geometry and Kinematics of Deformation

We shall now proceed to a theoretical study in which we shall emphasize


geometrical aspects of the problems; these problems will be dealt for a fixed t or
taking into account the time too [7].

2.1 Finite Deformations

To study the deformation of a body, we start from the general case of finite
deformations. After some general considerations, we pass to the geometric aspects
of the deformation, using material co-ordinates, as well as space ones.

2.1.1 General Considerations

Referring to a system of co-ordinates which are considered fixed (Fig.2.1a) the


particle of the body, coinciding with the geometric point Mðx1 ; x2 ; x3 Þ of position
vector r, before deformation, will be at the geometric point M  ðx1 ; x2 ; x3 Þ of
position vector r , after deformation. We shall admit that the external loads which
act upon the solid are in dynamic equilibrium at a given time t.

2.1.1.1 Displacement Vector

!
The displacement MM  ¼ u shall be a vector the components of which will be
denoted by u1 ; u2 ; u3 . We can write
xi ¼ xi þ ui ; i ¼ 1; 2; 3: ð2:1Þ

P. P. Teodorescu, Treatise on Classical Elasticity, Mathematical and Analytical 33


Techniques with Applications to Engineering, DOI: 10.1007/978-94-007-2616-1_2,
Ó Springer Science+Business Media Dordrecht 2013
34 2 Geometry and Kinematics of Deformation

x3 M *(x1*, x2*, x3* ) x3


u u+du
u3 N N*
M(x1, x2, x3) dv
u2 u1 v D D*
r v
r* M(x1, x2, x3) u v*
r
M *(x*1, x*2, x*3 )
r*
O x2 O x2
x1 x1
(a) (b)

Fig. 2.1 Displacement (a). Geometric mapping (b)

If the transition from the initial state to the actual (final) state does not occur
instantaneously (in theory) or in a very short time (in practice), as it is considered in
general, in the static case, then the phenomenon assumes a dynamic character and
the study of the deformation as a function of time becomes necessary. We will
consider to be in the general case; in the static case, the time variable will disappear.
If the body occupies the domain D before deformation (at the initial moment
t ¼ t0 ) and the domain D after deformation (at a moment t), we shall admit that
the transition from D to D is continuous, so that the equations characterizing the
deformation will take the form (Fig. 2.1b)
xi ¼ xi ðx1 ; x2 ; x3 ; tÞ; i ¼ 1; 2; 3; ð2:2Þ
corresponding to a geometric mapping. We shall also admit that the inverses of
these functions
xj ¼ xj ðx1 ; x2 ; x3 ; tÞ; j ¼ 1; 2; 3; ð2:20 Þ

are univocally determined. Thus, we must deal with single-valued displacements.


There are also cases, such as these of multiply connected domains, where multi-
valued displacements can be considered, but we shall not deal with them here.
So that the functions (2.20 ) be univocally determined, in at least one of the
neighbourhoods of the considered point, it is necessary and sufficient for the
function (2.2) to be of the class C1 in D and to have a non-vanishing Jacobian
      
oðx1 ; x2 ; x3 Þ ox
J  det ¼ det i 6¼ 0 ð2:3Þ
oðx1 ; x2 ; x3 Þ oxj

in this domain. This hypothesis is known under the name of axiom of continuity,
expressing the indestructibility of matter. No domain, to which corresponds a finite
positive volume, can be deformed into a domain of zero or infinite volume. This
implies moreover the impenetrability of matter. The motion (2.2) (or (2.20 ))
transforms any domain into another domain, any surface into another surface and
any curve into another curve.
2.1 Finite Deformations 35

2.1.1.2 State of Displacement. State of Deformation (State of Strain)

By taking into account the supporting condition of the solid body (at the sup-
porting points, lines or surfaces, the displacements are either zero or have imposed
values), we eliminate the rigid-body motion and the three functions
ui ¼ ui ðx1 ; x2 ; x3 ; tÞ; i ¼ 1; 2; 3; ð2:4Þ

will characterize entirely the state of displacement in the neighbourhood of the


point M. The state of deformation of the whole body shall be given by the states of
deformation in each of its points. Further, these states are called states of strain.
By a one-dimensional or linear state of strain we mean the state of strain in
which the points on the surfaces of normals perpendicular to a fixed direction
move along planes that are tangential to these surfaces. That is, for instance, the
case of the straight bars of any cross section subjected to simple torsion.
The two-dimensional or plane state of strain is a state of strain in which the
particles of a body move in planes the normals of which have a fixed directions,
such as the case of the elastic half-space uniformly acted upon along a given
direction on the separation plane or on a plane parallel to it, provided that the
external load has no tangential components along this direction.
The three-dimensional or spatial state of strain is the general case of defor-
mation of any body whatsoever.
For reasons that will be given subsequently, we shall admit that the functions
(2.4) are of class C3 with respect to the space variables and of class C2 with respect
to the time variable.

2.1.2 Material and Space Co-ordinates. Strains

It is necessary to study the variation of the state of strain when we pass from a
point M to a neighbouring one N (Fig. 2.1b). But we must remark from the outset
that this study can be made in two ways.
We can thus consider the variables x1 ; x2 ; x3 , corresponding to the domain D
(the solid in the initial state), as independent variables, named Lagrange
co-ordinates (material co-ordinates); this denomination is due to the fact that,
during the phenomenon of deformation, one follows a material particle of the solid
in its displacement. They are called referential co-ordinates too, because they
specify the position of the particle with respect to the initial state, considered to be
a referential frame. In this computational method, the co-ordinates of the point M  ,
which specify the domain D (the solid in the actual state) are the unknown
functions of the problem, that have to be determined; the same thing can be stated
about the displacements (2.4).
As well, we can consider the variables x1 ; x2 and x3 as independent variables,
under the name of Euler co-ordinates (space co-ordinates); this denomination is
36 2 Geometry and Kinematics of Deformation

due to the fact that, at a given point of the space, all the particles which pass
through this point at various moments, because of their displacements, are
examined. The problem is put to determine the position vector r (the unknown
functions (2.20 )) and the displacements in the form
ui ¼ ui ðx1 ; x2 ; x3 ; tÞ; i ¼ 1; 2; 3; ð2:40 Þ

where the co-ordinates are time functions


xj ¼ xj ðtÞ; j ¼ 1; 2; 3: ð2:5Þ

In the case of finite deformations, this will lead to two distinct formulations of
the problems, but these formulations will be identical with each other in the case of
small deformations and rotations.

2.1.2.1 Material Co-ordinates

Let us consider the variation of the state of strain when passing from the point
Mðx1 ; x2 ; x3 Þ to an adjoining point Nðx1 þ dx1 ; x2 þ dx2 ; x3 þ dx3 Þ; the corre-
sponding points after deformation will be M  and N  , respectively (Fig. 2.2).
We can write the vector relations
!  ! ! !
NN   MM  ¼ M  N   MN ¼ dr  dr; ð2:6Þ

it is easy to see that the displacement components will differ by the components of
!  !
the vector NN   MM  ¼ du, i.e.
oui
dui ¼ dxj ; i ¼ 1; 2; 3; ð2:7Þ
oxj

the relations being valid at a given time t.


We notice that
dr2 ¼ ds2 ¼ dxk dxk ¼ dij dxi dxj ð2:8Þ

Fig. 2.2 Material and space N


*

co-ordinates
n u+du dr*

M*
N
u
dr
M (x1, x2, x3)
r*
r

O
2.1 Finite Deformations 37

and
dr2 ¼ ds2 ¼ dxk dxk ¼ cij dxi dxj ; ð2:80 Þ

where
oxk oxk
cij ¼ ; i; j ¼ 1; 2; 3; ð2:9Þ
oxi oxj
is Green’s deformation tensor TG , represent the distance between the points M and
N (before deformation) and between the points M  and N  (after deformation),
respectively. We obtained thus

ds2  ds2 ¼ 2eij dxi dxj ; ð2:800 Þ

where
1 
eij ¼ cij  dij ; i; j ¼ 1; 2; 3; ð2:10Þ
2
is the Lagrangian deformation tensor TL .
Taking into account (2.1), it results
oxk ouk
¼ dki þ ; k; i ¼ 1; 2; 3;
oxi oxi
so that, using the expression (2.9), the relation (2.10) becomes
 
1 oui ouj 1 ouk ouk
eij ¼ þ þ ; i; j ¼ 1; 2; 3: ð2:100 Þ
2 oxj oxi 2 oxi oxj

Thus the functions


eij ¼ eij ðx1 ; x2 ; x3 ; tÞ; i; j ¼ 1; 2; 3; ð2:11Þ
are the components of a symmetric tensor of second order, which has been
introduced by G. Green and B. Saint-Venant and allows us to define the Green
deformation tensor; it can be expressed also in the matric form
2 1 1 3
e11 e12 e13
6 2 2 7
61 1 7
6
TL ¼ 6 e21 e22 e23 7 ð2:110 Þ
7:
42 2 5
1 1
e31 e32 e33
2 2

2.1.2.2 Space Co-ordinates

Taking (2.40 ) into account and referring to the results obtained in the preceding
!  !
subsections, the components of the vector NN   MM  ¼ du will be given by
38 2 Geometry and Kinematics of Deformation

oui 
dui ¼ dx ; i ¼ 1; 2; 3; ð2:70 Þ
oxj j

these relations are for the time t too.


We notice that

dr2 ¼ ds2 ¼ dxk dxk ¼ dij dxi dxj ð2:12Þ

and

dr2 ¼ ds2 ¼ dxk dxk ¼ cij dxi dxj ; ð2:120 Þ

where
oxk oxk
cij ¼ ; i; j ¼ 1; 2; 3; ð2:13Þ
oxi oxj

is Cauchy’s deformation tensor TC , represent the distance between the points M 


and N  (after deformation) and between the points M and N (before deformation),
respectively. We obtain thus

ds2  ds2 ¼ 2eij dxi dxj ; ð2:1200 Þ

where
1

eij ¼ dij  cij ; i; j ¼ 1; 2; 3; ð2:14Þ


2
is the Eulerian deformation tensor TE .
Taking into account (2.11), it results
oxk ouk
¼ dki   ; k; i ¼ 1; 2; 3;
oxi oxi

so that, using the expression (2.13), the relation (2.14) becomes


!
 1 oui ouj 1 ouk ouk
eij ¼ þ  ; i; j ¼ 1; 2; 3: ð2:140 Þ
2 oxj oxi 2 oxi oxj

Thus the functions


eij ¼ eij ðx1 ; x2 ; x3 ; tÞ; i; j ¼ 1; 2; 3; ð2:15Þ

are the components of a symmetric tensor of second order, which has been
introduced by A.-L. Cauchy for the infinitesimal case and by E. Almansi and G.
Hamel in the case of finite deformations and allows us to define the Cauchy
deformation tensor; it can be expressed also in the matric form
2.1 Finite Deformations 39

2 1  1  3
e e e
6 11 2 12 2 13 7
61  1  7
TE ¼ 6
6 2 e21 e22 e 7: ð2:150 Þ
4 2 23 75
1  1 
e e e33
2 31 2 32

2.1.2.3 Strains

We shall define by
ds  ds
en ¼ ð2:16Þ
ds
the linear strain (extension) of the element of a line at the point M ðx1 ; x2 ; x3 Þ, in
the direction n. If en [ 0, then we have a stretching, corresponding—in the case of
a one-dimensional state of strain and stress—to a normal tension stress, and, if
en \0, then we shall have a shortening, corresponding—under the same condi-
tions—to a normal compression stress.
Taking into account (2.16), it follows that
ds ¼ ð1 þ en Þds; ð2:160 Þ

while the relation (2.800 ) allowsto write,in material co-ordinates,


1
en 1 þ en ¼ eij ni nj ; ð2:17Þ
2

where we used the relations giving the direction cosines


dxi
ni ¼ cosðn; xi Þ ¼ ; i ¼ 1; 2; 3: ð2:18Þ
ds
If the direction n coincides with the direction Ox1 , then we obtain n1 ¼ 1,
n2 ¼ n3 ¼ 0, which leads to
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
e11 ¼ 1 þ 2e11  1; ð2:19Þ

therefore e11 , e22 , e33 represent the extensions in the directions of the axes of
co-ordinates, related to the components of the tensor TL by expressions of the form
(2.19) (we use a notation with two indices for reasons which will be seen in the
case of infinitesimal deformations).
Generally, for the direction n we can write
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
en ¼ 1 þ 2eij ni nj  1: ð2:190 Þ
! !
Let us now consider two elements of a line MN and MP , parallel to the co-ordinate
axes Ox2 and Ox3 , respectively; let u23 be the angle between these elements after
the deformation (Fig. 2.3). We shall define by
40 2 Geometry and Kinematics of Deformation

Fig. 2.3 Angular x3


deformation P*

P N*

23
dx3 M *

M (x1, x2,x3) dx2 N

O x2

p
c23 ¼  u23 ð2:20Þ
2
the variation of the right angle formed by two elements of a line, parallel to the
co-ordinate axes Ox2 and Ox3 . So, c23 , c31 and c12 characterize the angular strain
(shearing strain). They are positive when representing a diminution of the right
angle and correspond to positive tangential stresses; otherwise (an increase of the
right angle), the angular strains are negative.
From the triangle M  N  P we deduce (equally in material co-ordinates), by
using the cosine theorem,
! pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
N  P 2 ¼ ð1 þ 2e22 Þdx22 þ ð1 þ 2e33 Þdx23  2 ð1 þ 2e22 Þð1 þ 2e33 Þ cos u23 dx2 dx3 ;

on the other hand, the formula (2.800 ) allows us to write


! !
M  N  2 ¼ ð1 þ 2e22 Þdx22 ; M  P 2 ¼ ð1 þ 2e33 Þdx23 ;
!
N  P 2 ¼ ð1 þ 2e22 Þdx22 þ ð1 þ 2e33 Þdx23  4e23 dx2 dx3

and we remark that, by passing from N  to P , we obtain dx2 \0. This leads to
2e23
sin c23 ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffipffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi : ð2:21Þ
1 þ 2e22 1 þ 2e33
In general, for two directions of unit vectors n and m, which make the angle x,
we can write
n m n m
d r d r ¼ d r d r cos x;
n m n m
dr  dr ¼ dr  dr  cosðx  cnm Þ;

where we have introduced the angular strain cnm with respect to the angle x; taking
into account (2.9) and (2.10), we have
n m
n

m

n m ox ox n m n m
dr  dr ¼ dxk dxk ¼ k k dxi dx j ¼ ð2eij þ dij Þdxi dx j :
oxi oxj
2.1 Finite Deformations 41

Using also the formula (2.160 ), we get


n m n m
dr  dr ð2eij þ dij Þdxi dxj
cosðx  cnm Þ ¼ n m ¼ n m;
dr  dr  ð1 þ en Þð1 þ em Þdr dr

observing that
n m
dxi dxj
n ¼ ni ; m ¼ mj ; i; j ¼ 1; 2; 3;
dr dr
it results
ð2eij þ dij Þni mj 2eij ni mj þ cos x
cosðx  cnm Þ ¼ ¼ : ð2:210 Þ
ð1 þ en Þð1 þ em Þ ð1 þ en Þð1 þ em Þ
In particular, if the directions n and m are orthogonal one each other, then we
obtain
2eij ni mj
sin cnm ¼ : ð2:2100 Þ
ð1 þ en Þð1 þ em Þ
In space co-ordinates, we get—analogically—relations of the form
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
e11 ¼ 1  1  2e11 ; ð2:22Þ

2e23
sin c23 ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffipffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ; ð2:23Þ
1  2e22 1  2e33

the linear strain em being defined with respect to the distance between two adjacent
points after the deformation (Fig. 2.4a).
ds  ds
en ¼ ; ð2:1600 Þ
ds
obviously, the angular strain is defined with respect to the right angle between two
segments of a line after deformation (which before the deformation formed an
angle uyz ) (Fig. 2.4b).
Generally, we can write
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
en ¼ 1  1  2eij ni mj ; ð2:220 Þ

2eij ni mj


sin cnm ¼ ; ð2:230 Þ
ð1  en Þð1  em Þ

where the direction after deformation are emphasized.


42 2 Geometry and Kinematics of Deformation

Fig. 2.4 Linear (a) and n


angular (b) strain
*
N
x3
-dr * P*

N M* P
N*
-dr *
M

23
M r* N
r
M (x1, x2, x3)
O O x2
(a) (b)

Let us remark that, in this case, we can write the relation


ð1 þ en Þð1  en Þ ¼ 1: ð2:24Þ
The volume (cubical) strain in material co-ordinates (if there are not angular
strains in planes parallel to the axes of co-ordinates) takes the form (with respect to
the volume before deformation)
dx1 ð1 þ e11 Þdx2 ð1 þ e22 Þdx3 ð1 þ e33 Þ  dx1 dx2 dx3
ev ¼ ;
dx1 dx2 dx3
wherefrom

ev ¼ ð1 þ e11 Þð1 þ e22 Þð1 þ e33 Þ  1


pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffipffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffipffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
¼ 1 þ 2e11 1 þ 2e22 1 þ 2e33  1; ð2:25Þ

likewise, in space co-ordinates, we can write (with respect to the volume after
deformation)
dx1 dx2 dx3  dx1 ð1  e11 Þdx2 ð1  e22 Þdx3 ð1  e33 Þ
ev ¼ ;
dx1 dx2 dx3

so that

ev ¼ 1  ð1  e11 Þð1  e22 Þð1  e33 Þ


pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffipffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffipffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
¼ 1  1  2e11 1  2e22 1  2e33 : ð2:250 Þ

Referring to the mapping relations (2.2) and taking into account (2.1) and the
functional determinant (2.3), we can equally write
2.1 Finite Deformations 43

   
oxi oui
1 þ ev ¼ det ¼ det dij þ
oxj oxj

ou1 ou1 ou1
1 þ
ox1 ox2 ox3

ou2 ou2 ou2
¼ 1þ ; ð2:26Þ
ox1 ox2 ox3
ou ou3 ou3
3

ox1 ox2 ox3
analogically, we have
" # " #
oxi oui
1 ev ¼ det ¼ det dij þ 
oxj oxj

ou ou1 ou1
1  1
ox1 ox2 ox3

ou ou2 ou2
2
¼ 1  : ð2:260 Þ
ox1 ox2 ox3

ou3 ou3 ou3
1 
ox ox2 ox3
1

We can write a relation, analogue to the one in the case of extensions, in the
form
ð1 þ ev Þð1  ev Þ ¼ 1: ð2:27Þ
If we compute these determinants, then we are able to express the cubical
dilatation by means of linear and angular strain components; if the angular strains
vanish, then we shall obtain once again the relations (2.25), (2.250 ), which were
deduced under just these conditions (i.e., by admitting that an elementary paral-
lelepiped, the sides of which are parallel to the axes of co-ordinates, maintains its
form throughout the deformation). This occurs for the three directions, orthogonal
one to each other, for which the linear strains have extreme values; indeed, from
the formulae (2.19), (2.22) it follows that these directions coincide with the
principal direction of the tensors TL and TE , respectively, while the formulae
(2.21) and (2.23) show that the corresponding angular strains vanish.

2.2 Infinitesimal Deformations

In the case of infinitesimal deformations, we shall admit that the strain can be
neglected with regard to unity; also, the product of two such strains can be
neglected with respect to a third one.
44 2 Geometry and Kinematics of Deformation

We also consider the hypothesis of the infinitesimal rigid body local rotation
(which we will subsequently define); in this case, the components of the rigid body
local rotation vector are negligible with respect to unity. It is important to mention
that these hypotheses are independent from each other, since the possibility exists
of simultaneous infinitesimal deformations and finite rigid body local rotations or
inversely. On the other hand, the principle of superposition of effects ceases to be
valid, because the superposition order of infinitesimal deformations and finite
deformations, for example, should be taken into account.
In computations, products can also occur between strains and rigid body local
rotations, the omission of which with respect to unity should be made with great
care, observing especially that the ratio between these quantities is to be taken into
account.
The hypotheses of infinitesimal deformations and of infinitesimal rigid body
local rotations are equivalent to the hypotheses of neglecting of the displacement
gradient with respect to unity. The hypothesis corresponds to what we shall call,
shortly, the case of infinitesimal deformations, when the relations (2.100 ) and
(2.140 ) become linear; that is the case of a linear theory from a geometric
standpoint.
Let f ðx1 ; x2 ; x3 ; tÞ be a function of class C1 , expressed by means of material
co-ordinates; using space co-ordinates, we obtain the function in the form
f  ðx1 ; x2 ; x3 ; tÞ. The partial derivative of first order with respect to one of the space
variables, e.g., the variable x1 , will be written in the following form
 
of of  ou1 of  ou2 of  ou3
¼  1þ þ  þ ; ð2:28Þ
ox1 ox1 ox1 ox2 ox1 ox3 ox1

where we took (2.1) into account; neglecting the displacement gradient with
respect to unity, it follows that
of of 
¼ : ð2:29Þ
ox1 ox1
Therefore, in the linear theory, the same results, in the same form, will be found
whether the material co-ordinates or the space ones are used.
In what follows we will suppose to be in the case of infinitesimal deformations
and we will use only material co-ordinates.

2.2.1 Displacement Gradient

We will introduce the displacement gradient tensor as a sum of the strain tensor
and the rigid body local rotation tensor. By this occasion, we will put in evidence
the connection with the displacement vector, as well as with the volume strain.
2.2 Infinitesimal Deformations 45

2.2.1.1 Strain Tensor

In the case of the linear theory, we can stop at the first terms of an expansion in
series of the expression (2.19) and approximate the sine by means of the corre-
sponding arc in (2.21); we shall have
e11 ¼ e11 ; e22 ¼ e22 ; e33 ¼ e33 ; ð2:30Þ

c23 ¼ 2e23 ; c31 ¼ 2e31 ; c12 ¼ 2e12 : ð2:300 Þ


In space co-ordinates, we obtain similarly
e11 ¼ e11 ; e22 ¼ e22 ; e33 ¼ e33 ; ð2:31Þ

c23 ¼ 2e23 ; c31 ¼ 2e31 ; c12 ¼ 2e12 : ð2:310 Þ


Taking into account the previous remarks concerning the partial derivatives of
first order of a function of class C1 and the relations (2.100 ), (2.140 ), it follows that
the deformation tensors TG and TC coincide. We obtain thus the strain tensor
2 3
e11 e12 e13
Te ¼ ½eij  ¼ 4 e21 e22 e23 5; ð2:32Þ
e31 e32 e33

which is a symmetric tensor of second order due to Cauchy; its components


eij ¼ eij ðx1 ; x2 ; x3 ; tÞ; i; j ¼ 1; 2; 3; ð2:33Þ

will be considered as depending on the material variables, to which we shall refer


henceforth.
We notice that for i ¼ j we obtain the linear strains corresponding to the
co-ordinate axes; the corresponding angular strains are given by
cij ¼ 2eij : ð2:330 Þ

The relation (2.17) takes the form


en ¼ eij ni nj ð2:34Þ

and shows how the linear strains vary in the neighbourhood of a point; as a matter
of fact, this relation can also be deduced from the fact that the strains are the
components of a tensor, which allows us to write, by analogy,
cnm ¼ 2eij ni mj ; ð2:340 Þ

where cnm is the angular strain, corresponding to the directions n and m,


orthogonal with each other.
46 2 Geometry and Kinematics of Deformation

!
Let us consider the vector MP of modulus

! 1
MP ¼ pffiffiffiffiffiffiffiffi ; ð2:35Þ
en
directed along the external normal n to the considered area element; to simplify the
computation, we shall consider that the origin of the co-ordinate axes is at the
point M, so that
ni
xi ¼ pffiffiffiffiffiffiffiffi ; i ¼ 1; 2; 3: ð2:350 Þ
en
By means of the relation (2.34), we find that the locus of the point P is given by the
equation
eij xi xj ¼ 1; ð2:36Þ

which indicates how en varies; that is why the surface takes the name of Cauchy
linear strain quadric at the point M. We remark that one takes the sign + or -,
according to whether we have to deal with a tension or with a compression, so that
the quadric be real. It results from (2.35) that the extension of any linear element at
M is in inverse proportion with the square of the radius vector of the quadric along
the linear element; hence, the extreme values of these extensions will correspond
to the directions given by the axes of the quadric.
With M1; M2; M3 as principal axes, the equation of the quadric can be written
in the form

e1 x02 02 02
1 þ e2 x2 þ e3 x3 ¼ 1; ð2:360 Þ

where x0i ; i ¼ 1; 2; 3, are the new co-ordinates. If e1  e2  e3 [ 0 or


0 [ e1  e2  e3 , then the quadric is an ellipsoid; we take the sign of 1 so as to have
a real ellipsoid. Otherwise, one obtains a one-sheet or a two-sheet hyperboloid as
we choose the sign of 1 (we notice that both hyperboloids, which are conjugate,
form the considered locus). There occur also cases when the quadric is degenerate.
Taking into account the theory of quadrics, we see that there are three principal
directions along which the extensions e1  e2  e3 acquire extreme values, while
the angular strains vanish. So as to find these principal axes, we shall consider the
relation (2.34) and impose the condition that the function en of three variables,
linked by the relations
ni ni ¼ 1; ð2:37Þ

should have an extremum. By so doing, we introduce the Lagrange multiplier and


compute the partial derivatives of the function
U ¼ en  eðni ni  1Þ
2.2 Infinitesimal Deformations 47

with respect to the direction cosines, which we equate to zero


oU
¼ 0; i ¼ 1; 2; 3:
oni
We obtain thus a system of equations
ðeij  edij Þnj ¼ 0; i ¼ 1; 2; 3; ð2:38Þ

where e is an extreme strain; indeed, the relations (2.38) take place only for the
principal directions, and if we choose these directions as co-ordinate axes and
0
consider, e.g., n ¼ x1 ðn1 ¼ 1; n2 ¼ n3 ¼ 0Þ, then we get e ¼ e11 ¼ e1 .
So that this homogeneous system in direction cosines should possess non-trivial
solutions, the following determinant

e11  e e12 e13

det½eij  edij   e21 e22  e e23 ¼ 0 ð2:39Þ
e31 e32 e33  e

should be equal to zero, which leads to the third degree equation

e3  I1 e2 þ I2 e  I3 ¼ 0; ð2:40Þ

with
1
I1 ¼ ijk ljk eil ¼ dil eil ¼ eii ¼ e11 þ e22 þ e33 ¼ e1 þ e2 þ e3 ; ð2:400 Þ
2
1 1
I2 ¼ ijk lmk eil ejm ¼ ðeii ejj  eij eij Þ ¼ e22 e33 þ e33 e11 þ e11 e22
2 2
 ðe223 þ e231 þ e212 Þ ¼ e2 e3 þ e3 e1 þ e1 e2 ; ð2:4000 Þ

1
I3 ¼ ijk lmn eil ejm ekn ¼ det½eij  ¼ e11 e22 e33
6
 ðe11 e223 þ e22 e231 þ e33 e212 Þ þ 2e23 e31 e12 ¼ e1 e2 e3 : ð2:40000 Þ

The equation (2.40) supplies the values of the principal strains, while the
system (2.38) to which we associate the condition (2.37) supplies then the prin-
cipal direction corresponding to a given extension.
Let e1 ; e2 be the principal extensions, to which correspond the systems of
equations
1 1 2 2
e1 n i ¼ eij n j ; e2 n i ¼ eij n j ; i ¼ 1; 2; 3:
48 2 Geometry and Kinematics of Deformation

2 1
We multiply the first system by n i and the second one by n i ; we then subtract the
equations of the second system from the equations of the first system, take into
account the symmetry of the tensor Te and obtain the relation
1 2
ðe1  e2 Þ n i n i ¼ 0: ð2:41Þ
This condition is fulfilled in the case of a double root
e1 ¼ e2 ; ð2:410 Þ

the corresponding principal axis M3 is an axis of rotation, while the principal


directions M1; M2 can be whatever in a plane normal to M3. If e1 6¼ e2 , then we
shall have the relation
1 2
n i n i ¼ 0; ð2:4100 Þ

which shows that the principal directions M1 and M2 are orthogonal to each other.
Therefore, all three principal directions are orthogonal with one another and form a
trirectangular trihedron.
Let us admit that two of the roots of the Eq. (2.40) are complex conjugate
e1 ¼ A þ iB; e2 ¼ A  iB;

it results
e1  e2 ¼ 2iB:
Introducing in the system (2.38), we get for the direction cosines complex values
too, i.e.
1 2
n j ¼ aj þ ibj ; n j ¼ aj  ibj ; j ¼ 1; 2; 3:

The condition (2.41) takes the form


2iBðaj aj þ bj bj Þ ¼ 0;

that cannot be fulfilled, except when B ¼ 0; it follows that the roots of the
Eq. (2.40) are always real.
Since the principal directions at a point do not depend on the system of axes
Mx1 x2 x3 , from which we started, the coefficients (2.400 )–(2.40000 ) are three
invariants which play an important rôle in the geometry of the deformation (the
invariants of the strain tensor Te ).
Referring now to the principal axes determined above, the formulae (2.34),
(2.340 ) read
en ¼ e1 n21 þ e2 n22 þ e3 n23 ; ð2:42Þ

cnm ¼ 2ðe1 n1 m1 þ e2 n2 m2 þ e3 n3 m3 Þ: ð2:420 Þ


2.2 Infinitesimal Deformations 49

It is easy to show that the angular strains at the point M possess extreme values
for the planes bisecting the principal dihedrals 2M3, 3M1, 1M2; these values are
supplied by
c1 ¼ ðe2  e3 Þ; c2 ¼ ðe3  e1 Þ; c3 ¼ ðe1  e2 Þ; ð2:43Þ
while the corresponding linear strains will be
1 1 1
e1 ¼ ðe2 þ e3 Þ; e2 ¼ ðe3 þ e1 Þ; e3 ¼ ðe1 þ e2 Þ: ð2:430 Þ
2 2 2

2.2.1.2 Spherical Tensor. Deformation Deviator. Octahedral Strains

It is useful to decompose the tensor Te in a sum of two tensors: the spherical tensor
Te0 of components e0 dij , where
1 1
e0 ¼ eii ¼ I1 ; ð2:44Þ
3 3
and the deformation deviator tensor Te0 , defined in the form
e0ij ¼ eij  e0 dij ; i; j ¼ 1; 2; 3; ð2:45Þ

which is symmetric too.


The maximal and minimal values of the normal components of this tensor are
given by the equation
det½e0ij  e0 dij  ¼ 0; ð2:46Þ

which is of the form


e03  I20 e0  I30 ¼ 0: ð2:460 Þ
The first invariant of the deformation deviator vanishes
I10 ¼ e0ii ¼ e011 þ e022 þ e033 ¼ 0; ð2:47Þ

while the other two invariants can be expressed in the form


1
I20 ¼ e0ij e0ij ¼ 3e20  I2 ; ð2:470 Þ
2
1
I30 ¼ e0ij e0jk e0ki ¼ I3 þ I20 e0  e30 ; ð2:4700 Þ
3
with the aid of the invariants of the tensor Te .
50 2 Geometry and Kinematics of Deformation

0
We notice that for the invariant I2 one obtains
I20 ¼ ðe022 e033 þ e033 e011 þ e011 e022 Þ þ e223 þ e231 þ e212
1
¼ ðe02 þ e02 02 2 2
22 þ e33 Þ þ e23 þ e31 þ e12
2
2 11
1
¼ ½ðe22  e33 Þ2 þ ðe33  e11 Þ2 þ ðe11  e22 Þ2  þ e223 þ e231 þ e212
6
1
¼ ½ðe2  e3 Þ2 þ ðe3  e1 Þ2 þ ðe1  e2 Þ2 
6
1 1
¼ ðc21 þ c22 þ c23 Þ ¼ ðe02 þ e02 02 0 0 0 0 0 0
2 þ e3 Þ ¼ ðe2 e3 þ e3 e1 þ e1 e2 Þ; ð2:48Þ
6 2 1
0
the invariant I3 being given by
1
I30 ¼ ð2e1  e2  e3 Þð2e2  e3  e1 Þð2e3  e1  e2 Þ
27
8 1
¼ ðe1  e1 Þðe2  e2 Þðe3  e3 Þ ¼ ðe03 þ e03 03 0 0 0
2 þ e3 Þ ¼ e1 e2 e3 : ð2:480 Þ
27 3 1
In case of the bodies with plastic properties, an important rôle is played by the
equivalent strain (called intensity of deformations too), defined in the form
qffiffiffiffi
2 0
e ¼ pffiffiffi I2 : ð2:49Þ
3
Let us consider an octahedron bounded by eight planes equally inclined on the
three principal axes (four of the planes bounding the octahedron are put in evi-
dence in Fig. 2.5); let n be the unit vector of the external normal to one of these
planes, of direction cosines (with respect to the principal axes) given by
1
n21 ¼ n22 ¼ n23 ¼ :
3

Fig. 2.5 Octahedral strains x3

O x2

x1
2.2 Infinitesimal Deformations 51

The corresponding linear strain, given by the formula (2.42), is equal to e0 ; it


will be called octahedral linear strain.
The formula (2.420 ) shows that the angular strain in the octahedral plane is
given by
2
cnm ¼ pffiffiffi ðe1 m1 þ e2 m2 þ e3 m3 Þ
3
if we are in the first octant; the direction cosines m1 ; m2 ; m3 satisfy the conditions

m21 þ m22 þ m23 ¼ 1; m1 þ m2 þ m3 ¼ 0;

the last one being the orthogonality condition of the unit vectors n and m.
Searching the direction m corresponding to the maximal singular strain, we obtain
for that one
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 2
c0 ¼ ðe2  e3 Þ2 þ ðe3  e1 Þ2 þ ðe1  e2 Þ2 ¼ c21 þ c22 þ c23 ; ð2:50Þ
3 3
it is called octahedral angular strain. We get
pffiffiffi
c0 ¼ 2e: ð2:500 Þ
Taking into account (2.49), (2.500 ), we notice that, sometimes, it is useful to
express the invariant I20 in the form
3
I20 ¼ c20 ; ð2:51Þ
8
as well, for the invariant I2 we get
 
1
I2 ¼ 3 e20  c20 ; ð2:510 Þ
8

by means of the octahedral strains.

2.2.1.3 Relations Between Displacements and Strains

In the case of infinitesimal deformations, the formulae (2.100 ) or (2.140 ) lead to


linear relations between displacements and strains, by neglecting the terms of
second order; the Cauchy equations then read
1
eij ¼ ðui; j þ uj; i Þ ¼ uði; jÞ : ð2:52Þ
2
52 2 Geometry and Kinematics of Deformation

Fig. 2.6 Gradient of x3 u2 u2,3 dx3


displacement
P*

u3 + u3,3 dx3
α32

P N*

u3,2 dx2
M* α23

dx3
u3

u3
N
M ( x1, x2, x3) u2
dx2 u2 + u2,2 dx2

O x2

Shortly, one can write


Te ¼ def u ð2:520 Þ

where ‘‘def’’ is the differential operator used above; in an expanded form, we have
e11 ¼ u1; 1 ; e22 ¼ u2; 2 ; e33 ¼ u3; 3 ; ð2:53Þ

c23 ¼ u2; 3 þ u3; 2 ; c31 ¼ u3; 1 þ u1; 3 ; c12 ¼ u1; 2 þ u2; 1 : ð2:530 Þ
These results can be illustrated by studying the deformation of a right angle
NMP, the sides of which are parallel to the axes of co-ordinates, in a plane
x1 ¼ const (Fig. 2.6); we admit that the deformation does not depend on x1 and
occurs in this plane only, which means that we are in a linear case. After defor-
mation, this element becomes N  M  P :
Thus, we throw into relief the linear strains e22 ; e33 , as well as the angles
! ! ! !
(formed by M  N  and M  P with MN and MP , respectively, and measured from
the segments of a line in their undeformed state)
a23 ¼ u3; 2 ; a32 ¼ u2; 3 ; ð2:5300 Þ

which leads to shearing strains. With remark, for instance, that a23 represents the
sliding we respect to each other of two straight line elements parallel to the
Ox3 -axis; it results, immediately, the angular strain (taking into account the sign
convention adopted in Sect. 2.2.3)
c23 ¼ a23 þ a32 : ð2:53000 Þ
2.2 Infinitesimal Deformations 53

2.2.1.4 Rigid Body Local Rotation

We introduce the skew-symmetric tensor of second order


2 3
0 x12 x31
Tx  ½xij  ¼ 4 x12 0 x23 5; ð2:54Þ
x31 x23 0

where the components


xij ¼ xij ðx1 ; x2 ; x3 ; tÞ; i; j ¼ 1; 2; 3; ð2:55Þ

are given by
1
xij ¼ ðuj; i  ui; j Þ ¼ u½j; i ; i; j ¼ 1; 2; 3; ð2:56Þ
2
in a developed form, we can write
1 1 1
x23 ¼ ðu3; 2  u2; 3 Þ; x31 ¼ ðu1; 3  u3; 1 Þ; x12 ¼ ðu2; 1  u1; 2 Þ: ð2:560 Þ
2 2 2
To the tensor Tx we attach the axial vector x, the components of which are
given by
1
xi ¼ ijk xjk ; i ¼ 1; 2; 3; ð2:57Þ
2
conversely, we obtain univocally
xij ¼ ijk xk ; i; j ¼ 1; 2; 3; ð2:570 Þ

as it can be easily seen. Taking into account (2.56), we can also write
1 1 1
xk ¼ ijk u½j; i ¼ ijk ðu½j; i þ uðj; iÞ Þ ¼ ijk uj; i ; k ¼ 1; 2; 3;
2 2 2
so that
1
x ¼ curl u; ð2:5700 Þ
2
where x is the rigid body local rotation vector, while u is the displacement vector.
As well, Tx is the corresponding rigid body local rotation tensor (the justification
for these denominations will be seen further).
We notice that we can write the relations (2.7) in the form
dui ¼ aji dxj ð2:58Þ
54 2 Geometry and Kinematics of Deformation

too, putting in evidence an asymmetric tensor of second order Ta of components


aij ¼ uj; i ; ð2:59Þ

which corresponds to the gradient of the displacement u.


It follows that
Ta ¼ Grad u ¼ Te þ Tx ð2:60Þ

or
aij ¼ eij þ xij ; eij ¼ aði; jÞ ; xij ¼ a½i; j ; i; j ¼ 1; 2; 3; ð2:600 Þ

the symmetric part of this tensor being the strain tensor Te , while its antisymmetric
part is the rigid body local rotation tensor Tx .
The relations (2.800 ) and (2.1200 ) show that, if ds ¼ ds, then the tensors TL and
TE (hence the strain tensor Te too) vanish; the reciprocal is, as well, true. In other
words, the annulment of the tensor Te is the necessary and sufficient condition to
have a rigid body motion of the body in the neighbourhood of the point where this
tensor vanishes. In this case, the displacement gradient remains only with its
antisymmetric part Tx ; we can write
dui ¼ xji dxj ¼ kji xk dxj ; i ¼ 1; 2; 3; ð2:61Þ

or, in a vector form,


du ¼ x dr: ð2:610 Þ
This is just the rigid body (local) motion of a neighbourhood of the point
Mðx1 ; x2 ; x3 Þ of position vector r, due to an infinitesimal rotation jxj about an axis
along the direction x, which passes through the point M. The significance of the
tensor Tx is well specified. We mention that the rotation is ‘‘local’’ to can put in
evidence the fact that the motion is of rigid body only in a vicinity of the con-
sidered point.
From the formulae (2.60), (2.600 ) one sees that the hypothesis of neglecting the
displacement gradient with respect to unity works.
Using Fig. 2.6, we obtain (we take into account the signs of the angles a23 and a32 )
1
x23 ¼ ða32  a23 Þ:
2
!
Let be, in the most general case, the vector V ¼ MN before deformation,
!
which—after deformation—becomes V ¼ M  N  (Fig. 2.1b). The variation of this
vector dV ¼ V  V ¼ du is given by relations which generalize the relations
(2.58), i.e.
dVi ¼ aji Vj ; ð2:62Þ

the relations are valid in the case of infinitesimal deformations.


2.2 Infinitesimal Deformations 55

2.2.1.5 Volume Strain. Incompressible Deformations. Incompressible


Body

The formulae (2.25), (2.250 ) give the volume strain, with respect to the volume
before or after deformation, in material or space co-ordinates, respectively,
assuming that in planes parallel to the co-ordinate axes one has not angular strains.
We shall make some observations concerning these formulae.
The formula (2.19) shows that the directions for which e11 ; e22 ; e33 have extreme
values coincide with the directions for which e11 ; e22 ; e33 have a similar property;
but the latter quantities are the components of a symmetric tensor of second order.
It results that there exist three directions orthogonal one to each other for which the
linear strains have extreme values. The quantities e23 ; e31 ; e12 vanish for these
directions; the formula (2.21) shows that also the angular strains vanish for these
directions. Hence, the directions for which the linear strains have extreme values
specify, two by two, planes in which the angular strains vanish and reciprocally.
Thus, there exists a frame, characterized by three directions orthogonal one to each
other, which has only a rigid solid motion during the phenomenon of deformation;
we can also say that, at each point of the solid, there exists an elementary paral-
lelepiped in the initial state, which remains parallelepiped in the actual state.
The formulae (2.22) and (2.23) show, by an analogous reasoning, that—at each
point of the solid—there exists an elementary parallelepiped in the actual state,
which originates from an elementary parallelepiped in the initial state too. Obvi-
ously, the elementary parallelepiped which has this property is the same, either we
use material or space co-ordinates. It follows that the linear strains e11 ; e22 ; e33 and
e11 ; e22 ; e33 , respectively, have extreme values for the same three directions
orthogonal one to each other.
From the above considerations it results that the formulae (2.52) and (2.520 )
give the volume cubical strain expressed by means of the principal linear strains
(which have extreme values), We can thus write
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ev ¼ 1 þ 2I1 þ 4I2 þ 8I3  1; ð2:63Þ
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ev ¼ 1  1  2I1 þ 4I2  8I3 ; ð2:630 Þ

where I1 ; I2 ; I3 and I1 ; I2 ; I3 are the invariants of the tensor TL and of the tensor TC ,
respectively. We notice that, squaring the relation (2.25), we get

1 þ 2e11 2e12 2e13

ð1 þ ev Þ ¼ 2e21 1 þ 2e22 2e23 ;
2e31 2e32 1 þ 2e33

wherefrom it results the relation (2.63); analogously, we find again the relation
(2.630 ).
56 2 Geometry and Kinematics of Deformation

The relations (2.25), (2.250 ) or the relations (2.63), (2.630 ) show that, in case of
infinitesimal deformations, the volume strain will be given by
h ¼ ev ¼ e1 þ e2 þ e3 ¼ eii ¼ I1 : ð2:64Þ
Taking into account (2.52), we can also write
h ¼ ui;i : ð2:640 Þ
If we have
h¼0 ð2:65Þ
for a certain solid and for certain external loads, then we say that we have to do
with an incompressible state of deformation (called isochore state of deformation
too); if the relation (2.65) takes place for any external loads, we say that we have to
do with an incompressible solid (which is a physical property of the respective
solid).
0
We notice that in case of the deformations deviator, we have I1 ¼ 0. Thus, the
respective deviator leads to an incompressible state of deformation; otherwise, the
spheric tensor leads to a homogeneous deformation (without a change of the form,
but with a change of the volume).

2.2.2 Continuity Equations. Computation


of Displacements and Rotations

In what follows, we get the equations of continuity of deformations in case of


infinitesimal deformations. At the same time, we compute the components of the
displacement vector as functions of the components of the tensor Te .

2.2.2.1 Continuity Equations

We notice that we can consider the relations (2.52) between strains and
displacements as a system of 6 equations with 3 unknown functions ui ; i ¼ 1; 2; 3.
This system is compatible if the 6 distinct components of the tensor Te verify
certain condition of compatibility.
Starting from the relations (2.52) and (2.56), we can write the components of
the displacement gradient in the form
ui; j ¼ eij  xij ; i; j ¼ 1; 2; 3;

assuming that the displacements are functions at least of class C2 , the necessary
and sufficient conditions of compatibility are obtained equating the mixed deriv-
atives of second order of the functions ui . We obtain thus
2.2 Infinitesimal Deformations 57

eij; k  eik; j ¼ xij; k  xik; j ; i; j; k ¼ 1; 2; 3; ð2:66Þ

relation which can be written also in the form


jkl ðeij; k  xij; k Þ ¼ 0; i; l ¼ 1; 2; 3: ð2:660 Þ
By circular permutation of the indices i; j; k in the relation (2.66), we can write
ejk; i  eji;k ¼ xjk; i  xji;k ; i; j; k ¼ 1; 2; 3;
eki; j  ekj; i ¼ xki; j  xkj; i ; i; j; k ¼ 1; 2; 3;

summing in both members and taking into account the properties of symmetry and
of antisymmetry of the tensors Te and Tx , respectively, it results
xij; k þ xjk; i þ xki; j ¼ 0; i; j; k ¼ 1; 2; 3: ð2:67Þ

In this case, from the relation (2.66), we obtain


xkj; i ¼ eij; k  eik; j ; i; j; k ¼ 1; 2; 3: ð2:6600 Þ

Assuming, analogically, that the rigid body local rotations are functions at least of
the class C2 , the necessary and sufficient conditions of compatibility, expressed
only by means of the strains, are obtained equating the mixed derivatives of second
order of the functions xkj . We get thus

eij; kl þ ekl;ij ¼ eik; jl þ ejl;ik ; i; j; k; l ¼ 1; 2; 3: ð2:68Þ


These relations can be written also in the form
ilm ðeij; kl  eik; jl Þ ¼ 0; m ¼ 1; 2; 3;

or in the form
ilm jkn eij; kl ¼ 0; m; n ¼ 1; 2; 3; ð2:680 Þ

introducing the antisymmetric part of a tensor, we get


o½l ei½j; k ¼ 0: ð2:6800 Þ

In a tensor form, we have


Ink Te ¼ 0; ð2:68000 Þ

where ‘‘Ink’’ (incompatibility) corresponds to the differential operator introduced


above [4].
The relations (2.68), (2.680 ) or (2.6800 ) represent the conditions of compatibility
(of continuity) of B. de Saint-Venant [3].
Taking into account the symmetry of the tensor eij; kl with respect to the indices i
and j or k and l, respectively, we can write
58 2 Geometry and Kinematics of Deformation

* *
N1 N2
P N R P* R*

M*
Q M S Q* S*
(a) (b)

P*
P* R*

* R*
N2 N 1
*
N
*

Q* S* Q*

S*
M*
M*
(c) (d)

Fig. 2.7 Two neighbouring elements: undeformed state (a), separation (b), interpenetration (c),
continuity (d)

iln jkm eij; kl ¼ jkn ilm eji;lk ¼ ilm jkn eij; kl ; m; n ¼ 1; 2; 3;

hence the tensor of second order from the first member of the relation (2.680 ) is
symmetric with respect to the indices m and n; this tensor will have 6 distinct
components.
Hence, we can write six distinct equations of compatibility; they are of the form

e22; 33 þ e33; 22 ¼ 2e23; 23 ¼ c23; 23 ;


e33; 11 þ e11; 33 ¼ 2e31; 31 ¼ c31; 31 ; ð2:69Þ
e11; 22 þ e22; 11 ¼ 2e12; 12 ¼ c12; 12 ;

1
ðe23;1 þ e31;2 þ e12;3 Þ;1 ¼ ðc23;1 þ c31;2 þ c12;3 Þ;1 ¼ e11;23 ;
2
1
ðe31;2 þ e12;3 þ e23;1 Þ;2 ¼ ðc31;2 þ c12;3 þ c23;1 Þ;2 ¼ e22;31 ; ð2:690 Þ
2
1
ðe12;3 þ e23;1 þ e31;2 Þ;3 ¼ ðc12;3 þ c23;1 þ c31;2 Þ;3 ¼ e33;12 :
2
These conditions of compatibility have a particularly important physical
interpretations. Indeed, let be two neighbouring elements of the body, at the
beginning in a undeformed state (Fig. 2.7a). If the conditions of compatibility
would not be fulfilled, then the two elements would be separated after deformation
(Fig. 2.7b) or would be interpenetrated (Fig. 2.7c); if these conditions are fulfilled,
2.2 Infinitesimal Deformations 59

then the elements remain glued one to the other (Fig. 2.7d). Hence, the conditions
(2.68) or (2.680 ) or (2.69), (2.690 ) are, from a mechanical point of view, necessary
and sufficient conditions of continuity of the deformation for a simply connected
domain.

2.2.2.2 Cesàro’s Formulae

Using the results in the preceding subsection and placing ourselves in the case of a
simply connected domain, we can compute the displacements at an arbitrary point
P, supposing we know them at a point P0 (Fig. 2.8). Thus, starting from the
derivatives of the first order, we get

where we assumed that ui ; eij ; xij depend on the integration variables n1 ; n2 ; n3 ; but

Using the relation (2.6600 ), we get Cesàro’s formulae [2] in the form

ð2:70Þ

Analogically, one can show that


ð2:700 Þ

The conditions that the above integrals do not depend on the path, in case of a
simply connected domain, are just the continuity equations established in the
previous subsection.

Fig. 2.8 Calculation of the


displacement
P
0
P
D
60 2 Geometry and Kinematics of Deformation

2.2.2.3 Rigid Body Motion

In the particular case in which we assume that all the components of the tensor Te
vanish, the formula (2.70) is reduced to

ui ¼ u0i  ðxj  x0j Þx0ij ; i ¼ 1; 2; 3; ð2:71Þ

in an expanded form, we have

u1 ¼ u01  ðx2  x02 Þx03 þ ðx3  x03 Þx02 ;


u2 ¼ u02  ðx3  x03 Þx01 þ ðx1  x01 Þx03 ; ð2:710 Þ
u3 ¼ u03  ðx1  x01 Þx02 þ ðx2  x02 Þx01 ;

where we took into account the relation (2.570 ). This corresponds to a rigid body
motion, characterized by the displacement vector u0 and by the rigid body local
rotation vector x0 :
We notice, as well, that the formula (2.71) gives the general integral of the
system of Eq. (2.52), considered to be homogeneous (with vanishing components
of the tensor Te ).
We notice also that two states of displacement, to which corresponds the same
state of deformation, differ by a rigid body motion; this is the theorem of
G. R. Kirchhoff.
If we take the origin of the co-ordinate axes at the point P0 , then the rigid body
motion will be given by

ui ¼ u0i  xj x0ij ¼ u0i  ijk xj x0k ; i ¼ 1; 2; 3; ð2:7100 Þ

where u0i ; x0i ; i ¼ 1; 2; 3, are the displacements and the rigid local rotations,
respectively, corresponding to a zero tensor Te . In a vector form, we have
!
u ¼ u0 þ x P0 P: ð2:71000 Þ

2.2.3 Elementary States of Deformation

In the following, we present some elementary states of deformation which appear


frequently in applications, i.e.: uniform dilatation or contraction, extension or
shortening and simple shear. As well, we shall give some data concerning the case
of a plane state of strain.

2.2.3.1 Uniform Dilatation

We call uniform dilatation that state of deformation for which


2.2 Infinitesimal Deformations 61

e11 ¼ e22 ¼ e33 ¼ e [ 0; ð2:72Þ

e23 ¼ e31 ¼ e12 ¼ 0; ð2:720 Þ

if e\0, then we have to do with a uniform contraction.


Cauchy’s quadric is a sphere of equation
0 0 0 1
x12 þ x22 þ x32 ¼  ; ð2:73Þ
e
all the directions corresponding to principal axes.
The state of displacement is given by
ui ¼ exi ; i ¼ 1; 2; 3; ð2:74Þ

assuming that the rigid body displacement and local rotation vanish at the origin of
the co-ordinate axes, considered to be the centre of the phenomenon of dilatation.
We also have
xij ¼ ex½j; i ¼ 0; i; j ¼ 1; 2; 3: ð2:75Þ

The volume strain is also constant


ev ¼ ui;i ¼ exi;i ¼ 3e: ð2:76Þ

2.2.3.2 Simple Extension

We say to have a simple extension if the state of deformation is reduced to a linear


strain e1 ¼ e [ 0 along the direction of a principal axis (e.g., the O1-axis, which is
1
specified by the direction cosines n i ; i ¼ 1; 2; 3), with regard to an orthonormed
frame Oxi ; if e\0, then we have to do with a simple shortening. The corre-
sponding Cauchy quadric is given by the equation
0
ex12 ¼ 1 ð2:77Þ

and is reduced to two parallel planes.


With respect to a system of axes Oxi , we can write the state of strain in the form
1 1
eij ¼ eni nj ; i; j ¼ 1; 2; 3; ð2:78Þ

the state of displacement being given by


1 1
ui ¼ eni nk xk ; i ¼ 1; 2; 3; ð2:79Þ

where we assumed that u0 ¼ 0; x0 ¼ 0 at the origin of the co-ordinate axes.


62 2 Geometry and Kinematics of Deformation

We notice that the displacement gradient tensor


1 1
aij ¼ eni nj ; i; j ¼ 1; 2; 3;

is a symmetric one, so that


xij ¼ 0; i; j ¼ 1; 2; 3: ð2:80Þ
The volume strain is given by
ev ¼ e: ð2:81Þ

2.2.3.3 Simple Angular Strain

If the strain tensor has only one non-zero component, which is with different
indices, e.g., e12 ¼ e21 ¼ c=2 (Fig. 2.9a), then we have to do with a simple angular
strain, Cauchy’s quadric being a cylinder with an equilateral hyperbola directrix,
of equation
cx1 x2 ¼ 1; ð2:82Þ

the angular strain takes place in the Ox1 x2 -plane.


The state of displacement is given by
1 1
u1 ¼ cx2 ; u2 ¼ cx1 ; u3 ¼ 0; ð2:83Þ
2 2
assuming that at the origin u0 ¼ 0; x0 ¼ 0.
The rigid body local rotation vanishes
x ¼ 0: ð2:84Þ
By a rotation of p=4 in the Ox1 x2 -plane, one can express the quadric with
respect to the principal axes; so we get
1 02 0
cðx1  x22 Þ ¼ 1; ð2:820 Þ
2
hence, in case of a simple angular strain in the Ox1 x2 -plane, the principal linear
strains are (Fig. 2.9b)
1
e1 ¼ e2 ¼ c; e3 ¼ 0; ð2:85Þ
2
the reciprocal being true too.
The volume strain vanishes
ev ¼ 0: ð2:86Þ
2.2 Infinitesimal Deformations 63

x2 P P*

P* γ N*
Q 2 N
P Q
Q* Q*
N* 2 1
M
N M* M
M*
π
4
x1 O x1
O
(a) (b)

Fig. 2.9 Simple angular strain: in the Ox1 x2 -plane (a), in the O12-plane (b)

2.2.3.4 Plane State of Strain

We call plane state of strain that state of strain for which


e3i ¼ 0; i ¼ 1; 2; 3: ð2:87Þ
Cauchy’s quadric becomes

e11 x21 þ 2e12 x1 x2 þ e22 x22 ¼ 1 ð2:88Þ

or
0 0
e1 x12 þ e2 x22 ¼ 1: ð2:880 Þ
This leads to the state of displacement
ui ¼ ui ðx1 ; x2 Þ; i ¼ 1; 2; u3 ¼ 0: ð2:89Þ
The rigid body local rotation vector remains with only one non-zero component
1
x3 ¼ ðu2;1  u1; 2 Þ; ð2:90Þ
2
while the volume strain is given by
ev ¼ u1; 1 þ u2; 2 : ð2:91Þ

2.2.4 Displacements and Strains in Curvilinear Co-ordinates

It is often convenient to use curvilinear co-ordinates; taking into account the


particular form of the considered domains, these co-ordinates may facilitate the
statement of the boundary conditions.
64 2 Geometry and Kinematics of Deformation

For instance, in case of an elastic parallelepiped and in case of the infinite


domains derived from it by throwing to infinity one or more of its faces, we can
use the orthogonal Cartesian co-ordinates; on the contour of the respective domain
we have x1 ¼ const or x2 ¼ const or x3 ¼ const, the boundary conditions being
expressed in a simpler form. Likewise, in case of the oblique elastic parallelepiped
and in case of infinite domains derived from it, we use oblique Cartesian
co-ordinates. The axial symmetrical problems lead to the use of the cylindrical
co-ordinates r; h; z; the boundary conditions in case of an elastic circular cylinder
can therefore be written for r ¼ const or h ¼ const or z ¼ const. Likewise, in case
of problems with central symmetry, spherical (space polar) co-ordinates R; u; h are
used; e.g., in case of an elastic sphere, the boundary conditions are considered for
R ¼ const.
The above remarks lead to the idea of using, in general, any curvilinear
co-ordinates qa ; a ¼ 1; 2; 3, chosen as function of the given domain, so that the
frontier of it be formed by surfaces of co-ordinates; in such a case, the boundary
conditions refer to q1 ¼ const or q2 ¼ const or q3 ¼ const. The study of such a
system of co-ordinates was first undertaken by C. F. Gauss. But this theory was
developed systematically by G. Lamé [5, 14], who applied it to problems of
mathematical physics, in general, and to problems of the theory of elasticity, in
particular.
Studying the three-dimensional general problems, G. Lamé supplied formulae for
the extensions, as well as for the rotations, in orthogonal curvilinear co-ordinates.
C. W. Borchardt determined the extension by another way, W. Thomson studied the
problem of cubical dilatation, while E. Cesàro studied again the formulae concerning
rotations.
In what follows we present, in curvilinear co-ordinates, some of the results
previously obtained; we will use thus orthogonal curvilinear co-ordinates and, in
particular, cylindrical and spherical co-ordinates, as well as the corresponding
mathematical results in Sect. A.2.

2.2.4.1 Orthogonal Curvilinear Co-ordinates

Let be a point M, the position of which is specified by the orthogonal Cartesian


co-ordinates xi ; i ¼ 1; 2; 3, or by the orthogonal curvilinear co-ordinates qa ; a ¼
1; 2; 3 (by convention we shall use Greek indices for the curvilinear co-ordinates,
as well as for the components of various quantities in these co-ordinates).

!
The displacement vector MM  ¼ u (Fig. 2.10) has the components
ui ; i ¼ 1; 2; 3, or the components usa ; a ¼ 1; 2; 3.
Introducing the direction cosines of the directions qa , we can easily write the
relation
oxi
usa ¼ ui cosðqa ; xi Þ ¼ ui ; a ¼ 1; 2; 3; ð2:92Þ
osa
2.2 Infinitesimal Deformations 65

as well as the relations


oxi
ui ¼ usa cosðqa ; xi Þ ¼ us ; i ¼ 1; 2; 3: ð2:920 Þ
osa a
On the other hand, the passing from orthogonal Cartesian co-ordinates to
orthogonal curvilinear co-ordinates for strains can be made, starting from the
formulae (2.34), (2.340 ), by means of the relations
oxi oxj
esa sb ¼ eij cosðqa ; xi Þ cosðqb ; xj Þ ¼ eij ; a; b ¼ 1; 2; 3: ð2:93Þ
osa osb
If we take now into account (2.52) too, then we notice that
 
1 oxj ouj oxi ouj
eij ¼ þ ; i; j ¼ 1; 2; 3;
2 osc osc osc osc

replacing in the relations (2.93), using the formulae (2.920 ) and by the help of the
formula (A.134), we get
 
1 ousa ousb
es a s b ¼ þ
2 osb osa
 
1 oxi o2 xi oxi o2 xi
þ þ usc ; a; b ¼ 1; 2; 3: ð2:94Þ
2 osa osb osc osb osa osc
Taking into account the formula (A.1350 ), we may write (without summation
with respect to a:b ¼ 1; 2; 3)
 
1 ousa ousb
esa sb ¼ þ
2 osb osa
X3     
1 1 ohc ohc 1 oha 1 ohb
þ dac þ dbc  dab þ usc ð2:940 Þ
2 c¼1 hc osb osa ha osc hb osc

Fig. 2.10 Orthogonal x3


curvilinear co-ordinates q3
M*
q2
us q1
u 3
us
2 us
1

M
r

O x2

x1
66 2 Geometry and Kinematics of Deformation

too, where we use Kronecker’s tensor.


In particular, we get (without summation for a:b; c ¼ 1; 2; 3)
 
ousa 1 oha oha
es a s a ¼  us þ us ; a 6¼ b 6¼ c 6¼ a;
osa ha osb b osc c
ð2:95Þ
ousb ousc 1 ohb 1 ohc
cs b s c ¼ þ þ us b þ us ; a 6¼ b 6¼ c 6¼ a;
osc osb hb osc hc osb c

these relations generalize the formulae (2.52), (2.53), (2.530 ) of Cauchy.


The volume strain is given by (without summation with respect to b and c)
 
o ousa
ev ¼ hb hc ; a 6¼ b 6¼ c 6¼ a: ð2:96Þ
osa hb hc

2.2.4.2 Cylindrical Co-ordinates

We can write the relations to pass from the displacements in orthogonal Cartesian
co-ordinates to those in cylindrical co-ordinates in the form

ur ¼ u1 cos h þ u2 sin h;
uh ¼ u1 sin h þ u2 cos h; ð2:97Þ
u z ¼ u3

and inversely

u1 ¼ ur cos h  uh sin h;
u1 ¼ ur sin h þ uh cos h; ð2:970 Þ
u3 ¼ uz :

Applying the results in the preceding subsection, there result the relations between
strains and displacements
our
err ¼ ;
or
1 ouh 1 ouh 1
ehh ¼ þ ur ¼ þ ur ; ð2:98Þ
r oh r osh r
ouz
ezz ¼ ;
oz
2.2 Infinitesimal Deformations 67

ouh 1 ouz ouh ouz


chz ¼ þ ¼ þ ;
oz r oh oz osh
ouz our
czr ¼ þ ; ð2:980 Þ
or oz
1 our ouh 1 our ouh 1
crh ¼ þ  uh ¼ þ  uh :
r oh or r osh or r
The volume strain is given by
our ouh ouz 1
ev ¼ þ þ þ ur : ð2:99Þ
or osh oz r

2.2.4.3 Spherical Co-ordinates

The relations to pass from the displacements in orthogonal Cartesian co-ordinates


to those in spherical co-ordinates read

uR ¼ u1 sin u cos h þ u2 sin u sin h þ u3 cos u ¼ ur sin u þ uz cos u;


uu ¼ u1 cos u cos h þ u2 cos u sin h  u3 sin u ¼ ur cos u  uz sin u; ð2:100Þ
uh ¼ u1 sin h þ u2 cos h;

while those which make the passing from the spherical co-ordinates to the
orthogonal Cartesian ones are

u1 ¼ uR sin u cos h þ uu cos u cos h  uh sin h;


u2 ¼ uR sin u sin h þ uu cos u sin h þ uh cos h; ð2:1000 Þ
u3 ¼ uR cos u  uu sin u;

where we have put in evidence the linkage to the cylindrical co-ordinates.


The relations between strains and displacements will be of the form
ouR 1 ouu 1 ouR 1
eRR ¼ ; euu ¼ þ uR ¼ þ uR ; ð2:101Þ
oR R ou R osu R

1 ouh 1 cot u ouh 1 cot u


ehh ¼ þ uR þ uu ¼ þ uR þ uu ;
R sin u oh R R osh R R
1 ouu 1 ouh cot u ouh ouh cot u
cuh ¼ þ  uh ¼ þ þ uh ;
R sin u oh R ou R osh osu R
ð2:1010 Þ
ouh 1 ouR 1 ouh ouR 1
chR ¼ þ  uh ¼ þ  uh ;
oR R sin u oh R oR osh R
1 ouR ouu 1 ouR ouu 1
cRu ¼ þ  uu ¼ þ  uu :
R ou oR R osu oR R
68 2 Geometry and Kinematics of Deformation

The volume strain reads


ouR ouu ouh 2 cot u
ev ¼ þ þ þ uR þ uu : ð2:102Þ
oR osu osh R R

2.2.5 Kinematics of Deformation

Since the quantities characterizing the state of displacement and the state of strain
are function of time, a study of the kinematics of deformation becomes absolutely
necessary. We will make thus some general considerations and will analyse the
case of infinitesimal deformations.

2.2.5.1 General Considerations

Taking into account that the velocity is the derivative of the position vector (we
remark that one must consider the position vector in the actual state) with respect
to time (Fig. 2.11)
dr
v¼ ð2:103Þ
dt
and starting from (2.1), where xi are material co-ordinates, independent of time, and
xi ; i ¼ 1; 2; 3, are the space co-ordinates, supplied by (2.5), we can equally write
du ou dxi ou
v¼ ¼ þ : ð2:1030 Þ
dt oxi dt ot

In the same way, one can equally introduce the acceleration

dv d2 r d2 u ov dx ov
a¼ ¼ 2 ¼ 2 ¼  i þ : ð2:104Þ
dt dt dt oxi dt ot

Fig. 2.11 Velocity M* v


u
M
r*
r

O
2.2 Infinitesimal Deformations 69

2.2.5.2 Case of Infinitesimal Deformations

In the case of infinitesimal strains and rotations, we shall neglect the non-linear
terms and so we obtain the displacement velocity
ou
v¼ ¼ u;
_ ð2:105Þ
ot
where the partial derivatives with respect to time shall be marked by a point; thus,
the components of the displacement velocity are
vi ¼ u_ i ¼ u_ i ðx1 ; x2 ; x3 ; tÞ; i ¼ 1; 2; 3: ð2:1050 Þ
Neglecting the space derivative, thesubstantial derivative reduces itself to the time
derivative. Besides, we remark that, in the case of infinitesimal deformations, the
relations of the form (2.29) are valid and no distinction is to be made between
material co-ordinates and space ones. Therefore, we shall use only material co-
ordinates and we shall write relations such as (2.1050 ); likewise, in case of strains
and rigid body local rotations, we shall use the relations (2.33) and (2.55).
Concerning the displacement acceleration, we shall also have

ov o2 u
a¼ ¼ v_ ¼ 2 ¼ u
€; ð2:106Þ
ot ot
the components of the displacement acceleration will therefore be
ai ¼ v_ i ¼ €
ui ¼ €
ui ðx1 ; x2 ; x3 ; tÞ; i ¼ 1; 2; 3: ð2:1060 Þ
We can moreover introduce the strain velocities (the Euler tensor of the rate of
strain)
vij ¼ e_ ij ¼ e_ ij ðx1 ; x2 ; x3 ; tÞ; i; j ¼ 1; 2; 3; ð2:107Þ

as well as the rigid body local rotation velocities (theCauchy spin tensor of the
rate of rigid body local rotation)
x_ ij ¼ x_ ij ðx1 ; x2 ; x3 ; tÞ; i; j ¼ 1; 2; 3: ð2:108Þ
Starting from (2.52) and (2.56) and taking into account that we have to deal
with infinitesimal deformations (we use material co-ordinates, independent of
time; we have thus relations of the form o2 f =otox ¼ o2 f =oxot for functions f ¼
f ðx1 ; x2 ; x3 ; tÞ of class C2 ), we can also write the relations
1
e_ ij ¼ ðu_ i; j þ u_ j; i Þ ¼ u_ ði; jÞ ; i; j ¼ 1; 2; 3; ð2:109Þ
2
1
x_ ij ¼ ðu_ j; i  u_ i; j Þ ¼ u_ ½j; i ; i; j ¼ 1; 2; 3: ð2:110Þ
2
70 2 Geometry and Kinematics of Deformation

Obviously, for the tensors Te_ and Tx_ we can proceed to a study similar to that
dealt with in Sects. 3.1.1 and 3.1.4; as well, we can introduce the tensor Ta_
supplied by
Ta_ ¼ Gradu ¼ Te_ þ Tx_ ; ð2:111Þ
where the last two tensors represent its symmetrical and its antisymmetrical part,
respectively.
We notice that, in general, the principal axes of the tensor Te_ do not coincide
with the principal axes of the tensor Te , while the principal linear strain velocities
do not represent the derivatives with respect to time of the principal linear strains.
As well, after Euler’s criterion, if
e_ ij ¼ 0; i; j ¼ 1; 2; 3; ð2:112Þ

then we have to do with a rigid body motion

h_ ¼ vii ¼ e_ ii ¼ u_ i;i ¼ divu_ ¼ divv ¼ 0: ð2:113Þ


If for a certain loading of the solid, we have
h_ ¼ e_ ii ¼ 0 ð2:114Þ
then we say that the motion is isochore and we may write
o
div v ¼ div u_ ¼ ðdiv uÞ ¼ 0; ð2:1140 Þ
ot
hence, the displacement velocities field is solenoidal, so that these ones can be
represented by means of a field of curls
v ¼ u_ ¼ curl U; ð2:11400 Þ

expressed with the aid of the vector field U ¼ Uðx1 ; x2 ; x3 ; tÞ of class C2 .


_ corresponding to the Cauchy spin tensor, is called vorticity
The axial vector x,
vector, its components being given by the relation
1
x_ i ¼ ijk x_ jk ; i ¼ 1; 2; 3; ð2:115Þ
2
we also can write
1 1
x_ ¼ curl u_ ¼ curl v: ð2:1150 Þ
2 2
If
x_ ¼ 0; ð2:116Þ

then we have an irrotational motion; in this case, the field of the displacement
velocities is a field of gradients (the displacement velocities derive from a scalar
potential u ¼ uðx1 ; x2 ; x3 ; tÞ of class C2 ) and we may write
2.2 Infinitesimal Deformations 71

v ¼ u_ ¼ gradu: ð2:117Þ
All the consideration made in Sect. 2.2.2 concerning the continuity of the
deformation and of the rotation and concerning the computation of these quantities
can be repeated, in case of the infinitesimal deformations, for the strain velocities
and for the rotation velocities.
Likewise, in curvilinear co-ordinates, one can use the results supplied in
Sect. 2.2.4.

References

A. Books

1. Borchardt, C.W.: Gesammelte Werhe. Georg Reimer, Berlin (1888)


2. Cesàro, E.: Introduzione alla teoria matematica della elasticità. Fr. Bocca Ed., Torino (1894)
3. Clebsch, A.: Théorie de l’élasticité des corps solides (traduite par MM. Barré de Saint-Venant
et Flamant avec des notes étendues de M. de Saint-Venant), Paris (1883)
4. Kr}oner, E.: Kontinuumsteorie der Versetzungen und Eigenspannungen (Ergebn. der Angew.
Math.). Springer, Berlin (1958)
5. Lamé, G.: Leçons sur les coordonnées curvilignes. Mallet-Bachelier, Paris (1859)
6. Sudria, J.: L’action euclidienne de déformation et de mouvement (Mém. des Sci. Phys.), vol.
XXIX. Gauther-Villars, Paris (1935)
7. Teodorescu, P.P.: Probleme spatßiale in teoria elasticitătßii (Space Problems in the Theory of
Elasticity). Ed. Academiei, Bucuresßti (1970)
8. Truesdell, C., Noll, W.: The non-linear field theories of mechanics. In: Flügge, S. (ed.)
Encyclopedia of Physics, vol. III/3. Springer, Berlin (1965)
9. Truesdell, C., Toupin, R.A.: The classical field theories. In: Flügge, S. (ed.) Encyclopedia of
Physics, vol. III/1. Springer, Berlin (1960)
10. Wang, C.C., Truesdell, C.: Introduction to Rational Elasticity. Noordhoff International
Publication, Leyden (1973)

B. Papers

11. Beltrami, E.: Sur la théorie de la déformation infiniment petite d’un milieu. C. Rend. hebd. de
séance de l’Acad. Sci. 108, 502 (1889)
}
12. Golitsyn, B.: Uber die Dispersion und Dämpfung der seismischen Oberflächwellen. Bull. de
l’Acad. Imp. des Sci. de Saint-Petersburg, ser. 6, 219 (1912)
13. Galletto, D.: Sull’unicità in presenza di vincoli interni di una condizione cinematica
fondamentale nella teoria delle deformazioni finite. Atti dell’Ist. Veneto di Sci. Lett. e Arti,
Cl. di Sci. Mat. e Nat. 123, 197 (1965)
14. Lamé, G.: Mémoire sur les coordonnées curvilignes. J. de Math. Pures et Appl. 5, 313 (1840)
15. Noll, W.: A mathematical theory of mechanical behaviour of continuous media. Arch. Rat.
Mech. Anal. 2, 195 (1958)
16. Truesdell, C.: The mechanical foundations of elasticity and fluid dynamics. J. Rat. Mech.
Anal. 1, 125 (1952)
Chapter 3
Mechanics of Stresses

We shall continue now the theoretical study of the preceding chapter, by putting in
evidence the stresses which arise because of the deformation of a solid body; both
vector and tensor aspects of the stress are thus emphasized [4].

3.1 Stress Vector

So as to emphasize the stresses occuring within a solid body, stresses which can be
considered as variations of the internal forces of cohesion, we shall—in general—
carryout an arbitrary section S (which can be curved), open or closed, through this
body, that is thus divided into two parts (Fig. 3.1a). Applying the principle of
equilibrium of parts (static or dynamic equilibrium), we get the main properties of
the stress vector.

3.1.1 General Considerations

In what follows we make some general considerations concerning superficial


forces (stress vectors) which are introduced in the mathematical study of a
deformable solid and concerning the corresponding state of stress; as well, we will
define the volume forces and d’Alembert’s volume forces, corresponding to
dynamic problems.

3.1.1.1 Superficial Forces

An area element DA in the neighbourhood of the point M, in the plane tangent to


the section S at this point (e.g., on the left side of the body), will thus be acted upon

P. P. Teodorescu, Treatise on Classical Elasticity, Mathematical and Analytical 73


Techniques with Applications to Engineering, DOI: 10.1007/978-94-007-2616-1_3,
Ó Springer Science+Business Media Dordrecht 2013
74 3 Mechanics of Stresses

(a) A
II

(b) t (c) t
n n n
T n p
P

M M
I I
n n
N n x3 r dA n
A

x1 O x2

Fig. 3.1 Stress vector: section A (a), efforts (b), stress-vectors (c)

n n
by the effort D P, of components D N (along the external normal to the section at
n
the given point) and D T (in the tangent plane to the section at the point M), which
makes connection with the remote side (the right side) (Fig. 3.1b). We admit that
n
the effort D P is of the nature of a force. We obtain thus the mean stress vector in
the neighbourhood of the point M of an area element of external normal of unit
vector n, in the form
n
n DP
pmean ¼ : ð3:1Þ
DA
At the limit, admitting its existence, we get (Fig. 3.1c)
n
n n dP
p ¼ lim pmean ¼ ; ð3:2Þ
DA!0 dA
i.e. the stress vector at the point M, on an area element of external normal n; the
n
value p of this vector represents the stress at the point M.
The notion of stress should be introduced with certain precautions of mathe-
n
matical nature, because one cannot know how D P varies when DA ! 0. In regard
n
to the concentrated loads, for instance, such a definition is nonsense (p ! 1); the
theory of distributions plays a quite important rôle in this case.
We mention that the external (outer) normal n is used usually for solid bodies,
while in case of fluids it is used the internal (inner) normal.
3.1 Stress Vector 75

n
The component of the stress vector p along the direction m is usually denoted
n
by pm . Introducing the components along the three axes of co-ordinates, we may
write
n n n n
n n
p2 ¼ pi pi ¼ p21 þ p22 þ p23 : ð3:3Þ
n n
Ordinarily, we denote the component of p along the external normal n by r and
n
we call it normal stress; likewise, the component of p in the plane tangent at the
n
point M to the section surface S is denoted by s and is called tangential stress
n n
(Fig. 3.1c). These components result from D N and D T by a process of passing to
the limit, analogue to that mentioned above; evidently, we can introduce, in the
n n
same manner, the notions of normal stress vector r and tangential stress vector s.
We have
n n n
p ¼ r þ s; ð3:30 Þ
as well as
n n n
p2 ¼ r 2 þ s 2 : ð3:300 Þ
The normal stress is positive if it corresponds to a phenomenon of tension and
negative in the opposite case (phenomenon of compression). We shall admit
moreover that the tangential stress is positive if it corresponds to a positive angular
strain and negative in the contrary case; this fact should however be related to the
chosen system of co-ordinates.
The components of the stress vector will be functions of the form (we use
material co-ordinates, assuming that we are in the case of infinitesimal
deformations)
n n
pi ¼ pi ðx1 ; x2 ; x3 ; tÞ: ð3:4Þ
n
From a dimensional point of view, the stresses are expressed by ½p ¼ ML1 T2 .

3.1.1.2 State of Stress

n
By state of stress of the body at the point M we mean the totality of the stresses p
in all the directions n around that point. The state of stress of the whole body will
be given by the states of stress of all the points that form it.
By one-dimensional or linear state of stress (also called antiplane state of
stress) we shall mean that state of stress in which appear stress vectors only on the
area elements of a normal of fixed direction. This is the case of straight bars of
constant cross section, undergoing simple stresses.
76 3 Mechanics of Stresses

By two-dimensional or plane state of stress we shall mean the state of stress in


which on the area elements the normal of which has a fixed direction no stress
vector appears. This, for instance, is the case of a wall-beam, loaded in its middle
plane.
By three-dimensional or spatial state of stress we mean a state of stress cor-
responding to a general case of stress of any body whatever (that cannot be
reduced to any of the particular cases considered above).

3.1.1.3 Volume Force

We can also conceive a section of a closed-curved surface form whatever, that


would detach a volume element of an arbitrary form, from the rest of the body.
~ can appear, the value of which are in direct proportion
We consider that forces DF
to the volume element DV that we shall call mean value forces (Fig. 3.2a)
~
DF
Fmean ¼ : ð3:5Þ
DV
Analogically, by a limit process, when this exists, we can write (Fig. 3.2b)
~
dF
F ¼ lim Fmean ¼ ; ð3:6Þ
DV!0 dV
where F represents the volume force vector at the point M; the modulus F repre-
sents the volume force at this point. This notion shall be introduced with the same
precautions, from a mathematical standpoint, as the stress notion. From the
dimensional point of view, we have ½F ¼ ML2 T2 .
We denote by Fm the component of the vector F along the unit vector m. If we
introduce the components of this vector along the co-ordinate axes, then we may
write

F 2 ¼ Fi Fi ¼ F12 þ F22 þ F32 ; ð3:7Þ

(a) x3 (b) x3

~ ~
V F dV F
( M) (dM )
r M r M

O x2 O x2
x1 x1

Fig. 3.2 Volume force: on DV (a), on dV (b)


3.1 Stress Vector 77

where
Fi ¼ Fi ðx1 ; x2 ; x3 ; tÞ; i ¼ 1; 2; 3: ð3:8Þ
The component Fi is considered positive if it corresponds to the positive
direction of the co-ordinate axis and negative in the contrary case.
The volume forces can be of various origins; among the most important ones,
we mention the own weight (represented by the weight per unit volume c) and the
forces of inertia.
If DM is the mass corresponding to a volume element DV (Fig. 3.2a), then we
can similarly introduce the mean density
DM
qmean ¼ ; ð3:9Þ
DV
at the limit (Fig. 3.2b), provided it exists, we shall get
dM
q ¼ lim qmean ¼ ; ð3:10Þ
DV!0 dV
which gives the unit mass (density) at the point M, of dimension ½q ¼ ML3 .
The mass of the whole body is given by
ZZZ
M¼ qdV; ð3:11Þ
V

where, generally,
q ¼ qðr; tÞ ¼ qðx1 ; x2 ; x3 ; tÞ; ð3:12Þ

if q ¼ const or if q ¼ qðtÞ, then we have to deal with a homogeneous body and we


can write
M ¼ qV; ð3:13Þ
neglecting a variation in time of the domain of volume V.
If DF~ is reported to DM, then we shall obtain the mean mass force; at the limit,
when it exists, we shall get the mass force vector ð1=qÞF at the point M, where
(3.9), (3.10) were taken into account. Similarly, ð1=qÞF will be the mass force at
the same point. From a dimensional standpoint, we shall have ½ð1=qÞF ¼ LT2 .
As an example of mass force, we shall mention the own weight
c
¼ g; ð3:14Þ
q
where g is the gravity acceleration.
The notion of mass force is encountered especially in the case of fluid media,
where the density q must be brought into relief; in the case of deformable solids
the notion of volume force is chiefly used and so we shall make hereafter.
78 3 Mechanics of Stresses

3.1.1.4 Lost Force of d’Alembert

We remark that the dynamic problem is formally reduced to a static problem


(dynamic equilibrium) if, for the volume forces, we substitute d’Alembert’s lost
forces (that, in fact, are also of the nature of volume forces, since they must be
referred to the unit volume). That leads to the necessity of introducing the volume
force of inertia ðd=dtÞðqdu=dtÞ, where we took into account the second law of
Newton (by introducing the substantial derivative of the momentum), the fact that
the force must be referred to the unit volume (q is the density) and the formula
(2.1030 ) of the velocity.
If q ¼ const, then we get the volume force of inertia qd2 u=dt2 .
In case of infinitesimal deformations, the time derivatives can be substituted to
the substantial derivatives and so we obtain the volume force of inertia
ðo=otÞðqou=otÞ; if q ¼ qðx1 ; x21 ; x3 Þ, independent of time (therefore, eventually
a non-homogeneous body), then the volume force of inertia assumes the simpler
form q€ u.
In the last case, that we shall consider in this work, the lost forces of d’Alembert
will be used in the form
 i ¼ Fi  q€
F ui ; i ¼ 1; 2; 3: ð3:15Þ
These results correspond to the undamped small motions of the deformable
solids; in the case of the damped small motions, we must introduce certain
damping forces (of the same nature as the volume forces), resulting from certain
experimental laws. The most interesting case appears when the volume damping
force is in direct proportion to the velocity, i.e. it takes the form k du=dt where k
is the damping coefficient per unit volume. This coefficient can, eventually, take
the form k ¼ kðx1 ; x21 ; x3 ; tÞ; the sign ‘‘minus’’ appears owing to the fact that the
volume damping force is opposed to the motion.
In case of infinitesimal deformations, the volume damping force will be ku, _
while d’Alembert’s lost forces assume the form

Fi ¼ Fi  q€
ui  ku_ i ; i ¼ 1; 2; 3; ð3:150 Þ

hereafter we shall only consider the (3.15) form.

3.1.2 Variation of Stresses Around a Point

n
The stress vector p depends on the direction vector n of the area element upon
n
which it acts and of the point at which the computation is made. If p depends only
on the direction n, then we have to deal with a state of uniform stress; the contrary
case is the state of non-uniform stress.
3.1 Stress Vector 79

A uniform state of stress can be determined by simple algebraic equations of


dynamic equilibrium; in this case, if we do not take into account the action of the
volume loads, then we can confine ourselves to a global study of the body. A non-
uniform state of stress is determined by differential equations; besides the global
study, we have to enter upon a local study at an arbitrary point of the body under
consideration.

3.1.2.1 Cauchy’s Theorem

We shall cut out from the body, around the point M, an element of volume in the
form of a plate, of thickness h, the parallel faces being of area A; one of them has the
outer normal n and the other has the outer normal n0 (Fig. 3.3); obviously, we have

n þ n0 ¼ 0: ð3:16Þ
On the face Aþ , which passes through the point M, the centre of mass of the
respective surface, the stress vectors are reduced to a torsor at M, which has only a
n
force resultant p (we assume that the moment resultant is negligible); taking into
account the hypothesis of continuity of stresses, we can write the stress vector (the
n0
force resultant) at the point M 0 , centre of mass of the face A , in the form p þ g,
n0
where p is the corresponding stress vector, calculated at the point M 0 , while g ! 0
C
together with h ! 0. If F is the resultant of the lost forces of d’Alembert and if p
are the resultant stress vectors on the lateral surface along the contour C, on the
area element hds, then the dynamic equation of equilibrium reads
Z
n n0 C
p A þ ðp þ gÞA þ h p ds þ FAh ¼ 0;
C

making h ! 0 and simplifying by A, we get the relation


n n0
p þ p ¼ 0; ð3:160 Þ

which constitutes Cauchy’s theorem [6–11].

Fig. 3.3 Cauchy’s theorem n


n p
C
A+
M
h

M'
-
A

n' n'
p
80 3 Mechanics of Stresses

On the basis of this result, we can state that, by changing the sense (sign) on the
outer normal to the element of area upon which acts a stress vector, the sense
(sign) of the latter one changes too.
i
In particular, we consider the stress vectors p applied at the point O and cor-
responding to the co-ordinate axes Oxi ; i ¼ 1; 2; 3. We denote by
i
rij ¼ p ij ð3:17Þ
i
the component of the vector p along the co-ordinate axis Oxj of unit vector ij
(Fig. 3.4); the components rij ; i; j ¼ 1; 2; 3, will be called stresses. If i ¼ j, then
we get normal stresses, while if i 6¼ j, then we obtain tangential stresses.
We assume that the stresses are positive when the external normal of the face
upon which they act has the same direction as the corresponding co-ordinate axis
(or inversely) and when they are oriented in the positive (or negative) direction of
the axis of co-ordinates to which they are parallel. On the basis of this sign
convention, which corresponds to Cauchy’s theorem, the normal stresses
r11 ; r22 ; r33 will be always positive when they correspond to a tension and neg-
ative when they correspond to a compression. As to the tangential stresses
r23 ; r32 ; r31 ; r13 ; r12 ; r21 , their sign will depend on the chosen system of
co-ordinates; it is to be remarked however, that the above convention agrees with
the fact that the positive tangential stresses correspond to positive angular strains.

M3

x3
1
p

dx 3
11

12 G1
n
21
F3 p3
13
Gn n
G2 p n
22
M (x1,x2, x3) G n
F2 p2 dx2
2 n
p r p1 M2
23 F1 31
O 32 x2
G3
x1 dx1

3
p 33

M1

Fig. 3.4 Equilibrium of an infinitesimal tetrahedron


3.1 Stress Vector 81

3.1.2.2 Equilibrium of an Infinitesimal Tetrahedron

n
So as to be able to appreciate the manner in which the stress vector p varies when
the direction n varies, in the neighbourhood of the point Mðx1 ; x2 ; x3 Þ, we shall cut
out from the body an infinitesimal three-rectangular tetrahedron MM1 M2 M3 with
the faces MM2 M3 ; MM3 M1 ; MM1 M2 of areas
1
dAi ¼ dAni ¼ ijk dxj dxk ; dxj dxk ¼ dxk dxj ; i ¼ 1; 2; 3; ð3:18Þ
4
parallel to the planes of co-ordinates, the side edges having the dimensions
dxi ; i ¼ 1; 2; 3; the face M1 M2 M3 of area dA will be defined by the outer normal n,
of direction cosines ni ; i ¼ 1; 2; 3 (Fig. 3.4).
In case of a system of any curvilinear axes, we shall consider a tetrahedron the
faces of which are corresponding surfaces of co-ordinates.
On the faces MMj Mk act stress vectors the torsors of which at the gravity
centres Gi ðxi ; xj þ dxj =3; xk þ dxk =3Þ; i 6¼ j 6¼ k 6¼ i; i; j; k ¼ 1; 2; 3, are reduced
only to force resultants (we assume also that the moment resultants, which would
lead to micromoments, are negligible); taking into account the hypothesis of
continuity of stresses and Cauchy’s theorem, we can write the resultant stress
i i i
vector which acts at the point Gi , in the form  p þ g, where p is the stress vector
i
corresponding to the Oxi -axis, which acts at the point M, while g ! 0 for h ! 0,
where
h ¼ n1 dx1 ¼ n2 dx2 ¼ n3 dx3 ð3:19Þ
is the height of the tetrahedron with respect to the face M1 M2 M3 . On this face acts
a stress vector the torsor of which, at the gravity centre Gn ðxi þ dxi =3; i ¼ 1; 2; 3Þ,
n n n
leads to a force resultant of the form p þ g, where p is the stress vector which
corresponds to the area element of outer normal n, which acts at the point M,
n
observing that g ! 0 for h ! 0.
At the centre of gravity Gðxi þ dxi =4; i ¼ 1; 2; 3Þ of the tetrahedron act the lost
n
forces of d’Alembert F þ g, where F is the lost force of d’Alembert, corre-
sponding to the point M, while g ! 0 for h ! 0.
The equation of dynamic equilibrium of the tetrahedron element will be of the
form
i i n n  h
ð p þ gÞ dAi þ ðp þ gÞ dA þ F þ g dA ¼ 0;
3
taking into account (3.18) and making h ! 0, we get
n i
p ¼ p ni ; ð3:20Þ
82 3 Mechanics of Stresses

relation that shows that the stress vector on any direction around a point can be
expressed by means of the stress vectors corresponding to three directions
orthogonal one to each other. Projecting along the Oxi -axes, we may write
n
pi ¼ rji nj ; i ¼ 1; 2; 3; ð3:200 Þ

where we used the notation (3.17). These relation can be written in the developed
form
n
p1 ¼r11 n1 þ r21 n2 þ r31 n3 ;
n
p2 ¼r12 n1 þ r22 n2 þ r32 n3 ; ð3:2000 Þ
n
p3 ¼r13 n1 þ r23 n2 þ r33 n3 ;

being due to Cauchy; the nine quantities rij ; i; j ¼ 1; 2; 3, characterize thus,


entirely, the state of stress around a point, both statically and dynamically.

3.1.2.3 Symmetry of Tangential Stresses

We put the condition that the resultant moment of all the forces which act upon the
infinitesimal tetrahedron considered above be equal to zero; calculating this
moment with regard to the point Gn , we may write
X3

! i i !  h
Gg Gi  ð p þ gÞ dAi þ Gn G  F þ g dA ¼ 0:
i¼1
3

Projecting on the Oxj -axis, it results

X
3    
dxi i i dxk  h
jkl  dik ð l þ gl ÞdAi þ jkl 
p Fl þ gl dA ¼ 0; j ¼ 1; 2; 3;
i¼1
3 12 3

where dik is Kronecker’a symbol; taking into account (3.18) and (3.19), we get
k k dxk  
jkl ð pl þ gl Þ þ jkl Fl þ gl ¼ 0; j ¼ 1; 2; 3:
12
If we make h ! 0 and use the notation (3.17), we get the relation
jkl rkl ¼ 0; j ¼ 1; 2; 3; ð3:21Þ

equivalent to the theorem of symmetry of tangential stresses [20]


rkl ¼ rlk ; k; l ¼ 1; 2; 3: ð3:210 Þ
Hence, of the 9 quantities rij ; i; j ¼ 1; 2; 3, only 6 quantities are distinct.
3.1 Stress Vector 83

From a mechanical standpoint, we can assert that, in this case, the existence of a
tangential stress rnm on an area element of external normal n, within a body, entails
the appearance of a tangential stress rmn of a corresponding direction, upon an area
element normal to the first one (of outer normal m). Hence, we may write (Fig. 3.5a)
rmn ¼ rnm : ð3:22Þ
Taking into account the sign convention adopted in Sect. 2.2.1 for the tan-
gential stresses, we observe that one cannot have a situation as that in Fig. 3.5b.
We notice also that the formula (3.22) is valid also in the case in which one of the
area elements lies on the outer surface of the body, being acted upon by a
superficial loading.
This formula has many applications. Let thus be a straight cylinder of Ox3 -axis,
acted upon the lateral surface by a superficial load, which has no components
along the Ox3 -axis. We make a section normal to the considered bar; let now be
two elements of outer normals n and x3 , respectively (Fig. 3.6a). We notice that
n
r3n ¼ p3 ; ð3:23Þ
n
but the tangential stress p3 is equal to zero, so that
r3n ¼ 0 ð3:230 Þ
and we can state that the tangential stress r3t which arises in the cross section is
tangent to its contour (Fig. 3.6b).
n m
Let be now the stress vectors p and p, corresponding to the elements of area of
outer normals n and m, respectively. The components of these vectors along the
directions m and n, respectively, are given by
n n n
pm ¼ p m ¼ pi mi ¼ rji nj mi ;
m m m
pn ¼ p n ¼ pj nj ¼ rij mi nj ;

taking into account (3.210 ), it results


m n
pn ¼ pm ; ð3:24Þ

Fig. 3.5 a Symmetry of (a) (b)


m m
tangential stresses.
b Impossible situation m
p

mn n mn n
n
p
nm nm
84 3 Mechanics of Stresses

Fig. 3.6 a Equilibrium of the (a) (b) t


straight cylinder. b Action of
r3t 6¼ 0 n n
3t
n
p3
3n
O x3 O x3

Fig. 3.7 Symmetry relation


m
in case of arbitrary directions m
p
n and m m
pn n
p n

n
pm

relation which generalizes the theorem of symmetry of tangential stresses


(Fig. 3.7).
We can therefore assert that, in the neighbourhood of a point, the component of
the stress vector corresponding to an area element, along the outer normal to
another area element, is equal to the component of the stress vector corresponding
to the second area element, along the outer normal to the first area element.

3.2 Stress Tensor. Equations of Equilibrium and Motion

The fact that the stress vector acts on an oriented element of area shows that, to
deepen this notion, it is necessary to introduce a quantity of tensor character too,
which we shall call stress tensor. We present, in what follows, the main properties
of this tensor and we establish the equations of equilibrium and motion which are
verified by its components.

3.2.1 Stress Tensor

n
Projecting the stress vector p, corresponding to the element of area of outer normal
n, on the unit vector n (Fig. 3.1c), we obtain
3.2 Stress Tensor. Equations of Equilibrium and Motion 85

n n n
pn ¼ p n ¼ pi ni ¼ rji nj ni ;

hence the normal stress is given by


n
r ¼ rij ni nj ; ð3:25Þ
n
the diadic product ni nj is a tensor of second order, while r is a scalar, independent
on the system of co-ordinates used. In this case, the theorem of quotient shows that
rij are the components of a tensor of second order, which is called stress tensor,
2 3
r11 r12 r13
Tr ¼ ½rij   4 r21 r22 r23 5; ð3:26Þ
r31 r32 r33

the formula (3.210 ) show that this tensor is symmetric.


Obviously, the components of the tensor Tr depend on the space components
xi ; i ¼ 1; 2; 3; taking into account (2.28), we can assume that, in the differential
relations, in case of infinitesimal deformations, these components depend on the
material variables
rij ¼ rij ðx1 ; x2 ; x3 ; tÞ; i; j ¼ 1; 2; 3: ð3:27Þ
If n and m are two orthogonal unit vectors, then the necessary and sufficient
conditions that rij be the components of a tensor of second order read

rnm ¼ rij ni mj ; ð3:28Þ

thus, the relations (3.25) and (3.28) give the normal and tangential components,
n
respectively, of the stress vector p.

3.2.1.1 Quadric of Normal Stresses. Principal Stresses

!
Let MQ be the vector of modulus
 
 ! 1 ni
MQ ¼ pffiffiffiffiffiffiffinffi ; xi ¼ pffiffiffiffiffiffiffinffi ; i ¼ 1; 2; 3; ð3:29Þ
r r
directed after the outer normal n to the considered element of area. With the help
of the relation (3.25), we find that the locus of the point Q will be supplied by the
equation (for the sake of simplicity, the point M is considered to be the origin of
the axes of co-ordinates)
rij xi xj ¼ 1; ð3:30Þ
n
that shows how r varies; let us remark that we take the sign + or the sign -,
according to whether we have to deal with a tension or with a compression, so that
86 3 Mechanics of Stresses

the surface be real. This quadric is called the quadric of normal stresses (Cauchy’s
quadric) (Fig. 3.8).
Referring the quadric to its principal axes M1; M2; M3, its equation will take
the form

r1 x02 02 02
1 þ r2 x2 þ r3 x3 ¼ 1; ð3:31Þ

where x0i ; i ¼ 1; 2; 3, are the new co-ordinates. It follows that there exist three
principal directions along which the extreme normal stresses r1  r2  r3 develop
and for which the tangential stresses vanish. Generally, these three principal
directions do not coincide with the principal directions of the tensor Te .
If we have r1  r2  r3 [ 0 or if 0 [ r1  r2  r3 , then the quadric is an
ellipsoid. We choose the sign of 1 so that the ellipsoid be real; if two of the
principal normal stresses are equal, then we obtain an ellipsoid of revolution, while
if all the three principal normal stresses are equal, then the ellipsoid is a sphere.
If r1  r2 [ 0 [ r3 , e.g., then we obtain a one or a two-sheet hyperboloid as
we choose the sign of 1 (the locus is formed by both hyperboloids); if two of the
principal normal stresses are equal, then the two hyperboloids are of revolution.
We can have situations too, in which the quadric is degenerate (e.g., in case of a
n
state of plane stress). Imposing the condition r ¼ 0, we get only tangential stresses
along the asymptotic cone.
n
If the direction n is a principal one, to which corresponds the normal stress r,
n
and if we take into account (3.30 ) (where we take s ¼ 0), as well as (3.200 ), then we
are led to the equations
 
rij  rdij nj ¼ 0; i ¼ 1; 2; 3: ð3:32Þ
The homogeneous system has non-trivial solutions if the determinant
2 3

r11  r r12 r13
det rij  rdij  4 r21 r22  r r23 5 ¼ 0 ð3:33Þ
r31 r32 r33  r

vanishes; we are thus led to the equation of third degree

r3  J1 r2 þ J2 r  J3 ¼ 0; ð3:34Þ

Fig. 3.8 Quadric of normal


stresses

n
Q
M pn

2
3.2 Stress Tensor. Equations of Equilibrium and Motion 87

with
1
J1 ¼ ijk ljk ril ¼ dil ril ¼ rii ¼ r11 þ r22 þ r33 ¼ r1 þ r2 þ r3 ; ð3:340 Þ
2
1 1 
J2 ¼ ijk lmk ril rjm ¼ rii rjj  rij rij ¼ r22 r33 þ r33 r11 þ r11 r22
2 2
 ðr223 þ r231 þ r212 Þ ¼ r2 r3 þ r3 r1 þ r1 r2 ; ð3:3400 Þ

1
J3 ¼ ijk lmn ril rjm rkn ¼ det½rij  ¼ r11 r22 r33
6
 ðr11 r223 þ r22 r231 þ r33 r212 Þ þ 2r23 r31 r12 ¼ r1 r2 r3 : ð3:34000 Þ

The Eq. (3.34) supplies the values of the principal normal stresses, while the
system (3.32), with the additional condition (2.37), supplies the principal direc-
tions corresponding to the chosen principal normal stress.
As it has been specified in Sect. 2.2.1.1 for the tensor Te , one can show, in this
case too, that the three principal directions are three-orthogonal and that the three
roots of the Eq. (3.34) are always real.
The coefficients J1 ; J2 ; J3 are invariant at a change of co-ordinate axes; the first
of these invariants is denoted also by
H ¼ rii ¼ r1 þ r2 þ r3 : ð3:35Þ
Taking into account the above results, we can write the normal stress corre-
sponding to an area element of outer normal n, in the M123 system of axes, in the
form
n
r ¼ r1 n21 þ r2 n22 þ r3 n23 ; ð3:36Þ

the direction cosines being considered with respect to these axes; analogically, the
tangential stress corresponding to the directions of unit vectors n and m is given by
rnm ¼ r1 n1 m1 þ r2 n2 m2 þ r3 n3 m3 : ð3:360 Þ
n
The components of the stress p are
n n n
p1 ¼ r1 n1 ; p2 ¼ r2 n2 ; p3 ¼ r3 n3 ð3:37Þ

and the formula (3.300 ) shows that the tangential stress is given by
n  2
s2 ¼ r21 n21 þ r22 n22 þ r23 n23  r1 n21 þ r2 n22 þ r3 n23 : ð3:38Þ
Observing that the relation

n21 þ n22 þ n23 ¼ 1 ð3:39Þ


88 3 Mechanics of Stresses

takes place, we can write the expression (3.38) also in the remarkable form
n
s2 ¼ ðr2  r3 Þ2 n22 n23 þ ðr3  r1 Þ2 n23 n21 þ ðr1  r2 Þ2 n21 n22 : ð3:380 Þ
To obtain the principal tangential stresses, we equate to zero the derivatives of
first order of the form
n  
U ¼ s2  r2 1  n21  n22  n23 ð3:40Þ

with respect to the direction cosines, where r is a Lagrange’s multiplier of the


nature of a stress; we obtain thus the system of equations
h   i
n1 r21  2r1 r1 n21 þ r2 n22 þ r3 n23 þ r2 ¼ 0;
 

n2 r22  2r2 r1 n21 þ r2 n22 þ r3 n23 þ r2 ¼ 0; ð3:400 Þ


 

n3 r23  2r3 r1 n21 þ r2 n22 þ r3 n23 þ r2 ¼ 0:


One cannot have n1 ¼ n2 ¼ n3 ¼ 0, because the condition (3.39) would not be
fulfilled. Let us suppose now that all the direction cosines are non-zero. In this
case, the system (3.400 ) shows that the conditions

r21 þ r2 r22 þ r2 r23 þ r2


¼ ¼
r1 r2 r3
should be verified; because we have only one parameter, these conditions cannot
be simultaneously fulfilled, but only if two of the principal normal stresses are
equal. Hence, at least one of the direction cosines must vanish.
Let be n1 ¼ 0; n2 ; n3 6¼ 0. We notice that the system (3.40) is compatible if
r2 ¼ r2 r3 ; we get
1
r2 n22 þ r3 n23 ¼ ðr1 þ r2 Þ;
2
n22 þ n23 ¼ 1;
pffiffiffi
wherefrom n2 ; n3 ¼  2=2.
pffiffiffi
Analogically, we also determine the solutions n2 ¼ 0; n3 ; n1 ¼  2=2, n3 ¼ 0,
pffiffiffi
n1 ; n2 ¼  2=2. One thus sees that the principal tangential stresses take place in
planes which bisect the dihedral angles formed by the principal planes M23; M31
and M12; we notice that we can take all the combinations of signs for the direction
cosines, because the principal dihedrons have each one two bisector planes, and
these ones have different orientations.
The formula (3.38) leads to the principal tangential stresses
1 1 1
s1 ¼  ðr2  r3 Þ; s2 ¼  ðr3  r1 Þ; s3 ¼  ðr1  r2 Þ: ð3:41Þ
2 2 2
3.2 Stress Tensor. Equations of Equilibrium and Motion 89

We observe that, using the formula (3.360 ), we obtain, e.g.,


pffiffiffi
2
s1 ¼ ðr2 m2  r3 m3 Þ;
2
using the relation

m21 þ m22 þ m23 ¼ 1 ð3:390 Þ


pffiffiffi
and taking into account (3.41), it results that m1 ¼ 0; m2 ; m3 ¼  2=2, the last
two direction cosines having the sign specified by the condition of orthogonality
n1 m1 þ n2 m3 þ n3 m3 ¼ 0: ð3:3900 Þ
Hence, the principal tangential stress occurs along the intersection line of the
bisector plane with the principal plane.
If we take into account the order considered for the principal normal stresses, it
results that the maximal tangential stress is given by
1
smax ¼ ðr1  r3 Þ: ð3:42Þ
2
The normal stresses corresponding to the elements of area on which the prin-
cipal tangential stresses take place are given by (3.36) in the form
1 1 1
r1 ¼ ðr2 þ r3 Þ; r2 ¼ ðr3 þ r1 Þ; r3 ¼ ðr1 þ r2 Þ: ð3:410 Þ
2 2 2
The magnitudes of the corresponding stress vectors will be thus given by
pffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 2 2 2 2 2 2
p1 ¼ r2 þ r 3 ; p 2 ¼ r3 þ r 1 ; p 3 ¼ r21 þ r22 ; ð3:43Þ
2 2 2
these vectors being contained in the principal planes (Fig. 3.9).

Fig. 3.9 Principal tangential 3


stresses
3
4

M 4 p3

4 2
4
3

1
90 3 Mechanics of Stresses

3.2.1.2 Ellipsoid and Direction Surface of Stresses. Mohr’s Circles

By taking into consideration the relation (3.39), the relations (3.37) lead to the
equation

x21 x22 x23


þ þ ¼ 1; ð3:44Þ
r21 r22 r23

which represents Lamé’s ellipsoid of stresses; the radius vector of every point of
n
the ellipsoid is just the stress vector p. So as to establish the area element on which
such a stress is acting, we shall equally consider the direction surface of stresses

x21 x22 x23


þ þ ¼ 1: ð3:45Þ
r1 r2 r3
One can easily show that the stress vector corresponding to a radius vector of
Lamé’s ellipsoid acts on an area element the outer normal of which has the same
direction as the outer normal to the direction surface of stresses at the point where
the latter intersects the considered radius vector.
If jr1 j ¼ jr2 j, then the ellipsoid of stresses is of revolution, since the axis M3 is
an axis symmetry; if jr1 j ¼ jr2 j ¼ jr3 j, then the ellipsoid becomes a sphere.
Lastly, when one of the stresses (e.g., r3 ) vanishes, we have to deal with a plane
state of stress.
Starting from the Eq. (3.30) of Cauchy’s quadric, we introduce the function
2F ðx1 ; x2 ; x3 Þ ¼ rij xi xj ; ð3:46Þ

by computing the partial derivatives of first order of this function and taking into
account (3.200 ) and (3.27), we notice that we can write
! ! n
F;i ¼ rij xj ¼ jMQjrij nj ¼ jMQj pi ; i ¼ 1; 2; 3:
Finally, we have
n 1
pi ¼ ! F;i ; i ¼ 1; 2; 3; ð3:47Þ
jMQj
the function Fðx1 ; x2 ; x3 Þ playing thus the rôle of a potential function.
On the other hand, F;i ; i ¼ 1; 2; 3, are the direction parameters of the normal to
n
Cauchy’s quadric at the point Q, while pi ; i ¼ 1; 2; 3, are the components of the
n n
stress vector p. This offers the possibility of a simple construction of the vector p,
as concerns its direction; indeed, if we trace the plane, tangent at the point Q to
n
Cauchy’s quadric, the normal from M to this plane will supply the direction of p
(Fig. 3.8).
3.2 Stress Tensor. Equations of Equilibrium and Motion 91

Let be now the system of equations

n21 þ n22 þ n23 ¼ 1;


n
r1 n21 þ r2 n22 þ r3 n23 ¼ r; ð3:48Þ
n n n
r21 n21 þ r22 n22 þ r23 n23 þ ¼ p ¼ r þ s ;
2 2 2

solving with respect to the direction cosines, we obtain


n
n n
s2 þðr r2 Þðr r3 Þ
n21 ¼ ;
ðr1  r2 Þðr1  r3 Þ
n
n n
s2 þðr r3 Þðr r1 Þ ð3:480 Þ
n22 ¼ ;
ðr2  r3 Þðr2  r1 Þ
n
n n
s2 þðr r1 Þðr r2 Þ
n23 ¼ :
ðr3  r1 Þðr3  r2 Þ
Unlike the relations (3.36), (3.38), which gave the stress vector corresponding
to an element of area of given outer normal, the relations (3.480 ) specify the
element of area corresponding to a given stress vector.
Remarking that r1  r2  r3 and taking into account that the squared direction
cosines should be positive, it follows that
n
n n
s2 þðr r2 Þðr r3 Þ  0;
n
n n ð3:49Þ
s2 þðr r3 Þðr r1 Þ 0;
n
2 n n
s þðr r1 Þðr r2 Þ  0;
n n
referring this to a system of axes O r s we obtain, in the case of equality, three
circles: C1 of diameter r2  r3 and centre O1 ; C2 of diameter r1  r3 and centre
O2 and C3 of diameter r1  r2 and centre O3 . These circles have been introduced
by O. Mohr [2] and bear his name (Fig. 3.10).
The inequalities (3.49) are fulfilled by a point within the circle C2 and outside
the circles C1 and C3 or, at the limit, by a point on the contour of one of these three
circles. Mohr’s circles allow thus to represent, by a plane diagram, a three-
dimensional state of stress; besides, they allow to make a simple study of the
extreme value of the normal and tangential stresses. For example, the extreme
tangential stresses correspond to the radii of the three circles, as was shown by the
formula (3.41); as well, the corresponding normal stresses are the abscissae of the
centres of the three circles, as it has been shown by the formula (3.410 ).
Once the principal normal stresses are determined, the three circles can be
easily constructed. Because in the system of Eq. (3.48) the tangential stress
appears only by its square, it follows that we can use only the part of the diagram
n
for which s  0.
92 3 Mechanics of Stresses

Fig. 3.10 Mohr’s circles n


C2
C1
3 2 C3 1
O O1 O2 O3 n

An element of area of outer normal n being given, the relations (3.480 ) allow to
draw three arcs of circle with the centres at O1 ; O2 ; O3 , of radii given by
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1
r1 ¼ ðr2  r3 Þ2 þ n21 ðr1  r2 Þðr1  r3 Þ;
2 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1
r2 ¼ ðr3  r1 Þ2 þ n22 ðr2  r3 Þðr2  r1 Þ; ð3:50Þ
2 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1
r3 ¼ ðr1  r2 Þ2 þ n23 ðr3  r1 Þðr3  r2 Þ;
2
n
respectively (Fig. 3.11a); these arcs of circle meet at the point P of co-ordinates r
n
and s, obtaining the components of the corresponding stress vector.
We can use also an entirely graphic method. We draw thus the tangents to the
n
circle C2 at the points at which this one pierces the axis O r and construct semi-
straight lines, inclined to these tangents by the angles a2 and a3 , made by the unit
vector n with the principal axes, respectively (Fig. 3.11b); these semi-right lines
pierce the circles C2 and C3 at the points A2 and A3 and the circles C1 and C2 at the
points B1 and B2 , respectively. One can show easily that the arc of circle of radius
r1 with the centre at O1 passes through the points A2 and A3 , while the arc of circle
or radius r3 with the centre at O3 passes through the points B1 and B2 . The
construction is thus completely specified.

3.2.1.3 Spherical Tensor. Stress Deviator. Octahedral Stresses

As in case of strains, we decompose the tensor Tr into a sum of two tensors: the
spherical tensor Tr0 , given by r0 dij , where
1 1
r0 ¼ rii ¼ J1 ; ð3:51Þ
3 3
and the stress deviator tensor Tr0 , defined by the relation

r0ij ¼ rij  r0 dij ; ð3:52Þ

which is symmetric too.


3.2 Stress Tensor. Equations of Equilibrium and Motion 93

n n
(a) (b) A2 B2
C2
P C2
C1 n P r3 B1
r1 r2 C3 3
1

3 2 1 3 2 A3 1
O O1 O2 n O3 n O O1 O2 O3 n

C3
C1

Fig. 3.11 Mohr’s circles: semigraphic method (a), graphic method (b)

The extreme values of the normal components of the latter tensor are given by
the equation

det½r0ij  r0 dij  ¼ 0; ð3:53Þ

which is of the form

r03  J20 r0  J30 ¼ 0: ð3:530 Þ


The first invariant of the stress deviator vanishes

J10 ¼ r0ii ¼ 0; ð3:54Þ

while the other two invariants are given by


1
J20 ¼ r0ij r0ij ¼ 3r20  J2 ; ð3:540 Þ
2
1
J30 ¼ r0ij r0jk r0ki ¼ J3 þ J20 r0  r30 ; ð3:5400 Þ
3
by means of the invariant of the tensor Tr .
We notice that the invariant J20 can be expressed in the form
 
J20 ¼  r022 r033 þ r033 r011 þ r011 r022 þ r223 þ r231 þ r212
1 
¼ r02 þ r02 02 2 2
22 þ r33 þ r23 þ r31 þ r12
2
2 11
1h i
¼ ðr22  r33 Þ2 þðr33  r11 Þ2 þ ðr11  r22 Þ2 þ r223 þ r231 þ r212
6
1h i 2 
¼ ðr2  r3 Þ2 þðr3  r1 Þ2 þ ðr1  r2 Þ2 ¼ s21 þ s22 þ s23
6 3
1  02 02 02
  0 0 0 0 0 0

¼ r1 þ r2 þ r3 ¼  r2 r3 þ r3 r1 þ r1 r2 ; ð3:55Þ
2
94 3 Mechanics of Stresses

while the invariant J30 reads


1
J30 ¼ ð2r1  r2  r3 Þð2r2  r3  r1 Þð2r3  r1  r2 Þ
27
8 1 
¼ ðr1  r1 Þðr2  r2 Þðr3  r3 Þ ¼ r03 þ r03 03 0 0 0
2 þ r3 ¼ r1 r2 r3 : ð3:550 Þ
27 3 1
The equivalent stress (called the intensity of stresses too) is defined by the
relation
qffiffiffiffiffiffiffi
r ¼ 3J20 : ð3:56Þ

Considering the octahedron in Fig. 2.5, the formula (3.36) shows that the
octahedral normal stress is equal to r0 . Using the same procedure as in
Sect. 2.2.1.2, the octahedral tangential stress reads
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 2
s0 ¼ ðr2  r3 Þ2 þ ðr3  r1 Þ2 þ ðr1  r2 Þ2 ¼ s21 þ s22 þ s23 : ð3:57Þ
3 3
We mention also the relation
pffiffiffi
2
s0 ¼ r: ð3:570 Þ
3
Taking into account (3.56), (3.570 ), we can express the invariant J20 in the form
3
J20 ¼ s20 ; ð3:58Þ
2
while the invariant J2 is given by
 
2 1 2
J 2 ¼ 3 r 0  s0 ; ð3:580 Þ
2

where we used the octahedral stresses.


We notice that one can express the roots of the Eq. (3.530 ) in the form
pffiffiffi p
r01 ¼ 2s0 cos x  ;
3
0
pffiffiffi p ð3:59Þ
r2 ¼ 2s0 cos x þ ;
pffiffiffi 3
r03 ¼  2s0 cos x;

the angle 0 x p=3 being given by the relation


pffiffiffi J 0
cos 3x ¼ 2 33 ; ð3:590 Þ
s0
we will have r01  r02  r03 .
We can make analogous considerations for the Eq. (2.460 ).
3.2 Stress Tensor. Equations of Equilibrium and Motion 95

3.2.2 Equations of Equilibrium and Motion

The study up to this point refers to the behaviour of the stresses in the neigh-
bourhood of a point. In order to complete the local study, we must see what
happens when we pass from an area element through the point M to an area
element through an adjacent point. Thus, we deduce the differential equations of
equilibrium and motion which must be fulfilled by the components of the tensor Tr
and by the components of the vector u (the last ones in the dynamic case).

3.2.2.1 Equations of Equilibrium

Let us assume that the considered solid body occupies a domain of volume V,
bounded by the surface S; let be also an arbitrary subdomain of volume V 0 ,
bounded by the surface S0 (Fig. 3.12). Using the principle of equilibrium of parts
for this subdomain (principle which has been used till now too for the dynamic
equilibrium for the elementary domains in form of a plate or of a tetrahedron,
which have been cut from the body), one can write the vector equations of
equilibrium in the form
ZZ ZZZ
n 0
pdS þ FdV 0 ¼ 0; ð3:60Þ
S0 V0
ZZ ZZZ
n
r  pdS0 þ r  FdV 0 ¼ 0: ð3:61Þ
S0 V0

We suppose that we have to do with infinitesimal deformations, so that we may


apply the principle of rigidity for the considered elementary subdomain; in this
case, the condition that the torsor of the stress vectors which act upon the surface
S0 and of the volume forces which act upon the volume V 0 be zero is a necessary
and sufficient condition of equilibrium.
Projecting on the Oxi -axis, we have
ZZ ZZZ
n
pi dS0 þ Fi dV 0 ¼ 0; i ¼ 1; 2; 3; ð3:600 Þ
S0 V0

Fig. 3.12 Equilibrium of an S


arbitrary subdomain
S' V'

V
r

O
96 3 Mechanics of Stresses

ZZ ZZZ
n
ijk xj pk dS0 þ ijk xj Fk dV 0 ¼ 0: i ¼ 1; 2; 3: ð3:610 Þ
S0 V0

Starting from the Eq. (3.600 ) and using the formula (3.200 ) and the flux-diver-
gence formula (A.1.960 ), which transforms a surface integral in a volume one, it
follows
ZZZ
 
rji;j þ Fi dV 0 ¼ 0: i ¼ 1; 2; 3;
V0

assuming that the stresses are functions of class C1 and that the volume forces are
functions of class C0 and observing that the subdomain V 0 is arbitrary, we obtain
the equations of equilibrium in the form
rji;j þ Fi ¼ 0; i ¼ 1; 2; 3: ð3:62Þ
The relation (3.610 ) can be written, analogically, in the form
ZZZ h i

ijk xj rlk ;l þxj Fk dV 0 ¼ 0: i ¼ 1; 2; 3;
V0

differentiating and taking into account the equations of equilibrium (3.62), it


results
ZZZ
ijk rjk dV 0 ¼ 0: i ¼ 1; 2; 3:
V0

By considerations analogue to those above, we are led to the theorem of symmetry


of tangential stresses (3.21). Taking into account this result, we may write the
equations of equilibrium also in the form
rij;j þ Fi ¼ 0; i ¼ 1; 2; 3: ð3:620 Þ
The notation (3.17), by which we have introduced the components of the stress
tensor, allows to write the Eq. (3.620 ) in the form
i
div p þFi ¼ 0; i ¼ 1; 2; 3; ð3:6200 Þ
i
observing that div p are the components of the divergence of the tensor Tr , the
equation of equilibrium also read
DivTr þ F ¼ 0: ð3:62000 Þ
The linear partial differential Eq. (3.62)–(3.62000 ) represent the mechanical
conditions which must be fulfilled by the stresses (the mechanical aspect of the
problem).
To can put better in evidence the action of the components of the stress tensor,
we repeat the above considerations on a subdomain of volume V 0 of a parallel-
epipedical form and dimensions dxi ; i ¼ 1; 2; 3, having the faces parallel to the
3.2 Stress Tensor. Equations of Equilibrium and Motion 97

x3
33 + 33,3 dx3
32 + dx3
32,3

31 + 31,3 dx3
12 11
dx3
23 + 23,2 dx2
F3
13

21 F2 22 + dx2
22,2
F1
22 13 + 13,1 dx1 G + dx2
21 21,2
23
M (x1,x2,x3)
11 + 11,1 dx1 12 + 12,1 dx1
31
r dx1
32

O 33 x2

x1
dx2

Fig. 3.13 Equilibrium of an elastic parallelepiped

co-ordinate planes (Fig. 3.13). Taking into account the variations of the stresses
from a face of the parallelepiped to a neighbouring parallel face by formulae of
Taylor type and observing that we must pass from the stresses at the point
Mðx1 ; x2 ; x3 Þ to the stresses at the centres of gravity of each adjacent face, as in
Sect. 2.2.2, we obtain the equations of equilibrium (3.62).
If we choose a subdomain bounded by co-ordinate surfaces, in a system of
arbitrary curvilinear co-ordinates, then we obtain equivalent equations of equi-
librium in the considered system of co-ordinates.
The above considerations are valid for an arbitrary deformable continuous
medium, hence also for a fluid one [3, 16–19].

3.2.2.2 Equations of Motion

In the dynamical case, we must replace the volume forces by d’Alembert’s lost
forces, defined in Sect. 3.1.1.4. Thus, the Eq. (3.62) allow to write the equations of
motion in the form
rji; j þ Fi ¼ q€
ui ; i ¼ 1; 2; 3; ð3:63Þ

corresponding to the case of infinitesimal deformations and to undamped motions.


As in the statical case, one can write these equations in the form
rij; j þ Fi ¼ q€
ui ; i ¼ 1; 2; 3; ð3:630 Þ
98 3 Mechanics of Stresses

too, or in the form


i
div p þFi ¼ q€
ui ; i ¼ 1; 2; 3; ð3:6300 Þ
in a developed form, we have

r11;1 þ r12;2 þ r13;3 þ F1 ¼ q€u1 ;


r21;1 þ r22;2 þ r23;3 þ F2 ¼ q€u2 ; ð3:63000 Þ
r31;1 þ r32;2 þ r33;3 þ F3 ¼ q€u3 :
We can also write
DivTr þ F ¼ q€
u: ð3:64Þ
In the case of finite deformations, the equations of motion take the form
 
d du
DivTr þ F ¼ q ; ð3:640 Þ
dt dt

where we must take into consideration the substantial derivative.


Introducing a damping in direct proportion to the displacement velocity as we
have shown in Sect. 3.1.1.4, we may write the Eq. (3.63) in the form
rji; j þ Fi ¼ q€
ui þ ku_ i ; i ¼ 1; 2; 3; ð3:65Þ

as well, the correspondent of the vector Eq. (3.64) is


DivTr þ F ¼ q€
u þ ku:
_ ð3:650 Þ

3.2.3 Elementary States of Stress

In that follows we present some elementary states of stress which appear fre-
quently in applications, i.e.: the hydrostatic stress, the simple tension or com-
pression and the simple shear. As well, we give some data concerning a state of
plane stress.
We mention that these states of stress correspond to the elementary states of
deformation considered in Sect. 2.2.3.

3.2.3.1 Hydrostatic Stress

We call simple state of normal stress that state of stress for which
rij ¼ rdij ; r [ 0; i; j ¼ 1; 2; 3; ð3:66Þ
3.2 Stress Tensor. Equations of Equilibrium and Motion 99

we have, in this case,


r11 ¼ r22 ¼ r33 ¼ r; ð3:660 Þ

r23 ¼ r31 ¼ r12 ¼ 0: ð3:6600 Þ


Hence, the stress vector has the same direction as the outer normal to the
elementary area on which it acts. If r\0, then we have to do with a hydrostatic
stress.
The quadric of normal stresses is a sphere of equation
1
x21 þ x22 þ x23 ¼  ; ð3:67Þ
r
where, in case of a hydrostatic stress, one takes the sign -. All the directions
which pass through the centre of the sphere are principal directions.

3.2.3.2 Simple Tension and Compression

We say that we have to do with a simple tension if the state of stress is reduced to
the normal stress r1 ¼ r [ 0 along the direction of a principal axis (e.g., O1,
1
which is specified by the direction cosines ni ; i ¼ 1; 2; 3, with respect to an ort-
honormed frame Oxi ; i ¼ 1; 2; 3); if r\0, then we have to do with a simple
compression.
The quadric of normal stresses is given by the equation

rx02
1 ¼ 1 ð3:68Þ

and is reduced to two parallel planes.


The state of stress may be written also in the form
1 1
rij ¼ r ni nj ; i; j ¼ 1; 2; 3; ð3:69Þ

with regard to the orthonormed frame Oxi .

3.2.3.3 Simple Shear

If the stress tensor has only one component with distinct indices, e.g., the tan-
gential stress r12 ¼ r21 ¼ s (Fig. 3.14), then we have to do with a simple shear;
the quadric of normal stresses is given by

2sx01 x02 ¼ 1; ð3:70Þ

being an equilateral hyperbola, the state of stress taking place in the Ox1 x2 -plane.
100 3 Mechanics of Stresses

Fig. 3.14 Simple shear 1


x2

O 2 2
x1

By a rotation of p=4 in the Ox1 x2 -plane we can express the quadric with respect
to the principal axes, obtaining the equation
 
s x02 02
1  x2 ¼ 1; ð3:700 Þ

hence, in case of a simple shear in the Ox1 x2 -plane, the principal normal stresses
are given by
r1 ¼ r2 ¼ s; r3 ¼ 0; ð3:71Þ
the reciprocal being true.
We can thus state that the simple shear in one direction is equivalent to a simple
tension along a direction which makes an angle p=4 with the first direction and to a
simple compression, of the same magnitude, along a direction normal to that of the
simple tension; this result is put in evidence by considering the equilibrium of the
triangular element in Fig. 3.14.

3.2.3.4 Plane State of Stress

We call plane state of stress that state of stress for which the principal normal
stress r3 vanishes
r3i ¼ 0; i ¼ 1; 2; 3: ð3:72Þ
The quadric of normal stresses is written in the form

r11 x21 þ 2r12 x1 x2 þ r22 x22 ¼ 1 ð3:73Þ

or in the form

r1 x02 02
1 þ r2 x2 ¼ 1; ð3:730 Þ

being a cylinder the normal section of which is an ellipse or a hyperbola.


In particular, we can have a simple tension or compression or a simple shear.
3.2 Stress Tensor. Equations of Equilibrium and Motion 101

3.2.4 Finite Deformations

In the study on the geometry and kinematics of deformation made in Chap. 2 we


have seen that the obtained results have a different form, as one uses material or
space co-ordinates.
On the other hand, the stress tensor referring to the actual state is a natural
physical concept; the equations of equilibrium and motion obtained in Sect. 3.2.2
correspond just to such an idea. But if, in case of finite deformations, we express
the state of strain with respect to the initial state, then it is difficult to make a
connection between that one and the state of stress.
Hence, it is necessary to express the state of stress with respect to the initial state
too. Besides the Eulerian stress tensor of Cauchy, corresponding to the actual state
and considered above, we introduce the Lagrangian Piola-Kirchhoff stress tensors
of first and second kind, respectively, which correspond to the initial state. Thus,
one can give new forms for the equations of equilibrium and motion. But, first of all,
it is necessary to establish some results concerning the continuity of mass.

3.2.4.1 Continuity of Mass

The density q, introduced in Sect. 2.1.3, must behave so that the mass M of the
solid which we follow in its motion be invariant in time; this is the conservation
principle of mass.
At the actual moment, the mass is expressed in the form
ZZZ

 
M ¼ q
x
1 ; x
2 ; x
3 ; t dV
; ð3:74Þ
V

while, at the initial moment, it is given by


ZZZ
M¼ qðx1 ; x2 ; x3 ; tÞdV; ð3:740 Þ
V

effecting a change or variable of the form (2.2), we can write


ZZZ

M ¼ q
J dV; ð3:7400 Þ
V

where we have introduced the Jacobian (2.3). Equating the two masses, we get
ZZZ
ðq
J  qÞdV ¼ 0;
V

assuming that the function under the integral is continuous and observing that the
volume V is arbitrary, it results the relation
Jq
¼ q; ð3:75Þ
102 3 Mechanics of Stresses

which represents d’Alembert’s continuity condition of mass.


As it results from the relations (2.26), (2.260 ), in the case of infinitesimal
deformations, we have
J ¼ 1; q
¼ q; ð3:76Þ
if we have such a relation in time, in case of finite deformations, it results that the
motion is incompressible.
We notice also that

   
dJ d oxi d ox
i ovi ovi ox
k
¼ det ¼ det ¼ det ¼ det

dt dt oxj dt oxj oxj oxk oxj



 
ovi ox ovi ovi
¼ det
det k ¼ Jdik det
¼ J
;
oxk oxj oxk oxi

we obtain thus Euler’s theorem, according to which the relation


dJ
¼ Jdivv ð3:77Þ
dt
holds for a particle in motion.
We notice that this relation can be written also in the form
d
ln J ¼ divv: ð3:770 Þ
dt
We can put the condition of invariance of the mass in time also by equating to
zero the derivative with respect to time of the mass (3.7400 ); we obtain
ZZZ ZZZ 

dM
d
dq
dJ
¼ q J dV ¼ Jþq dV
dt dt V dt dt
ZZZ 
V ZZZ 

dq
dq
¼ þ q divv JdV ¼ þ q divv dV
¼ 0;

V dt V
dt

where the volume V is arbitrary. If the function under the integral is continuous,
then we can write Euler’s continuity condition of mass in the form
dq

þ q
divv ¼ 0: ð3:78Þ
dt
Observing that
dq
oq
dr
oq

¼ þ gradq
 ¼ þ gradq
 v;
dt ot dt ot
divðq
vÞ ¼ q
divv þ gradq
 v;
3.2 Stress Tensor. Equations of Equilibrium and Motion 103

we can write the condition (3.78) also in the form


oq

þ divðq
vÞ ¼ 0: ð3:780 Þ
ot
If the density q
is constant in time, we obtain
divv ¼ 0; ð3:79Þ
hence the field of displacement velocities is solenoidal.
The relation (3.78) can be expressed also in the form
d
ln q
+ divv ¼ 0: ð3:7800 Þ
dt
From the relation (3.770 ), (3.7800 ), we get
d
lnðJq
Þ ¼ 0;
dt
hence, the product Jq
is constant in time, being equal to q (corresponding to the
initial state). We find thus again the continuity condition of mass given by
d’Alembert.

3.2.4.2 The Piola-Kirchhoff Stress Tensors

We consider that on an element of area dA of outer normal n, in the initial state,


acts the stress vector
n
n ^
dP

^ ; ð3:80Þ
dA
called the Piola-Kirchhoff stress vector of first kind or the stress vector
n
n dP
p¼ ; ð3:800 Þ
dA
called the Piola-Kirchhoff stress vector of second kind (Fig. 3.15).
As well, on the corresponding element of area dA
, of outer normal n
, in the
actual state, acts the stress vector
n


d P

p ¼
; ð3:81Þ
dA
called Cauchy’s stress vector. As a matter of fact, till now, in the frame of this
chapter, we have used the latter stress vector (being in the case of the infinitesimal
deformations, it was not necessary to use the asterisk for the actual state).
104 3 Mechanics of Stresses

Fig. 3.15 The Piola- x3 n* n*


Kirchhoff and the Cauchy n n dP*
dP
stress vector
M*
M dA
*
dA

r*
r

O x2

x1

The Piola-Kirchhoff stress vector of first kind is defined, by means of Cauchy’s


stress vector, by the relation
n n

^ ¼ d P
;
dP ð3:82Þ
equivalent with
n n

^ i = d P
i ;
dP i ¼ 1; 2; 3: ð3:820 Þ
Taking into account (3.200 ), (3.80) and (3.81), the relation (3.820 ) leads to
^ji nj dA ¼ r
ji ; n
j dA
;
r i ¼ 1; 2; 3; ð3:8200 Þ

where r^ij are the components of Piola-Kirchhoff stress tensor of first kind, while r
ij
are the components of Cauchy’s stress tensor, which is a tensor of second kind.
Observing that two elements of volume constructed on the elements of area
dA and dA
, respectively, are given by
oxj

dV ¼ dAn  dr ¼ nj dAdxj ¼ nj dA dx ; ð3:83Þ


ox
i i

ox
i
dV
¼ dA
n
 dr
¼ n
i dA
dx
i ¼ n
i dA
dxj ð3:830 Þ
oxj

and that
q
dV
¼ JdV ¼ dV; ð3:84Þ
q

3.2 Stress Tensor. Equations of Equilibrium and Motion 105

we obtain
q oxj
n
i dA
¼

nj dA
; i ¼ 1; 2; 3; ð3:85Þ
q oxi

q


ox
j
ni dA ¼ n dA ; i ¼ 1; 2; 3: ð3:850 Þ
q j oxi
The relations (3.82’’) and (3.85) allow to express the components of the tensor
Tr^ in the form
q oxi

^ij ¼
r r ; i; j ¼ 1; 2; 3; ð3:86Þ
q
ox
k kj

by means of the components of the tensor Tr


; analogically, the relations (3.8200 )
and (3.850 ) lead to
q
ox
i
r
ij ¼ r
^kj ; i; j ¼ 1; 2; 3: ð3:87Þ
q oxk
It follows that the tensor Tr^ is, in general, asymmetric.
Taking into account (2.1), one can write
 
q
oui

^ij ¼
rij 
rij ; i; j ¼ 1; 2; 3;
r ð3:860 Þ
q oxk
 

q
oui
rij ¼ ^ij þ
r ^kj ; i; j ¼ 1; 2; 3:
r ð3:870 Þ
q oxk
The Piola-Kirchhoff stress vector of second kind is defined by means of
Cauchy’s stress vector by the relations
n oxi n

d Pi ¼ d Pj ; i ¼ 1; 2; 3; ð3:88Þ
ox
j

which are of the same form as the relations


oxi

dxi ¼ dx ; i ¼ 1; 2; 3:
ox
j j

The relations (3.200 ), (3.800 ) and (3.81) allow to write the relation (3.88) in the
form
oxi

rij nj dA ¼ r n dA ; i ¼ 1; 2; 3; ð3:880 Þ
ox
j jk k

where rij are the components of the Piola-Kirchhoff stress tensor of second kind
(which is called the Piola-Kirchhoff stress tensor too).
106 3 Mechanics of Stresses

Using the relations (3.84), we can express the components of the tensor Tr in
the form
q oxi oxj

rij ¼ r ; i; j ¼ 1; 2; 3; ð3:89Þ
q
ox
k ox
l kl

analogically, using the relations (3.850 ), we get

q
ox
i ox
j
r
ij ¼ rkl ; i; j ¼ 1; 2; 3: ð3:90Þ
q oxk oxl
Unlike the tensor Tr^ , the tensor Tr is a symmetric tensor.
By means of the relations (2.1), we can also write
 
q ouj oui oui ouj
rij ¼
r
ij 
r
il 
r
kj þ

r
kl ; i; j ¼ 1; 2; 3; ð3:890 Þ
q oxl oxk oxk oxl
 

q
ouj oui oui ouj
rij ¼ rij þ ril þ rkj þ rkl ; i; j ¼ 1; 2; 3: ð3:900 Þ
q oxl oxk oxk oxl
Using the relations (3.86), (3.87), (3.89) and (3.90), we can express the Piola-
Kirchhoff tensor one with respect to the other by the relations
oxj
rij ¼ r
^ik ; i; j ¼ 1; 2; 3; ð3:91Þ
ox
k

ox
j
^ij ¼
r rik ; i; j ¼ 1; 2; 3; ð3:92Þ
oxk
which may be written also in the form
ouj
rij ¼ r
^ij  r
^ik ; i; j ¼ 1; 2; 3; ð3:910 Þ
ox
k

ouj
^ij ¼ rij þ
r rik ; i; j ¼ 1; 2; 3: ð3:920 Þ
oxk
The Piola-Kirchhoff stress tensors are thus entirely defined.
The formulae (3.900 ) show that, in case of the infinitesimal deformations, the
Piola-Kirchhoff tensor becomes Cauchy’s tensor.

3.2.4.3 Equations of Equilibrium and Motion

To establish the equations of equilibrium which must be verified by the compo-


nents of the tensor Tr^ , one uses the method introduced in Sect. 3.2.1. Thus, for an
elementary subdomain V 0 in the actual state, we write the relations corresponding
to the relations (3.600 ) in the form
3.2 Stress Tensor. Equations of Equilibrium and Motion 107

ZZ n
ZZZ
p
i dS0
þ Fi
dV 0
¼ 0; i ¼ 1; 2; 3;
S0
V 0

we obtain thus the relations


ZZ n ZZZ
0
pi dS þ
^ Fi dV 0 ¼ 0; i ¼ 1; 2; 3;
S0 V0

for an elementary subdomain V 0 in the initial state (Fig. 3.12), where, for the
change of co-ordinates, we took into account the relations (3.80), (3.81) and (3.82)
and the fact that the volume forces (which are given forces), expressed with
respect to the initial state and to the actual state, respectively, verify the relations
1 1
Fi ¼
Fi
; i ¼ 1; 2; 3; ð3:93Þ
q q
corresponding to the equality of the mass forces with regard to the two states.
Using the flux-divergence formula (A.1.960 ), we obtain
ZZZ  
o^
rji
þ Fi dV 0 ¼ 0; i ¼ 1; 2; 3;
V 0 oxj

wherefrom, assuming that r ^ji are functions of class C1 , while Fi are functions of
0
class C , we get the equations of equilibrium
o^
rji
þ Fi ¼ 0; i ¼ 1; 2; 3; ð3:94Þ
oxj

the volume V 0 being arbitrary.


In the dynamical case, we can write the equations of motion in the form
o^
rji
þ Fi ¼ q€
ui ; i ¼ 1; 2; 3; ð3:940 Þ
oxj

where we took into account the fact that, in the initial state, using the material
co-ordinates, the derivatives with respect to time are partial derivatives.
Using the formula (3.52), we can write the equations of equilibrium with the aid
of the tensor Tr in the form
 
o ox
i
rkj þ Fi ¼ 0; i ¼ 1; 2; 3; ð3:95Þ
oxj oxk

or in the form

orji oui orjk o2 ui


þ þ rjk þ Fi ¼ 0; i ¼ 1; 2; 3; ð3:950 Þ
oxj oxk oxj oxj oxk

where we have introduced the displacement gradient.


108 3 Mechanics of Stresses

We obtain forms analogue to (3.940 ) for the equations of motion, by replacing


the volume forces by the lost forces of d’Alembert.

3.2.5 Stresses in Curvilinear Co-ordinates

In what follows, we present, in curvilinear co-ordinates, the main results previ-


ously obtained for the stresses, in the linear case. We use thus orthogonal curvi-
linear co-ordinates and, in particular, cylindrical and spherical co-ordinates, as
well as the corresponding results of mathematical character given in the Appendix;
we use, as well, the results obtained in Sect. 3.1.2.4.
We mention that J. Larmor obtained the equations of equilibrium and motion in
curvilinear co-ordinates by a variational method.

3.2.5.1 Orthogonal Curvilinear Co-ordinates

Let us consider a point M the position of which is specified by the orthogonal


Cartesian co-ordinates xi ; i ¼ 1; 2; 3, or by the orthogonal curvilinear co-ordinates
qa ; a ¼ 1; 2; 3 (Fig. 3.16).
Starting from the formula (3.28), we find that the passing from stresses in
orthogonal Cartesian co-ordinates to stresses in orthogonal curvilinear co-ordinates
can be made by means of relations such as
  oxi oxj
rsa sb ¼ rij cosðqa ; xi Þ cos qb ; xj ¼ rij ; a; b ¼ 1; 2; 3: ð3:96Þ
osa osb
Analogically, we can express the stresses in orthogonal Cartesian co-ordinates
by means of the stresses in orthogonal curvilinear co-ordinates by the relations
  oxi oxj
rij ¼ rsa sb cosðqa ; xi Þ cos qb ; xj ¼ rs s ; i; j ¼ 1; 2; 3: ð3:960 Þ
osa osb a b
Using the relations (3.960 ) and the differential operator (A.128), the equations of
equilibrium (3.62) become
 
oxi o oxi oxj oxi
rs s þ Fs ¼ 0; i ¼ 1; 2; 3;
osc osc osa osb b a osc c

where we have used the relations (A.100 ) of transformation of the components of a


vector too; in a developed form, we have
 2 
oxi orsb sa o xi oxi oxj o2 xj oxi
þ þ rs b s a þ Fs ¼ 0; i ¼ 1; 2; 3:
osa osb osb osa osa osc osc osb osc c
3.2 Stress Tensor. Equations of Equilibrium and Motion 109

Fig. 3.16 Stresses in


orthogonal curvilinear co- s3 s3
ordinates s3 s2
s3 s1
Fs3
q3
Fs2
Fs1 s2s3
q1 q2
s1s3
s2 s2
M (x1 , x2, x3 )
s2s1
s1s2
s1s1

Multiplying by oxi =osc , summing with respect to the index i and using the
formula (A.134), we can write the equations of equilibrium in the form

orsb sa oxi o2 xi oxi o2 xi


þ rsb sc þ rs s þ Fsa ¼ 0; a ¼ 1; 2; 3; ð3:97Þ
osb osa osb osc osc osc osb b a

observing that the stress tensor is symmetric ðrsb sc ¼ rsc sb Þ, it results that the
product of this tensor by the skew-symmetric part of the tensor o2 xi =osb osc van-
ishes, so that we can write the equations of equilibrium in the form
 
orsb sa 1 oxi o2 xi o2 x i oxi o2 xi
þ þ rs b s c þ rs s þ Fsa ¼ 0; a ¼ 1; 2; 3:
osb 2 osa osb osc osc osb osc osc osb b a
Taking into account the formula (A.13600 ). (A.137), we may write the
Eq. (3.96) also in the form (without summation with respect to a)
orsb sa 1 oha 1 ohc
 rs s  rs s
osb ha osb b a hc osb b a
X3  
1 ohb ohb
þ rsb sa þ rsb sb þ Fsa ¼ 0; a ¼ 1; 2; 3: ð3:970 Þ
h osb
b¼1 b
osa

Effecting the sums above, we may also write (without summation)


orsa sa orsa sb orsc sa 1 o 1 ohb
þ þ  ðhb hc Þrsa sa þ rs s
osa osb osc hb hc osa hb osa b b
1 ohc 1 o 2 1
þ rsc sc  2 ðha hc Þrsa sb  2 ðh2a hb Þrsc sa þ Fsa ¼ 0;
hc osa ha hc osb ha hb
a 6¼ b 6¼ c 6¼ a; a ¼ 1; 2; 3: ð3:9700 Þ

These equations can be written also in the form (without summation)


110 3 Mechanics of Stresses

orsa sa orsa sb orsc sa 1 ohb


þ þ þ ðrsb sb  rsa sa Þ
osa osb osc hb osa
 
1 ohc 2 oha ohb
þ ðrsc sc  rsa sa Þ  rsa sb þ rsc sa
hc osa ha osb osc
1 ohc 1 ohb
 rs s  rs s þ Fsa ¼ 0;
hc osb a b hb osc c a
a 6¼ b 6¼ c 6¼ a; a ¼ 1; 2; 3: ð3:97000 Þ
In the dynamic case, we use the same equations, replacing the volume forces
Fsa by the lost forces of d’Alembert
 sa ¼ Fsa  q€
F us a ; a ¼ 1; 2; 3: ð3:98Þ
We obtain the same equations if we put the conditions of equilibrium for an
elementary domain, bounded by co-ordinate surfaces, cut from the body. We have
represented only the stresses on three adjacent three-orthogonal faces (with their
positive directions), in Fig. 3.16, the stresses on the other faces being obtained by
means of formula of Taylor type.

3.2.5.2 Cylindrical Co-ordinates

The relations by which one can pass from stresses in orthogonal Cartesian co-
ordinates to stresses in cylindrical co-ordinates read
1 1
rrr ¼ ðr11 þ r22 Þ þ ðr11  r22 Þ cos 2h þ r12 sin 2h;
2 2
1 1 ð3:99Þ
rhh ¼ ðr11 þ r22 Þ  ðr11  r22 Þ cos 2h  r12 sin 2h;
2 2
rzz ¼ r33 ;

rhz ¼ r23 cos h  r31 sin h;


rzr ¼ r23 sin h þ r31 cos h; ð3:990 Þ
1
rrh ¼ ðr11  r22 Þ sin 2h  r12 cos 2h:
2
The equations of equilibrium corresponding to an elementary subdomain rep-
resented in Fig. 3.17a will be of the form
orrr 1 orrh orzr 1
þ þ þ ðrrr  rhh Þ þ Fr ¼ 0;
or r oh oz r
orrh 1 orhh orhz 2
þ þ þ rrh þ Fh ¼ 0; ð3:100Þ
or r oh oz r
orzr 1 orhz orzz 1
þ þ þ rzr þ Fz ¼ 0:
or r oh oz r
3.2 Stress Tensor. Equations of Equilibrium and Motion 111

Fig. 3.17 Elementary (a) zz


equilibrium in cylindrical
(a) and spherical (b) co-
ordinates z

zr

dz
rz
Fz F
r
r Fr

d z z rr

M r

(b)

R
R
FR RR
M
R
R
d

F
d

dR

3.2.5.3 Spherical Co-ordinates

We can pass from stresses in orthogonal Cartesian co-ordinates to stresses in


spherical co-ordinates, by formula of the form

rRR ¼ rrr sin2 u þ rzz cos2 u þ rzr sin 2u


1 1
¼ ðr11 þ r22 Þ sin2 u þ ðr11  r22 Þ sin2 u cos 2h
2 2
þ r33 cos u þ r23 sin 2u sin h þ r31 sin 2u cos h þ r12 sin2 u sin 2h;
2

ruu ¼ rrr cos2 u þ rzz sin2 u  rzr sin 2u


1 1
¼ ðr11 þ r22 Þ cos2 u þ ðr11  r22 Þ cos2 u cos 2h ð3:10Þ
2 2
þ r33 sin2 u  r23 sin 2u sin h  r31 sin 2u cos h þ r12 cos2 u sin 2h;
1 1
rhh ¼ ðr11 þ r22 Þ  ðr11  r22 Þ cos 2h  r12 sin 2h;
2 2
112 3 Mechanics of Stresses

1
ruh ¼  rhz sin u  rrh cos u ¼  ðr11  r22 Þ cos u sin 2h
2
 r23 sin u cos h þ r31 sin u sin h þ r12 cos u cos 2h;
1
rhR ¼ rhz cos u  rrh sin u ¼ ðr11  r22 Þ sin u sin 2h
2
þ r23 cos u cos h  r31 cos u sin h þ r12 sin u cos 2h; ð3:1010 Þ
1 1
rRu ¼ ðrrr  rzz Þ sin 2u þ rzr cos 2u ¼ ðr11 þ r22 Þ sin 2u
4 4
1 1
þ ðr11  r22 Þ sin 2u cos 2h  r33 sin 2u
4 2
1
þ r23 cos 2u sin h þ r31 cos 2u cos h þ r12 sin 2u sin 2h;
2
where we have put in evidence the linkage with the cylindrical co-ordinates too.
For an elementary subdomain, represented in Fig. 3.17b, the equations of
equilibrium read
orRR 1 orRu 1 orhR
þ þ
oR R ou R sin u oh
1
þ ð2rRR  ruu  rhh þ rRu cot uÞ þ FR ¼ 0;
R
orRu 1 oruu 1 oruh
þ þ
oR R ou R sin u oh
ð3:102Þ
1
þ ½ðruu  rhh Þ cot u þ 3rRu  þ Fu ¼ 0;
R
orhR 1 oruh 1 orhh
þ þ
oR R ou R sin u oh
1
þ ð2ruh cot u þ 3rhR Þ þ Fh ¼ 0:
R

References

A. Books

1. Grioli, G.: Mathematical Theory of Elastic Equilibrium (Recent Results). Ergebnisse der
angewandten Mathematik. Springer, Berlin (1962)
2. Mohr, O.: Technische Mechanik, 2nd edn. Verlag von W. Ernst & Sohn, Berlin (1914)
3. Navier, L.M.H.: Résumé des leçons sur l’application de la mécanique à l’établissement des
constructions et des machines, 3rd edn. Avec des notes et des appendices de B. de Saint-
Venant, Paris (1864)
4. Teodorescu, P.P.: Dynamics of Linear Elastic Bodies (Ed. Academiei, Bucuresßti). Abacus
Press, Tunbridge Wells (1975)
References 113

B. Papers

5. Borchardt, C.W.: Uber} die Transformation der Elasticitätsgleichungen in allgemeine


orthogonale Coordinaten. J. f} ur die reine und angew. Math. (Crelle) 76, 45 (1873)
6. Cauchy, A.-L.: Recherches sur l’équilibre et le mouvement intérieur des corps solides ou
fluides, élastiques ou non élastiques. Bull. Soc. Philomath., Paris, 9 (1823)
7. Cauchy, A.-L.: De la pression ou tension dans un corps solide. Ex. de Math. 2, 42 (1827)
8. Cauchy, A.-L.: Sur les équations qui expriment les conditions d’équilibre ou les lois du
mouvement intérieur d’un corps solide, élastique ou non élastique. Ex. de Math. 3, 160
(1828)
9. Cauchy, A.-L.: Sur l’équilibre et le mouvement intérieur des corps considérés comme des
masses continues. Ex. de Math. 4, 293 (1829)
10. Cauchy, A.-L.: Sur les diverses méthodes à l’aide desquelles on peut établir les équations qui
représentent les lois d’équilibre on le mouvement intérieur des corps solides ou fluides. Bull.
Sci. Math. Soc. Prop. Conn. 13, 169 (1830)
11. Cauchy, A.-L.: Note sur l’équilibre et les mouvements vibratoires des corps solides. C. Rend.
hebd. des séances de l’Acad. Sci. 32, 323 (1851)
12. Gwither, R.F.: The formal specification of the elements of stress in Cartesian and in
cylindrical and spherical polar co-ordinates. Mem. Manchester Lit. Phil. Soc. 56, 10 (1912)
13. Gwither, R.F.: The specification of the elements of stress. Part II. A simplification of the
specification given in Part I. Mem. Manchester Lit. Phil. Soc. 57, 5 (1913)
14. Lamé, G., Clapeyron, B.-P.-E.: Mémoire sur l’équilibre des corps solides homogènes. Mém.
divers savants Acad. Sci. 4, 465 (1833)
15. Moisil, Gr. C.: Sur les petits mouvements des solides élastiques. Disquisitiones Mat. et Phys.
1, 83 (1940)
16. Navier, C.-L.-M.-H.: Sur les lois de l’équilibre et du mouvements des corps solides
élastiques. Bull. Soc. Philomath. 177 (1833)
17. Navier, C.-L.-M.-H.: Mémoire sur les lois de l’équilibre et du mouvements des corps solides
élastiques. Mém. Acad. Sci. Inst. France 7, 375 (1827)
18. Poisson, S.-D.: Mémoire sur l’équilibre et le mouvement des corps élastiques. Mém. Acad.
Sci. 8, 356 (1829)
19. Poisson, S.-D.: Addition au mémoire sur l’équilibre et le mouvement des corps élastiques.
Mém. Acad. Sci. 8, 623 (1829)
20. Reissner, E.: Note on the theorem of symmetry of the stress tensor. J. Math. Phys. 23, 192
(1944)
Chapter 4
Mathematical Models in Mechanics
of Deformable Solids

The theoretical elements of a model in mechanics of deformable solids were


described previously; it can be seen that these elements are generally closer to the
ideal models of second order, but they must be supplemented by some data of
experimental nature.
The state of deformation (the state of displacement) and the state of stress are
components of the model we are constructing. The latter is completed by intro-
ducing a relation between these states. The modelling of the physical behaviour of
the real bodies shall thus be put in a concrete form by introducing a dependence
such as cause-effect, force-deformation, stress-strain, i.e. by establishing a con-
stitutive law of the respective deformable solid.
For idealized types of materials, various ideal models of the first order are
conceived, which can be divided into two categories: dynamic models and kine-
matic models; the first category comprises the models that, besides the notion of
displacement, make use of the notion of force, while the second category com-
prises the models which make use of only the notion of displacement.

4.1 Elastic Models

By the ‘‘ut tensio sic vis’’ anagram, published by R. Hooke in 1678, the first model
of a deformable solid, the Hookean model, the ideal (perfect) elastic solid appears;
it is assumed that the solid undergoes a deformation in direct proportion to the
acting force, which disappears at the same time as the latter one (reversible
deformation).
In what follows, we make a general study concerning the construction of a
constitutive law of elastic solids, taking also into account considerations of ther-
modynamical nature, laying stress on the case of linear elasticity.

P. P. Teodorescu, Treatise on Classical Elasticity, Mathematical and Analytical 115


Techniques with Applications to Engineering, DOI: 10.1007/978-94-007-2616-1_4,
Ó Springer Science+Business Media Dordrecht 2013
116 4 Mathematical Models in Mechanics of Deformable Solids

4.1.1 Construction of the Elastic Model

Let us introduce, first of all, a linear relation, independent of time, corresponding


to a one-dimensional experiment, of the form (Hooke’s law)
r ¼ Ee; e  0; ð4:1Þ
or
r ¼ Eeþ ¼ EehðeÞ; ð4:10 Þ
where r is the normal stress, e is the linear strain, E is the modulus of longitudinal
elasticity (Fig. 4.1a) and where we used the functions introduced by the formulae
(A.164) and (A.166). The corresponding dynamical model (ideal, of first order) is a
spring, fixed at one end and acted upon by a force P at the other end; this force and
the corresponding displacement D will represent the stress and the strain,
respectively, which arise (Fig. 4.1b). We make, as well, some considerations with
a general non-linear character.

4.1.1.1 General Considerations

It is as a consequence of experimental results that the ideal model of the elastic


body then appeared. Despite the fact that no such real bodies exist, the Hooke
model was, and still is, necessary in the process of knowledge; it reflects the reality
since, within certain limits, the bodies do have this property. The theory of elas-
ticity was created on the basis of this model. The solutions obtained for the
problems which have arisen within the frame of this theory are compared with the
physical reality; thus, we shall be able to see how well the model reflects objec-
tively the reality and what are the limits of its applicability.
The Hooke model considered above corresponds to linearly elastic, isotropic
and homogeneous bodies. In the case of anisotropic bodies, this model must be
improved by assuming a general linear relation between the components of the
strain tensor and the components of the stress one. Moreover, in the case of non-
homogeneous bodies, the elastic coefficients are no longer constant, but become
functions of a point; they can be also functions of time in the case of bodies the
properties of which change in time.

Fig. 4.1 Hooke’s law:


geometric representation (a);
technical model (b)

P
arctan E
O
(a) (b)
4.1 Elastic Models 117

Fig. 4.2 Non-linear


+( )
unidimensional constitutive
law

We mention the equal possibility of occurence of a physical non–linearity


which, in the one-dimensional case considered above, is expressed by a relation of
the form (Fig. 4.2)
r ¼ rðeÞ; e [ 0; ð4:100 Þ

which also reads


r ¼ rþ ðeÞ ¼ rðeÞhðeÞ: ð4:1000 Þ
In the following, we consider only isotropic and homogeneous, linearly elastic
bodies.

4.1.1.2 Generalized Hooke’s Law

We shall now specify the mathematical model of the elastic body, by adding to the
geometrical and mechanical aspect of the problem, an aspect of physical and
experimental nature; for this end, we shall introduce a constitutive law, a relation
between stresses and strains, a relation that (to avoid computation difficulties),
should be as simple as possible (obviously, within the limits of the admitted
hypotheses concerning the body under consideration).
We shall thus admit as constitutive law Hooke’s law corresponding to the one-
dimensional case. Generally, a stress on a direction yields deformations on all
directions; these deformations can be acquired by superposition of effects (the
linear case), by using the generalized Hooke law (which we shall simply call
Hooke’s law) that shall therefore have a linear form.
Let be a parallelepipedical infinitesimal element within the body, the faces of
which are parallel to the planes of co-ordinates. Let us suppose that only the
normal stress r11 appears (Fig. 4.3a); owing to this stress, a linear strain is born
1
e11 ¼ r11 ; ð4:2Þ
E
where E bears the name of modulus of longitudinal elasticity, as it has been
mentioned in the previous subsection. This elastic constant was introduced and
discussed by Euler; it is known by the name of Young’s modulus, although he did
118 4 Mathematical Models in Mechanics of Deformable Solids

23
11

dx3

dx3
23

23

11
dx 1 32 dx 1
dx2 dx2
(a) (b)

Fig. 4.3 Generalized Hooke’s law: normal stress (a); tangential stress (b)

not define it. Assuming that the stress r11 corresponds to a tension, along the Ox2 -
axis and the Ox3 -axis appear linear strains (shortenings), expressed by
m
e22 ¼ e33 ¼ me11 ¼  r11 ; ð4:20 Þ
E
where m is a coefficient of transverse contraction (Poisson’s ratio). We assume
that, due to the normal stress, there do not appear angular strains.
If on the faces of the parallelepiped appear only the tangential stresses r23 ¼
r32 (Fig. 4.3b), in a plane parallel to the plane Ox2 x3 appears the angular strain
1
c23 ¼ 2e23 ¼ r23 ; ð4:200 Þ
G
where G is the modulus of transverse elasticity. This tangential stress does not give
rise to other strains.
The stresses r22 ; r33 ; r31 ¼ r13 ; r12 ¼ r21 lead to analogue strains. Using the
principle of superposition of effects, the constitutive law (Hooke’s law) reads
1
e11 ¼ ½e11  mðe22 þ e33 Þ;
E
1
e22 ¼ ½e22  mðe33 þ e11 Þ; ð4:3Þ
E
1
e33 ¼ ½e33  mðe11 þ e22 Þ;
E
1
c23 ¼ 2e23 ¼ r23 ;
G
1
c31 ¼ 2e31 ¼ r31 ; ð4:30 Þ
G
1
c12 ¼ 2e12 ¼ r12 ;
G
4.1 Elastic Models 119

where are involved three elastic constants of the material; we shall see that these
constants are linked by a relation, only two of them being independent.
For this, we assume that the principal directions of the stress tensor coincide
with the principal directions of the strain tensor (we will show in Sect. 4.1.3.1 that
this takes place always in case of linearly elastic, isotropic bodies).
In case of a simple shear (see Sect. 3.2.3.3), the state of stress with respect to
the principal axes is expressed in the form
r1 ¼ r2 ¼ s; r3 ¼ 0;

where s is a tangential stress; there corresponds a simple angular strain (see Sect.
2.2.3.3) given by
1
e1 ¼ e2 ¼ c; e3 ¼ 0;
2
where c is an angular strain. Taking into account Hooke’s law (4.3, 4.30 ), we may
write
1 1þm 1
e1 ¼ ðr1  mr2 Þ ¼ s; c¼ s;
E E G
which leads to the relation
E
G¼ ; ð4:4Þ
2ð1 þ mÞ
that takes place between the three elastic constants.
Hooke’s law (4.3), (4.30 ) can be written in the form
 
1 1 m
eij ¼ ½ð1 þ mÞrij  mrkk dij  ¼ rij  rkk dij ; i; j ¼ 1; 2; 3; ð4:300 Þ
E 2G 1þm

or
 
1 m
Te ¼ Tr  trTr d : ð4:3000 Þ
2G 1þm

4.1.1.3 Experiments in the One-Dimensional Case. The Characteristic


Curve of the Material

We make a simple experiment for a one-dimensional loading on a cylindrical


sample of a given material, where the relation between the stress and the strain is a
direct one; the state of stress of the sample is a simple tension (Fig. 4.4a) or a
simple compression (Fig. 4.4b), being obtained with the help of a general machine
for mechanical experiments. The curve which represents geometrically this
relation in the plane Oer is called the characteristic curve of the material.
120 4 Mathematical Models in Mechanics of Deformable Solids

l l*
l* l
(a) (b)

Fig. 4.4 One-dimensional experiment: tension (a); compression (b)

The respective diagrams may be conventional or real as, in the computation, the
stress is obtained with respect to the initial cross section (the most used case) or to
the actual cross section. In these diagrams, the strain can be considered in the sense
of Cauchy (e ¼ ðl  lÞ=l, where l is the initial length, while l is the actual one) or
in the sense of Hencky (e ¼ lnðl =lÞ); we will consider the first sense (Fig. 4.5a).
The experiments are made in certain standard conditions, that is at the same
temperature, humidity and at the same radioactive radiations; the loading is made
with the same loading velocity, the dimensions of the sample are the same etc.
Only in such conditions the results which are obtained are comparable.
On the characteristic curve, we consider some remarkable points, which mark
the limits of significant parts from the mechanical point of view. We mention thus
the limit stress of elasticity, denoted by re , beyond which the plastic deformations,
which are smaller than the elastic ones, can no more be neglected; theoretically, till
this point the deformation is reversible. The position of this point depends on the
approximation requested by each experiment.
At certain materials, one can put in evidence also a limit stress of propor-
tionality, denoted by rp , till which the curve may be approximated by a straight
line.
Beyond re we find a point called critical stress of plasticity (yield point),
denoted by rY . The deformation process beyond this point, where the plastic
deformations are approximately of the same order of magnitude as the elastic ones,
is called plastic deformation. If the plastic deformations become very large, so that
the elastic ones can be neglected with respect to the first ones, then the corre-
sponding process is called plastic flow. To these two processes correspond two
portions of the characteristic curve, separated by a point of a somewhat conven-
tional position, denoted by rf and called limit stress of flow. For some materials,
the elastic deformations can be neglected after the beginning of the process of
plastic deformation; in this case, the limit of flow coincides practically with the
yield point. This coincidence takes place, e.g., for the materials the characteristic
curve of which has a long horizontal bearing.
Passing from the elastic domain to the plastic one is, customarily, progressive,
gentle. But there exist also some materials for which this passing is sudden; in this
case, the limit of elasticity has no more a conventional character, but corresponds
to a well definite physical phenomenon.
If beyond the yield point the characteristic curve of a material is strictly
increasing, then the respective material has the property of stress hardening;
4.1 Elastic Models 121

Fig. 4.5 Characteristic curve C'


of the material (a); hysteresis B
phenomenon (b) A
y
f C A
Y
e
p

A'
O p e
A'
O
(a) (b)

during the plastic deformation, the limit stress of elasticity is increasing for such
materials. Indeed, for an unloading from the point A, e.g., this one is linearly
elastic; beginning again the loading, the diagram which is obtained is perfectly
superposed over that corresponding to the unloading, the new limit stress of
elasticity (corresponding to the point A) being greater than the initial one.
The perfectly elastic unloading represents an idealized schema of the unloading
phenomenon, being—the most times—in a good harmony with the experiment.
Although it does not correspond to the reality in some case, either because an
important phenomenon of hysteresis of the material (by transformation of the
mechanical energy into a caloric one) is put in evidence (Fig. 4.5b) or because the
unloading is an elastic-plastic process.
To represent the property of hardening of a material, the function (4.100 ) must be
continuous, monotone increasing and have a derivative with a finite number of
points of discontinuity.
In some cases, the elastic deformations can be neglected with respect to the
plastic deformations; if the body is a hardening one, then it is called a plastic rigid/
hardening body.
If the stress increases and we reach the point B on the characteristic curve
(Fig. 4.5a), then—the most times—appears a striction (a sudden reduction of the
cross section of the sample), while at the point C takes place the breakage
(corresponding to a conventional diagram); in case of a real diagram, the breakage
is produced at the point C0 . In the considered case, the breakage is a viscous one; if
there is no striction, then the breaking is fragile.
In general, the characteristic curves have their concavity directed towards the
Oe-axis, as in Fig. 4.5; there are also materials for which the concavity is directed
entirely or partially towards the Or-axis, e.g., the natural rubber (Fig. 4.6a).
We have considered above a phenomenon of tension; if the sample is subjected
to compression, then one obtains, the most times, a similar characteristic curve.
But there exist bodies which resist better to compression than to tension, e.g., the
rocks, the concrete etc. (Fig. 4.6b).
122 4 Mathematical Models in Mechanics of Deformable Solids

Fig. 4.6 Non-conventional


characteristic curves: natural
rubber (a); concrete (b)

O
(a) (b)

The study of the characteristic curve of a material is particularly complex and puts
in evidence the most parts of its mechanical properties; in the above considerations
we have made only a short presentation of these properties, being interested to
specify the limit stress of elasticity re , corresponding to the passing from the
elastic zone to the plastic one.
From the Fig. 4.5a, we notice that
e ¼ ee þ ep ; ð4:5Þ

where we put in evidence the elastic and the plastic part of the strain. Taking into
account the relative order of magnitude of the strains, there results various pos-
sibilities of study of the respective bodies in the plastic zone.

4.1.2 Elastic Potential. Green’s Theory

In the following, we present Hooke’s law in the general case of an arbitrary


anisotropic, linearly elastic body. We admit thus that between the stress tensor and
the strain tensor there exists a linear relation of the form
Tr ¼ TH Te ; ð4:6Þ
where TH is the tensor of elastic coefficients (Hooke’s tensor) of fourth degree. It
remains to show that such a law corresponds to the conditions of thermodynamical
order.

4.1.2.1 Elastic Potential. Green’s Formulae

The hypothesis of the local dependence of the stresses on the strains leads, both for
loading and unloading, to a relation of the form
rij ¼ rij ðe11 ; e22 ; e33 ; e23 ; e31 ; e12 ; x1 ; x2 ; x3 ; TÞ; i; j ¼ 1; 2; 3; ð4:7Þ

where the functions are considered to be uniform (to a state of strain corresponds a
state of stress), while the absolute temperature T is a parameter.
4.1 Elastic Models 123

The relations (4.7) must be invertible for the elastic bodies, hence
eij ¼ eij ðr11 ; r22 ; r33 ; r23 ; r31 ; r12 ; x1 ; x2 ; x3 ; TÞ; i; j ¼ 1; 2; 3; ð4:70 Þ

this takes place if the Jacobian of the respective relations is non-zero and bounded
 
oðr11 ; r22 ; . . .; r12 Þ

0 6¼  \1: ð4:700 Þ
oðe11 ; e22 ; . . .; e12 Þ 
Thus, the elastic deformation is reversible; if a body is passing for the second
time through a given state of stress, then it passes, at the same time, through the
corresponding state of strain.
If the above functions do not depend on the point, then the body is
homogeneous.
The elastic bodies for which, in certain given conditions (see Sect. 4.1.2.6 too),
there exists an elastic potential are called hyperelastic bodies. The hyperelastic
state depends both on the material and on the conditions of the deformation
process, which can be described only in mechanical terms, obviously the same for
the loading and the unloading of the body. We notice that the theory of elasticity
was developed especially as a theory of hyperelastic bodies, but not any elastic
body is a hyperelastic one. But in the frame of the linear theory, as we will see, any
elastic body is hyperelastic too; such bodies will be called, shortly, elastic bodies.
We mention the existence of some more complex constitutive laws, which
represent more general relations between aspects of geometric-kinematic nature
and aspects of mechanical nature (static or dynamic aspects). Thus, the hypoelastic
bodies are characterized by a one-to-one dependence between the variations of the
stress and the variations of the strain at any moment; for such bodies there does not
exist a natural state, while the components of the stress variation depend on the
component of the strain variation, but not on their velocity variation, because the
hypoelastic bodies do not have viscous properties. The hypoelastic deformation is
reversible and in the neighbourhood of a state without stresses, in case of infini-
tesimal deformations, the respective body has elastic properties. As well, there
exist anelastic bodies for which the deformations are reversible and linearly linked
to stresses, the temporal variable being involved too; the correspondence between
stresses and strains is no more one-to-one.
Another interesting case is that in which the stress tensor is asymmetric (case of
non-Boltzmannian bodies); this leads to the apparition of a supplementary free
rotation, each particle of the body having six degrees of freedom. Thus is born the
theory of asymmetric elasticity; the elastic bodies of Cosserat type are situated in
the frame of this theory.
We admit, in what follows, the existence of a potential W, which depends only
on the components of the tensor Te (and on point, in case on non-homogeneous
bodies), being a function of class C1 with respect to these six variables
W ¼ Wðe11 ; e22 ; . . .; e12 Þ; ð4:8Þ
124 4 Mathematical Models in Mechanics of Deformable Solids

this is Green’s elastic potential (for a hyperelastic body). This form corresponds to
the reversible character of the deformation and to the hypothesis of local depen-
dence. We assume, as well, that the state of stress is given by Green’s formulae [3]
1  oW
rij ¼ 1 þ dij ð!Þ; i; j ¼ 1; 2; 3; ð4:9Þ
2 oeij

in a developed form, we have


oW oW oW
r11 ¼ ; r22 ¼ ; r33 ¼ ; ð4:90 Þ
oe11 oe22 oe33
1 oW oW 1 oW oW 1 oW oW
r23 ¼ ¼ ; r31 ¼ ¼ ; r12 ¼ ¼ : ð4:900 Þ
2 oe23 oc23 2 oe31 oc31 2 oe12 oc12
One can say that the state of stress derives from the elastic potential W:
The differential of the elastic potential W reads
oW
dW ¼ deij ; i  j; ð4:10Þ
oeij

the condition imposed to the indices is put because the elastic potential depends
only on six independent variables, the tensor Te being symmetric. Taking into
account Green’s formulae (4.9), we may write
dW ¼ rij deij ð4:100 Þ

too, the sum involving now all the values i; j ¼ 1; 2; 3.


Because to know the components of the strain tensor is equivalent to the
knowing of the principal linear strains (or to the invariants of the tensor Te ), as
well as of the position of the principal axes, it results
W ¼ Wðe1 ; e2 ; e3 ; w; h; uÞ ð4:80 Þ

or
W ¼ WðI1 ; I2 ; I3 ; w; h; uÞ; ð4:800 Þ
where w; h; u are Euler’s angles which specify the principal axes; if W does not
depend explicitly on these angles, then the body is isotropic.
The formulae (4.8) (or (4.80 ) or (4.800 )) and (4.9) constitute a constitutive law for
the elastic bodies (law of hyperelastic nature).

4.1.2.2 Elementary Work of Deformation

The elementary work of a force P, which acts upon the elastic body at a point of
position vector r (in the actual state), is given by
4.1 Elastic Models 125

dW ¼ P  dr ; ð4:11Þ

taking into account (2.1) and observing that this work is expressed by means of
space co-ordinates, we may also write
dW ¼ P  du ¼ Pi dui : ð4:12Þ
Let be a state of strain and stress characterized by the quantities ui ; eij ; rij ,
i; j ¼ 1; 2; 3, due to arbitrary external loads; in case of a differential variation, the
new state of strain and stress is characterized by the quantities
ui þ dui ; eij þ deij ; rij þ drij ; i; j ¼ 1; 2; 3:
Let us consider a parallelepipedic infinitesimal element, with the faces parallel
to the co-ordinate planes, acted upon only by the stress r11 (Fig. 4.3a); in this case,
the elementary work reads
ðr11 dx2 dx3 Þdðe11 dx1 Þ ¼ ðr11 dx2 dx3 Þde11 dx1 ¼ r11 de11 dV;

where we took into account that the distance between the two faces of the par-
allelepiped which are parallel to the plane Ox2 x3 vary by e11 dx1 . Analogically, if
on the faces of the elementary parallelepiped appear only the tangential stresses
r23 ¼ r32 (Fig. 4.3b), then the elementary work is given by

ðr23 dx1 dx3 Þdða23 dx2 Þ þ ðr32 dx1 dx2 Þdða32 dx3 Þ
¼ r23 dc23 dV ¼ 2r23 de23 dV ¼ ðr23 de23 þ r32 de32 ÞdV;

where we took into account the considerations made in Sect. 2.2.1.3 (see Fig. 2.6
too). By analogous procedure for the stresses r22 ; r33 ; r31 ¼ r13 ; r12 ¼ r21 , we get
the elementary work, corresponding to an element of volume equal to unity, in the
form
dW ¼ rij deij ; ð4:13Þ

this is called the elementary work of deformation (the volume density of strain
energy).
Taking into account (4.100 ), we notice that (4.13) has also another mechanical
significance (corresponding to the elastic potential W); for this reason, one uses the
same notation.
The dual expression
dWc ¼ eij drij ð4:14Þ

is called the elementary complementary work of deformation; if eij and rij are seen
as independent parameters, we notice that
dW þ dWc ¼ dðeij rij Þ: ð4:15Þ
126 4 Mathematical Models in Mechanics of Deformable Solids

4.1.2.3 Considerations on Thermodynamics of Deformation

We make some considerations on thermodynamics of deformation, assuming that


we study processes of deformation in which the time variable does not play any
rôle. However, we can imagine various processes of deformation, various models
to pass from the initial state to the actual one. It is obvious that, in a theory which
neglect the history of the process of deformation, the various modes to pass from a
state to another one must be equivalent [1, 13].
We have seen that, for the deformation of the body, a work of deformation
(possibly also other forms of energy: thermic, electromagnetic etc.) is conserved.
We cannot assert, in general, that the elementary work of deformation is a total
differential; one cannot state, in general, that the work of deformation depends
only on the initial and actual states and does not depend on how the actual state has
been reached, hence on the process of deformation.
To pass beyond this difficulty, one can replace the purely mechanical study of
the problem by the study of a problem of phenomenological thermodynamics,
which can give supplementary information on the initial and actual states. Another
way is that of imposing supplementary (restrictive) conditions; one is thus led to
the consideration of hyperelastic bodies, for which the elementary work of
deformation is a total differential (the work of deformation depends only on the
initial and actual states).
A system which can receive from (or give to) the surrounding medium work,
heat, electric or magnetic charges, matter etc., is called thermodynamic system.
Such a system which changes with the surrounding medium energy and matter is
called an open system; if the system changes only energy, but not matter, then it is
called a closed system, while if it does not change neither energy, nor matter, then
it is called an isolated system. The solid body considered till now constitutes just
such a system.
The thermodynamic parameters are the quantities the knowledge of which at a
given moment is equivalent to the knowledge of the state of the system. A ther-
modynamic process is a phenomenon which leads to the change of the values of
these parameters.
A function which depends on the state of the body at a given moment, but not
on how this state has been reached, is called function of state. The state with regard
to which one can measure the variation of these functions is called natural (or null
or normal) state.
To establish some equations of state (relations between the thermodynamic
parameters) is equivalent to the modelling of certain type of bodies, the models
obtained being valid between certain limits of the parameters. The independent
parameters that are not involved are called generalized thermodynamic co-ordi-
nates. For instance, the relations (4.7)–(4.700 ) represent the equations of state of the
elastic body, being valid both for its loading and its unloading; such, seven gen-
eralized co-ordinates may be eij ; T or rij ; T; i; j ¼ 1; 2; 3.
4.1 Elastic Models 127

In what follows, we will consider only quantities with respect to the unit
volume, without specifying this; the values corresponding to the whole domain
occupied by the body are obtained by integration on V. The elementary variations
from a state to another one are denoted by d, being differentials (exact or not ones);
the non-elementary variations are obtained by integrating elementary variations on
paths C in the thermodynamic parameters space (figurative space).

4.1.2.4 Principles of Classical Thermodynamics

If we exclude the friction, the phenomenon of hysteresis (the loss of energy by


loading or unloading of the body), the viscosity, the temperature gradients, the
dependence of the properties of the body on the temperature, then the properties of
the energy can be described in two different modes. Thus, if several systems act
one upon the other in various modes, the set of these systems being isolated from
the rest of the universe, then the sum of the corresponding energies remains
constant. As well, if only one system is interacting with the rest of the universe,
then the increase of energy of this system is equal to the work effected upon it by
the rest of the universe. These two descriptions of the phenomenon are equivalent
in the above mentioned idealized conditions. The first formulation corresponds to a
conservation of the total energy, based on the classical experiments of Joule.
Let us consider in detail the interaction between two systems, supposing they
are isolated from the rest of the universe; if we use the indices 1 and 2 to specify
the two systems, then we may write
du1 þ du2 ¼ 0; ð4:16Þ
where u is the total energy. If the idealized conditions mentioned above are not
fulfilled, then between the two systems there exists also another change of energy
that the change by the work; such a change of energy is determined by a difference
of temperature and is called heat. If w21 is the work effected by the system 2 upon
system 1 and if we denote by q21 the flux of heat from 2 to 1 and conversely, then
we may write
du1 ¼ dw21 þ dq21 ; du2 ¼ dw12 þ dq12 ; ð4:17Þ

dw12 þ dw21 ¼ 0; dq12 þ dq21 ¼ 0: ð4:170 Þ


These relations constitute the first principle of thermodynamics.
The sign of q is determined by the difference of temperature between 1 and 2;
we define, by convention, the sign of a difference of temperature so that the flux of
temperature be directed from the highest temperature to the lowest one.
Passing now from the case of a general interaction between a thermodynamic
system and the rest of the universe, we can write a relation of the form

dU þ dT þ dV ¼ dW þ dQ þ dQ0 ; ð4:18Þ
128 4 Mathematical Models in Mechanics of Deformable Solids

where U is s function of state called internal energy, while dQ is the elementary


flux of heat. We have put in evidence the kinetic energy T and the potential energy
V of the macroscopic motion; as well, we mention the elementary work dW
effected upon the system and the elementary work of non-mechanical and non-
thermical nature (e.g., electromagnetical nature) dQ0 .
If we restrict ourselves to the case of quasistatic processes of bodies in mac-
roscopic equilibrium for which the elementary work of non-mechanical and non-
thermical nature is negligible, we may write the relation (4.18) in the form
dU ¼ dW þ dQ: ð4:19Þ

For the whole system we have


ZZZ ZZZ ZZZ
dUdV ¼ dWdV þ dQdV: ð4:190 Þ
V V V

In particular, in the case of an isolated system, there results, for any elementary
process,
ZZZ
dUdV ¼ 0; ð4:20Þ
V

while the internal energy does not vary; the result remains valid for a non-ele-
mentary process (by integration on an arbitrary path in the figurative space)
Z ZZZ
dUdV ¼ 0: ð4:200 Þ
C V

If two systems are separated by a frontier which allows the heat to pass freely
from one to the other, then we say that the two systems are in thermal contact.
Otherwise, if the frontier is preventing any change of heat, then one says that it is
thermic insulating; if a system is bounded by a thermic insulating frontier, then it
is thermic insulated.
A thermodynamic process which takes place in a thermic insulated system is
called adiabatic; we may write
dQ ¼ 0; dU ¼ dW; ð4:21Þ
as well as
ZZZ ZZZ ZZZ
dQdV ¼ 0; dUdV ¼ dWdV: ð4:210 Þ
V V V

In this case, there will not exist a flux of heat, the system being in thermic
equilibrium.
We notice also that the first principle of thermodynamics allows to establish the
mechanical equivalent of heat; thus from a dimensional point of view, we have
½calorie ¼ ML2 T2 .
4.1 Elastic Models 129

After Planck, we can consider three types of thermodynamic processes: natural


(irreversible) processes, non-natural processes and reversible processes. The
irreversible processes are those which take place in nature in the actual state; they
lead to an equilibrium. A non-natural process is that which removes us form the
equilibrium; such a process takes never place. The reversible processes constitute
a limit case; they correspond to a passing, in any sense, through several states of
equilibrium. Reversible processes do not take place in the nature but, by a small
variation of the given conditions, one can always obtain a natural process, differing
very little from a reversible one. A process of passing from the initial state to the
actual one is reversible if it can be developed in the inverse sense too, passing
through the same intermediate states. The quasistatic deformation is a reversible
process.
We call phase a system or a part of a system which is completely homogeneous.
An extensive property is a property the value of which for the whole system is
equal to the sum of the values corresponding to each phase; its value for a given
phase is in direct proportion to the magnitude of the phase. The mass, the volume,
the total energy are extensive properties. An intensive property is a property the
value of which is constant for a phase; its value for a given phase is independent of
its magnitude. The density, the pressure, the temperature are intensive properties.
After Clausius, to describe a thermodynamic process we introduce an extensive
property called entropy and denoted by S. The variation dS of this function of state
is formed by the variation de S due to the interaction with the rest of the universe
and by the variation di S due to the phenomena which take place in the interior of
the system itself; we may write
dS ¼ de S þ di S: ð4:22Þ
One admits that the variation de S is given by
dQ
de S ¼ ; ð4:23Þ
T
where T is the absolute temperature (with regard to -273.15 °C as origin of the
temperature (the lowest possible temperature, zero absolute) at the Kelvin scale).
We have
di S  0 ð4:230 Þ

for the variation di S, the equality taking place only for the reversible processes; in
case of the inequality the process is irreversible.
From (4.22)–(4.230 ) it result that
dQ
dS ¼ ; ð4:24Þ
T
for a reversible system, as well as
130 4 Mathematical Models in Mechanics of Deformable Solids

dQ
dS [ ; ð4:240 Þ
T
for an irreversible system; these relations represent the second principle of
thermodynamics.
For the whole system, one can write
ZZZ ZZZ
dQ
dSdV  dV: ð4:2400 Þ
V V T

In case of a reversible system, the relation (4.24) for the function of state S
shows, from the mathematical point of view, that the elementary flux of heat dQ
admits as integral factor the inverse of the absolute temperature T.
This principle can be obtained starting from various hypotheses of physical
nature. Thus, Clausius obtains this result assuming that it is not possible without
the intervention of another process. But Kelvin obtains it starting from another
idea: it is impossible that the only consequence of a sequence of changes of a
system does not consist in the whole transformation of a quantity of heat in work.
Carathéodory uses a more general hypothesis; he assumes that it is impossible to
reach all the thermodynamic states in the neighbourhood of a given arbitrary initial
state by an adiabatic process. The second principle of thermodynamics is equiv-
alent to the impossibility to construct a perpetum mobile.
The two principles stated above form the basis of classical thermodynamics;
they are experimentally justified for a real system at a macroscopic scale.
The thermodynamic processes for which T ¼ const, hence
dT ¼ 0; ð4:25Þ

are called isothermic; in case of an isothermic reversible process, dQ is a total


differential.
The second principle of thermodynamics puts in evidence a sense of devel-
opment of the irreversible processes, i.e. that corresponding to the increasing of
the entropy in any thermic isolated system; indeed, we have
ZZZ
dQdV ¼ 0; ð4:26Þ
V

while in case of an isothermic process, e.g., we may write


ZZZ
dSdV [ 0: ð4:260 Þ
V

The relation dS\0 is possible only for non-isolated thermic systems.


The entropy S as well as the internal energy U are quantities which can be
calculated only by neglecting an additive constant S0 or U0 , respectively; this is
sufficient because the mentioned constants disappear. As a matter of fact, the
constant S0 may be determined by means of Nernst’s theorem, which states that, at
the zero absolute temperature, the entropy S vanishes.
4.1 Elastic Models 131

A thermodynamic process for which the entropy is constant, hence


dS ¼ 0; ð4:27Þ
is called isentropic; the second principle of thermodynamics shows that the
reversible and adiabatic processes are isentropic.

4.1.2.5 Helmholtz’s Function. Gibbs’s Function

In case of a reversible process we have


dQ ¼ TdS; ð4:28Þ
which allows to write the relation (4.19) in the form
dU ¼ dW þ TdS: ð4:29Þ
We introduce a function of state F, called free energy (Helmholtz’s function),
by the relation
F ¼ U  TS; ð4:30Þ
where the product TS is called bound energy. We obtain, in this case,
dF ¼ dW  S dT; ð4:31Þ

for a reversible process, and


dF \ dW  ST; ð4:310 Þ
for an irreversible process. In particular, in case of an isothermic process, we can
write
dF ¼ dW; ð4:32Þ
for a reversible process, and
dF \ dW; ð4:320 Þ

for an irreversible one.


From (4.29) it results that, in an isentropic process, the elementary work is a
total differential and we have
W ¼ U; S ¼ const, ð4:33Þ

neglecting an additive constant.


Analogically, we introduce Gibbs’s function of state (called thermodynamical
potential too) by the relations
G ¼ H  TS; ð4:34Þ
132 4 Mathematical Models in Mechanics of Deformable Solids

where H is a quantity of energetical nature called enthalpy, which, in case of the


deformation of solid bodies, is given by the relation
H ¼ U  rij eij : ð4:35Þ

We get
dH ¼ dU  dW  dWc ¼ dQ  dWc ¼ dWc þ TdS; ð4:36Þ

where we took into account the formula (4.15) and the relations (4.19), (4.24);
moreover, this is a new form of the first principle of thermodynamics. We notice
that the elementary work is, in our case, just the elementary work of deformation
introduced in Sect. 4.1.2.2.
We obtained thus
dG ¼ dWc  S dT; ð4:37Þ
for a reversible process, and
dG\  dWc  S dT; ð4:370 Þ

for an irreversible process. In particular, in case of an isothermic process, we can


write
dG ¼ dWc ; ð4:38Þ
for a reversible process, and
dG\  dWc ; ð4:380 Þ

for an irreversible process.


It results, from (4.36), that, in an isentropic process, the complementary ele-
mentary work is a total differential, so that one has
Wc ¼ H; S ¼ const; ð4:39Þ

neglecting an additive constant. Analogically, the relation (4.37) shows that, in a


reversible and isothermic process, the complementary elementary work is an exact
differential too; we also have
Wc ¼ G; T ¼ const; ð4:390 Þ
where we neglect once more an additive constant.

4.1.2.6 Elastic Potentials. Green Type Formulae. Castigliano Type


Formulae

We notice that, in case of a quasistatic process of elastic deformation one can write
dU ¼ rij deij þ TdS; ð4:40Þ
4.1 Elastic Models 133

as well as
dF ¼ rij deij  SdT: ð4:400 Þ
We observe that, in each of these two cases, we have to deal only with seven
independent variables e11 ; e22 ; e33 ; e23 ¼ e32 ; e31 ¼ e13 ; e12 ¼ e21 and S or T,
respectively. The process is reversible, so that in the above formulae we have total
differentials; it results that
1 oU oU
rij ¼ ð1 þ dij Þ ð!Þ; i; j ¼ 1; 2; 3; T¼ ; ð4:41Þ
2 oeij oS

or
1 oF oF
rij ¼ ð1 þ dij Þ ð!Þ; i; j ¼ 1; 2; 3; S¼ ; ð4:410 Þ
2 oeij oT

which are Green type formulae [3].


In general, dW is not a total differential; both the work and the quantity of heat
absorbed by the body depend on the history of the thermodynamic process.
From (4.33), (4.330 ) it results that, in case of an adiabatic process (being
reversible, it is isentropic too) or in case of an isothermic process, we may build a
function of state
U ¼ Uðe11 ; e22 ; . . .; e12 ; S0 Þ ð4:42Þ

or a function of state
F ¼ Fðe11 ; e22 ; . . .; e12 ; T0 Þ; ð4:420 Þ
where S0 and T0 are the constant values of the entropy and of the absolute tem-
perature, respectively, corresponding to the conditions in which the process is
developed.
The first case is of interest for the dynamic problems, where rise elastic
vibrations of small amplitude, the process being sufficiently quick, so that the
change of heat be practically not possible; the second case, important for static
problems, corresponds to slow loadings and unloadings of the body (quasistatic
processes), when it exists time for absorption or release of heat, so that the tem-
perature does remain constant.
In both cases (for S ¼ S0 or for T ¼ T0 ) there exists a function of state called
elastic potential or volume density of strain energy; as state of reference one
chooses that state in which both the stresses and the strains vanish at a temperature
T ¼ T0 . The elastic bodies which have this property are called hyperelastic.
If the process of unloading of an elastic solid is effected on the same way as the
process of loading, then the variation of the work on the whole cycle vanishes; but,
in general, the process of unloading takes place on another way, dW is not an exact
differential, while the first principle of thermodynamics allows to state that
134 4 Mathematical Models in Mechanics of Deformable Solids

Z Z
dW ¼  dQ 6¼ 0; ð4:43Þ
C C

where the integral is taken along the way travelled through. This leads to the
increase of the entropy and to decrease of the free energy, as well as to the
transformation of a part of the work of loading in heat (dissipation of the energy).
In case of the hyperelastic deformation, the existence of a function of state
which plays the rôle of elastic potential shows that for any cycle one has
Z Z
dW ¼ dQ ¼ 0; ð4:430 Þ
C C

the phenomenon of deformation taking place without dissipation of energy.


The formulae (4.41) and (4.410 ) are formulae of Green type, which, together
with (4.42) or (4.420 ), may constitute constitutive laws of the elastic solid. Because
the elastic potentials are different, it results that the respective constitutive laws are
not the same.
If the thermodynamic process is not adiabatic (hence, isentropic) or isothermic,
then it does not exist an elastic potential and one cannot build a constitutive law on
this way; but if the variation of temperature or of entropy have a very small
contribution with respect to dW, then we can practically assume the existence of
an elastic potential. Much more, if the bound energy TS is negligible with respect
to the internal energy U, then we may write F ffi U, assuming the existence of only
one elastic potential, which will be denoted by W and will not depend on thermal
factors; in various studies of experimental nature one has seen that the error thus
made is not greater than 1 %. This elastic potential (Green’s potential) has been
introduced in Sect. 4.1.2.1.
If the process of deformation is associated with great changes of temperature or
with an important quantity of heat, without any isentropic or isothermic reversible
character, the problem can no more be studied only by mechanical means; it
becomes a real problem of thermodynamics in this case.
Using the enthalpy and Gibb’s functions, we may write, analogically,
dH ¼ eij drij þ TdS; ð4:44Þ

as well as
dG ¼ eij drij  SdT; ð4:440 Þ

in each of these cases we have to deal with seven independent variables, i.e.:
r11 ; r22 ; r33 ; r23 ¼ r32 ; r31 ¼ r13 ; r12 ¼ r21 and S or T, respectively. The process
is reversible, so that in the above formulae we have exact differentials and we may
write
1 oH oH
eij ¼  ð1 þ dij Þ ð!Þ; T¼ ð4:45Þ
2 orij oS
4.1 Elastic Models 135

or
1 oG oF
eij ¼  ð1 þ dij Þ ð!Þ; S¼ : ð4:450 Þ
2 orij oT
In case of an adiabatic process (hence, isentropic), it results, from (4.39), that
one can build a function of state
H ¼ Hðr11 ; r22 ; . . .; r12 ; S0 Þ; ð4:46Þ

where S0 is the constant value of the entropy, corresponding to the conditions in


which the process is developed; as well, in case of an isothermic process, it results,
from (4.390 ), that one can set up a function of state
G ¼ Gðr11 ; r22 ; . . .; r12 ; T0 Þ; ð4:460 Þ

where T0 is a constant which has an analogous interpretation.


The formulae (4.45) and (4.450 ) are formulae of Castigliano type, which,
together with the elastic potentials (4.46) and (4.460 ), can constitute constitutive
laws of the elastic solid.
If we assume that the two elastic potentials are equal (H ¼ G), then we agree—
taking into account (4.39), (4.390 )—to choose an elastic potential, which will be
denoted by Wc and will be called complementary elastic potential; this potential is
of the form
Wc ¼ Wc ðr11 ; r22 ; . . .; r12 Þ: ð4:47Þ
Castigliano’s formulae are
1 oWc
eij ¼ ð1 þ dij Þ ð!Þ; i; j ¼ 1; 2; 3; ð4:48Þ
2 orij

in a developed form, we have


oWc oWc oWc
e11 ¼ ; e22 ¼ ; e33 ¼ ; ð4:480 Þ
or11 or22 or33
oWc oWc oWc
c23 ¼ 2e23 ¼ ; c31 ¼ 2e31 ¼ ; c12 ¼ 2e12 ¼ : ð4:4800 Þ
or23 or31 or12
We can also say that the state of strain derives from the complementary elastic
potential Wc .
The differential of this potential is given by
oWc
dWc ¼ drij ; i  j; ð4:49Þ
orij

where the condition imposed to the indices is due to the fact that the comple-
mentary elastic potential depends only on six independent variables, the tensor Tr
being symmetric; taking into account Castigliano’s formulae (4.48), one may write
136 4 Mathematical Models in Mechanics of Deformable Solids

Fig. 4.7 Geometrical


interpretation of W and Wc in
case of a one-dimensional
experiment P( , )
Wc

dWc ¼ eij drij ; ð4:490 Þ

summing now for all the value i; j ¼ 1; 2; 3. The introduction of the dual expres-
sion (4.14) is thus justified.
In the case of a one-dimensional experiment, the elastic potentials W and Wc
have an interesting significance, in connection with the characteristic curve of the
material. Thus, if we stop at a point Pðe; rÞ on this curve (Fig. 4.7), then the elastic
potential W represents the area of the domain between the characteristic curve, the
Oe-axis and the parallel at P to the Or-axis; as well, the complementary elastic
potential Wc represents the area of the domain between the characteristic curve, the
Or-axis and the parallel at P to the Oe-axis. These results are valid excepting an
additive constant; we assumed that the elastic potentials are equal to zero for a
vanishing state of strain and stress. We notice that, in this case,
W þ Wc ¼ re; ð4:50Þ

we have Wc ¼ W only in the case of a linear constitutive law (Hooke’s law),


assuming that the characteristic curve cannot change its concavity.
In general, if the two elastic potentials are equal
Wc ¼ W; ð4:51Þ

then we have
eij drij ¼ rij deij ;

wherefrom
 
1
rij deij ¼ d rij eij :
2

Taking into account (4.9), (4.10), we must have


1
W ¼ rij eij ; ð4:510 Þ
2
excepting an additive constant; the reciprocal is easily proved.
4.1 Elastic Models 137

4.1.2.7 Gibbs’s Thermodynamic Conditions of Equilibrium

We have seen that a reversible thermodynamic process is produced, in a certain


sense, corresponding to the second principle of thermodynamics; it must lead to a
thermodynamic equilibrium, where no one function of state varies in time.
A system is in thermodynamic equilibrium if no change in the limit conditions
occurs and no spontaneous process, consistent with these conditions, takes place.
To obtain the necessary and sufficient conditions of thermodynamic equilib-
rium, we compare the system in equilibrium with the neighbouring system the
functions of state of which differ only a little from those of the state of equilibrium,
all of them being subjected to the same limit conditions; this comparison must be
made in the sense of the variational calculus.
The second principle of thermodynamics shows in what conditions a system in
a neighbourhood of a system in equilibrium can be transformed in the latter one,
but does not specify if this transformation must take place; if this should occur, the
respective state of equilibrium would be a stable state of equilibrium (in the sense
in which a neighbouring perturbed state has the tendency to return to the state of
equilibrium). A not so large condition is necessary for the stability of the ther-
modynamic equilibrium.
After Gibbs, a state of thermodynamic equilibrium is stable if the entropy S of
the system represents a maximum with respect to all the neighbouring states which
have the same internal energy, hence if
S ¼ max; U ¼ const; ð4:52Þ

this is equivalent to
dS  0; U ¼ const, ð4:520 Þ
for any neighbourhood, in which we will have an entropy S þ dS smaller than that
corresponding to the stable state of equilibrium.
One can further state that for the stable thermodynamic equilibrium of an
isolated system it is necessary and sufficient that the internal energy be a minimum
with respect to all the neighbouring states which have the same entropy; one can
write
U ¼ min; S ¼ const; ð4:53Þ

which is equivalent with


dU  0; S ¼ const; ð4:530 Þ
for any neighbourhood.
Gibbs demonstrated the equivalence of the two above conditions, the second
one being closer to the necessities of mechanics of solids.
In case of the quasistatic deformations of an elastic body (reversible process),
the condition (4.530 ) will be fulfilled for an isentropic process; it results that the
corresponding elastic potential, which is obtained neglecting an arbitrary constant
138 4 Mathematical Models in Mechanics of Deformable Solids

and the minimum of which can be thus considered equal to zero, must be a positive
definite function.
If we take into account (4.52) or (4.53), then the relation (4.30) shows that
Helmholtz’s function F reaches also a minimum for the stable equilibrium of the
thermodynamic system. In this case too, the corresponding elastic potential will be
a positive definite function.
As a consequence of these results, we assume, in what follows, that the elastic
potential W, corresponding to the elastic body with hyperelastic properties, is a
positive definite function. This corresponds to the fact that the deformation process
needs an expense of work.

4.1.3 Hooke’s Law

In what follows, we make some considerations concerning the linear case of a


hyperelastic constitutive law, case in which the elastic potential W is a positive
definite quadratic form. We exclude the existence of an initial state of strain and
stress, so that we cannot have linear terms. Starting from the general form of
Hooke’s law, in case of a linear anisotropy, we deduce the law which corresponds
to an isotropic body, making then a detailed study of this important case.

4.1.3.1 General Considerations

We assume that the elastic potential is of the form


1
W ¼ Hijkl eij ekl ; ð4:54Þ
2
where Hijkl is the tensor of fourth order of the elastic coefficients (Hooke’s tensor);
from a dimensional point of view, we have ½Hijkl  ¼ ML1 T2 . Using Euler’s
theorem on homogeneous functions, we may write
oW
2W ¼ eij ; i  j; ð4:55Þ
oeij

where we have observed that the elastic potential depends on six independent
variables. Taking into account Green’s formulae (4.9), we may write this potential
in the form (4.510 ), the sum being made for all the values i; j ¼ 1; 2; 3; using the
considerations in Sect. 4.1.2.6, it results the relation (4.51).
Comparing the relations (4.510 ) and (4.54), one obtains the constitutive law
rij ¼ Hijkl ekl ; i; j ¼ 1; 2; 3; ð4:56Þ

of the form (4.6); one can obtain this result also by using Green’s formulae (4.9).
Observing that the tensors Tr and Te are symmetric, it results that
4.1 Elastic Models 139

Hjikl ¼ Hijkl ; Hijlk ¼ Hijkl ; i; j; k; l ¼ 1; 2; 3; ð4:57Þ

hence, from the 34 ¼ 81, components of the tensor TH remain only 36 independent
coefficients. As well, from the expression (4.54), one sees that
Hklij ¼ Hijkl ; i; j; k; l ¼ 1; 2; 3; ð4:570 Þ

remaining only 21 independent elastic coefficients; in case of a homogeneous body


the elastic coefficients are constant, the body being with a general linear
anisotropy.
In a developed form, we may write this law in the form

rij ¼ Hij11 e11 þ Hij22 e22 þ Hij33 e33


þ 2Hij23 e23 þ 2Hij31 e31 þ 2Hij12 e12 ; i; j ¼ 1; 2; 3; ð4:560 Þ

taking into account the symmetry of the tensor Te , the matrix of the tensor TH
reads
2 3
H1111 H1122 H1133 2H1123 2H1131 2H1112
6 H2211 H2222 H2233 2H2223 2H2231 2H2212 7
6 7
6 H3311 H3322 H3333 2H3323 2H3331 2H3312 7
TH 66 H2311 H2322 H2333
7: ð4:5600 Þ
6 2H2323 2H2331 2H2312 7
7
4 H3111 H3122 H3133 2H3123 2H3131 2H3112 5
H1211 H1222 H1233 2H1223 2H1231 2H1212
Assuming that
det½Hijkl  6¼ 0; ð4:58Þ

the relations (4.56) are one-to-one and we can write


eij ¼ Cijkl rkl ; i; j ¼ 1; 2; 3; ð4:56000 Þ

where Cijkl are 81 coefficients of elastic deformation (of compliance), from which
21 are distinct; in this case, the elastic potential can be written in the form
1
W ¼ Cijkl rij rkl ð4:540 Þ
2
too, as a positive definite quadratic form in stresses, corresponding to the com-
plementary elastic potential.
Castigliano’s formulae
1 oW
eij ¼ ð1 þ dij Þ ð!Þ; i; j ¼ 1; 2; 3; ð4:59Þ
2 orij

take place.
140 4 Mathematical Models in Mechanics of Deformable Solids

The necessary and sufficient condition for the quadratic form (4.54) be positive
definite is given be Sylvester’s criterion; the diagonal minors of the matrix of the
coefficients Hijkl must be positive, i.e.
 
1 1  H1111 H1122  1 2
H1111 [ 0; ¼ ðH1111 H2222  H1122 Þ [ 0; . . .; ð4:60Þ
2 4  H2211 H2222  4

this leads to certain conditions for the elastic coefficients.


Concerning the constitutive law considered above, we must remark a certain
softness. Indeed, by introducing the components of the tensor Te , one has admitted
that the components ui of the displacement vector are functions of the material co-
ordinates; on the other hand, one has used space co-ordinates to define the com-
ponents of the stress tensor Tr . We have seen that, in case of infinitesimal
deformations, one can neglect this difference, so that we will use everywhere only
material co-ordinates [6, 8–11, 13, 15, 17, 18].
We put now the problem to find the types of anisotropic elastic bodies and the
conditions in which the principal axes of the stress tensor coincide with the
principal axes of the strain tensor. We mention that for the principal axes of
the tensor Tr the tangential stresses must vanish, while for the principal directions
of the tensor Te the angular strains must vanish; assuming that the principal
directions coincide at the point M and denoting them by M1; M2; M3, the relations
(4.56) take the form
ri ¼ Hi;1 e1 þ Hi;2 e2 þ Hi;3 e3 ; i ¼ 1; 2; 3; ð4:61Þ

0 ¼ Hij;1 e1 þ Hij;2 e2 þ Hij;3 e3 ; i; j ¼ 1; 2; 3; ð4:610 Þ

Hi;j ; Hij;k ; i 6¼ j; Hk;ij ; i 6¼ j; Hij;kl ; i 6¼ j; k 6¼ l; i; j; k; l ¼ 1; 2; 3; being the elas-


tic coefficients which correspond to these directions.
To verify identically the relations (4.610 ) (we assume that det½Hijkl  6¼ 0; i 6¼ j),
the nine elastic coefficients Hij;k ; i 6¼ j; i; j; k ¼ 1; 2; 3, must vanish, i.e.

H23;1 ¼ H23;2 ¼ H23;3 ¼ H31;1 ¼ H31;2 ¼ H31;3


¼ H12;1 ¼ H12;2 ¼ H12;3 ¼ 0: ð4:62Þ
On the basis of the symmetry relations (4.570 ) one can state that a second group
of elastic coefficients Hk;ij ; i 6¼ j; i; j; k ¼ 1; 2; 3, must vanish, i.e.

H1;23 ¼ H2;23 ¼ H3;23 ¼ H1;31 ¼ H2;31 ¼ H3;31


¼ H1;12 ¼ H2;12 ¼ H3;12 ¼ 0: ð4:620 Þ
Hooke’s law, written for the principal directions common to the two tensors Te
and Tr is thus reduced to the relation (4.61).
If we take into account the relations to pass from the elastic coefficients in the
system of axes M123 to the elastic coefficients in the system of axes Mx1 x2 x3
(relations of tensor nature, because Hijkl are the components of a tensor of fourth
4.1 Elastic Models 141

order), one cannot write interesting relations for the coefficients in the latter system
of axes; assuming that the system of axes Mx1 x2 x3 is just the system of principal
axes, there result the conditions

H2311 ¼ H2322 ¼ H2333 ¼ H3111 ¼ H3122 ¼ H3133


¼ H1211 ¼ H1222 ¼ H1233 ¼ 0: ð4:63Þ

H1123 ¼ H1131 ¼ H1112 ¼ H2223 ¼ H2231 ¼ H2212


¼ H3323 ¼ H3331 ¼ H3312 ¼ 0; ð4:630 Þ

so that the matrix of the tensor TH reads


2 3
H1111 H1122 H1133 0 0 0
6 H2211 H2222 H2233 0 0 0 7
6 7
6 H3311 H3322 H3333 0 0 0 7
TH ¼ 6
6 0
7; ð4:64Þ
6 0 0 2H2323 2H2331 2H2312 7
7
4 0 0 0 2H3123 2H3131 2H3112 5
0 0 0 2H1223 2H1231 2H1212

containing 2  6 ¼ 12 independent elastic coefficients; thus, in the considered


system of orthogonal Cartesian co-ordinates, the most general linearly elastic
constitutive law for which the principal directions of the tensors Te and Tr
coincide is given by
rii ¼ Hii11 e11 þ Hii22 e22 þ Hii33 e33 ð!Þ; i ¼ 1; 2; 3; ð4:65Þ

rij ¼ 0; eij ¼ 0; i 6¼ j; i; j ¼ 1; 2; 3: ð4:650 Þ

These results are valid only for the point M and may not be valid at another point.
Among the types of the most important anisotropic elastic bodies which have
this property we mention the orthotropic bodies, for which the planes of elastic
symmetry are principal planes, at the point of intersection of them, as well as the
bodies with transverse isotropy for which the monotropy axis is a principal one at
all the points of the axis; as well, for all the points of the isotropic bodies the
principal axes of the tensors Te and Tr coincide. For the bodies with a plane of
elastic symmetry, as well as for the bodies with an axis of elastic symmetry of the
third or of the fourth order such a property does no more hold.

4.1.3.2 Bodies with a Plane of Elastic Symmetry

If the mechanical properties of an elastic body maintain their form after symmetric
directions with respect to a plane, we say that this body is with a plane of elastic
symmetry. Choosing the plane x3 ¼ 0 as a plane of elastic symmetry, it results that,
passing to the system of axes Ox01 x02 x03 , specified by the relations
142 4 Mathematical Models in Mechanics of Deformable Solids

x01 ¼ x1 ; x02 ¼ x2 ; x03 ¼ x3 ;

Hooke’s law (4.56) must be written in the form

r0ij ¼ Hijkl e0kl : ð4:66Þ

Using formulae of type (A.12), we get

e011 ¼ e11 ; e022 ¼ e22 ; e033 ¼ e33 ;


ð4:67Þ
e023 ¼ e23 ; e031 ¼ e31 ; e012 ¼ e12 ;

r011 ¼ r11 ; r022 ¼ r22 ; r033 ¼ r33 ;


ð4:670 Þ
r023 ¼ r23 ; r031 ¼ r31 ; r012 ¼ r12 :

Making, e.g., i ¼ j ¼ 1 in the relations (4.56) and (4.66), we have

r11 ¼ H1111 e11 þ H1122 e22 þ H1133 e33


þ 2H1123 e23 þ 2H1131 e31 þ H1112 e12 ; ð4:68Þ

r011 ¼ H1111 e011 þ H1122 e022 þ H1133 e033


þ 2H1123 e023 þ 2H1131 e031 þ H1112 e012 : ð4:680 Þ

Taking into account (4.67, 4.670 ), we notice that the relations (4.68, 4.680 ) hold
simultaneously if and only if H1123 ¼ H1131 ¼ 0; repeating this reasoning for the
other relations which form Hooke’s law too, we state that the elastic coefficients
must verify the conditions
Hii23 ¼ Hii31 ¼ H23ii ¼ H31ii ¼ 0ð!Þ; i ¼ 1; 2; 3; ð4:69Þ

H2312 ¼ H3112 ¼ H1223 ¼ H1231 ¼ 0: ð4:690 Þ

The matrix of the tensor TH is then of the form


2 3
H1111 H1122 H1133 0 0 2H1112
6 H2211 H2222 H2233 0 0 2H2212 7
6 7
6 H3311 H3322 H3333 0 0 2H3312 7
TH ¼ 6 6 0
7; ð4:70Þ
6 0 0 2H2323 2H2331 0 7 7
4 0 0 0 2H3123 2H3131 0 5
H1211 H1222 H1233 0 0 2H1212

containing 13 independent elastic coefficients.

4.1.3.3 Orthotropic Bodies

We assume now that the plane x1 ¼ 0 is also a plane of elastic symmetry; in this
case, the elastic coefficients must verify the conditions
4.1 Elastic Models 143

Hiijk ¼ Hjkii ¼ 0ð!Þ; j 6¼ k; i; j; k ¼ 1; 2; 3; ð4:71Þ

H2331 ¼ H3123 ¼ H3112 ¼ H1231 ¼ H1223 ¼ H2312 ¼ 0: ð4:710 Þ


The matrix of the tensor TH will be of the form
2 3
H1111 H1122 H1133 0 0 0
6 H2211 H2222 H2233 0 0 0 7
6 7
6 H3311 H3322 H3333 0 0 0 7
TH ¼ 6 6 0
7; ð4:72Þ
6 0 0 2H 2323 0 0 7 7
4 0 0 0 0 2H3131 0 5
0 0 0 0 0 2H1212

containing 9 independent elastic coefficients [12].


We notice that, in this case, also the plane x2 ¼ 0 is a plane of elastic symmetry.
Hence, a body which has two orthogonal planes of elastic symmetry, admits also a
third plane of elastic symmetry, orthogonal to the first two ones; such a body is
called orthotropic.

4.1.3.4 Hooke’s Law for an Isotropic Body

In case of a body with central symmetry, the mechanical properties must be the
same in any direction around a point of the body. In what follows we start from the
case of the orthotropic body, considered above.
Effecting a rotation of p=2 with respect to the Ox1 -axis, specified by the
relations

x01 ¼ x1 ; x02 ¼ x3 ; x03 ¼ x2 ;

as well as a rotation of p=2 with respect to the Ox2 -axis, specified by the relations

x01 ¼ x3 ; x02 ¼ x2 ; x03 ¼ x1 ;

one can show, as at the Sect. 4.1.3.2, that the conditions

H2233 ¼ H3311 ¼ H1122 ; H1111 ¼ H2222 ¼ H3333 ;


H2323 ¼ H3131 ¼ H1212 ð4:73Þ

hold, so that the body does remain with the same mechanical properties with
respect to the new obtained directions.
Now, effecting a rotation of p=4 with respect to the Ox3 -axis, specified by the
relations
pffiffiffi pffiffiffi
2 2
x01 ¼ ðx1 þ x2 Þ; x02 ¼  ðx1  x2 Þ; x03 ¼ x3 ;
2 2
one states, analogously, that the conditions
144 4 Mathematical Models in Mechanics of Deformable Solids

1
H2323 ¼ ðH1111  H1122 Þ ð4:730 Þ
2
must be fulfilled, so that the mechanical properties do remain the same with
respect to the new directions. Denoting
H1111 ¼ k; H2323 ¼ l; ð4:74Þ
we observe that one can write the constitutive law of a linearly elastic body with
central symmetry (Hooke’s law) in the form
rij ¼ kell dij þ 2leij ; i; j ¼ 1; 2; 3; ð4:75Þ

depending only on two elastic coefficients of the material; in case of a homoge-


neous body, the coefficients k and l are known as Lamé’s elastic constants.
As a matter of fact, in the latter case, all the points of the linearly elastic body
are points of central symmetry; in this case, the body is an isotropic one.
We notice that one cannot reduce the number of elastic constants to less than
two distinct ones (e.g., k and l). Indeed, effecting a general change of orthogonal
Cartesian co-ordinate axes and taking into account the formula (A.12), which
allows to pass from the components of a tensor of second order with respect to a
system of orthogonal Cartesian axes to the components with respect to another
system of orthogonal Cartesian axes, we may write

r0kl ¼ rij aki alj ; e0kl ¼ eij aki alj ; k; l ¼ 1; 2; 3;

in this case, Hooke’s law (4.75) becomes

r0kl ¼ ke0ii dkl þ 2le0kl ; k; l ¼ 1; 2; 3;

maintaining the same form. The above statement is thus proved.


Introducing the volume strain (2.64), one can write Hooke’s law also in the
form
rij ¼ khdij þ 2leij ; i; j ¼ 1; 2; 3; ð4:750 Þ

or in the developed form

r11 ¼ kh þ 2le11 ;
r22 ¼ kh þ 2le22 ; ð4:7500 Þ
r33 ¼ kh þ 2le33 ;

r23 ¼ 2le23 ¼ lc23 ;


r31 ¼ 2le31 ¼ lc31 ; ð4:75000 Þ
r12 ¼ 2le12 ¼ lc12 :

Summing in both members, we find that the first invariants of the two tensors are
linked by the relation
4.1 Elastic Models 145

H ¼ ð3k þ 2lÞh; ð4:76Þ

we remember that the principal axes of the stress tensor coincide with the principal
axes of the strain tensor.
Solving with respect to the strains, one can write Hooke’s law also in the form
k 1
eij ¼  Hdij þ rij ; i; j ¼ 1; 2; 3: ð4:77Þ
2lð3k þ 2lÞ 2l

4.1.3.5 Considerations Concerning the Elastic Constants

We assume that a simple axial tension, as it has been defined in Sect. 3.2.3.2, takes
place; in this case, the state of stress is expressed in the form
r11 ¼ r ¼ const; r22 ¼ r33 ¼ r23 ¼ r31 ¼ r12 ¼ 0: ð4:78Þ

We notice that, in the absence of the volume forces, the equations of equilibrium
(3.610 ) are identically satisfied.
Hooke’s law (4.77) leads to the state of strain
kþl k
e11 ¼ r; e22 ¼ e33 ¼  r; e23 ¼ e31 ¼ e12 ¼ 0;
lð3k þ 2lÞ 2lð3k þ 2lÞ
ð4:79Þ
which satisfies Saint-Venant’s conditions of continuity (2.68).
Let be a finite right cylinder of axis Ox1 , subjected to the simple axial load
considered above (Fig. 4.8). We notice that this takes place if the cylinder is acted
upon only on the end cross sections by external loads of intensity equal to r. If
r [ 0, (e11 [ 0; e22 ¼ e33 \0), then we have to do with a simple tension, for which
the cylinder is extended, while the cross section is diminished (the case in
Fig. 4.8); if r\0, (e11 \0; e22 ¼ e33 [ 0), then there corresponds a simple com-
pression, case in which the cylinder is shortened, while its cross section increases.
Taking into account the form (4.2), (4.20 ) of Hooke’s law, it results that
lð3k þ 2lÞ k
E¼ ; m¼ ; ð4:80Þ
kþl 2ðk þ lÞ
these are the technical elastic constants of the material.

Fig. 4.8 Straight cylinder x3


subjected to a simple axial x2
load

σ
σ O x1
146 4 Mathematical Models in Mechanics of Deformable Solids

One obtains also the relations


mE E
k¼ ; l¼ : ð4:800 Þ
ð1 þ mÞð1  2mÞ 2ð1 þ mÞ
The relations (4.77) take also the form
 
1þm m
eij ¼ rij  Hdij ; i; j ¼ 1; 2; 3: ð4:770 Þ
E 1þm
Assuming a state of stress corresponding to a simple shear, so as it has been
defined in Sect. 3.2.3.3, we may write
r12 ¼ r21 ¼ s; r11 ¼ r22 ¼ r33 ¼ r23 ¼ r31 ¼ 0; ð4:81Þ
it results the state of strain
1
e12 ¼ e21 ¼ s; e11 ¼ e22 ¼ e33 ¼ e23 ¼ e31 ¼ 0: ð4:82Þ
2l
Both the equations of equilibrium, in the absence of the volume forces, and the
equation of continuity of deformations are identically verified.
It results, from (4.4), that l corresponds to the modulus of transverse elasticity
Gðl ¼ GÞ.
Analogically, in case of a hydrostatic state of stress, defined in Sect. 3.2.3.1, we
have
r11 ¼ r22 ¼ r33 ¼ p; p [ 0; r23 ¼ r31 ¼ r12 ¼ 0; ð4:83Þ
the equations of equilibrium being identically verified, in the absence of the
volume forces; the state of strain is given by
p
e11 ¼ e22 ¼ e33 ¼  ; e23 ¼ e31 ¼ e12 ¼ 0; ð4:84Þ
3k þ 2l
the equations of continuity of deformation being identically verified.
Let be
2 E
K ¼kþ l¼ ð4:85Þ
3 3ð1  2mÞ
the modulus of elastic compression (bulk modulus) of the body; it results that
p
h¼ ; ð4:86Þ
K
the mechanical significance of the constant K being thus put in evidence.
From the above considerations it results that E  0; K  0; l  0; m  0. Taking
into account (4.85), we notice that 0  m  1=2; in this case, we have k  0 too. As
well, we have 2G  E  3G.
4.1 Elastic Models 147

Sometimes it is useful to introduce the elastic constant


1
m¼  2; ð4:87Þ
m
which is called Poisson’s number.
From the dimensional point of view, the elastic constants k; l ¼ G; E and K are
of the nature of a stress ½k ¼ ½l ¼ ½G ¼ ½E ¼ ½K ¼ ML1 T2 , while m and m
are numbers.
The fundamental ideas that led to this constitutive law were offered in 1829 by
Fresnel. This form of the relation was deduced by means of a molecular model of
Navier and Poisson. The general relation in the isotropic case was given by
Cauchy. Stokes shows equally that, in case of an isotropic body, the state of strain
and stress can be entirely characterized by means of the constants K and G ¼ l
corresponding to a uniform or to a phenomenon of simple shear, respectively.
In the limit case in which m ¼ 0 (m ¼ 1) (we have not a transverse contrac-
tion), we have k ¼ 0, E ¼ 2G ¼ 3K, while Hooke’s law takes the form
rij ¼ 2leij ; i; j ¼ 1; 2; 3; ð4:88Þ

one observes that this case corresponds to a superposition of one-dimensional


problems.
Numerical data concerning the elastic constants of some important materials are
given in Table 4.1.
We notice that, in case of an isotropic elastic body, the elastic coefficients can
be expressed in the form
Hijkl ¼ kikm jlm þ ðk þ 2lÞdil dkj ¼ kdij dkl þ 2ldil djk ; i; j; k; l ¼ 1; 2; 3; ð4:89Þ

the matrix of these coefficients being given by

Table 4.1 Numerical data concerning the elastic constants of some important materials
Material 105 E 105 G 105 K 105 k m
ðMPa) ðMPa) ðMPa) ðMPaÞ
Grey cast iron 1.20 0.480 0.800 0.480 0.25
Carbon steel 2.10 0.840 1.400 0.840 0.25
Tough copper 1.10 0.416 1.018 0.740 0.32
Bronze 1.15 0.432 1.127 0.838 0.33
Brass 0.95 0.349 1.130 0.900 0.36
Laminated aluminium 0.69 0.259 0.676 0.502 0.33
Lead 0.17 0.059 0.354 0.314 0.42
Glass 0.56 0.224 0.373 0.224 0.25
Concrete 200 0.21 0.089 0.106 0.089 0.17
Rubber 0.00008 0.00003 0.00040 0.00040 0.47
Celluloid 0.00020 0.00007 0.00022 0.00017 0.35
148 4 Mathematical Models in Mechanics of Deformable Solids

2 3
k þ 2l k k 0 0 0
6 k k þ 2l k 0 0 0 7
6 7
6 k k k þ 2l 0 0 0 7
TH ¼ 6
6 0
7: ð4:890 Þ
6 0 0 2l 0 0 7
7
4 0 0 0 0 2l 0 5
0 0 0 0 0 2l

Replacing (4.89) in (4.54), we get


1
W ¼ kh2 þ leij eij ; ð4:90Þ
2
with the aid of the first two invariants (2.400 , 2.4000 ) of the tensor Te , we may write
1
W ¼ ðk þ 2lÞI12  2lI2 : ð4:900 Þ
2
According to Sylvester’s criterion, the successive diagonal minors are
1 1
D1 ¼ ðk þ 2lÞ; D2 ¼ lðk þ lÞ; D3 ¼ l2 ð3k þ 2lÞ;
2 2
D4 ¼ l3 ð3k þ 2lÞ; D5 ¼ 2l4 ð3k þ 2lÞ; D6 ¼ 4l5 ð3k þ 2lÞ;

all these minors are positive if


2
l [ 0; K ¼ k þ l [ 0:
3
These conditions are always fulfilled, so that the elastic potential W is a positive
definite quadratic form.
The compliances Cijkl can be written in the form
1
Cijkl ¼ ðdik djl  m ilm jkm Þ
E
1
¼ ½ð1 þ mÞdik djl  mdij dkl ; i; j; k; l ¼ 1; 2; 3; ð4:91Þ
E
while the elastic potential, expressed in stresses, becomes
1
W¼ ½ð1 þ mÞrij rij  mH2 ; ð4:9000 Þ
2E
by means of the invariants of the tensor Tr one can also write
1 2
W¼ ½J  2ð1 þ mÞJ2 : ð4:90000 Þ
2E 1
4.1 Elastic Models 149

4.1.3.6 Deformation of Form. Deformation of Volume

Using the spheric tensors of strain and stress, as well as the deviators of strain and
stress, which have been introduced in Sects. 2.2.1.2 and 3.2.1.3, we notice that
Hooke’s law can be decomposed in the form
r0 ¼ 3Ke0 ; ð4:92Þ

r0ij ¼ 2le0ij ; i; j ¼ 1; 2; 3; ð4:920 Þ

thus, some important properties of the elastic constants of the material, as well as
an interesting decomposition of its deformations are put into evidence. One can
state that the constitutive law (4.92) corresponds to a change of volume of the
elastic solid, without a change of form, while the constitutive law (4.920 ) corre-
sponds to a change of form of the elastic solid, the volume remaining constant
(because I10 ¼ 0).
We notice that one can write the elastic potential in the form
 
1 2l 2 1 h
W¼ kþ h þ l ðe22  e33 Þ2 þ ðe33  e11 Þ2
2 3 3
i
2
þðe11  e22 Þ þ 6ðe223 þ e231 þ e212 Þ

too; taking into account the results given in Sect. 2.2.1.2 concerning the deviator
of strains, it results
W ¼ Wt þ Wf ; ð4:93Þ

where
1
Wt ¼ KI12 ð4:94Þ
2
is the unit work of deformation in case of a change of volume without a change of
form, while

Wf ¼ 2lI20 ð4:95Þ

is the unit work of deformation in case of a change of form, the volume remaining
constant.
Analogically, we can express the two works, in stresses, in the form
1 2
Wt ¼ J ; ð4:940 Þ
18K 1
1 0
Wf ¼ J : ð4:950 Þ
2l 2
We have the relations
150 4 Mathematical Models in Mechanics of Deformable Solids

J1 ¼ 3KI1 ; ð4:96Þ

J20 ¼ 4l2 I20 ; ð4:960 Þ

which result from the constitutive law (4.92), (4.920 ).


Introducing the equivalent strain e and the equivalent stress r, we get
3 1 2
Wf ¼ le2 ¼ r ; ð4:9500 Þ
2 6l
as well, between the two quantities takes place the relation
r ¼ 3le; ð4:97Þ

corresponding to the constitutive law of change of form.


Analogically, introducing the octahedric strains and the octahedric stresses one
can express the two elastic potentials in the form
9 1 2
Wv ¼ Ke20 ¼ r; ð4:98Þ
2 2K 0
3 3 2
Wf ¼ lc20 ¼ s; ð4:980 Þ
4 4l 0
as well, there results the relation (4.92), as well as the relation
s0 ¼ lc0 : ð4:9200 Þ

4.1.3.7 Incompressible Bodies. Incompressible State of Deformation

For an incompressible body we have m ¼ 1=2 (m ¼ 2), so that ev ¼ h ¼ 0, what-


ever be the state of stress undergone by the solid, as it results from the relations
(4.76), (4.85). In this limit case, we get E ¼ 3G ¼ 3l and k ¼ 1.
Hooke’s law becomes
1
eij ¼ ð3rij  rll dij Þ; i; j ¼ 1; 2; 3; ð4:99Þ
6l
with respect to the principal directions, it results
1
ei ¼ ðri  ri Þ; i ¼ 1; 2; 3: ð4:990 Þ
3l
As well, we may write Hooke’s law in the form
rij  r0 dij ¼ 2leij ; i; j ¼ 1; 2; 3: ð4:9900 Þ
4.1 Elastic Models 151

If, for a given state of stress of the deformable body, we have H ¼ 0, then it
follows, from the formula (4.76), that h ¼ 0 whatever Poisson’s ratio m would be;
we have to do with a state of incompressible deformation (state of isochore
deformation). But this is no more a property depending on the nature of the
material, as in the preceding case; such a state of deformation may occur in any
body, depending on the respective state of strain and stress.

4.1.3.8 Non-homogeneous Bodies

Let us admit that the hypothesis of homogeneity of the body is no longer valid.
Generally, we can have to deal with two types of non-homogeneous bodies: bodies
with a discontinuous non-homogeneity (e.g., stratified bodies) and bodies with a
continuous non-homogeneity [22, 23].
In the first of these cases, distinct elastic constants are introduced for every
subdomain occupied by a homogeneous part of the given elastic body (which
occupies the whole considered domain). The coefficients referring to other prop-
erties of the material should be similarly specified.
In the case of bodies with a continuous non-homogeneity, the elastic constants
become functions of point; thus, we have E ¼ Eðx1 ; x2 ; x3 Þ, m ¼ mðx1 ; x2 ; x3 Þ and,
consequently, k ¼ kðx1 ; x2 ; x3 Þ, l ¼ lðx1 ; x2 ; x3 Þ. Besides, the density can be
q ¼ qðx1 ; x2 ; x3 Þ.
Because the coefficient of transverse contraction varies, in general, between
close limits, one considers often that m ¼ const, remaining with only one elastic
coefficient function of point.
It is useful to introduce the reduced stresses
1
rij ¼ rij ; i; j ¼ 1; 2; 3; ð4:100Þ
E
which are non-dimensional quantities. Hooke’s law may be written in the form
eij ¼ ð1 þ mÞrij  mrll dij ; i; j ¼ 1; 2; 3; ð4:101Þ

the relations having now constant coefficients.


Consequently, one introduces also the reduced volume forces
1
Fi ¼ Fi ; i ¼ 1; 2; 3; ð4:102Þ
E
and the reduced density
1
q¼ q: ð4:103Þ
E
From the computational standpoint, it is frequently convenient to express the
modulus of longitudinal elasticity in the form
152 4 Mathematical Models in Mechanics of Deformable Solids

E ¼ E0 e f ; ð4:104Þ

where f ¼ f ðx1 ; x2 ; x3 Þ is a continuous function, differentiable as many times as


necessary (generally, it suffices to be a function of class C4 ); here E0 is a constant
modulus of longitudinal elasticity (a constant rigidity). In this manner, E ¼ const
on the surface f ðx1 ; x2 ; x3 Þ ¼ const.
Particularly, if the function f has the form
f ðx1 ; x2 ; x3 Þ ¼ ai xi ; ai ¼ const; i ¼ 1; 2; 3; ð4:105Þ
we shall have to deal with a body of constant rigidity in parallel planes; in this
case, the equations of the problem will have constant coefficients.
The density can also be expressed as follows
q ¼ q0 eg ; ð4:106Þ

where g ¼ gðx1 ; x2 ; x3 Þ is a continuous function, differentiable as many times as


necessary.
If g ¼ f , i.e. if q possesses the same type of non-homogeneity as the modulus of
longitudinal elasticity E, we can see that q ¼ const, which leads to important
simplifications of computation.
Similarly, we can introduce the reduced coefficient of damping
1
k¼ k ð4:107Þ
E
and we can consider

k ¼ k0 eh ; ð4:108Þ
where h ¼ hðx1 ; x2 ; x3 Þ is a continuous function, differentiable as many time as
necessary; as in the previous case, if h ¼ f , then we shall have k ¼ const.

4.1.3.9 Hooke’s Law in Curvilinear Co-ordinates

In case of orthogonal curvilinear co-ordinates, we can start from Hooke’s law


(4.770 ), by using the mapping relations (2.95), (2.950 ), (3.95); the constitutive law
becomes
1
esa sb ¼ ½ð1 þ mÞrsa sb  mrsc sc dab ; a; b ¼ 1; 2; 3: ð4:109Þ
E
All the considerations in the previous subsections can be repeated in this case.
The particularizations for cylindrical and spherical co-ordinates are immediate.
Thus, in the case of cylindrical co-ordinates, we shall have
4.1 Elastic Models 153

1
err ¼ ½rrr  mðrhh þ rzz Þ;
E
1
ehh ¼ ½rhh  mðrzz þ rrr Þ; ð4:110Þ
E
1
ezz ¼ ½rzz  mðrrr þ rhh Þ;
E
1 1 1
chz ¼ rhz ; czr ¼ rzr ; crh ¼ rrh ; ð4:1100 Þ
G G G
while in the case of spherical co-ordinates we can write
1
eRR ¼ ½rRR  mðruu þ rhh Þ;
E
1
euu ¼ ½ruu  mðrhh þ rRR Þ; ð4:111Þ
E
1
ehh ¼ ½rhh  mðrRR þ ruu Þ;
E
1 1 1
cuh ¼ ruh ; chR ¼ rhR ; cRu ¼ rRu : ð4:1110 Þ
G G G

4.1.3.10 Influence of the Temperature Variation

In case of an isotropic body, the influence of the temperature variation is put in


evidence by the linear strain which appear; in case of hyperelastic bodies, hence in
the presence of an elastic potential, this influence appear as a linear form in this
potential. In what follows we remain in the case of linearly elastic, homogeneous,
isotropic, classical bodies, subjected to infinitesimal deformations [2, 5].
An element of length ds becomes, after a deformation due to a temperature
variation, of the form ds ¼ ð1 þ a#Þds, where # ¼ #ðx1 ; x2 ; x3 Þ is the temperature
variation T ¼ T0 þ #, while a is a coefficient of linear dilatation, which does not
depend on the point and on the direction in case of homogeneous isotropic bodies.
In this case, an infinitesimal parallelepiped cut up of the body maintains its
parallelepipedic form after deformation too; we get
eij ¼ a#dij ; i; j ¼ 1; 2; 3: ð4:112Þ

The constitutive law becomes the Duhamel-Neumann law, of the form


1
eij  a#dij ¼ ½ð1 þ mÞrij  mrll dij ; i; j ¼ 1; 2; 3: ð4:113Þ
E
Summing, we obtain
1  2h
h  3a# ¼ H: ð4:114Þ
E
154 4 Mathematical Models in Mechanics of Deformable Solids

Solving with respect to stresses, we get


rij ¼ khdij þ 2leij  b#dij ; i; j ¼ 1; 2; 3; ð4:115Þ

with
E
b¼ a: ð4:116Þ
1  2m
One can use the results obtained above for anisotropic bodies too; we take different
coefficients of linear dilatation for the directions which are considered.

4.1.3.11 Initial Strains. Initial Stresses

More general, if a body is subjected to initial strains e0ij , then the linear constitutive
law becomes
1
eij  e0ij ¼ ½ð1 þ mÞrij  mrll dij ; i; j ¼ 1; 2; 3: ð4:117Þ
E
We can express these relations also in the form

rij ¼ kðell  e0ll Þdij þ 2lðeij  e0ij Þ; i; j ¼ 1; 2; 3: ð4:1170 Þ

If the initial strains verify the equations of continuity, then do not appear
stresses due to them. Otherwise, they represent incompatibilities and there appear
stresses, called initial stresses.
The results above are used for the distribution of stresses in the multiply
connected bodies, with the aid of the cuts. As well, they can be used also for the
bodies with initial stresses due to the processing of the material of which is made
the body, due to the contraction during the solidification, due to errors by
assembling etc. They can occur also in case of defects in crystals (e.g., disloca-
tions), which introduce incompatibilities.
In the following we do not take into consideration the initial stresses, assuming
that to a unloaded body corresponds a state of zero stress. In general, to can speak
about the state of stress of a body, one must admit that there exists a state of that
body—the natural state of stress—for which all the stresses vanish; as well, to can
speak about the state of strain of a body, one must choose a certain form of the
body—the initial state of strain—for which all the quantities which characterize
the deformation are, by definition, equal to zero.
It is necessary to suppose that neither the external loads vanish for the natural
state of stress, nor that this state corresponds to the initial state of strain. However,
this is the hypothesis assumed by Lamé and which we admit also in that follows.
4.1 Elastic Models 155

4.1.3.12 Case of Finite Deformations

In case of finite deformations, we can take again the considerations of thermo-


dynamical nature made above; one obtains interesting results concerning the
elastic bodies, as well as the hyperelastic and the hypoelastic ones.
Thus, in material co-ordinates, an elastic body can be described by a consti-
tutive law of the form
^ij ¼ f ðeij Þ;
r i; j ¼ 1; 2; 3; ð4:118Þ

where we introduce the Lagrangian deformation tensor TL and the Piola-Kirchhoff


stress tensor of first kind Tr^ ; in the most simple case (linear case) of an isotropic
elastic body, we get

^ij ¼ k0 ell dij þ 2l0 eij ;


r i; j ¼ 1; 2; 3: ð4:119Þ

where k0 ; l0 are the elastic coefficients of the material.


In space co-ordinates, we can express this law in the form
rij ¼ kell dij þ 2leij ; i; j ¼ 1; 2; 3; ð4:1190 Þ

while the Eulerian deformation tensor TE and Cauchy’s stress tensor Tr have
been introduced.

4.2 Inelastic Bodies

The deformable solids which, by unloading, do not return to the initial form are
inelastic solids (the thermodynamic process is irreversible); their properties of
plasticity and viscosity are thus put in evidence. In what follows we make a concise
presentation of some inelastic deformable solids, returning then later to these
problems; the scope of these considerations is only that to have a general view on
the constitutive laws of deformable solids, laws which complete their mathemat-
ical model.

4.2.1 Bodies with Plastic Properties

In the model considered in the preceding paragraph, we have taken out from the
complex phenomenon the principal part, neglecting other aspects; thus, the influ-
ence of peak loads, hence the consideration of plastic properties of the material
can play an essential rôle.
156 4 Mathematical Models in Mechanics of Deformable Solids

4.2.1.1 The Perfect Plastic Body. The Elastic-Plastic/Linearly


Hardening Body

We introduce some models in which appear plastic properties, in the frame of a one-
dimensional loading, presenting characteristic curves of the idealized material.
Thus, the perfect plastic (rigid plastic) body (Fig. 4.9a) will be represented by a
rigid solid, who slides with friction on a horizontal plane (Fig. 4.9b); when the force
P attains the maximal intensity of the sliding friction, the body begins to move,
arising a displacement D (which corresponds to the limit stress of elasticity re ).
One obtains thus Saint-Venant’s model; the respective constitutive law reads
r ¼ re ; e  0: ð4:120Þ
Linking in parallel a Hooke model with a Saint-Venant model, one obtains the
model which corresponds to the plastic rigid/linearly hardening body (the hard-
ening modulus being just the elasticity modulus E1 of the spring of the model)
(Fig. 4.10a and b); one may write the constitutive law
r ¼ re þ E1 e; e  0: ð4:121Þ
As well, linking in series the above elementary models, one obtains the model
of the elastic-perfect plastic body (Fig. 4.11a and b); it results the constitutive law

Ee for 0  e  rEe ;
r¼ ð4:122Þ
re for e  rEe :

Finally, by linking in parallel a Hooke law to an elastic–plastic model one


obtains the model of the elastic-perfect plastic/linearly hardening body (Fig. 4.12a
and b); the corresponding constitutive law is
 re

Ee  re
 for 0  er  E ; ð4:123Þ
re þ E1 e  E for e  Ee :

Using the functions introduced by the formulae (A.165) and (A.167), we can
write the above constitutive laws, in the following remarkable forms
r ¼ re hðeÞ; ð4:1200 Þ

r ¼ ðre þ E1 eÞhðeÞ ¼ re hðeÞ þ E1 eþ ; ð4:1210 Þ


re
re

r ¼ EehðeÞ  E e  h e
h Er
i E
re

e
¼ Ee hðeÞ  h e  þ re h e 
E E
re

¼ E eþ  e  ; ð4:1220 Þ
E þ
4.2 Inelastic Bodies 157

Fig. 4.9 Perfect plastic


model: characteristic curve
(a); technical model (b) e
P

O
(a) (b)

Fig. 4.10 Plastic rigid/


linearly hardening model:
characteristic curve (a); arctan E 1
technical model (b)
e
P

O
(a) (b)

Fig. 4.11 Linearly elastic/


perfect plastic model: e
characteristic curve (a);
technical model (b)

arctan E 1 P

O
(a) (b)

Fig. 4.12 Linearly elastic-


perfect plastic/linearly
hardening model:
characteristic curve (a); arctan E 1
technical model (b) e

arctan E
O
(a) (b)
158 4 Mathematical Models in Mechanics of Deformable Solids

re
re

r ¼ EehðeÞ  ðE  E1 Þ e  h e
E E
re

¼ Eeþ  ðE  E1 Þ e  ; ð4:1230 Þ
E þ
which can be useful in various practical computations.
Although their intuitivity, these models could not be developed for bodies with
more complex plastic properties, because between the phenomenon of plastic flow
and the sliding phenomenon there exist important differences.
More complex constitutive laws are of the form (4.100 ) (or (4.1000 )) and can
correspond to bodies with elastic and plastic properties.

4.2.1.2 Plasticity Theories

In the considerations made in the preceding subsection we notice that the defor-
mation of a solid body is composed by an elastic part and by a plastic one, being of
the form (4.5). A special rôle is played by the moment at which appears the plastic
deformation ep , hence the position of the point of ordinate re . The construction of a
theory of plasticity in the one-dimensional case contains thus a constitutive law, as
well as the specification of the point at which one passes from the elastic zone to
the plastic one, by means of the limit stress of elasticity re .
Sometimes, the plastic deformation can be neglected with respect to the elastic ones;
other times, they are of the same order of magnitude, but at times the plastic defor-
mations are greater then the elastic ones. The study of these states of plastic deformation
can be effected using various theorems of plasticity. In the actual state, each theory
can be applied to a small number of materials or phenomena and only between certain
limits of variation of the parameters on which depend the deformations.
A basic concept in a three-dimensional theory of plasticity is that of loading
surface, which represents the frontier between the elastic and the plastic domains
in the stress space; this surface is correspondent of the limit stress of elasticity of
the one-dimensional case. We notice that the stress space rij is a space with 9
dimensions; one can use also a space with only 6 dimensions, because the tensor
Tr is symmetric. Frequently, one uses the deviators of stresses and strains; one is
thus led to a space of deviators which has 5 dimensions. The most times, one uses
the three-dimensional space of the principal normal stresses r1 ; r2 ; r3 . In this
space, the loading surface, which is an open surface, puts in evidence the domain
of the elastic state, containing the origin, and an external domain, which has not
mechanical interpretation in case of a perfect plastic material or which represents
possible plastic states in the future, in case of a hardening material; the points of
the loading surface correspond to the plastic states of stress.
The components of the strain tensor will be written in the form

eij ¼ eeij þ epij ; ð4:124Þ


4.2 Inelastic Bodies 159

where the elastic part and the plastic one of the strain are put in evidence.
The equation of the loading surface is given by

f ðrij ; epij ; kÞ ¼ 0; ð4:125Þ

where k is a hardening parameter (constant, if the material is perfect plastic); a


condition of loading is associated to this equation. If
of
f ¼ 0; r_ ij \0; ð4:126Þ
orij

then a unloading takes place, the increasing of stresses leading to the passing from
a plastic state to an elastic one, the plastic deformation remaining the same. If
of
f ¼ 0; r_ ij ¼ 0; ð4:1260 Þ
orij

then one has a neutral loading, which corresponds to the passing from a plastic
state to another plastic one, the plastic deformation remaining the same. If
of
f ¼ 0; r_ ij [ 0; ð4:12600 Þ
orij

then a loading takes place, in which one passes from a plastic state of stress to
another state of plastic stress, the plastic deformation varying too; the last con-
dition is without sense in case of perfect plastic materials.
A variation drij of the stress is represented by a vector with a point of appli-
cation on the loading surface. In case of a unloading, this vector is directed
towards the interior of the surface, in case of a neutral loading it is tangent to the
loading surface, while in case of a loading it is directed towards its exterior; the
vector of =orij is always directed towards the exterior of the loading surface
(Fig. 4.13).
Initially, for the plastic non-deformed material, the loading surface depends
only on the components of the stress tensor. Because for the most materials the
hydrostatic pressure has a negligible influence on the plastic deformation, the
loading depends, especially, on the deviator tensor of stresses r0ij . For an isotropic
body, the equation of this surface depends only on the invariants of the deviator,
being of the form

f ðJ20 ; J30 Þ ¼ 0; ð4:127Þ

this equation represents, in the space Or1 r2 r3 , a cylinder with the generatrices
parallel to the straight line r1 ¼ r2 ¼ r3 , normal to the octahedral plane
r1 þ r2 þ r3 ¼ 0. A relation of the form (4.127), corresponding to the initial
surface of loading, represents a condition of plasticity.
In particular, one can use Mises’s condition, valid for many polycrystalline
materials, for which—during the plastic flow—the relation
160 4 Mathematical Models in Mechanics of Deformable Solids

Fig. 4.13 The loading f =0


surface d ij

0 unloading ij
ij

neutral loading

J20 ¼ s2e ; ð4:128Þ


pffiffiffi
where se ¼ re = 3 is the elasticity limit for simple shear, takes place; this rep-
resents a circular cylinder in the space Or1 r2 r3 . Another condition often used is
Tresca’s condition, in which the extreme tangential stress is constant during the
plastic flow; hence, one of the three relations

ðri  rj Þ2 ¼ 4s2e ; i 6¼ j; i; j ¼ 1; 2; 3; ð4:129Þ

where se ¼ re =2, takes place.


The relation between re and se are established in case of a simple shear for
which the relations (3.71) are valid.
Tresca’s condition reads

4J203  27J302  36s2e J202 þ 96s4e J20 þ 64s6e ¼ 0; ð4:1290 Þ

by means of the invariants of the stress deviator.


We represent in the octahedral plane, sections in the loading surface, which are
a circle, in case of Mises’s condition, or a hexagon inscribed in this circle, in case
of Tresca’s condition (Fig. 4.14).
Excepting the properties mentioned above, the loading surfaces have also the
property to be non-concave; this is a consequence of the basic postulate of
Drucker. Let thus be an elastic-plastic body in equilibrium under the action of the
external loads; if upon an element of the body, which is in a certain state of stress
r0ij , is applied a supplementary stress rij  r0ij , which afterwards is taken off,
Drucker’s principle states: During the action of supplementary stresses and in a
cycle loading-unloading, the work is not negative.
To understand better the sense of this principle, we will consider firstly the case
of a one-dimensional loading. If r  r0 [ 0 and, correspondingly, e  e0 [ 0
(loading), then we have ðr  r0 Þðe  e0 Þ [ 0; the unit work of deformation, hat-
ched in Fig. 4.15a will be positive, the corresponding material being called stable.
As well, if r  r0 \0 and e  e0 \0 (unloading) (Fig. 4.15b), then the unit work
will be positive too. If r  r0 \0 and e  e0 [ 0 (Fig. 4.16a) or if r  r0 [ 0 and
e  e0 \0 (Fig. 4.16b), then we have ðr  r0 Þðe  e0 Þ\0; in the first of these
cases the corresponding material is unstable. In the last case (Fig. 4.16b) the
second principle of thermodynamics is not respected; in reality, such a case cannot
take place. The above relations correspond to Drucker’s principle.
4.2 Inelastic Bodies 161

Fig. 4.14 Intersection of the


3
loading surface with the
octahedral plane

1 2

Fig. 4.15 Drucker’s


principle: loading (a);
unloading (b)
0 0

O 0 O 0

(a) (b)

Fig. 4.16 Unstable material


(a). Second principle of
thermodynamics is not 0
0
respected (b)

O 0 O 0

(a) (b)

Fig. 4.17 Case of a f =0


hardening material d ij C
p
d ij
B ij
A
0
ij

In case of a three-dimensional state of strain and stress we shall consider a


loading surface and a loading path ABC (Fig. 4.17); we can express Drucker’s
principle on the path A ! B ! C in the form
162 4 Mathematical Models in Mechanics of Deformable Solids

Z
ðrij  r0ij Þ deij  0: ð4:130Þ
_
ABC

Because the work corresponding to the strains eeij vanishes along the closed
cycle A ! B ! C ! A, it results (stability in great)
I
ðrij  r0ij Þ depij  0: ð4:1300 Þ

In case of a hardening material, the equality takes place only in the absence of
plastic deformations. Because the plastic deformation\takes place only on the path
B ! C, the last inequality becomes (stability in small)

ðrij  r0ij Þ depij  0: ð4:13000 Þ

We consider now another cycle with the initial point B. For the loading process
B ! C we may write
drij deij  0; ð4:131Þ

in case of infinitesimal deformations, this condition becomes


r_ ij e_ ij  0: ð4:1310 Þ
For the cycle loading–unloading B ! C ! B it results

drij depij  0; ð4:132Þ

while, in case of the infinitesimal deformations, we have

r_ ij e_ pij  0: ð4:1320 Þ

The equality takes place only in case of neutral loadings; as well, from the
above relations, it results that the loading surface can never be concave.
We notice also that, beside regular points, the loading surface can have sin-
gular points (vertices etc.); concerning these points must be made a special study.
To complete the constitutive law of a body with plastic properties, we intro-
duce, beside a condition of plasticity, a law of hardening too, that is relations
which describe the mode of deformation and displacement of the loading surface
during the plastic deformation. The most used hardening law is that of isotropic
hardening, on the basis of which, during the plastic deformation, the loading
surface is uniformly deformed in all directions; analytically, one can write such a
law in the form

f ðJ_ 2 ; J_ 3 Þ ¼ k2 ; ð4:133Þ
where k ¼ kðqÞ is an increasing variable parameter, while q is a positive scalar
(Fig. 4.18a). For instance, if the loading surface is that corresponding to Mises’s
4.2 Inelastic Bodies 163

Fig. 4.18 Law of hardening: f = k2


isotropic (a); translated (b) f = k2
f =0
f =0
O'
O O

(a) (b)

condition, hence if the surface is a circular cylinder, then—immaterial on the path


of loading which provokes plastic deformations—the radius of the director circle
of the cylinder is increasing during the plastic deformations. We mention also the
schema of the kinematic hardening in Prager’s or in Ziegler’s variant, corre-
sponding to a translation; the respective law is called the law of translated
hardening and reads

f ðr0ij  aij Þ ¼ k2 ; ð4:1330 Þ

where aij are the co-ordinates of the centre of the loading surface while k ¼ const
(Fig. 4.18b); e.g., we can have aij ¼ cepij , where c is a characteristic constant of the
material. One can use also a hardening law which corresponds to an expansion and
a translation of the form

f ðr0ij  aij Þ ¼ k2 ðqÞ; ð4:13300 Þ

which can often correspond better to the reality; however, no one of the mentioned
hardening laws covers all the forms taken by the physical reality.
As well, the third element which characterizes the constitutive law is a law of
flow, i.e. a relation between stresses and strains; in such a relation often appear the
deviators of the respective tensors.
A law of flow is, in general, a differential (non-integrable) relation between
stresses and strains; thus, it constitutes a relation between the infinitesimal increase
of the stresses and strains, stresses and some parameters corresponding to the
plastic state. We make following hypotheses:
(i) The body is isotropic.
(ii) The volume strain is elastic and small, being proportional to the mean nor-
malstress; we have

r0 ¼ 3Ke0 ð4:134Þ

or
dr0 ¼ 3Kde0 : ð4:1340 Þ
164 4 Mathematical Models in Mechanics of Deformable Solids

(iii) The increases of the strain are of the form.

deij ¼ deeij þ depij ; i; j ¼ 1; 2; 3; ð4:135Þ

we can write the relations


 
1 3m
deeij ¼ drij  dij dr0 ; i; j ¼ 1; 2; 3; ð4:136Þ
2l 1þm

and

depij ¼ dep0 ¼ 0: ð4:137Þ

(iv) The stress deviator Dr is in direct proportion with the deviator of the plastic
increasing strains

Dpde ¼ Dr dk; ð4:138Þ

where dk is an infinitesimal scalar coefficient. This hypothesis shows that the state
of stress determines the instantaneous increase of the plastic strain.
Taking into account (4.137), we notice that e0p p
ij ¼ eij , so that one can write the
relation (4.138) in the form

depij ¼ r0ij dk; i; j ¼ 1; 2; 3: ð4:139Þ

For the plastic unit deformation work we have

dW p ¼ rij depij ; ð4:140Þ

introducing (4.139) and using the relation (2.55), (2.57), (2.570 ), we get

dW p ¼ 2J20 dk: ð4:1400 Þ

From dW p  0 it results dk  0 too, because J20 [ 0.


In this case, the relation (4.135) leads to the law of flow
dW p 0
deij ¼ deeij þ r ; i; j ¼ 1; 2; 3; ð4:141Þ
2J20 ij

taking into account (4.136) and introducing the stress and strain deviators, we can
write the law of flow (4.141) in the form
1 dW p
de0ij ¼ dr0ij þ 0 r0ij ; i; j ¼ 1; 2; 3; ð4:142Þ
2l 2J2
4.2 Inelastic Bodies 165

1
de0 ¼ dr0 : ð4:1420 Þ
3K
In case of infinitesimal deformations, it results the law

W_p 0
e_ ij ¼ e_ 0ij þ r ; i; j ¼ 1; 2; 3; ð4:143Þ
2J20 ij

which also reads


1 0 W _p
e_ 0ij ¼ r_ ij þ 0 r0ij ; i; j ¼ 1; 2; 3; ð4:144Þ
2l 2J2

1
e_ 0 ¼ r_ 0 : ð4:1440 Þ
3K
These relations constitute the Prandtl-Reuss law of flow. In this case, it is cus-
tomary to use Mises’s condition of plasticity; the law of flow (4.144) becomes

1 0 W _p
e_ 0ij ¼ r_ ij þ 2 r0ij ; i; j ¼ 1; 2; 3; ð4:14400 Þ
2l 2se
_ p is given by
where W
_ p ¼ rij e_ pij ¼ r0ij e_ 0p
W ij : ð4:145Þ

These relations are useful in the case in which the elastic deformations cannot be
neglected.
Assuming that the modulus of transverse elasticity tends to infinite ðl ! 1Þ,
that is if the material becomes rigid in the elastic domain, one obtains the Saint-
Venant-Lévy-Mises law of flow in the form

W_p 0
e_ 0ij ¼ r ; i; j ¼ 1; 2; 3; ð4:146Þ
2s2e ij

e_ 0 ¼ 0: ð4:1460 Þ
From (4.146) it results that the principal axes of the stress and of the strain velocity
tensors coincide. One obtains
W_p 0 0
r0ij e_ 0ij ¼ r r
2s2e ij ij
too; using Mises’s condition of plasticity, it results
_ p ¼ r0ij e_ 0ij ;
W ð4:1450 Þ

_ e ¼ 0, where W e is the unit work which corresponds to


a relation equivalent to W
the elastic deformation.
166 4 Mathematical Models in Mechanics of Deformable Solids

The relation (4.146) allows to write


_ p Þ2
ðW
e_ 0ij e_ 0ij ¼ ;
2s2e

in this case, wherefrom we get


qffiffiffiffi
_ p ¼ 2se I20 ;
W ð4:147Þ

I20 being the quadratic invariant of the strain velocity deviator.


The law of flow (4.146) may be thus written in the form
0 0
e_ ij rij
pffiffiffiffi0 ¼ ; i; j ¼ 1; 2; 3: ð4:14600 Þ
I2 se

This law is applied when the plastic deformations are much more greater than
the elastic ones. Thus, it can be applied to describe the processes of metal working:
rolling, drawing, cutting etc.
Assuming that the plastic work of deformation is given by a relation of the form

dUð
dW p ¼ r;
d ð4:148Þ
r
d
where we made to intervene the intensity of stresses, and denoting
1 rÞ
dUð
0 ¼ rÞ;
Fð
J2 r
d

we obtain
dk ¼ Fð
rÞd
r;
the law of flow (4.141) becomes

rÞr0ij d
deij ¼ deeij þ Fð r; i; j ¼ 1; 2; 3: ð4:1410 Þ

This law, corresponding to a solid hardening body, is valid if dr  0; in case of


equality we get a neutral loading.
Instead of the differential theories presented above, one uses often a finite
theory, where the physical relations represent a generalization of the physical
relations of the elastic case, appearing as an extrapolation of them beyond the
elastic state. The theory of plastic deformation can be applied if, in the loading
process, following conditions are satisfied.
(i) The principal directions of the stress tensor remain unchanged.
(ii) The configuration of the stress deviator remains unchanged.
These conditions are fulfilled if the variation of the components of the tensor
Tr is proportional; such a loading is called simple.
4.2 Inelastic Bodies 167

The hypotheses at the basis of the physical law corresponding to the theory of
plastic deformation are:
(i) The body is isotropic.
(ii) The mean normal stress is in direct proportion with the mean strain, the
coefficient of proportionality being the same both in the elastic and in the
plastic state
(iii) The deviators of the stress and strain tensors are in direct proportion.
The Nadai-Hencky-Ilyushin law can be written in the form

e0ij ¼ wr0ij ; i; j ¼ 1; 2; 3; ð4:149Þ

1
e0 ¼ r0 ; ð4:1490 Þ
3K
where w ¼ wð r; eÞ is a function which must be specified on an experimental way
for each case. The law may be written in the form
1
eij ¼ r0 dij þ wr0ij ; i; j ¼ 1; 2; 3; ð4:14900 Þ
3K
too; solving with respect to the stresses, it results
1 0
rij ¼ 3Ke0 dij þ e ; i; j ¼ 1; 2; 3: ð4:149000 Þ
w ij
One has

e0ij e0ij ¼ w2 r0ij r0ij ;

from (4.149), which leads to

I20 ¼ w2 J20 ð4:150Þ

or to the important relation


2
e ¼ w
r: ð4:1500 Þ
3
The unit work of deformation is given by
 
1
dW ¼ rij deij ¼ rij d r0 dij þ wr0ij ;
3K

from which we have


dW ¼ 3r0 de0 þ r
de: ð4:151Þ
168 4 Mathematical Models in Mechanics of Deformable Solids

Taking
1
w¼ ¼ const; ð4:152Þ
2l
one obtains Hooke’s linearly elastic law, for which the unit work of deformation is
given by
3
W ¼ ð3Ke20 þ le2 Þ ð4:153Þ
2
and for which Green’s formulae (4.9) are valid.
Using Mises’s condition of plasticity, one gets
pffiffiffi
3 e
w¼ : ð4:154Þ
2 se
The unit work of deformation corresponding to a deformation potential reads
9 pffiffiffi
W ¼ Ke20 þ 3see: ð4:155Þ
2
The relation between stresses and strain takes the form
2 se
rij ¼ 3Ke0 dij ¼ pffiffiffi e0ij ; i; j ¼ 1; 2; 3; ð4:156Þ
3 e
in case of an incompressible body, we have
2 se
rij  r0 dij ¼ pffiffiffi eij ; i; j ¼ 1; 2; 3: ð4:1560 Þ
3 e
The formulae of (4.9) type remain, further, valid.
In case of a hardening body, we assume that
1
w¼ gð
rÞ; ð4:157Þ
2
rÞ is a function which must be specified; from (4.1500 ) it results
gð
where 
1
e ¼ r gð
 rÞ: ð4:158Þ
3
Between stresses and strains takes place the relation
1 1
eij ¼ r0 dij þ  rÞr0ij ;
gð i; j ¼ 1; 2; 3; ð4:159Þ
3K 2
introducing a new function gðeÞ, defined by the relation
1
gðeÞ ¼ ; ð4:160Þ
gð
 rÞ
4.2 Inelastic Bodies 169

we have

rij ¼ 3Ke0 dij þ 2gðeÞe0ij ; i; j ¼ 1; 2; 3: ð4:1590 Þ

The unit work of deformation is given by


Z
9
W ¼ Ke20 þ 3 gðeÞede: ð4:161Þ
2
We notice that one can use further formulae of (4.9) type.

4.2.2 Bodies with Viscous Properties

Another factor which must be taken into consideration is the time factor. Direct
observations have shown, e.g., that the earth is settled in time under the action of
the constructions that it supports (which are deformed too). Various bodies con-
sidered to be solid (as steel, concrete, mountain glacier etc.) flow. From intuition
one passes to abstraction, to considerations of new models [1, 7, 24].
To can describe the phenomena in which appears the time variable we will use
a relation between stresses and strains of the form
f ðr; e; tÞ ¼ 0; ð4:162Þ

in a one-dimensional case. In what follows, we consider the medium as visco-


elastic, hence formed by two media, from which one is perfect elastic and the
second one has viscous properties. Thus, beside Hooke’s model we consider
Newton’s model (linear relation between the stress and the deformation velocity)
for viscous fluids. The Newtonian viscous model can be represented by a damper
(Fig. 4.19), in the one-dimensional case, and by a law of the form
rðtÞ ¼ g_eðtÞ; ð4:163Þ

where g is a coefficient of viscosity.

Fig. 4.19 Newton’s


technical model
170 4 Mathematical Models in Mechanics of Deformable Solids

From the linear theory of elasticity and the dynamics of viscous fluids one
obtains thus the linear theory of viscoelasticity, which implies two hypotheses:
(i) The bodies undergo infinitesimal deformations.
(ii) The mechanical properties of the bodies are subjected to Boltzmann’s prin-
ciple of superposition of effects.
Thus, the relation (4.162) represents a differential, integral or integro-differ-
ential equation.

4.2.2.1 Creep. Relaxation

Let be a cylindrical sample of viscoelastic material, subjected to a simple axial


loading. At a moment t ¼ t0 one has the stress r0 and the strain e0 in the elastic
domain; assuming that the stress r0 remains constant and taking, for the sake of
simplicity, t0 ¼ 0 (Fig. 4.20a), the strain e increases in time. This phenomenon is
called creep; to study it, we consider the function
1
uðt; r0 Þ ¼ eðtÞ; ð4:164Þ
r0
which we call creep function and where we have taken the moment t0 ¼ 0 as
origin for the time variable.
At the moment t ¼ 0 þ 0 the strain increases instantaneously till the value
1
eð0 þ 0Þ ¼ r0 ; ð4:165Þ
E
where E is the modulus of longitudinal elasticity; then the strain increases in time
and is given by (Fig. 4.20b)
1
eðtÞ ¼ eð0 þ 0Þ þ ef ðtÞ ¼ r0 þ r0 cðtÞ; ð4:166Þ
E
where cðtÞ is a function which is experimentally determined. We have thus
1
uðt; r0 Þ ¼ þ cðtÞ ¼ uð0 þ 0; r0 Þ þ cðtÞ; ð4:1640 Þ
E
if t0 [ 0, then we can write
uðt  t0 ; r0 Þ ¼ uðt0 þ 0; r0 Þ þ cðt  t0 Þ: ð4:16400 Þ
The creep function is independent on the form of the body, depending only on
time; this function increases with t, while its derivative decreases. For cðtÞ ¼ 0 we
have to deal with elastic deformations.
If at the moment t ¼ t1 we are unloading the body, then the strain decreases
suddenly, at the beginning, from t ¼ t1 þ 0 tending to the residual strain er ; the
respective curve is called returning curve (or curve of inverse creep). If er ¼ 0,
4.2 Inelastic Bodies 171

Fig. 4.20 Creep: rðtÞ vs t (t)


(a); eðtÞ vs t (b)

O t1 t
(a)

(t) = 0 (t; 0)

1
0
(t )
E
0

O t1 t
(b)

hence if after a long time the sample has no more a residual deformation, then the
body is called with delayed elasticity.
Let us consider the same loading of the bar at the moment t ¼ t0 , where we
take, analogically, t0 ¼ 0; assuming now that the strain e0 remains constant
(Fig. 4.21a), then the stress r decreases with time. This phenomenon is called
relaxation; to study it, we consider the function
1
wðt; e0 Þ ¼ eðtÞ; ð4:167Þ
e0
which we call relaxation function and where we considered that the moment
t0 ¼ 0 is origin for the time variable.
The stress decreases with time after a relation of the form (Fig. 4.21b)
rðtÞ ¼ rð0 þ 0Þ þ rr ðtÞ ¼ Ee0 þ e0 #ðtÞ; ð4:168Þ

where #ðtÞ is a function which must be experimentally determined. It results


wðt; e0 Þ ¼ E þ #ðtÞ ¼ wð0 þ 0; e0 Þ þ #ðtÞ; ð4:1670 Þ
if t0 [ 0, then we have
wðt  t0 ; e0 Þ ¼ wðt0 þ 0; e0 Þ þ #ðt  t0 Þ: ð4:16700 Þ
172 4 Mathematical Models in Mechanics of Deformable Solids

Fig. 4.21 Relaxation: eðtÞ (t)


vs t (a); rðtÞ vs t (b)

O t1 t
(a)

(t) = 0 (t; 0)

(t) 0
E 0

O t1 t

(b)

The relaxation curve is decreasing, its derivative being negative.


If at the moment t ¼ t1 we are unloading the body, then the normal stress
changes of sign and decreases; in case of bodies with delayed elasticity one has not
residual stresses for t ! 1.
The functions u and w play an important rôle in the theory of viscoelasticity.
If we apply the stresses r1 ; r2  r1 ; r3  r2 ; . . .; rn  rn1 at the moments
s1 ; s2 ; . . .; sn , afterwards maintaining them constant, and if we use Boltzmann’s
principle, then we get
X
n
eðtÞ ¼ ðri  ri1 Þuðt  si Þ; r0 ¼ 0; ð4:169Þ
i¼1

if the stress rðsÞ is a continuous function in the interval ð1; t and if we replace
the sum by an integral, then we can express the strain at the moment t in the form
Z t
drðsÞ
eðtÞ ¼ uðt  sÞds: ð4:170Þ
1 ds

If the phenomenon is studied for t  0, then we may write


4.2 Inelastic Bodies 173

Z t Z t
drðsÞ drðt  sÞ
eðtÞ ¼ hðtÞ uðt  sÞds ¼ hðtÞ uðtÞds; ð4:1700 Þ
0 ds 0 ds
where hðtÞ is Heaviside’s distribution, given by (A.165); this relation represents
the correspondent of Hooke’s law in the viscoelastic case.
Analogically, using the relaxation function, we get
Z t
deðsÞ
rðtÞ ¼ wðt  sÞds ð4:171Þ
1 ds

or
Z t Z t
deðsÞ deðt  sÞ
rðtÞ ¼ hðtÞ wðt  sÞds ¼ hðtÞ wðsÞds: ð4:1710 Þ
0 ds 0 ds
These relations, which correspond to Boltzmann’s principle, are known as
Boltzmann’s integral relations [1] and can be considered as integral equations if
one of the quantities eðtÞ, rðtÞ is unknown, while the other one appears as given. In
this case, the functions u and w are the nuclei of Volterra’s type equations.
We can express the loading at the moment t0 ¼ 0 in the form
rðtÞ ¼ r0 hðtÞ; ð4:172Þ
replacing in (4.170) and taking into account (A.166), we find again the relation
(4.164). An analogous result is obtained with a strain of the form
eðtÞ ¼ e0 hðtÞ; ð4:173Þ
In case of the loading-unloading process previously described (creep-inverse
creep), we may write (Fig. 4.20a)
rðtÞ ¼ r0 ½hðtÞ  hðt  t1 Þ; ð4:174Þ
obtaining
eðtÞ ¼ r0 ½uðtÞhðtÞ  uðt  t1 Þhðt  t1 Þ: ð4:175Þ

4.2.2.2 Models of Viscoelastic Bodies

Although the additivity of elastic deformations and viscous ones is a nonsense at a


microscopic level, it is extremely useful at a macroscopic one. Thus appears the
Kelvin [7] (or Voigt [24]) model, by linking a Hooke model with a Newton one
(Fig. 4.22), to which corresponds a constitutive law of the form
 
 d
rðtÞ ¼ EeðtÞ þ g_eðtÞ ¼ E 1 þ t eðtÞ; ð4:176Þ
dt
174 4 Mathematical Models in Mechanics of Deformable Solids

where the parameter t ¼ g=E represents the retardation time (or the creep time).
Assuming that the body is in the natural state for t  0, hence putting the initial
conditions eð0Þ ¼ rð0Þ ¼ 0, we can integrate the differential equation of first order
(4.176); we get thus
Z Z
hðtÞ t  hðtÞ t 
eðtÞ ¼ rðsÞeðtsÞ=t ds ¼ rðt  sÞes=t ds: ð4:177Þ
g 0 g 0
Putting rðtÞ given by (4.172) in (4.177), we find the creep function in the form
eðtÞ 1 
uðtÞ ¼ ¼ ð1  et=t ÞhðtÞ: ð4:178Þ
r0 E
The strain eðtÞ will vary after the law
r0  r0 
eðtÞ ¼ ð1  et=t ÞhðtÞ ¼ ð1  et=t Þþ ; ð4:179Þ
E E
its graphic being given in Fig. 4.23a. For t ! 1 we obtain e ¼ r0 =E.
If we admit a loading-unloading process to which corresponds a creep-inverse
creep phenomenon, then we take rðtÞ of the form (4.174) in (4.177) and obtain
(r 
0
ð1  et=t Þ for t  t1 ;
eðtÞ ¼ E 
ð4:180Þ
e1 et=t for t  t1 ;
where the strain
r0 t1 =t
e1 ¼ ðe  1Þ ð4:1800 Þ
E
is constant; the corresponding graphic is given in Fig. 4.23b. The strain tends to
zero for t ! 1; the rapidity by which the strain tends to zero depends on the
retardation time t , justifying thus its name.

Fig. 4.22 The Kelvin-Voigt


technical model

P
4.2 Inelastic Bodies 175

Fig. 4.23 Creep (a); creep- (t) (t)


inverse creep (b)
1
0
E
1

O t O t1 t
(a) (b)

Fig. 4.24 Maxwell’s


technical model

To can find the relaxation function corresponding to Kelvin’s model, we take


eðtÞ of the form (4.173) in (4.176); taking into account the relation of definition
(4.167) and (A.172), we find the function (in fact, a distribution) in the form
wðtÞ ¼ E½hðtÞ þ t dðtÞ: ð4:181Þ
When the strain is applied, the stress rðtÞ ¼ e0 wðtÞ tends to infinite, explaining
thus this result; indeed, due to the properties of viscosity, the body cannot be
subjected to a finite instantaneous deformation for a finite instantaneous variation
of stress.
Analogically, by linking in series a Hooke model with a Newton one
(Fig. 4.24), one obtains a Maxwell model to which corresponds the law
 
1 1 1 1 d
_ þ rðtÞ ¼
e_ ðtÞ ¼ rðtÞ þ rðtÞ: ð4:182Þ
E g E t dt
Assuming that the body is in the natural state for t  0, one can integrate the
differential equation of first order (4.182); one obtains
Z t
deðsÞ ðtsÞ=t
rðtÞ ¼ EhðtÞ e ds: ð4:183Þ
0 ds
176 4 Mathematical Models in Mechanics of Deformable Solids

Taking eðtÞ in the form (4.173), we find the relaxation function corresponding
to Maxwell’s model

wðtÞ ¼ EhðtÞet=t ; ð4:184Þ
t plays here the rôle of a relaxation time. The graphic of this function is given in
Fig. 4.25a.
Taking rðtÞ of the form (4.172) in the constitutive law (4.182), one obtains the
creep function corresponding to Maxwell’s model; observing that
d d
tþ ¼ ½thðtÞ ¼ hðtÞ þ tdðtÞ ¼ hðtÞ; ð4:185Þ
dt dt
on the basis of the relations (A.167) and (A.169), we get
1 t
1h tþ i
uðtÞ ¼ 1 þ  hðtÞ ¼ hðtÞ þ  ; ð4:186Þ
E t E t
the creep function being defined by the relation (4.164). Its graphic is a straight
line of slope g, which, at the moment t ¼ 0, leads to uð0 þ 0Þ ¼ 1=E (Fig. 4.25b).
If we take rðtÞ of the form (4.174), we obtain
1 h t
t  t1
i
uðtÞ ¼ 1 þ  hðtÞ  1 þ  hðt  t1 Þ
E t t
1 n t
t1 o
¼ 1 þ  ½hðtÞ  hðt  t1 Þ þ  hðt  t1 Þ ; ð4:187Þ
E t t
there corresponds the graphic in Fig. 4.25b.
Although these models have been obtained in an intuitive way, they are par-
ticularly useful, facilitating the scientific imagination, to may build up more
complex models, reflecting also other aspects of the reality. It is important that
these models can explain the phenomena of creep and relaxation too, which occur
with time.

(t)
(t)

E
1
E
1

E E
e t1
O t1 t O t1 t
(a) (b)

Fig. 4.25 Maxwell’s model: relaxation (a); creep (b)


4.2 Inelastic Bodies 177

Fig. 4.26 The generalized E1 E2 En


Kelvin technical model
P

1 2 n

A qualitatively superior model is the generalized Kelvin model, obtained by


grouping into series n simple Kelvin models, with the retardation times ti ¼ gi =Ei ;
i ¼ 1; 2; . . .; n (Fig. 4.26); the total strain is given by
X
n
eðtÞ ¼ ei ðtÞ; ð4:188Þ
i¼1

but in every element one has the same normal stress


 
 d
rðtÞ ¼ Ei 1 þ ti ei ðtÞ; i ¼ 1; 2; . . .; n: ð4:189Þ
dt
We get
Xn   1
 d
eðtÞ ¼ rðtÞ Ei 1 þ ti ð4:1880 Þ
i¼1
dt

from (4.188), (4.189); this is an ordinary differential equation of order n, written in


an operational form, which leads to
Xn Z
1 t 
eðtÞ ¼ hðtÞ rðtÞeðtsÞ=ti ds; ð4:18800 Þ
i¼1
g i 0

where we assumed that, at the moment t ¼ 0, the body is free of loading.


If we take rðtÞ of the form (4.172) in the above relation, then we get the creep
equation
Xn
1 

uðtÞ ¼ hðtÞ 1  et=t ; ð4:190Þ


E
i¼1 i

which can better approximate the real phenomenon.


We may write, in general, Ei ¼ Ei ðti Þ; we introduce the quantities Ki ðti Þ given
by
1 1
¼ K Ki ; i ¼ 1; 2; . . .; n; ð4:191Þ
Ei E0

with

1 Xn
1 X n
¼ ; Ki ¼ 1; ð4:1910 Þ
E0K EK i¼1
i¼1 i
178 4 Mathematical Models in Mechanics of Deformable Solids

passing from a discrete spectrum, characterized by the discrete parameter ti , to a


continuous parameter k, we may introduce the notation
 
d 1
KðkÞ ¼ EK ; ð4:192Þ
dk E

with
Z 1 Z 1
1 dk
¼ ; KðkÞdk ¼ 1: ð4:1920 Þ
EK 0 EðkÞ 0

Thus, the creep function corresponding to a generalized Kelvin model with con-
tinuous retardation spectrum is given by
Z
hðtÞ 1
uðtÞ ¼  KðkÞð1  et=k Þdk: ð4:193Þ
EK 0
In an analogous way, one can group in parallel n simple Maxwell models of
relaxation times ti ¼ gi =Ei ; i ¼ 1; 2; . . .; n (Fig. 4.27), obtaining thus the gen-
eralized Maxwell model; the total stress is given by
X
n
rðtÞ ¼ ri ðtÞ; ð4:194Þ
i¼1

the strain being the same for each element


eðtÞ ¼ ei ðtÞ; i ¼ 1; 2; . . .; n: ð4:195Þ
One can thus write
Xn  1
1 d
rðtÞ ¼ e_ ðtÞ E
i¼1 i
þ ; ð4:1940 Þ
ti dt

obtaining thus an ordinary differential equation of order n, written in an opera-


tional form; assuming that the body is subjected to no one loading at the moment
t ¼ 0, it results

Fig. 4.27 The generalized


Maxwell technical model
E1 E2 En-1 En

1 2 n-1 n

P
4.2 Inelastic Bodies 179

X
n Zt
deðsÞ ðtsÞ=ti
rðtÞ ¼ hðtÞ Ei e ds: ð4:19400 Þ
i¼1
ds
0

If one takes eðtÞ of the form (4.173), then one obtains the relaxation function
X
n

wðtÞ ¼ hðtÞ Ei et=ti : ð4:196Þ
i¼1

Starting from the quantities Fi ¼ Fi ðti Þ, given by


Ei
Fi ¼ ; i ¼ 1; 2; . . .; n; ð4:197Þ
E0M
X
n X
n
E0M ¼ Ei ; Fi ¼ 1; ð4:1970 Þ
i¼1 i¼1

we can pass, as in the preceding case, from a discrete spectrum to a continuous


one; we introduce the notation
1 dEðkÞ
FðkÞ ¼  ; ð4:198Þ
EM dk

with
Z 1 Z 1

EM ¼ EðkÞdk; FðkÞdk ¼ 1: ð4:1980 Þ
0 0

The relaxation function corresponding to a generalized Maxwell model with a


continuous relaxation spectrum will be thus given by
Z 1

wðtÞ ¼ EM hðtÞ FðkÞ et=k dk: ð4:199Þ
0

Starting from the simple Kelvin and Maxwell models, to which we associate
Hooke and Newton models, we can build up other models of viscoelastic bodies,
useful for various particular bodies too; these models will be models with more
than two parameters.
We can thus set up a model with three parameters, obtained by grouping into
series a Hooke model with a Kelvin one (Fig. 4.28a); the corresponding consti-
tutive law reads
   
E1 þ E2 d 1 d
þ rðtÞ ¼ E1  þ eðtÞ; ð4:200Þ
g2 dt t1 dt
180 4 Mathematical Models in Mechanics of Deformable Solids

Fig. 4.28 Grouping in series


a Hooke technical model and
a Kelvin one (a); Zener’s
E1
model (b)
E2
E1
E2 2 2

P P
(a) (b)

while the function of creep is given by



1 1 t=t2
uðtÞ ¼ þ ð1  e Þ hðtÞ; ð4:201Þ
E1 E 2

where t2 ¼ g2 =E2 ; we notice that


1 1 1
uð0 þ 0Þ ¼ ; uð1Þ ¼ þ ;
E1 E1 E2
ð4:2010 Þ
1 t1 =t
uðt1  0Þ ¼ uð1Þ  e 2:
E2
Taking rðtÞ of the form (4.174) in (4.200), we get
eðtÞ ¼ r0 ½uðtÞhðtÞ  uðt  t1 Þhðt  t1 Þ; ð4:202Þ

the corresponding graphic being given in Fig. 4.29; we notice that we have a jump
equal to r0 =E1 at the moment t ¼ t1 , the strain eðtÞ tending exponentially to zero.
Analogically, one obtains Zener’s model, by grouping in parallel a Hooke
model and a Maxwell one (Fig. 4.28b); the corresponding constitutive law reads
   
1 d 1 E2 d
þ rðtÞ ¼ E 1  þ 1 þ eðtÞ; ð4:203Þ
t1 dt t2 E1 dt

the creep function being given by



uðtÞ ¼ ðE1 þ E2 et=t2 ÞhðtÞ: ð4:204Þ
Grouping into series a Newton model and a Kelvin one (Fig. 4.30a), one
obtains a model the constitutive law of which is of the form
   
1 1 g2 d d  d
þ 1þ rðtÞ ¼ 1 þ t2 eðtÞ; ð4:205Þ
g 1 E2 g1 dt dt dt
4.2 Inelastic Bodies 181

Fig. 4.29 Grouping in series (t) = 0 (t)


a Hooke technical model and
a Kelvin one: eðtÞ vs t (a);
rðtÞ vs t (b) ( )
(t1-0)

0
E1
1
1
0
E1

O t1 t
(a)

(t)

O t1 t
(b)

Fig. 4.30 Grouping in series


a Newton technical model
1
and a Kelvin one (a);
grouping in parallel a Newton
technical model and a E2
Maxwell one (b)
1
E2 2 2

P P
(a) (b)

the corresponding creep function is


hðtÞ h g 
i
uðtÞ ¼ 1 þ 1 ð1  et=t2 Þ : ð4:206Þ
g1 E
Grouping in parallel a Newton model and a Maxwell model (Fig. 4.30b), we
obtain a constitutive law of the form
   
1  d d g1  d
1 þ t2 rðtÞ ¼ 1þ 1 þ t2 eðtÞ: ð4:207Þ
g2 dt dt g2 dt
182 4 Mathematical Models in Mechanics of Deformable Solids

Fig. 4.31 Burgers’s


technical model
E1

E2 2

A model often used is the Burgers model obtained by linking into series a
Kelvin model and a Maxwell one (Fig. 4.31); in an operational way, the consti-
tutive law is of the form
 
1  d d=dt
e_ ðtÞ ¼ 1 þ t1 þ rðtÞ: ð4:208Þ
g1 dt E2 ð1 þ t2 d=dtÞ
The creep function reads

1  1 t=t2
uðtÞ ¼ ðt þ tÞ þ ð1  e Þ hðtÞ: ð4:209Þ
g1 1 E2
If rðtÞ is given by (4.174), then one obtains eðtÞ in the form (4.207) (Fig. 4.32).
On the same line, one can link into series a generalized Kelvin model and a
Newton one or one can link in parallel a generalized Maxwell model with a Hooke
model and a Newton one; the corresponding curves of creep and relaxation may be
thus closer to those obtained experimentally for some particular bodies.
We can successfully effect a study of the models presented above by means of
the Laplace transform; using the results in Sect. A.3.4.2, we can introduce the
transforms of the stress and of the strain in the form
Z 1
L½rðtÞ ¼ r~ðpÞ ¼ rðtÞept dt;
0
Z 1 ð4:210Þ
L½eðtÞ ¼ ~eðpÞ ¼ eðtÞept dt;
0

where p is the new complex variable, corresponding to the time variable. Thus,
e.g., the constitutive law (4.176), corresponding to Kelvin’s model, becomes
~ðpÞ
r
~eðpÞ ¼ ; ð4:211Þ
gð1=t þ pÞ

effecting the inverse Laplace transform, we get the constitutive law (4.177).
Analogically, the Laplace transforms of the creep and relaxation functions
will be
4.2 Inelastic Bodies 183

Fig. 4.32 Burgers’s model. (t)


Creep: uðtÞ vs t (a); rðtÞ vs t
(b)

E1
1
1
E1

1t
1
1

O t1 t
(a)

(t)

O t1 t
(b)

1
L½uðtÞ ¼ u
~ ðpÞ ¼ ; ð4:212Þ
Epð1 þ t pÞ

~ E
L½wðtÞ ¼ wðpÞ ¼ ð1 þ t pÞ: ð4:2120 Þ
p
We notice that the relation

p2 u ~
~ ðtÞwðpÞ ¼ 1; ð4:213Þ

by means of which one can pass from the creep function to the relaxation one and
conversely; this property is valid for all the bodies which admit Boltzmann’s
principle of superposition and which verify the relations (4.1700 ), (4.1710 ).

4.2.2.3 General Constitutive Laws

To can represent the particular models considered above in a general form, we


introduce the operators
X
n X
n
d
PðDÞ ¼ ai Di ; QðDÞ ¼ bj D j ; D¼ ; ð4:214Þ
i¼1 j¼1
dt
184 4 Mathematical Models in Mechanics of Deformable Solids

where ai ; i ¼ 1; 2; . . .; n, bj ; j ¼ 1; 2; . . .; m, are constant coefficients of the mate-


rial; the corresponding constitutive law can be written in the form
PðDÞrðtÞ ¼ QðDÞeðtÞ; ð4:215Þ

which take also the form of Boltzmann’s integral relations (4.1700 ), (4.1710 ).
Applying the Laplace transform, it results
~ðpÞ QðPÞ
r
¼ ; ð4:216Þ
~eðpÞ PðpÞ
starting from Boltzmann’s relations, one obtains the Laplace transforms of the
creep and relaxation functions
1 ~eðpÞ 1 PðpÞ
~ ðpÞ ¼
u ¼ ; ð4:217Þ
pr~ðpÞ p QðpÞ

~ 1r~ðpÞ 1 QðpÞ
wðpÞ ¼ ¼ ; ð4:2170 Þ
p ~eðpÞ p PðpÞ
which verify the relation (4.213). Using the decomposition in simple fractions, we
obtain, easily, the inverse Laplace transforms too.
We consider now the case of a three-dimensional loading, where appear the
tensors Tr and Te with all their components; we notice, from the very beginning,
that it is important to use the decomposition of these tensors in their spheric and
deviator parts. Thus, in case of Hooke’s model, we use the constitutive law in the
form (4.92), (4.920 ).
Using the mode of construction of a Kelvin model in the one-dimensional case,
we obtain a viscoelastic model of generalized Kelvin type, specified by the con-
stitutive law
 
0  d
rij ðtÞ ¼ 2l 1 þ t e0 ðtÞ; i; j ¼ 1; 2; 3; ð4:218Þ
dt ij

where t ¼ g=l is the retardation time; we associate to it the relation (4.92).


Analogically, we may construct a model of viscoelastic body with the constitutive
law
 
1 d 0
þ r ðtÞ ¼ 2l_e0ij ðtÞ; i; j ¼ 1; 2; 3; ð4:219Þ
t dt ij

adding the relation (4.92), we obtain a viscoelastic model of generalized Maxwell


type, where t ¼ g=l is a relaxation time.
Analogically, one may consider the ‘‘standard’’ model of the viscoelastic body,
specified by the constitutive law
4.2 Inelastic Bodies 185

   
d 0 d 0
1 þ t1 rij ðtÞ ¼ 2l 1 þ t2 e ðtÞ; i; j ¼ 1; 2; 3; ð4:220Þ
dt dt ij

where l; tk ; k ¼ 1; 2, are the constants of the material, to which is associated the
elastic relation (4.92); this model corresponds to the Zener model of the one-
dimensional case.
Using the operators
n0
X n00
X d
P1 ðDÞ ¼ a0i Di ; P2 ðDÞ ¼ a00i Di ; D¼ ; ð4:221Þ
i¼1 i¼1
dt

m0
X m00
X d
Q1 ðDÞ ¼ b0j D j ; Q2 ðDÞ ¼ b00j D j ; D¼ ; ð4:2210 Þ
j¼1 j¼1
dt

where a0i ; a00i ; b0j ; b00j are the constant coefficients of the material, we may write the
general differential constitutive law of a linear viscoelastic body in the form

P1 ðDÞr0ij ðtÞ ¼ Q1 ðDÞe0ij ðtÞ; i; j ¼ 1; 2; 3; ð4:222Þ

P2 ðDÞr0 ðtÞ ¼ Q2 ðDÞe0 ðtÞ: ð4:2220 Þ


In particular, in case of Hooke’s model, we have

P1 ðDÞ ¼ P2 ðDÞ ¼ 1;
ð4:223Þ
Q1 ðDÞ ¼ 2l; Q2 ðDÞ ¼ 3K;

as well, in case of a model of generalized Kelvin type, e.g., it results

P1 ðDÞ ¼ P2 ðDÞ ¼ 1;
 
d ð4:224Þ
Q1 ðDÞ ¼ 2l 1 þ t ; Q2 ðDÞ ¼ 3K;
dt

while in case of a model of generalized Maxwell type, we may write


1 d
P1 ðDÞ ¼ þ ; P2 ðDÞ ¼ 1;
t dt ð4:225Þ
d
Q1 ðDÞ ¼ 2l ; Q2 ðDÞ ¼ 3K:
dt
Applying the Laplace transform to the relations (4.222, 4.2220 ), we get

~0ij ðpÞ ¼ 2~
r lðpÞ~e0ij ðpÞ; i; j ¼ 1; 2; 3; ð4:226Þ

~ e0 ðpÞ;
~0 ðpÞ ¼ 3KðpÞ~
r ð4:2260 Þ
where we used the notations
186 4 Mathematical Models in Mechanics of Deformable Solids

Q1 ðpÞ ~ Q2 ðpÞ
~ðpÞ ¼
l ; KðpÞ ¼ ; ð4:22600 Þ
2P1 ðpÞ 3P2 ðpÞ

this constitutive law has the same form as that in case of linearly elastic bodies,
being written in the space of Laplace transforms. Obviously, we may write the
constitutive law in this space also in the form

~ij ðpÞ ¼ ~
r rll ðpÞdij þ 2~
kðpÞ~ lðpÞ~eij ðpÞ; i; j ¼ 1; 2; 3; ð4:227Þ

where

~ 1 Q2 ðpÞ Q1 ðpÞ Q1 ðpÞ
kðpÞ ¼  ~ðpÞ ¼
; l ; ð4:2270 Þ
3 P2 ðpÞ P1 ðpÞ 2P1 ðpÞ

analogue to the form of Hooke’s law.


Writing the relations (4.226), (4.2260 ) also in the form
~e0ij ðpÞ ¼ p~ r0ij ðpÞ;
u1 ðpÞ~ i; j ¼ 1; 2; 3; ð4:228Þ

~e0 ðpÞ ¼ p~
u2 ðpÞ~
r0 ðpÞ; ð4:2280 Þ
where
P1 ðpÞ P2 ðpÞ
~ 1 ðpÞ ¼
u ~ ðpÞ ¼
;u ; ð4:22800 Þ
pQ1 ðpÞ 2 pQ2 ðpÞ

we are led to the relations of Boltzmann type


Z t 0
drij ðsÞ
e0ij ¼ hðtÞ u1 ðt  sÞds; i; j ¼ 1; 2; 3; ð4:229Þ
0 ds
Z t
dr0 ðsÞ
e0 ðtÞ ¼ hðtÞ u2 ðt  sÞds; ð4:2290 Þ
0 ds
the function uk ðtÞ; k ¼ 1; 2 being creep functions.
As well, the constitutive law can be represented in the form

r ~ ðpÞ~e0 ðpÞ; i; j ¼ 1; 2; 3;
~0ij ðpÞ ¼ pw ð4:230Þ
1 ij

~ ðpÞ~e0 ðpÞ;
~0 ðpÞ ¼ pw
r ð4:2300 Þ
2

where

~ ðpÞ ¼ Q1 ðpÞ ; w
w ~ ðpÞ ¼ Q2 ðpÞ ; ð4:23000 Þ
1
pP1 ðpÞ 2 pP2 ðpÞ
4.2 Inelastic Bodies 187

the corresponding relations of Boltzmann type are


Z t 0
deij ðsÞ
r0ij ðtÞ ¼ hðtÞ w ðt  sÞds; i; j ¼ 1; 2; 3; ð4:131Þ
0 ds 1
Z t
0 de0 ðsÞ
r0 ðtÞ ¼ hðtÞ w ðt  sÞds; ð4:1310 Þ
0 ds 2
the functions wk ðtÞ; k ¼ 1; 2, being relaxation functions.
We notice that the relations

p2 u ~ ðpÞ ¼ 1;
~ k ðpÞw k ¼ 1; 2; ð4:232Þ
k

take place.
One can use much more complex linear models too, where the constitutive law
is an integro-differential one.
We notice that one can combine the property of viscosity with that of plasticity,
obtaining thus models of viscoplastic bodies. We mention thus Bingham’s model
for which the relation between stresses and strains is of the form

r0ij ðtÞ ¼ 2g1 e_ pij ðtÞ; i; j ¼ 1; 2; 3; ð4:233Þ

where g1 is a coefficient of viscosity corresponding to a non-linear viscous fluid,


given by
 
1
g1 ¼ g 1 þ ; ð4:2330 Þ
4kl

1=4kl being a characteristic of viscoplastic behaviour. Usually, one uses Mises’s


condition of plasticity.
Associating also an elastic component to this body, one obtains a model of
elastic-viscoplastic body; the relation between stresses and strains is of the form
1 0 1
e_ 0ij ðtÞ ¼ r_ ij ðtÞ þ rij ðtÞ; i; j ¼ 1; 2; 3; ð4:234Þ
2l 2g1
with

J20 [ k2 ; k ¼ const: ð4:2340 Þ


We will not make a detailed study of these bodies; we wish only to put in
evidence the possibility to build up much more complex models of deformable
solids.
188 4 Mathematical Models in Mechanics of Deformable Solids

References

A. Books

1. Boltzmann, L.: Populäre Schriften, 3rd edn. Leipzig (1925)


2. Carlslaw, H.S., Jaeger, J.C.: Conduction of Heat in Solids. Oxford University Press, Oxford
(1948)
3. Green, G.: Mathematical Papers. London (1871)
4. Lamé, G.: Leçons sur la théorie mathématique de l’élasticité des corps solides. Paris (1852)
5. Neumann, F.: Vorlesungen } uber die Theorie der Elasticität der festen K} orper und des
Lichtäters. Leipzig (1885)
6. Sokolnikoff, I.S.: Mathematical Theory of Elasticity, 2nd edn. McGraw-Hill, New York
(1956)
7. Thomson, W.: (Lord Kelvin) Mathematical and Physical Papers, I–III, (1882, 1884, 1890)
8. Timoshenko, S.P., Goodier, J.N.: Theory of Elasticity, 2nd edn., McGraw-Hill, New York

B. Papers

9. Adomit, G.: Determination of elastic constants of structured materials. In: IUTAM


Symposium, 1967, Freudenstadt–Stuttgart, Mechanics of Generalized Continua, vol. 80
(1968)
10. Bekhterev, P.: Analiticheskoe issledovanie obobshchenogo zakona Guka (Analytical
application of the generalized law of Hooke). J. Russkogo fiz.-khim. obshchestva, I. 7, 34
(1925), 8, 3 (1926)
11. Cauchy, A.L.: Mémoire sur les systèmes isotropes de points matériels. Mém. Acad. Sci. 22,
605 (1850)
12. Chentsov, N.G.: Issledovanie fanery, kak ortotropno plastinki (Applications of wood layers
as orthotropic plates). Tekhn. zam, vol. 91 (1936)
13. Duhamel, J.-M.-C.: Mémoire sur le calcul des actions moléculaires développées par les
changements de température dans les corps solides. Mém. Acad. savants étrangers, 5, 440
(1838)
14. Euler, L.: Determinatio onerum, quae columne gestare valent. Act. Acta Acad. Sci.
Petrograd, 2, 121 (1780)
15. Green, G.: On the laws of reflection and refraction of light at the common surface of two non-
crystallized media. Trans. Cambridge Phil. Soc. 7, 1 (1839)
16. Hoppman, W.H., Shahman, F.O.F.: Physical model of a 3-constant isotropic elastic material.
Trans. ASME. Ser. E J. Appl. Mech. 32, 837 (1965)
17. Joel, N., Wooster, W.A.: Number of elastic constants required in crystal elasticity. Nature
182, 1078 (1958)
18. Krishnan, R.S., Rajagopal, E.S.: The atomistic and the continuum theories of crystal
elasticity. Ann. der Physik 8, 121 (1961)
19. Lamé, G.: Mémoire sur les surfaces isostatiques dans les corps solides homogènes, en
èquilibre d’élasticité. J. Math. Pureset Appl. 6, 37 (1841)
20. Lamé, G., Clapeyron, B.-P.-E.: Mémoire sur l’équilibre intérieur des corps solides
homogènes. Mém. prés. divers sav. étr., vol. 4 (1883)
21. Rabinovich, A.L.: Ob uprugikh postoyannykh i prochnosti anizotropnykh materyalov (On
elastic constants and strength of isotropic materials). Trudy CAGI, vol. 582 (1946)
References 189

}
22. Teodorescu P.P.: Uber das kinetische Problem nichthomogener elastischer K} orper. Bull.
Acad. Pol. Sci., sér. Sci. Technol. 12, 867 (1964)
23. Teodorescu, P.P., Predeleanu, M.: Quelques considérations sur le problème des corps
élastiques hétérogènes. In: Proceedings of IUTAM-Symposium. Non-Homogeneity in
Easticity and Plasticity, Warszawa, 1958, Perganon Press, 31 (1959) [Bull. Acad. Pol. Sci.,
sér. Sci. Technol. 7, 81 (1959)]
24. Voigt, W.: Theoretische Studien } uber die Elastizitätsverhältnisse der Krystalle. I, II. Abh. der
K}onigl. Ges. Wiss., G}ottingen, 34 (1887)
Chapter 5
General Equations of the Theory of
Elasticity. Formulation of Problems

The local study made up to the present time allows us to define the state of strain
and stress at an arbitrary point of any solid body (the set of all the possible states of
strain and stress), without taking into account the specific conditions of the given
problem (external loads acting upon the given body, as well as the conditions in
which the latter happens to be at a given time). The aim of this chapter is just to
complete this study by a global one.

5.1 General Equations of the Theory of Elasticity

If we refer to a given body and a given state of strain and stress, then we must take
into account the limit conditions of the problem, that, in the static case, are
boundary conditions, while, in the dynamic case, are boundary conditions and
initial conditions. These conditions state precisely the solution of the problem,
Therefore, the local study must be finished off by a global study, liable to make the
connection between the considered body and the other bodies at any time, also
taking into account the motion of the respective body at a given time.
Moreover, we shall state the formulations for both static and dynamic problems
of the theory of elasticity in the classical case: homogeneous, isotropic, linearly
elastic bodies, subjected to an infinitesimal state of deformation.
A preliminary study of the system of partial differential equations of elasticity
was made by Gr. C. Moisil [28], who used the method of matrices associated with
systems of partial differential equations, elaborated in [7]; thus, various differential
equations to be verified by each of the unknown functions by means of certain
potential functions are emphasized.

P. P. Teodorescu, Treatise on Classical Elasticity, Mathematical and Analytical 191


Techniques with Applications to Engineering, DOI: 10.1007/978-94-007-2616-1_5,
Ó Springer Science+Business Media Dordrecht 2013
192 5 General Equations. Formulation of Problems

5.1.1 Statical Problems

We shall deal firstly with the case of a static state of strain and stress, specifying
the basic system of equations; the three fundamental problems are thus put in
evidence.

5.1.1.1 Basic System of Equations

In the case of the classical theory of elasticity we must determine 15 unknown


functions: 3 displacements ui ; 6 strains eij and 6 stresses rij ; to this end, we can use
15 equations, i.e. 3 differential equations of equilibrium
rij;j þ Fi ¼ 0; i ¼ 1; 2; 3; ð5:1Þ

6 differential relations between strains and displacements


1
eij ¼ ðui;j þ uj;i Þ; i; j ¼ 1; 2; 3; ð5:2Þ
2
and 6 finite relations between stresses and strains (Hooke’s constitutive law) in the
form
rij ¼ kell dij þ 2leij ; i; j ¼ 1; 2; 3; ð5:3Þ

or in the form
1
eij ¼ ½ð1 þ mÞrij  mrll dij ; i; j ¼ 1; 2; 3: ð5:30 Þ
E
These equations, which must be verified on the domain occupied by the elastic
body, form the fundamental system of equations of the linear elasticity; if we add
the boundary conditions, then they form the complete system of equations of the
classical theory of elasticity in the static case.

5.1.1.2 Boundary Conditions. Basic Problems

In displacements (first basic problem of the theory of elasticity) the conditions are
put on the boundary (the frontier C of the domain D) in the form (Fig. 5.1a)
n
ui ¼ ui ; i ¼ 1; 2; 3; ð5:4Þ
n n
where the displacement u; of components ui ; i ¼ 1; 2; 3; of the point of external
normal n is given by the linkage with other bodies. The cases in which the
conditions in displacements can be put on all the contour are uncommon; even-
tually, one puts conditions for certain derivatives of the displacements (e.g., for
strains).
5.1 General Equations of the Theory of Elasticity 193

(a) x3 (b) x3
n
u n p
n n

M M
r dA r dA

x1 x2 x1 x2

Fig. 5.1 Boundary conditions: in displacements (a), in stresses (b)

One must mention that these conditions are distinct from the conditions of fixity
(conditions of support) of the body, which are put at various points to determine its
rigid body motion.
The most times we know the external loads which act on the contour. At each
point of the contour must take place a mechanical equilibrium between the internal
n n
stresses rij and the external loads p (of components pi ; i ¼ 1; 2; 3; on the element
of area of external normal n); hence, we must put the condition that, on the contour
C, the stresses, obtained by computation, be equated by the given external loads
(Fig. 5.1b), using the relations
n
p i¼ rji nj ; i ¼ 1; 2; 3; ð5:5Þ

or the relations
n
p i ¼ ri ni ð!Þ; i ¼ 1; 2; 3; ð5:50 Þ

corresponding to the principal directions. If the normal and the tangential com-
ponents of the external load are given, then we can use relations of the form (3.25),
(3.28); we notice also that, if on an element of area of external normal n we have
not tangential components of the external load, then the direction n is a principal
one, the other principal directions being contained in the tangent plane.
If we succeed to integrate the system of equations considered above, verifying
the boundary conditions too, then the state of strain and stress is completely
determined, the problem of elastostatics being thus solved. The above problem, in
which the conditions on the boundary are put in stresses, is called the second basic
problem of the theory of elasticity.
As well (and this happens frequently) one can put conditions in displacements
on some portions of the contour (frontier Cu ) and conditions in stresses on other
portions of it (frontier Cr ; Cu [ Cr ¼ C; Cu \ Cr ¼ ;) (the mixed problem of the
theory of elasticity). One can conceive also more elaborate mixed problems, in
which, e.g., at the same point both conditions in displacements and in stresses are
put.
The problems of the theory of elasticity can be easier solved by choosing as
unknowns only the displacements or only the stresses, as the boundary conditions
194 5 General Equations. Formulation of Problems

are put, and eliminating the other unknowns between the 15 equations mentioned
above. One is thus led to a solution in displacements or to a solution in stresses of
these problems.
On can imagine a solution in strains of the problem (introducing, eventually, the
local rotations of rigid body) too, using also the equations of continuity (2.68); but
this solution is not interesting, from a practical point of view, because the con-
ditions on the contour are not put, customarily, in this form.
In a practical solution of the problems, we use that representation which is
better fitted to the boundary conditions; thus, in case of the first basic problem as
well as in case of the mixed one, a solution in displacements is convenient, while
in case of the second basic problem one can use successfully a solution in stresses.
In case of the mixed problem, it is useful to can express the conditions in
stresses by means of the displacements; starting from Hooke’s law, written in the
form
   
m 1 m
rij ¼ 2l eij þ ell dij ¼ 2l ðui;j þ uj;i Þ þ ul;l dij ;
1  2m 2 1  2m

it results
 
n n 1 m
p ¼ pk ik ¼ rjk nj ik ¼ 2l ðuj;k  uk;j Þ nj þ uk;j nj þ ul;l djk nj ik ;
2 1  2m

so that the stress vector along the direction of unit vector n is given by
 
n ou 1 m
p ¼ 2l þ n x curl u þ n div u : ð5:500 Þ
on 2 1  2m
The use of potential functions is particularly important to formulate the prob-
lems of the theory of elasticity; in case of a formulation in displacements one has
to do with displacement functions, while in case of a formulation in stresses with
stress functions. We will pay a special attention to such formulations; most of the
problems will be studied on this way.
In a solution in displacements of the static problem, we mention, first of all, the
formulation given by B. G. Galerkin [16], who uses three biharmonic displace-
ment functions; these functions may be considered as the components of a vector
(Galerkin’s vector). Using Almansi’s formula (A.100), one can state that in this
representation intervene six harmonic functions which must be determinate with
the aid of six conditions at a point of the contour. A triple functional indetermi-
nation occurs; the state of displacement is univocally determined, but the dis-
placement functions have not the same property.
Another formulation in displacements has been obtained by P. F. Papkovich
[33] in 1932, using four harmonic functions; it has been previously found by G.
D. Grodski, but published only in 1935, after being mentioned in 1927 by the
occasion of a conference. The conference has been forgotten, although Papkovich
recognized later the priority. This result has been independently formulated by
5.1 General Equations of the Theory of Elasticity 195

H. Neuber [31]; it is known as the Papkovich-Neuber representation. E. Sternberg


and R. A. Eubanks [46] dealt with the completeness of this representation (any
state of displacement can be expressed in this form; in any case, any representation
of this form expresses a possible state of displacement). Other investigations have
been made by M. E. Gurtin [2] and by E. Sternberg [45]. Using the method of
associated matrices, Gr. C. Moisil [26] did find again Galerkin’s vector on a simple
and elegant way. With connected problems dealt V. I. Blokh [13].
Relations between various kind of representations are considered by K. Mar-
guerre [24].
Dealing firstly only with the equations of equilibrium (5.1), we mention the
representations of J. C. Maxwell and G. Morera [29, 30], who use every one three
stress functions of class C3 ; which must not verify any other condition. B. Finzi
[1, 15] showed that these representations are particular cases of a representation
(the Beltrami–Finzi representation) by means of six arbitrary functions of class
C3 ; components of a symmetric tensor of second order (Finzi’s tensor). C. Weber
[56] finds again these results.
Starting from Finzi’s tensor, H. Schaefer [39–41] gives a representation by means
of six harmonic functions (components of a symmetric tensor of second order) and of
a biharmonic one, related with them. E. Kr} oner [22] and K. Marguerre [24] introduce
three supplementary conditions of divergence, so that the components of Finzi’s
tensor must be all biharmonic functions. H. Schaefer [39] puts in evidence the
relation between his representation and the Papkovich-Neuber one.
An important problem is that of the boundary conditions. In case of a two-
dimensional problem there intervene three stress functions (a function of G.-B.
Airy for the state of plane stress of the plate and two functions corresponding to
the bending of it), as it has been shown by H. Schaefer [41]; if one considers a
fictitious bar along the contour of the given domain, it results that the values of
these functions on it are given by the components of the moment vector on the bar
axis, subjected to the action of the external loads. In case of the plane problem the
value of the stress function corresponds to the bending moment, while the values
of the normal and tangential derivatives correspond to the axial and the shear
forces, respectively. The determination of the efforts in the bar leads, in general, to
a sextuple static indetermination, which plays an important rôle in case of multiply
connected domains. In the case of a simple connected domain one can make a cut,
which leads to the study of an isostatic bar; it remains to be seen how the given
informations may be used to separate the boundary conditions in case of a shell.
G. Peretti [34] puts in evidence a statical meaning of the components of Finzi’s
stress tensor, finding results similar to those obtained for Airy’s function.
H. Schaefer [41] finds the analogy between the values of the components of Finzi’s
tensor on the contour and the components of the efforts tensor in a shell along the
boundary. On this way one may make (perhaps) a uncoupling of the boundary
conditions, which, added to a uncoupling of the equations, could lead to a new
formulation of the space problems of the theory of elasticity (a formulation by
means of six stress functions, knowing the boundary conditions for each of these
196 5 General Equations. Formulation of Problems

functions); the problems of the theory of elasticity could be thus reduced to


problems of the classical equations of mathematical physics. It is true, one must
solve the problem of a three times hyperstatic shell. As in the two-dimensional
case, this could be of importance only in the case of multiply connected domains;
in case of a simply connected domain, by a complete cut one can obtain an
isostatic shell, being easier to obtain the values of the potential functions, without
losing anything from the generality of the result. H. Schaefer [41] considered thus
the half-space and the torsion of an axisymmetric body; the problem remains open
in the general case.
Topological considerations have been made by Rieder [35–38].
We have given in 1956 [47, 48] a representation by means of three biharmonic
stress functions in three variables, to which three harmonic functions, each one in
two variables, are added; one obtains thus a representation by means of four (or
three if m 6¼ 1=4) harmonic functions in three variables, to which the mentioned
functions in two variables are added.

5.1.2 Dynamical Problems

The basic problems of the theory of elasticity are the same in the dynamic case as
in the static one; but we must notice that both the displacements and strains, as
well as the stresses are functions not only of point but also of the time variable,
which involves the initial conditions too.

5.1.2.1 The Basic System of Equations

To determine the 15 unknown functions of the classical elastic problems in the


dynamic case, we dispose as well of 15 equations, i.e. 3 equations of motion
rji;j þ Fi ¼ q€
ui ; i ¼ 1; 2; 3; ð5:6Þ

6 relations between strains and displacements (5.2) and 6 finite relations between
stresses and strains (5.3) or (5.30 ).
These equations form the basic system of equations of linear elastodynamics; if
we add also the limit conditions, they form the complete system of equations of the
classical theory of elasticity in the dynamic case.

5.1.2.2 Limit Conditions. Basic Problems

The three basic problems of the theory of elasticity are defined as in the static case;
the conditions on the contour are put in the form (5.4), in case of the first basic
problem, or in one of the forms (5.5), (5.50 ), in case of the second basic problem,
5.1 General Equations of the Theory of Elasticity 197

for t  t0 ; where t0 is the initial moment. In case of the mixed problem, one
proceeds analogically.
To the boundary conditions one adds the initial conditions
ui ¼ u0i ; u_ i ¼ u_ 0i ; i ¼ 1; 2; 3; ð5:7Þ

which must be fulfilled in the domain D occupied by the body and on the contour
C, where they must connect to the boundary conditions mentioned above; one can
put initial conditions also for the stress tensor Tr as well as for the velocity stress
tensor Tr_ , in the form
rij ¼ r0ij ; r_ ij ¼ r_ 0ij ; i; j ¼ 1; 2; 3: ð5:8Þ

We have denoted by u0i ; u_ 0i the displacements and the displacement velocities at the
initial moment t ¼ t0 ; respectively; analogically, r0ij and r_ 0ij represent the stresses
and the stress velocities, at the same initial moment, respectively. These functions
depend on the space variables x1 ; x2 ; x3 : To these conditions one adds also the
conditions of support (conditions of fixity) of the body.
The conditions on the contour, the initial conditions and the fixity conditions
form the limit conditions of the problem.
If we succeed to integrate the above considered system of equations, putting the
limit conditions too, then the state of strain and stress is completely determined;
the problem is thus solved.
As in the static case, we are led to a solution in displacements or to a solution in
stresses of the problem, as the unknown functions are mostly involved in the limit
conditions.
The formulation of the problems of elastodynamics constituted, since the
beginning of nineteenth century, the object of many researches; especially, general
formulations were sought for, in the case of homogeneous, isotropic, linearly
elastic bodies, of G. Lamé’s equations [6] in the dynamic case. These equations are
to be found by C. L. M.-H. Navier in the particular case when k ¼ l; as well as
later by G. Lamé and B. Clapeyron and by G. G. Stokes. In the mechanics of fluids
they are known as the Navier-Stokes equations.
In this direction we mention, firstly, the representations with the help of a vector
and of a scalar displacement function, based upon the decomposition of a vector
into the sum of a gradient and a curl. The first of such a representation was given by
S. D. Poisson in 1829 in the particular case in which m ¼ 1=4; without emphasizing
the completeness of the representation or the fact that the solenoidal part of the
representation is a curl; it was found later on by G. Lamé [6] in 1852, for an
arbitrary Poisson’s ratio. A. Clebsch resumed the problem in 1863 and gave a
demonstration, yet insufficient to ensure the completeness of the representation; the
subsequent demonstration of W. Thomson did not conclude either. Rigorous
demonstrations of this result were given later on by C. Somigliana [43] in 1892 and
by O. Tedone in 1897; we mention also the corresponding demonstration of
P. Duhem made in 1898 and resumed in 1960 in a more concise form by
E. Sternberg [45]. We shall call this formulations the Lamé-Clebsch representation.
198 5 General Equations. Formulation of Problems

M. E. Gurtin [17] calls it the Green-Lamé representation. We consider however


that the contribution of Clebsch, who laid the completeness problem, is more
important than that of Green, who considered it in a particular case, just like
Poisson did. Similar representations were given by B. A. Bondarenko [14] and P.
D. Chadwick and E. A. Trowbridge.
G. G. Stokes demonstrated in 1849 that the cubical dilatation and the rigid body
local rotation must satisfy the first and the second simple ware equation, respec-
tively, in the absence of volume forces.
In 1892, C. Somigliana [43] gave a new representation with the help of three
scalar potential functions, each one verifying the double wave equation. We
mention that, in various forms, this representation is equally found in the previous
studies of A. L. Cauchy and S. Kowalewski. This representation remained
unutilized and somewhat forgotten for a long time until, in 1949, M. Iacovache
[18], independently, found it again in vector form with the help of an analogy to B.
G. Galerkin’s vector, which verifies now the double waves equation; the demon-
stration of M. Iacovache does not warrant the completeness of the representation.
E. Sternberg and R. A. Eubanks [46] demonstrated its completeness, in 1957. It is
to be remarked that, during a whole decade, this representation was known, in the
literature, as the M. Iacovache representation. It is still M. Gurtin [17] that calls it
the Cauchy-Kowalewski-Somigliana representation. In our opinion, the contribu-
tion of M. Iacovache, who rediscovered this result and gave it a more elegant form,
that gave rise to many particularly important subsegment researches, gives a
satisfactory account of the name (the Somigliana-Iacovache representation) given
by us to it. We mention moreover that we gave [52] of this representation a new
demonstration that ensures its completeness.
E. Soós gave an analogous of the Galerkin vector for the dynamic problem of
the isotropic and non-homogeneous elastic bodies and thus generalized certain
results given by P. P. Teodorescu and M. Predeleanu for the static case; we also
gave another generalization of the problems [49] in 1964. An analogue of the
Galerkin vector for the dynamics of viscous liquids was found by Gr. C. Moisil
[26].
Starting from the Somigliana-Iacovache representation, E. Sternberg and
R. A. Eubanks [46] proposed a representation with the help of four scalar potential
functions, one of which verifies the longitudinal waves equation, while the other
three verify the transverse waves equation. This representation for the Lamé
equations in the dynamic case generalizes the Papkovich-Neuber one given for
Lamé’s equations in the static case. The two authors emphasized also, for a simply
connected domain, the completeness of their representation, that we shall call the
Sternberg-Eubanks representation.
E. Sternberg [45] gave moreover a new form of the last representation.
Gr. C. Moisil [27] and D. L. Ionescu [21] dealt with the decomposition into
primary and secondary waves.
An extension to the dynamic problem of the Beltrami-Finzi representation
(including the representations of Maxwell and Morera), as well as of Schaefer’s
representation was made by us [11, 19, 51–54]; as to the last representation, for
5.1 General Equations of the Theory of Elasticity 199

which we brought also into relief the completeness, in case of a simply connected
domain, we started, as Schaefer did, from the equations of motion, imposing the
condition that the other equations should be verified too.
Another method of tackling these problems, by solving them in stresses, consist
of performing in the opposite direction the method indicated above: i.e. starting
from the Beltrami type equations and imposing the condition that the resulted
representation should verify both the motion equations and the constitutive law.
We [11] gave thus a new representation in stresses either with the help of three
functions verifying the double waves equation or with the help of four functions,
each of them verifying a simple waves equation. The passage from a representa-
tion to another one can be made by means of formulae of Boggio type (see A.2.7).
These representations generalise those previously given by us [47], in the static
case. General results on this subject can be found in the synthesis work [49].

5.2 Statical Problem. Potential Functions

The use of the potential functions is particular important for the formulation of the
problems of the theory of elasticity; in case of a formulation in displacements, we
have to deal with displacement functions, while in case of a formulation in stresses,
we introduce stress functions. We mention that one cannot represent the general
solution of a problem of static or dynamic elasticity by means of only one potential
function.
We shall attach a special importance to the formulations in displacements or in
stresses, the most of the limit problems which will be considered being studied in
such a manner.

5.2.1 Formulations in Displacements

In what follows, we present the formulation in displacements of the problem of


elastostatics by means of Lamé’s equations, as well as Galerkin’s representation
and the Papkovich-Neuber representation.

5.2.1.1 Lamé’s Equations

If we replace in the equations of equilibrium (5.1) the stresses as functions or strains,


by means of Hooke’s law (5.3), then we get a system of three equations of the form
kell;i þ 2leij;j þ Fi ¼ 0; i ¼ 1; 2; 3: ð5:9Þ

which, associated with the equations of continuity (2.68), allows to formulate a


solution in strains of the static problems of the theory of elasticity.
200 5 General Equations. Formulation of Problems

Such a possibility to solve the problems is not sufficiently practical; therefore,


in the three equations thus obtained, we replace the strains by their expressions in
displacements (5.2). We obtain thus Lamé’s equations
lDui þ ðk þ lÞh;i þ Fi ¼ 0; i ¼ 1; 2; 3; ð5:10Þ

where we have introduced the volumic strain


h ¼ div u ¼ uj;j : ð5:11Þ
If we express Lamé’s elastic constants by means of the technical ones, then we
get Lamé’s equations in the form
1 1
Dui þ h;i þ Fi ¼ 0; i ¼ 1; 2; 3; ð5:100 Þ
1  2m G
we notice thus that, in the absence of the volume forces, Lamé’s equations depend
only on Poisson’s ratio m; and become
1
Dui þ h;i ¼ 0; i ¼ 1; 2; 3: ð5:12Þ
1  2m
In a developed form, Lamé’s equations read
1 1
Du1 þ ðu1;1 þ u2;2 þ u3;3 Þ;1 þ F1 ¼ 0;
1  2m G
1 1
Du2 þ ðu1;1 þ u2;2 þ u3;3 Þ;2 þ F2 ¼ 0; ð5:1000 Þ
1  2m G
1 1
Du3 þ ðu1;1 þ u2;2 þ u3;3 Þ;3 þ F3 ¼ 0:
1  2m G
Vectorially, these equations are of the form
lDu þ ðk þ lÞ grad div u þ F ¼ 0 ð5:13Þ

or of the form
1 1
Du þ grad div u þ F ¼ 0: ð5:130 Þ
1  2m G
In the absence of the volume forces, it results
1
Du þ grad div u ¼ 0: ð5:120 Þ
1  2m
If we apply the operator div to the Eq. (5.13) or to the Eq. (5.130 ) and if we take
into consideration (5.11), then we find that the volume strain must verify the
equation
1 1  2m
Dh ¼  div F ¼  div F; ð5:14Þ
k þ 2l 2ð1  mÞG
5.2 Statical Problem. Potential Functions 201

in the absence of a volume force or in the case of a solenoidal field of volume


forces ðdiv F ¼ 0Þ; the cubic strain becomes a harmonic function
Dh ¼ 0: ð5:15Þ
Using the formula (A.101) and taking into account (5.14) and (5.15), the Eqs.
(5.130 ) and (5.120 ) read
   
1 1 1
D uþ r div u ¼  Fþ r div F ð5:1300 Þ
2ð1  2mÞ G 4ð1  mÞ

and
 
1
D uþ r div u ¼ 0; ð5:1200 Þ
2ð1  2mÞ

respectively.
Applying Laplace’s operator to the Eq. (5.13) and taking into account (5.14),
we find that the displacement vector must verify the equation
kþl 1
DDu ¼  grad div F  DF; ð5:16Þ
lðk þ 2lÞ l
taking into account the relation (A.920 ) between the differential operators of sec-
ond order, we can write the Eq. (5.16) in the form
1 1
DDu ¼  grad divF þ curl curl F ð5:160 Þ
k þ 2l l
too, which can be useful in some cases. In case of a solenoidal field of volume
forces we may write
1
DDu ¼ curl curl F; ð5:17Þ
l
while, in case of an irrotational field of volume forces, we get
1
DDu ¼  grad div F: ð5:170 Þ
k þ 2l
In the absence of the volume forces, the displacement vector is a biharmonic
function
DDu ¼ 0: ð5:18Þ
We must mention that the set of integrals of the Eq. (5.16) includes the set of
integrals of the Eq. (5.13), the first equation being not sufficient to solve the
problem. We must integrate the Eq. (5.13) or the Eq. (5.130 ) to get a solution in
displacements.
202 5 General Equations. Formulation of Problems

We notice that the components of the displacement must be functions of class


C2 to verify the Eq. (5.10). Because replacing the expressions of the strains by
means of displacements in the equations of continuity (2.68) one must obtain
identities, it results that the displacements must be functions of class C3 ; but the
components of the displacement vector verify the Eq. (5.16) only if they are of
class C4 : The components of the volume force may be of class C0 if we do not take
into account the Eq. (5.14); otherwise, they must be of class C1 ; while the Eq.
(5.16) shows that they must be even of class C2 :

5.2.1.2 Galerkin’s Representation

Starting from the homogeneous Eq. (5.100 ), one can introduce a function by means
of the relation
2G div u ¼ ð1  2mÞDu; ð5:19Þ

the Eq. (5.120 ) leads to


Dð2G u þ grad uÞ ¼ 0;
wherefrom
2G u ¼ v  grad u; ð5:20Þ

with
Dv ¼ 0: ð5:21Þ
Taking into account the relation (5.19), we get
div v  div grad u ¼ ð1  2mÞDu
or
2ð1  mÞDu ¼ div v: ð5:22Þ
We introduce now a vector C ¼ Cðx1 ; x2 ; x3 Þ; with the aid or the relation
v ¼ 2ð1  mÞDC; ð5:23Þ

the Eq. (5.21) shows that this vector must verify the biharmonic equation
DDC ¼ 0: ð5:24Þ
We are led to
Dðu  div CÞ ¼ 0;
wherefrom
u ¼ div C þ C0 ;
5.2 Statical Problem. Potential Functions 203

with
DC0 ¼ 0:
By means of Almansi’s formula (A.100), which allows the representation of a
biharmonic function with the aid of two harmonic functions, we can introduce the
function C0 in the harmonic part of the biharmonic components of the biharmonic
vector C.
The representation (5.20) leads now to Galerkin’s representation
2Gu ¼ 2ð1  mÞD C  grad div C; ð5:25Þ

based on a potential vector of class C4 ; Galerkin’s vector C. We may write in


components
2Gui ¼ 2ð1  mÞD Ci  Cj;ji ; ð5:250 Þ

where the functions Ci ¼ Ci ðx1 ; x2 ; x3 Þ of class C4 must be biharmonic


DD Ci ¼ 0; i ¼ 1; 2; 3: ð5:240 Þ
In a developed form, we may write

2Gu1 ¼ 2ð1  mÞD C1  ðC1;1 þ C2;2 þ C3;3 Þ;1 ;


2Gu2 ¼ 2ð1  mÞD C2  ðC1;1 þ C2;2 þ C3;3 Þ;2 ; ð5:2500 Þ
2Gu3 ¼ 2ð1  mÞD C3  ðC1;1 þ C2;2 þ C3;3 Þ;3 :

The above given representation is complete for a simply connected domain (any
displacement vector which verifies Lamé’s Eq. (5.120 ) can be represented in the
form (5.25)). One must make a supplementary study of the relations (5.19 and
5.23) for a multiply connected domain, in other words one must specify the
conditions in which the functions u and C are univocally determined; certain
complementary terms, corresponding to some singularities, are thus introduced.
Using Cauchy’s relations (5.2) and Hooke’s law (5.3), we obtain the state of
stress in the form
rij ¼ 2ð1  mÞDCði;jÞ þ ðmDdij  oi oj ÞCk;k ; i; j ¼ 1; 2; 3: ð5:26Þ

If the volume forces are non-zero, then we can use the same representation
(5.25); Galerkin’s vector must verify the equation
1
DDC þ F ¼ 0; ð5:27Þ
1m
as it can be easily seen, by replacing the representation (5.25) in the Eq. (5.130 ).
We notice that a particular integral is sufficient to introduce the influence of the
volume forces.
204 5 General Equations. Formulation of Problems

Taking into account Almansi’s formulae (A.100), (A.1000 ), which allow the
representation of a biharmonic function by means of two harmonic ones, it results
that in Galerkin’s representation intervene six harmonic functions, which must be
determined with the aid of three conditions at one point of the contour. One has
thus a triple functional indetermination; the state of displacement is univocally
determined, as it will be seen in the next chapter, but the displacement function has
not this property.

5.2.1.3 The Papkovich-Neuber Representation

Taking into account the formula (A.1030 ) and observing that the vector v verifies the
harmonic Eq. (5.21), it results that the relation (5.22) allows to write u in the form
1
u¼ ðv þ r  vÞ; ð5:28Þ
4ð1  mÞ 0
with
Dv0 ¼ 0; ð5:29Þ

where r is the radius vector. Replacing in (5.20), we find the Papkovich-Neuber


representation in the form
1
2Gu ¼ v  gradðr  v þ v0 Þ; ð5:30Þ
4ð1  mÞ
where a harmonic vector potential v ¼ vðx1 ; x2 ; x3 Þ and a harmonic scalar potential
v0 ¼ v0 ðx1 ; x2 ; x3 Þ have been used; these two potentials must be functions of class
C3 : As the preceding representation, also this representation is complete for a
simply connected domain.
In components, we may write
1
2Gui ¼ vi  ðxj vj þ v0 Þ;i ; i ¼ 1; 2; 3; ð5:300 Þ
4ð1  mÞ
where the functions vj ¼ vj ðx1 ; x2 ; x3 Þ are harmonic

Dvj ¼ 0; j ¼ 1; 2; 3: ð5:210 Þ

In a developed form, we have


1
2Gu1 ¼ v1  ðx1 v1 þ x2 v2 þ x3 v3 þ v0 Þ;1 ;
4ð1  mÞ
1
2Gu2 ¼ v2  ðx1 v1 þ x2 v2 þ x3 v3 þ v0 Þ;2 ; ð5:3000 Þ
4ð1  mÞ
1
2Gu3 ¼ v3  ðx1 v1 þ x2 v2 þ x3 v3 þ v0 Þ;3 :
4ð1  mÞ
5.2 Statical Problem. Potential Functions 205

Introducing the notation


1 1
vi ¼ v ; i ¼ 1; 2; 3; v0 ¼ v; ð5:31Þ
4ð1  mÞ i 4ð1  mÞ 0

where vi ¼ vi ðx1 ; x2 ; x3 Þ; v0 ¼ v0 ðx1 ; x2 ; x3 Þ are harmonic functions


Dvi ¼ 0; i ¼ 1; 2; 3; Dv0 ¼ 0; ð5:32Þ
the representation (5.300 ) may be written also in the form
2Gui ¼ ð3  4mÞvi  ðxj 
vj;i þ v0;i Þ: ð5:3000 Þ

It results the state of stress


1
rij ¼ vði;jÞ þ ðmDdij  oi oj Þðxk vk þ v0 Þ; i; j ¼ 1; 2; 3: ð5:33Þ
4ð1  mÞ
If the volume forces are non-zero, then we may use the same representation
(5.30). Where the vector potential must verify the equation
Dv þ 2F ¼ 0; ð5:34Þ

the scalar potential being given by


Dv0 ¼ 2r  F: ð5:35Þ
As in the case of the preceding representation, a particular solution of the
Eqs. (5.34) and (5.35) is sufficient to introduce the influence of the volume forces.
Excepting the case m ¼ 1=4; one of the four scalar harmonic functions may be
equated to zero or may be expressed by means of the other functions, as it is more
convenient from the point of view of practical computation; hence, excepting the
case m ¼ 1=4; the state of displacement of the body is determined with the aid of
three scalar harmonic functions. Thus, we have to determine three harmonic
functions, for which three conditions at an arbitrary point of the contour are given;
the three functions are univocally determined for a finite simply connected domain
if the loads on the contour are in equilibrium. We do not know any demonstration
of this statement (nor any counter-example); it is probably true, by analogy with
the plane problem, case in which the statement is proved.

5.2.2 Formulations in Stresses

In what follows, we present the formulation in stresses of the problem of elasto-


statics, by introducing the Beltrami-Michell equations, as well as the Beltrami-
Finzi, Maxwell, Morera, Schaefer and Teodorescu representations.
206 5 General Equations. Formulation of Problems

5.2.2.1 The Beltrami-Michell Equations

If we eliminate the displacements between the relations (5.2), then we obtain the
equations of continuity in the form
mij nkl eik; jl ¼ 0; m; n ¼ 1; 2; 3; ð5:36Þ

where we have introduced Ricci’s permutation tensor. Replacing, in the latter


equations, the strains as functions of stresses, by means of Hooke’s law (5.30 ), we
get
ð1 þ mÞmij nkl rik;jl  mmij nkl rpp;jl dik ¼ 0;

wherefrom, taking into account (A.38), there results


ð1 þ mÞmij nkl rik;jl  mðdmn djl  dml djn Þrpp;jl ¼ 0

or
ð1 þ mÞmij nkl rik;jl  mDrll dmn þ mrll;mn ¼ 0; m; n ¼ 1; 2; 3; ð5:37Þ

this represents a form in stresses of the equations of continuity of deformations.


Multiplying the Eq. (5.37) by dmn and summing, we obtain, analogically,
ð1 þ mÞðdik djl  dil djk Þrik;jl  3mDrll þ mDrll ¼ 0

or
ð1  mÞDrll  ð1 þ mÞrij;ij ¼ 0; ð5:38Þ

taking into account the equation of equilibrium (5.1) too, we may write
rij;ij ¼ Fk;k ; ð5:39Þ

so that we obtain the equation verified by the sum of normal stresses in the form
1þm 1þm
Drii ¼ DH ¼  Fi;i ¼  div F: ð5:40Þ
1m 1m
We notice that one could obtain this result starting from (5.14) and using
Hooke’s law.
In the absence of volume forces or in the case of a solenoidal field of volume
forces, the sum of the normal stresses becomes a harmonic function
DH ¼ 0: ð5:41Þ
Taking into account (A.37), we can write the Eq. (5.37) in the form

ð1 þ mÞ½Drll dmn  rij;ij dmn  Drmn þ 2rkðm;nÞk  rll;mn 


 mDrll dmn þ mDrll;mn ¼ 0
5.2 Statical Problem. Potential Functions 207

or in the form

ð1 þ mÞDrmn þ rll;mn  Drll dmn þ ð1 þ mÞrij;ij dmn


 2ð1 þ mÞrkðm;nÞk ¼ 0; m; n ¼ 1; 2; 3: ð5:370 Þ

Using the relations


rkði;jÞk ¼ Fi;j ; ð5:390 Þ

which results from the equations of equilibrium (5.1), as well as the relations
(5.39), (5.40), the Eq. (5.370 ) read
1 m
Drij þ H;ij ¼ 2Fði;jÞ  Fk;k dij ; i; j ¼ 1; 2; 3; ð5:42Þ
1þm 1m
one obtains thus the Beltrami-Michell equations of continuity in stresses (obtained
by Beltrami in the absence of the volume forces and by Michell in case of arbitrary
volume forces).
In a developed form, we have
1 m
Dr11 þ ðr11 þ r22 þ r33 Þ;11 ¼ 2F1;1  ðF1;1 þ F2;2 þ F3;3 Þ;
1þm 1m
1 m
Dr22 þ ðr11 þ r22 þ r33 Þ;22 ¼ 2F2;2  ðF1;1 þ F2;2 þ F3;3 Þ; ð5:43Þ
1þm 1m
1 m
Dr33 þ ðr11 þ r22 þ r33 Þ;33 ¼ 2F3;3  ðF1;1 þ F2;2 þ F3;3 Þ;
1þm 1m
1
Dr23 þ ðr11 þ r22 þ r33 Þ;23 ¼ ðF2;3 þ F3;2 Þ;
1þm
1
Dr31 þ ðr11 þ r22 þ r33 Þ;31 ¼ ðF3;1 þ F1;3 Þ; ð5:430 Þ
1þm
1
Dr12 þ ðr11 þ r22 þ r33 Þ;12 ¼ ðF1;2 þ F2;1 Þ:
1þm
Applying Laplace’s operator to the Eq. (5.42), we obtain the equations
1
DDrij ¼ 2DFði;jÞ þ ðoi oj  mDdij ÞFk;k ; i; j ¼ 1; 2; 3; ð5:44Þ
1þm
which are verified by the components of the stress tensor; in the absence of the
volume forces, the stresses are biharmonic functions
DDrij ¼ 0; i; j ¼ 1; 2; 3; ð5:45Þ

while Beltrami’s equations read


1
Drij þ H;ij ¼ 0; i; j ¼ 1; 2; 3: ð5:420 Þ
1þm
208 5 General Equations. Formulation of Problems

We notice that the stresses must be functions of class C2 ; to can verify the
Eq. (5.42); if we wish the Eq. (5.44) be verified too, then these functions must be
of class C4 .
We also must notice that the set of integrals of the Eq. (5.44) contains the set of
integrals of the Eq. (5.42), the first equations being not sufficient to solve the
problem. For a solution in stresses one must integrate the system formed by the
equations of equilibrium (5.1) and by the Eq. (5.42). One cannot use only the latter
equations because they do not contain the equations of equilibrium (these differ-
ential equations are of the first order and represent conditions much more
restrictive that the ones imposed by the Beltrami-Michell equations, which are
differential equations of second order).

5.2.2.2 The Beltrami-Finzi Representation for the Equations


of Equilibrium

In the absence of the volume forces, the equations of equilibrium have the form
rji;j ¼ 0; i ¼ 1; 2; 3: ð5:10 Þ

We notice that these equations correspond to the conditions of equating to zero the
divergence of the tensor Tr ; hence, this tensor must be the curl of another tensor
and we may write
rij ¼ jkl fli;k ; ð5:46Þ

where fli is an asymmetric tensor.


The condition that the stress tensor be symmetric leads to
ijm rij ¼ 0; ð5:47Þ

so that
ijm jkl fli;k ¼ 0; ð5:470 Þ

analogically, we get
ijm jkl fli ¼ knp gnm;p ;

where gnm is an asymmetric tensor too. Taking into account (A.38), we may write
this relation in the form
fmk ¼ fii dkm þ kpn gnm;p

too or, changing the indices, in the form


fij ¼ fll dij þ jkl gli;k ; ð5:48Þ
5.2 Statical Problem. Potential Functions 209

contracting the tensor, there results


1
fll ¼  ijk gij;k ; ð5:480 Þ
2
so that
1
fij ¼  klm gkl;m dij þ jkl gli;k : ð5:4800 Þ
2
The stress tensor becomes
1
rij ¼ ijk mnp gmn;pk þ ikl jmn gnl;mk : ð5:460 Þ
2
Introducing the skewsymmetric part of the tensor gij ; we notice that the relation

ijk mnp g½mn;pk ¼ 2ikl jmn g½nl;mk ð5:49Þ

is verified for any i and j: If we denote


gðijÞ ¼ Fij ; ð5:50Þ

then we obtain the Beltrami-Finzi representation


rij ¼ ikl jmn Fkm;ln ; i; j ¼ 1; 2; 3; ð5:51Þ

where Fij is Finzi’s symmetric tensor, the components of which must be functions
of class C3 : This representation has been given by Beltrami, but its tensor form is
due to Finzi. The given demonstration ensures its completeness for the equations
of equilibrium (5.10 ), in case of a simply connected domain.
In case of non–zero volume forces, the Eq. (5.1) allow particular integrals of the
form
Z Z Z
r11 ¼  F1 dx1 ; r22 ¼  F2 dx2 ; r33 ¼  F3 dx3 ; ð5:52Þ

r23 ¼ r31 ¼ r12 ¼ 0; ð5:520 Þ


where Fi ; i ¼ 1; 2; 3; must be integrable functions.

5.2.2.3 Maxwell’s and Morera’s Representations

We obtain Maxwell’s representation


r11 ¼ F22;33 þ F33;22 ; r22 ¼ F33;11 þ F11;33 ; r33 ¼ F11;22 þ F22;11 ; ð5:53Þ

r23 ¼ F11;23 ; r31 ¼ F22;31 ; r12 ¼ F33;12 ð5:530 Þ


210 5 General Equations. Formulation of Problems

if Fij ¼ 0; i 6¼ j; while if Fij ¼ 0; i ¼ j; i; j ¼ 1; 2; 3: Then we get Morera’s


representation
r11 ¼ 2F23;23 ; r22 ¼ 2F31;31 ; r33 ¼ 2F12;12 ; ð5:54Þ

r23 ¼ ðF23;1 þ F31;2 þ F12;3 Þ;1 ;


r31 ¼ ðF31;2 þ F12;3 þ F23;1 Þ;2 ; ð5:540 Þ
r12 ¼ ðF12;3 þ F23;1 þ F31;2 Þ;3 ;

these two representations depend anyone only on three potential functions.


Let Fij0 be a symmetric tensor, the components of which verify equations of
Saint-Venant’s type
0
ikl jmn Fkm;ln ¼ 0; ð5:55Þ

it results, necessarily,
1
Fij0 ¼ kði;jÞ ¼ ðki;j þ kj;i Þ; ð5:550 Þ
2
where ki ; i ¼ 1; 2; 3; are the components of an arbitrary vector of class C3
We notice thus that if the potential functions Fij lead to a certain state of stress,
then the potential functions Fij þ Fij0 ; i; j ¼ 1; 2; 3; give the same state of stress.
We can choose always the functions ki so that Fij þ Fij0 ¼ 0 for i 6¼ j or for i ¼ j;
obtaining Maxwell’s representation or Morera’s one, respectively; indeed, this
0 0 0
corresponds to the fact that if the components F23 ; F31 ; F12 of the symmetric
0 0 0 0
tensor Fij are given, then the other components F11 ; F22 ; F33 are determined by the
Eq. (5.55) and conversely. Hence, Maxwell’s and Morera’s representations are, as
well, complete for a simply connected domain.
We also mention that between Maxwell’s and Morera’s potential functions one
has the relations (the other relations are obtained by circular permutations)
F22;33 þ F33;22 ¼ 2F23;23 ; ð5:56Þ

ðF23;1 þ F31;2 þ F12;3 Þ;1 ¼ F11;23 ; ð5:560 Þ

which are relations of the Saint-Venant type; thus, if Maxwell’s representation is


given, then one can obtain Morera’s one and conversely.
Starting from the relations (5.56), we can express Morera’s functions in the form
Z Z 
1
F23 ¼  F22;3 dx2 þ F33;2 dx3 ; ð5:57Þ
2

as well, the relations (5.560 ) allow to express Maxwell’s functions by


ZZ Z Z
F11 ¼ F23;11 dx2 dx32  F12;1 dx2  F31;1 dx3 : ð5:570 Þ
5.2 Statical Problem. Potential Functions 211

One observes that the representations (5.57) and (5.570 ) verify the Eqs. (5.560 ) and
(5.56), respectively. Thus, one can also determine the functions of two variables
which appear by the indefinite integration (which may be effected, the potential
functions being of class C3 ) and which have not been put in evidence yet.
Hence, if we take into account only the equations of equilibrium (5.10 ), then we
can represent the state of stress by means of only three potential functions of class
C3 :

5.2.2.4 Schaefer’s Representation

We choose the components of Finzi’s tensor in the form


Fij ¼ Hij þ ðHll  XÞdij ; i; j ¼ 1; 2; 3; ð5:58Þ

where the potential functions Hij ¼ Hij ðx1 ; x2 ; x3 Þ; i; j ¼ 1; 2; 3; are the compo-
nents of Schaefer’s tensor; both these functions and the function X ¼ Xðx1 ; x2 ; x3 Þ
will be of class C3 : From
1
Hij ¼ Fij þ ðFll þ XÞdij ð5:580 Þ
2
one observes that the components of Schaefer’s tensor TH are univocally deter-
mined, excepting the arbitrary function X:
Introducing the expressions (5.58) in the representation (5.51), we get the
representation
rij ¼ ikl jmn Hkm;ln þ DðHll  XÞdij  ðHll  XÞ;ij ; i; j ¼ 1; 2; 3; ð5:59Þ

as well as the relation


rll ¼ Hij;ij þ DðHll  2XÞ; ð5:590 Þ

where we took into account the formula (A.38). If the stresses given by (5.59)
verify the Beltrami Eq. (5.420 ) too, then we may write

 ikl jmn DHkm;ln þ DDðHll  XÞdij  DðHll  XÞ;ij


1
þ ðDHll;ij  2DX;ij þ Hkl;klij Þ ¼ 0;
1þm
therefore, it results

 ikl jmn DHkm;ln  DðHll  XÞ;ij


1
þ ðDHll;ij  2DX;ij þ Hkl;klij Þ ¼ 0; ð5:60Þ
1þm
DDðHll  XÞ ¼ 0: ð5:600 Þ
212 5 General Equations. Formulation of Problems

If Schaefer’s tensor components are harmonic functions


DHij ¼ 0; i; j ¼ 1; 2; 3; ð5:61Þ

then the conditions (5.60), (5.600 ) take the form


 
1
DX þ ðHkl;kl  2DXÞ ¼ 0; ð5:62Þ
1þm ;ij

DDX ¼ 0: ð5:620 Þ
We notice that the condition (5.62) is satisfied if the function X verifies the
equation
1
DX ¼ Hkl;kl ; ð5:63Þ
1m
taking into account (5.61), the condition (5.620 ) is verified too.
Returning to the relations (5.58) between Finzi’s and Schaefer’s tensor, we
notice that for i 6¼ j we have
Fij ¼ Hij ; i 6¼ j; ð5:64Þ

while for i ¼ j we may write


Fii ¼ Hkk þ Hll  X ð!Þ; i 6¼ k 6¼ l 6¼ i: ð5:640 Þ
The first case corresponds to Morera’s stress functions, which must be harmonic
DFij ¼ 0; i 6¼ j; ð5:65Þ

while the second case corresponds to Maxwell’s stress functions, which must be
biharmonic
DDFij ¼ 0; i ¼ j: ð5:650 Þ
Taking into account (5.61), (5.63) and (A.37), the representation (5.59) reads
rij ¼ mDXdij þ X;ij  2Xkði;jÞk ; ð5:66Þ

this being Schaefer’s representation; the seven stress functions must be of class
C4 .
We notice that the sum of normal stresses is given by
rll ¼ ð1 þ mÞDX; ð5:67Þ

taking into account (5.30 ), we can write the strains in the form
1
eij ¼ ðX;ij  2Hkði;jÞk Þ; i; j ¼ 1; 2; 3: ð5:68Þ
2G
5.2 Statical Problem. Potential Functions 213

In this case, Cauchy’s relations (5.2) show that the state of displacement is given
by
2Gui ¼ X;i  2Hij;j : ð5:69Þ
Although supplementary conditions have been put in the above demonstration,
the representation (5.66), (5.69) is complete for a simply connected domain; to put
this in evidence, we will show the connection which can be made between this
representation and the representations in displacements given above.
Introducing the notation
Ui ¼ Hij;j ; i ¼ 1; 2; 3; ð5:70Þ

the state of stress reads


rij ¼ mDXdij þ X;ij  2Uði;jÞ ; i; j ¼ 1; 2; 3; ð5:71Þ

the state of displacement being given by


2Gui ¼ X;i  2Ui : ð5:72Þ

The functions Ui ¼ Ui ðx1 ; x2 ; x3 Þ; i ¼ 1; 2; 3; of class C3 ; are the components of a


vector U which satisfies the harmonic equation
DU ¼ 0; ð5:73Þ
while the function X must satisfy the equation
1
DX ¼ divU: ð5:74Þ
1m
Using the notation
U ¼ ð1  mÞDC; ð5:75Þ

we notice that
X ¼  divC; ð5:76Þ
where we have neglected a function which satisfies the harmonic equation,
because it can be included in the biharmonic vector C.
If in the representation
2Gu ¼ gradX  2U ð5:720 Þ
we take into account (5.75), (5.76), then we find again Galerkin’s representation
(5.25); because this latter representation is complete for a simply connected
domain, it results that Schaefer’s representation is complete too, in the same
conditions.
214 5 General Equations. Formulation of Problems

In a developed form, we can express the state of displacement in the form

2Gu1 ¼ X;1  2U1 ;


2Gu2 ¼ X;2  2U2 ; ð5:7200 Þ
2Gu3 ¼ X;3  2U3 ;

the state of stress reads

r11 ¼ X;11  mDX  2U1;1 ;


r22 ¼ X;22  mDX  2U2;2 ; ð5:7100 Þ
r33 ¼ X;33  mDX  2U3;3 ;

r23 ¼ X;23  ðU2;3 þ U3;2 Þ;


r31 ¼ X;31  ðU3;1 þ U1;3 Þ; ð5:7100 Þ
r12 ¼ X;12  ðU1;2 þ U2;1 Þ:

If the volume forces are non-zero, then we may use the same representations
(5.66), (5.69) or (5.71), (5.72). In the latter case, we notice that the equations of
equilibrium and Beltrami-Michell equations are verified if the vector U is given by
the equation
DU ¼ F; ð5:730 Þ

while the function X is given by the Eq. (5.74); applying the operator D to this
equation, we notice that the function X must satisfy the equation
1
DDX ¼ divF; ð5:740 Þ
1m
From the point of view of the particular solution which corresponds to given
volume forces, we observe that one can replace the Eq. (5.74) by the Eq. (5.740 ).
Indeed, the Eq. (5.740 ) is equivalent to the equation
D½ð1  mÞDX  divU ¼ 0;
which leads to
ð1  mÞDX  divU ¼ u;

where u is a harmonic function. Taking into account (5.730 ), we can write

U ¼ U þ U0 ; ð5:7300 Þ

where U is a vector harmonic function and U0 is a particular integral. We notice


that the function u which may be written in the form
u ¼ divu;
5.2 Statical Problem. Potential Functions 215

where u is a harmonic vector function, may be included in the components of the


vector U; hence, we can use the Eq. (5.740 ) to obtain the particular integral
corresponding to the given volume forces.
An important particular case is that of the conservative volume forces of the
form
F ¼ gradv: ð5:77Þ

where v ¼ vðx1 ; x2 ; x3 Þ is a function of class C2 ; in this case, we can choose a


particular integral
U ¼ grad w: ð5:78Þ

where w ¼ wðx1 ; x2 ; x3 Þ is a function of class C4 ; given by


Dw ¼ v: ð5:79Þ
The Eq. (5.74) becomes
1
DX ¼ Dw ð5:80Þ
1m
and leads to
1
DX ¼  Dx; ð5:800 Þ
1  2m
where we have introduced the scalar potential x ¼ xðx1 ; x2 ; x3 Þ by the relation
x ¼ X  2w; ð5:81Þ
taking into account (5.78), (5.81), the state of displacement will be given by
2Gu ¼ grad x; ð5:82Þ

while the state of stress reads


m
rij ¼ Dxdij þ x;ij ; i; j ¼ 1; 2; 3; ð5:83Þ
1  2m
corresponding to the conservative volume forces (5.77).
The relations (5.79)–(5.800 ) show that the function x must be a particular
integral of Poisson’s equation
1  2m
Dx þ v ¼ 0; ð5:84Þ
1m
that is a function of class C2 ; if we wish to satisfy the Beltrami-Michell equations
too, then the function x must be of class C4 :
216 5 General Equations. Formulation of Problems

5.2.2.5 Teodorescu’s Representation for the Beltrami-Michell


Equations

Another possibility to obtain a representation in stresses of the static problem of


the theory of elasticity consists in the finding of a representation of Beltrami’s
Eq. (5.420 ), putting then the conditions that the equations of equilibrium be verified
too.
Let us introduce the function F by the relation
H ¼ ð1 þ mÞDF; ð5:85Þ
in this case, the Eq. (5.420 ) show that the stresses must be of the form
rij ¼ F;ij þ uij ; i; j ¼ 1; 2; 3; ð5:86Þ

where uij ¼ uji are harmonic functions

Duij ¼ 0; i; j ¼ 1; 2; 3: ð5:87Þ

Taking into account (5.85), we get


ð2 þ mÞDF ¼ ukk ; ð5:88Þ
while the condition (5.87) shows that F must be a biharmonic function
DDF ¼ 0: ð5:89Þ
We also introduce the functions Fij ¼ Fij ðx1 ; x2 ; x3 Þ; i; j ¼ 1; 2; 3, as compo-
nents of a stress symmetric tensor, by the relations
uij ¼ ð2 þ mÞDFij ; ð5:90Þ

the Eq. (5.87) show that these functions are biharmonic


DDFij ¼ 0; i; j ¼ 1; 2; 3: ð5:91Þ
The condition (5.89) leads, in this case, to
F ¼ Fkk þ u0 ;
where u0 is a harmonic function; we notice that one can take u0 ¼ 0, because this
harmonic function can be introduced in the harmonic part of each of the bihar-
monic functions F11 ; F22 ; F33 ; having no influence on the state of stress given by
the formula (5.86).
We obtain thus Teodorescu’s representation
rij ¼ ð2 þ mÞDFij  Fkk;ij ; i; j ¼ 1; 2; 3; ð5:92Þ

which is complete for Beltrami’s system of equations in case of a simply con-


nected domain; the six stress functions Fij ; i; j ¼ 1; 2; 3; must be of class C4 :
5.2 Statical Problem. Potential Functions 217

In a developed form, we can write

r11 ¼ð2 þ mÞDF11  ðF11 þ F22 þ F33 Þ;11 ;


r22 ¼ð2 þ mÞDF22  ðF11 þ F22 þ F33 Þ;22 ; ð5:920 Þ
r33 ¼ð2 þ mÞDF33  ðF11 þ F22 þ F33 Þ;33 ;

r23 ¼ð2 þ mÞDF23  ðF11 þ F22 þ F33 Þ;23 ;


r31 ¼ð2 þ mÞDF31  ðF11 þ F22 þ F33 Þ;31 ; ð5:9200 Þ
r12 ¼ð2 þ mÞDF12  ðF11 þ F22 þ F33 Þ;12 :

In the case of non-zero volume forces, one can use the same representation,
adding to the biharmonic functions Fij particular integrals of the equations
1 h m i
DDFij ¼  2Fði;jÞ þ Fk;k dij ; i; j ¼ 1; 2; 3; ð5:910 Þ
2þm 1m
thus, the Beltrami-Michell Eq. (5.42) are identically verified.
On the basis of Almansi’s formula (A.100), we may write the biharmonic
functions Fij in the form
 
1 1
Fij ¼ Wij þ ðxi Uj þ xj Ui Þ ; i; j ¼ 1; 2; 3; ð5:93Þ
2ð2 þ mÞ 2

where Ui and Wij are harmonic functions

DUi ¼ DWij ¼ 0; i; j ¼ 1; 2; 3; ð5:930 Þ

in this case, we get a new representation for the state of stress, in the form
1
rij ¼ Uði;jÞ  ðxk Uk þ U0 Þ;ij ; i; j ¼ 1; 2; 3; ð5:94Þ
2ð2 þ mÞ
where we have introduced the function U0 ¼ U0 ðx1 ; x2 ; x3 Þ given by
U0 ¼ Wll : ð5:95Þ
We notice that the functions Ui ¼ Ui ðx1 ; x2 ; x3 Þ; i ¼ 1; 2; 3, are the components
of the vector U ¼ Uðx1 ; x2 ; x3 Þ; we must have
DU ¼ 0; DU0 ¼ 0 ð5:96Þ

both the vector potential and the scalar one being functions of class C4 :
The formula (5.30 ) allows to write the state of strain in the form
 
1 1 m
eij ¼ Uði;jÞ  ðxk Uk þ U0 Þ;ij  Uk;k dij ; i; j ¼ 1; 2; 3; ð5:97Þ
2G 2ð2 þ mÞ 2þm
218 5 General Equations. Formulation of Problems

Cauchy’s relations (5.2) lead to the state of displacement


Z
1 m
2Gui ¼ Ui  ðxk Uk þ U0 Þ;i  Uk;k dxi ; i ¼ 1; 2; 3; ð5:98Þ
2ð2 þ mÞ 2þm

where we have not put in evidence functions of two variables which appear by
integration.
To find the conditions in which the state of stress given by (5.93) corresponds to
the basic system of equations of the theory of elasticity, this representation must
verify the equations of equilibrium (5.10 ) too; we get thus the equations
D½ð2 þ mÞFij;j  Fkk;i  ¼ 0; i ¼ 1; 2; 3; ð5:99Þ

which lead to
ð2 þ mÞFij;j  Fkk;i ¼ ui ; i ¼ 1; 2; 3; ð5:990 Þ

where ui ; i ¼ 1; 2; 3; are harmonic functions. We may write this system also in the
form
1
Fij;j þ Fik;k ¼ ½F11 þ F22 þ F33  ð2 þ mÞFii ;i ð!Þ;
2þm ð5:9900 Þ
i 6¼ j 6¼ k 6¼ i; i ¼ 1; 2; 3;

the function ui being contained in the harmonic part of the biharmonic functions
Fij ; i 6¼ j; i; j ¼ 1; 2; 3:
One can show that the general integral of the system of homogeneous equations
Fij;j þ Fik;k ¼ 0 ð!Þ; i 6¼ j 6¼ k 6¼ i; i ¼ 1; 2; 3; ð5:100Þ

is of the form
Fij ¼ ðfik  fjk Þ;k ð!Þ; i 6¼ j 6¼ k 6¼ i, i; j ¼ 1; 2; 3; ð5:1000 Þ

where the functions fji ¼ fij ¼ fij ðxi ; xj Þ; i 6¼ j; i; j ¼ 1; 2; 3, of class C5 depend each
one on only two variables and are biharmonic too
DDfij ¼ 0; i; j ¼ 1; 2; 3: ð5:10000 Þ
One adds a particular integral of the system of Eq. (5.10000 ), which can be easily
obtained, observing that this system is equivalent to the system
1
Fij;ij ¼ Fkk;kk  ðFii;ii þ Fjj;jj þ Fkk;kk Þ
2 
1 1
þ DðF11 þ F22 þ F33 Þ  ðF11 þ F22 þ F33 Þ;kk ð!Þ;
2þm 2
i 6¼ j 6¼ k 6¼ i; i ¼ 1; 2; 3; ð5:10000 Þ

neglecting functions of only two variables.


5.3 Dynamical Problem. Potential Functions 219

5.3 Dynamical Problem. Potential Functions

As in the static case, the use of potential functions is very important in the dynamic
case too. In the following, we present formulations in displacements, by means of
displacement functions, and formulations in stresses, with the aid of stress func-
tions. In this study, we restrain us to the case of small nondamped motions.

5.3.1 Formulations in Displacements

Eliminating the stresses and the strains, one obtains a formulation in displacements
of the problems of elastodynamics, by means of Lamé’s equations. Starting from
these equations, we deduce the Somigliana-Iacovache, Sternberg-Eubanks and
Lamé-Clebsch representations.

5.3.1.1 Lamé’s Equations

Eliminating the stresses between the equations of motion (5.6) and Hooke’s law
(5.3), we get the equations
kell;i þ2leij;j þFi ¼q€
ui ; i ¼ 1; 2; 3; ð5:101Þ

which, associated to Cauchy’s relations (5.2), lead to a formulation in strain and


displacements of the problem. Eliminating now the strains between these relations,
we obtain Lamé’s equations for the dynamic case in the form
lh2 ui þ ðk þ lÞh;i þ Fi ¼ 0; i ¼ 1; 2; 3; ð5:102Þ

where the volume strain (5.5) is involved too.


In these equations we have introduced d’Alembert’s operators

1 o2
hi ¼ D  ; i ¼ 1; 2; ð5:103Þ
c2i ot2

c1 and c2 being the wave propagation velocities given by


k þ 2l 2ð1  mÞ l 1m E
c21 ¼ ¼ ¼ ; ð5:104Þ
q 1  2m q ð1 þ mÞð1  2mÞ q
l 1 E
c22 ¼ ¼ : ð5:1040 Þ
q 2ð1 þ mÞ q

We notice that a relation of the form


 
2ð1  mÞ 2 1
c21 ¼ c ¼ 1þ c2 ð5:105Þ
1  2m 2 1  2m 2
220 5 General Equations. Formulation of Problems

takes place; thus, the propagation velocity of the longitudinal waves c1 is strictly
greater than the propagation velocity of the transverse waves c2 (c1 [ c2 ). In the
pffiffiffiffiffiffiffi
limit cases, one obtains c1 ¼ 2c2 for m ¼ 0 or c1 ¼ 1 for m ¼ 1=2 but we remark
that, in these limit cases, we must start from Hooke’s law written in the corre-
sponding form.
We notice that between the operators (5.104) introduced above there exists a
relation of the form
2ð1  mÞh1  ð1  2mÞh2 ¼ D: ð5:106Þ
The vector form of the Eq. (5.102) reads
lh2 u þ ðk þ lÞ grad div u þ F ¼ 0: ð5:1020 Þ
Using the technical elastic constants, we have
1 1
h 2 ui þ h;i þ Fi ¼ 0; i ¼ 1; 2; 3; ð5:107Þ
1  2m G
or
1 1
h2 u þ grad div u þ F ¼ 0: ð5:1070 Þ
1  2m G
In a developed form, Lamé’s equations become
1 1
h 2 u1 þ ðu1;1 þ u2;2 þ u3;3 Þ;1 þ F1 ¼ 0;
1  2m G
1 1
h 2 u2 þ ðu1;1 þ u2;2 þ u3;3 Þ;2 þ F2 ¼ 0; ð5:10700 Þ
1  2m G
1 1
h 2 u3 þ ðu1;1 þ u2;2 þ u3;3 Þ;3 þ F3 ¼ 0:
1  2m G
In the absence of volume forces, Lamé’s equations depend only on an elastic
constant of the material, as in the static case; they read
1
h2 ui þ h;i ¼ 0; i ¼ 1; 2; 3; ð5:108Þ
1  2m
or
1
h2 u þ grad div u ¼ 0: ð5:1080 Þ
1  2m
If we apply the operator div to the Eq. (5.1020 ) or to the Eq. (5.1070 ) and take
into account the relation (5.106), then it results the equation
1 1  2m
h1 h ¼  div F ¼  div F; ð5:109Þ
k þ 2l 2ð1  mÞ
5.3 Dynamical Problem. Potential Functions 221

which must be verified by the volume strain; in the absence of the volume forces or
in case of a solenoidal field of volume forces, the cubic strain will verify the
equation of longitudinal waves
h1 h ¼ 0: ð5:110Þ
Applying the operator h1 to the Eq. (5.1020 ) and taking into account (5.109),
one sees that the displacement vector must verify the equation
kþl 1
h1 h2 u ¼ grad div F  h1 F; ð5:111Þ
lðk þ 2lÞ l
taking into account the relation (A.920 ) between the differential operators of sec-
ond order; one can make observations analogue to those in Sect. 5.2.1.1.
In the absence of volume forces, the displacement vector verifies the double
waves equation
h1 h2 u ¼ 0: ð5:112Þ
We notice that the index 1 corresponds to the primary waves (P waves); as it
will be later seen, these waves are irrotational, longitudinal waves of dilatation.
The index 2 corresponds to the secondary waves (S waves); these waves are shear,
transverse waves which correspond to a deformation under constant volume, hence
to an incompressible (isochore) deformation.
We mention that the set of integrals of the Eq. (5.111) includes the set of
integrals of the Eq. (5.1020 ); but the Eq. (5.111) is not sufficient to solve the
problem. In a solution in displacements one must integrate the equation (5.1020 ).
To verify the Eq. (5.102), it is sufficient that the displacements be functions of
class C2 : The equations of continuity (2.68) are identically verified, after replacing
the strains by Cauchy’s relations (5.2), if the displacements are functions of class
C3 ; finally, to can verify the double wave Eq. (5.112) the displacements must be
functions of class C4 : The components of the volume force can be of class C0 ;
provided we do not raise the problem of Eq. (5.109) or of Eq. (5.111); otherwise,
they must be of class C1 or of class C2 ; respectively.
We remark that the Eq. (5.1020 ) can be equally written in the form
1
c22 h2 u þ ðc21  c22 Þ grad div u þ F ¼ 0; ð5:113Þ
q
where we introduced the velocities of wave propagation and the mass force
ð1=qÞF.
Taking into account the relation (A.920 ), we can write Lamé’s equation in the form
ðk þ 2lÞh1 u þ ðk þ lÞ curl curl u þ F ¼ 0 ð5:114Þ
or in the form
1 1
h1 u þ curl curl u þ F¼0 ð5:1140 Þ
2ð1  mÞ k þ 2l
222 5 General Equations. Formulation of Problems

or with the help of the wave propagation velocities


1
c21 h1 u þ ðc21  c22 Þ curl curl u þ F ¼ 0: ð5:11400 Þ
q
Applying the divergence operator to the Eq. (5.11400 ), one obtains the equation
1
c21 h1 h þ F ¼ 0 ð5:1090 Þ
q
for the cubic strain. Likewise, applying the operator curl to the Eq. (5.113) and
taking into account the relation (2.5700 ), we obtain the equation
1
c22 h2 x þ curl F ¼ 0 ð5:115Þ
2q
for the rigid body local rotation vector; in the absence of the volume forces, the
rigid body local rotation vector verifies the transverse wave equation
h2 x ¼ 0: ð5:1150 Þ

5.3.1.2 The Somigliana-Iacovache Representation

As in the static case, starting from the homogeneous Eq. (5.1080 ), one can intro-
duce a function u by means of the relation
2G div u ¼ ð1  2mÞh2 u; ð5:116Þ

the Eq. (5.1080 ) leads to


2G u ¼ v  grad u; ð5:117Þ
with
h2 v ¼ 0: ð5:118Þ
Taking into account the relations (5.106) and (5.116), we get
2ð1  mÞh1 u ¼ div v: ð5:119Þ
We introduce now a vector C ¼ Cðx1 ; x2 ; x3 ; tÞ by means of the relation
v ¼ 2ð1  mÞh1 C; ð5:120Þ
the Eq. (5.118) shows that this vector must verify the double wave equation
h1 h2 C ¼ 0: ð5:121Þ
We are thus led to
u ¼ div C þ C0 :
5.3 Dynamical Problem. Potential Functions 223

with
h1 C0 ¼ 0:
Taking into account Boggio’s formula (A.107), which allows to write the vector
C as a sum of two vectors, verifying each one a simple wave equation, one can
introduce the function C0 in one of the components of the vector which verifies the
waves equation of index 1.
The representation (5.118) leads now to the Somigliana-Iacovache representation
2G u ¼ 2ð1  mÞh1 C  grad div C; ð5:122Þ

based on a potential vector of class C4 ; i.e. the Somigliana-Iacovache vector


(analogue of Galerkin’s vector), which verifies the double wave Eq. (5.121). The
above demonstration shows that this representation is complete for a simply
connected domain.
In components, we can write
2Gui ¼ 2ð1  mÞh1 Ci  Cj;ji ; i ¼ 1; 2; 3; ð5:1220 Þ

where the functions Ci ¼ Ci ðx1 ; x2 ; x3 ; tÞ of class C4 must verify the double waves
equations
h1 h2 Ci ¼ 0; i ¼ 1; 2; 3: ð5:1210 Þ

In a developed form, we get

2Gu1 ¼ 2ð1  mÞh1 C1  ðC1;1 þ C2;2 þ C3;3 Þ;1 ;


2Gu2 ¼ 2ð1  mÞh1 C2  ðC1;1 þ C2;2 þ C3;3 Þ;2 ; ð5:12200 Þ
2Gu3 ¼ 2ð1  mÞh1 C3  ðC1;1 þ C2;2 þ C3;3 Þ;3 :

Using Cauchy’s relations (5.2) and Hooke’s law (5.3), one obtains the state of
stress in the form
rij ¼ 2ð1  mÞh1 Cði;jÞ þ ðmh2 dij  oi oj ÞCk;k ; i; j ¼ 1; 2; 3: ð5:123Þ

In the case in which the volume forces are non-zero, one can use the same
representation (5.122), where the Somigliana-Iacovache vector must verify the
equation
1
h1 h2 C þ F ¼ 0; ð5:124Þ
1m
as it can be easily seen by replacing the representation (5.122) in the Eq. (5.1070 ); a
particular integral is sufficient to introduce the influence of the volume forces.
We notice that, taking into account (5.106) and (A.920 ), we may write the
representation (5.122) also in the form
2G u ¼ ð1  2mÞh2 C  curl curl C: ð5:12200 Þ
224 5 General Equations. Formulation of Problems

As in the static case, if one uses Boggio’s formula (A.107), then we notice that
in the Somigliana-Iacovache representation intervene six functions which verify
simple wave equations and must be determined by means of three conditions at a
point of the contour. There appears thus a triple functional indetermination; the
state of displacement is univocally determined, but the displacement functions do
not have the same property.

5.3.1.3 The Sternberg-Eubanks Representations

As in the Sect. 5.2.1.3, we notice that the function in the representation (5.117) is
of the form (5.120) and get the first Sternberg-Eubanks representation in the form
(5.122). In this representation one uses a vector potential v ¼ vðx1 ; x2 ; x3 ; tÞ, which
verifies the transverse wave Eq. (5.118), as well as a scalar potential
v0 ¼ v0 ðx1 ; x2 ; x3 ; tÞ, which, taking into account the relation (A.108), must verify
the equation
h1 v0 þ r  h1 v ¼ 0; ð5:125Þ
thus, the two potentials cannot be uncoupled, as in the static case. The first
Sternberg-Eubanks representation is, as well, complete for a simply connected
domain.
The state of stress is given by the formulae (5.125).
If the volume forces are non-zero, then one can use the same representation
(5.125), the vector potential verifying the Eq. (5.126), while the scalar potential is
given by the Eq. (5.125) too.
Taking into account the expressions (5.103) of d’Alembert’s operators and the
Eq. (5.126), one may write the Eq. (5.125) also in the form
1þmq
h 1 v0 þ r€
v ¼ 2r  F: ð5:1250 Þ
1  mE
Using the formula (A.107), one can write the Somigliana-Iacovache vector in
the form

C ¼ , þ ,0 ð5:126Þ
too, where

h2 , ¼ 0; h1 ,0 ¼ 0: ð5:1260 Þ
Denoting by

,0 ¼ div,0 ð5:127Þ
a scalar potential which verifies the longitudinal wave equation
h1 ,0 ¼ 0; ð5:1270 Þ
5.3 Dynamical Problem. Potential Functions 225

introducing the functions (5.126) in the representation (5.122) and taking into
account (5.1250 ), we obtain a new representation in displacements of the form
2Gu ¼ 2ð1  mÞh1 ,  gradðdiv, þ ,0 Þ; ð5:128Þ
where , is a vector potential, which verifies the longitudinal wave equation, while
,0 is a scalar potential, which verifies the transverse wave equation.
Taking into account the relation (5.106), the expression (5.103) of d’Alembert’s
operator and the Eq. (5.1270 ), one gets the second Sternberg–Eubanks represen-
tation in the form
q
2Gu ¼ ,€  gradðdiv , þ ,0 Þ: ð5:129Þ
l
One obtains Sternberg’s representation too, equivalent with the above one, in
the form
2Gu ¼ D,  gradðdiv , þ ,0 Þ: ð5:130Þ
All these representations are complete for a simply connected domain. The
potential functions , ¼ ,ðx1 ; x2 ; x3 ; tÞ and ,0 ¼ ,0 ðx1 ; x2 ; x3 ; tÞ must be of class C4 :
In components, we have
2Gui ¼ D,i  ð,j;j þ ,0 Þ;i ; i ¼ 1; 2; 3; ð5:1300 Þ

and, in a developed form, it results

2Gu1 ¼ D,1  ð,1;1 þ,2;2 þ,3;3 þ,0 Þ;1 ;


2Gu2 ¼ D,2  ð,1;1 þ,2;2 þ,3;3 þ,0 Þ;2 ; ð5:13000 Þ
2Gu3 ¼ D,3  ð,1;1 þ,2;2 þ,3;3 þ,0 Þ;3 :

One must notice that, in what concerns these last results, one cannot obtain the
static case by a simple particularization of the dynamic problem. It is true that
some results can be obtained by using such a particularization, but nor all results
can be obtained in this way. This happens because a representation of Almansi
type is not a representation of Boggio type, the two representations being essen-
tially different.
Observing that the cubic strain is given by
2Gh ¼ D,0 ; ð5:131Þ
we get the state of stress in the form
m
rij ¼ D,ði;jÞ  ð,k;k þ,0 Þ;ij  D,0 dij ; i; j ¼ 1; 2; 3: ð5:132Þ
1  2m
If the volume forces are not zero, then we can express them univocally in the form

F ¼ DF0 ; ð5:133Þ
226 5 General Equations. Formulation of Problems

for a simply connected domain; for a particular solution, corresponding to the


given volume forces, we may use the representation (5.130), where the vector
potential must verify the equation
h2 , þ 2F0 ¼ 0; ð5:134Þ

while the scalar potential is given by the equation


1  2m
h1 ,0 ¼ div F0 : ð5:1340 Þ
1m

5.3.1.4 The Lamé-Clebsch Representation

Taking into account the notation (5.127), the representation (5.130) reads

2Gu ¼ D,  grad divð, þ ,0 Þ: ð5:135Þ

The differential relation (A.920 ) leads to

2Gu ¼ curl curl ,  grad div ,0 ; ð5:1350 Þ


introducing the scalar potential u ¼ uðx1 ; x2 ; x3 ; tÞ by means of the relation

u ¼  div ,0 ð5:136Þ
and the vector potential W ¼ Wðx1 ; x2 ; x3 ; tÞ with the aid of the relation
W ¼  curl ,; ð5:137Þ

we get the Lamé-Clebsch representation


2Gu ¼ gradu þ curl W: ð5:138Þ
We obtain the same result starting from the first Sternberg-Eurbanks representation
(5.122), where we take the vector potential in the form
, ¼ curl W ð5:139Þ
and where we take into account the formula (A.1100 ).
We notice that the potentials thus introduced must be functions of class C3 ; the
scalar potential must verify the longitudinal wave equation
h1 u ¼ 0; ð5:140Þ
while the vector potential must verify the transverse wave equation
h2 W ¼ 0: ð5:1400 Þ

We also notice that the vector potential must verify the condition
div W ¼ 0: ð5:14000 Þ
5.3 Dynamical Problem. Potential Functions 227

Hence, the corresponding field is a solenoidal one. But this condition is not nec-
essary to verify identically Lamé’s equations; taking into account this notice and
the fact that the representation in the previous subsection is complete, it results that
the representation (5.138), with or without the condition (5.14000 ) is complete.
In components, we may write
2Gui ¼ u;i þ ijk wk;j ; i ¼ 1; 2; 3; ð5:1380 Þ

in a developed form, we have

2Gu1 ¼u;1  w2;3 þw3;2 ;


2Gu2 ¼u;2  w3;1 þw1;3 ; ð5:13800 Þ
2Gu3 ¼u;3  w1;2 þw2;1 :

We notice that the cubic strain is given by


2Gh ¼ Du; ð5:141Þ

in this case, it result the state of stress in the form


1 m
rij ¼ u;ij  ðikl wk;j þjkl wk;i Þ;l þ Dudij ; i; j ¼ 1; 2; 3: ð5:142Þ
2 1  2m
If the volume forces are non-zero, then one obtains the same representation
(5.138); replacing in Lamé’s Eq. (5.1070 ), we get the condition
2ð1  mÞ
gradh1 u þ curlh2 W þ 2F ¼ 0; ð5:143Þ
1  2m
where we took into account the relation (5.106). Expressing the volume forces in
the form
F ¼ gradU þ curlW; ð5:144Þ

where the functions U ¼ Uðx1 ; x2 ; x3 ; tÞ and W ¼ Wðx1 ; x2 ; x3 ; tÞ are given, the


equations
1  2m
h1 u þ ¼ 0; ð5:145Þ
1m
h2 W þ 2W ¼ 0 ð5:1450 Þ

may provide the corresponding particular integrals.

5.3.2 Formulations in Stresses

In what follows, we present the formulation in stresses of the problems of elas-


todynamics, deducing the equations of the Beltrami-Michell type, and the repre-
sentations of Beltrami-Finzi, Maxwell and Morera type, as well as Teodorescu’s
representations.
228 5 General Equations. Formulation of Problems

5.3.2.1 Equations of Beltrami-Michell Type

If we wish to express the conditions of continuity in stresses, in the dynamic case,


then we can follow the considerations made in Sect. 2.2.1. Taking into account the
equations of motion (5.6) and Hooke’s law (5.3), (5.30 ), we write the relations
q
rij;ij ¼ ð1  2mÞ r€ll  Fl;l ; ð5:146Þ
E
q
rkði;jÞk ¼ ½ð1 þ mÞ€
rij  r
€ll dij   Fði;jÞ ; ð5:1460 Þ
E
starting from the Eq. (5.130) and using the relation (5.146), we get the equation
which is verified by the sum of normal stresses in the form
1þm 1þm
h1 rii ¼ h1 H ¼  Fi;i ¼  divF; ð5:147Þ
1m 1m
result to which one could arrive also starting from the Eq. (5.109), with the aid of
Hooke’s law.
In the absence of volume forces or in case of a solenoidal field of volume
forces, the sum of normal stresses verifies the longitudinal wave equation
h1 H ¼ 0: ð5:148Þ
0
Using the relations (5.146), (5.146 ) and the Eq. (5.147), the Eq. (5.129) take
the form
 
1 mð1 þ mÞ q o2
h2 rij þ oi o j þ d ij H
1þm 1  m E ot2
m
¼ 2Fði;jÞ  Fk;k dij ; i; j ¼ 1; 2; 3; ð5:149Þ
1m
these are the continuity equations in stresses of Beltrami-Michell type, obtained
independently by Ignaczak [20] and Teodorescu [48].
In a developed form, one has
 
1 mð1 þ mÞ q € m
h2 r11 þ H;11 þ H ¼  2F1;1  ðF1;1 þ F2;2 þ F3;3 Þ;
1þm 1m E 1m
 
1 mð1 þ mÞ q € m
h2 r22 þ H;22 þ H ¼  2F2;2  ðF1;1 þ F2;2 þ F3;3 Þ; ð5:150Þ
1þm 1m E 1m
 
1 mð1 þ mÞ q € m
h2 r33 þ H;33 þ H ¼  2F3;3  ðF1;1 þ F2;2 þ F3;3 Þ;
1þm 1m E 1m
5.3 Dynamical Problem. Potential Functions 229

1
h2 r23 þ H;23 ¼  ðF2;3 þ F3;2 Þ;
1þm
1
h2 r31 þ H;31 ¼  ðF3;1 þ F1;3 Þ; ð5:1500 Þ
1þm
1
h2 r12 þ H;12 ¼  ðF1;2 þ F2;1 Þ:
1þm
Applying the operator h1 to the Eq. (5.149) and taking into account d’Alem-
bert’s operators (5.103), we get the equations
1
h1 h2 rij ¼ 2h1 Fði;jÞ þ ðoi oj  mh2 dij ÞFk;k ; i; j ¼ 1; 2; 3; ð5:151Þ
1m
which are verified by the components of the stress tensor; in the absence of volume
forces, the stresses verify the double wave equation
h1 h2 rij ¼ 0; i; j ¼ 1; 2; 3; ð5:152Þ

while the equations of Beltrami type are of the form


 
1 mð1 þ mÞ q o2
h2 rij þ oi oj þ d ij H ¼ 0; i; j ¼ 1; 2; 3: ð5:1490 Þ
1þm 1  m E ot2

We notice that the stresses must be functions of class C2 to can verify the
Eq. (5.149); if we wish to verify the Eq. (5.151) too, then these functions must be
of class C4 :
We also notice that the set of integrals of the Eq. (5.151) contains the set of
integrals of the Eq. (5.149), the first equations being not sufficient to solve the
problem. For a solution in stresses one must integrate the system formed by the
equations of Beltrami-Michell type (5.149) and by the equations of motion (5.6)
(one must take into account the equations of motion too, because they represent
more restrictive conditions than those imposed by the Beltrami-Michell type
equations). To this system one must associate Hooke’s law (5.30 ), where we take
into account Cauchy’s relations (5.2), because in the equations of motion appear
the displacements too.

5.3.2.2 The Beltrami-Finzi Type Representation

We start from the equations of motion (5.6), in the absence of the volume forces,
and express the displacement vector in the form
ui ¼ uji;jkk ¼ DUji;j ; i ¼ 1; 2; 3; ð5:153Þ

where uij is an asymmetric tensor; obviously, this representation has a general


character. The Eq. (5.6) become
230 5 General Equations. Formulation of Problems

ðrij  q€
uki;kj Þ;j ¼ 0; i ¼ 1; 2; 3;

wherefrom it results, as in the static case,


rij ¼ ikl flj;k þ q€
ukj;ki ; i ¼ 1; 2; 3; ð5:154Þ

where fij is an asymmetric tensor.


Putting the condition (5.140) and proceeding as in Sect. 5.2.2.2, we get
fij ¼ fll dij þ jkl gli;k þ qikl u
€ jk;l ; ð5:155Þ

where gij is an asymmetric tensor too; by contraction,


1
fll ¼  ijk ðgij þ q€
uij Þ;k ; ð5:1550 Þ
2
so that
1
fij ¼  klm ðgkl þ q€
ukl Þ;m dij þ jkl gli;k þ qikl u
€ jk;l : ð5:15600 Þ
2
The tensor (5.1550 ) becomes
1
rij ¼ ijk mnp gmn;pk þ ikl jmn gnl;mk
2   ð5:1540 Þ
1
þ q ijk mnp u € mn;pk þ ikl lmn u
€ jm;nk þ u
€ kj;ki ; i; j ¼ 1; 2; 3;
2

using the considerations made in Sect. 5.2.2.2 and taking into account (A.38), we
obtain
 
1
rij ¼ ikl jmn Fkm;ln þ q D€
uij þ 2€
u½kj;ki þ ijk mnp u
€ mn;pk ; ð5:15400 Þ
2

where we have put in evidence the antisymmetric part of the tensor uij .
Introducing the notations

DuðijÞ ¼ F ij ; ð5:156Þ

u½ij;l ¼ ijk Fkl0 ; ð5:1560 Þ

which correspond to the symmetric and to the antisymmetric part of uij ; respec-
tively, we may write
1 0 0 0
Duji þ 2u½kj;kl þ ijk mnp umn;pk ¼ F ij  ijk Fkl;l þ 2jkl Fki;l þ ijk Fli;k :
2
We also introduce the notation
5.3 Dynamical Problem. Potential Functions 231

0 0 0
ikl F kj;l ¼ ijk Fkl;l þ 2jkl Fki;l þ ijk Fll;k ;

so that the stresses (5.15400 ) will be given by

€ €
 Þ; i; j ¼ 1; 2; 3;
 ij þ ikl F
rij ¼ ikl jmn Fkm;ln þ qðF kj;l ð5:157Þ

and the displacements by


 ij;j  ijk Fkl;lj
ui ¼ DuðjiÞ;j þ u½ji;llj ¼ F 0
¼F  ; i; j ¼ 1; 2; 3:
 ij;j þ ikl F ð5:158Þ
kj;lj

We notice that one can write these displacements in the form


 ij;j þ ðikl F
ui ¼ F  Þ ; i ¼ 1; 2; 3;
 kj;l þ jkl F ð5:1580 Þ
ki;l ;j

and the stress (5.158) in the form

€ €
 þ F
 ij þ ikl F €

rij ¼ ikl jmn Fkm;ln þ qðF kj;l jkl ki;l Þ; i; j ¼ 1; 2; 3; ð5:1570 Þ

too, because the new term introduced leads to rij;j ¼ 0 and does not influence the
equations of motion; but the tensor
 þ F
ikl F 
kj;l jkl ki;l

is symmetric with respect to the indices i and j; so that it can be contained in the
 ij ; without losing the generality. Finally we obtain the Beltrami-Finzi type
tensor F
representation, given by Teodorescu,

 ij ; i; j ¼ 1; 2; 3;
rij ¼ ikl jmn Fkm;ln þ qF ð5:159Þ
 ij;j ; i ¼ 1; 2; 3:
ui ¼ F ð5:1590 Þ

which is complete for a simply connected domain.


This representation depends on 2  6 ¼ 12 potential functions; the functions Fij
must be of class C3 with respect to space variables and of class C0 with respect to
 ij must be of class C1 with respect to space variables
time, while the functions F
2
and of class C with respect to time.
In the case of non-zero volume forces, we can use particular integrals of the
form
Z
r11 ¼  F1 dx1 þ qF € ;
1
Z
r22 ¼  F2 dx2 þ qF € ; ð5:160Þ
2
Z
r33 ¼  F3 dx3 þ qF € ;
3
232 5 General Equations. Formulation of Problems

r23 ¼ r31 ¼ r12 ¼ 0; ð5:1600 Þ


 1;1 ; u2 ¼ F
u1 ¼ F  2;2 ; u3 ¼ F
 3;3 ; ð5:16000 Þ

where F1 ; F
2; F
 3 are functions of class C1 with respect to space variables and of
2
class C with respect to time, while F1 ; F2 ; F3 are integrable functions.

5.3.2.3 Representations of Maxwell and Morera Type

 ij ¼ 0; i 6¼ j; i; j ¼ 1; 2; 3;, one obtains a representation of


If one makes Fij ¼ 0; F
Maxwell type
€ ;
r11 ¼F22;33 þ F33;22 þ qF 11

r ¼F þF €

þ qF ; ð5:161Þ
22 33;11 11;33 22
€ ;
r33 ¼F11;22 þ F22;11 þ qF 33

r23 ¼ F11;23 ; r31 ¼ F22;31 ; r12 ¼ F33;12 ; ð5:1610 Þ


 11;1 ; u2 ¼ F
u1 ¼ F  22;2 ; u3 ¼ F
 33;3 ; ð5:16100 Þ
 ij ¼ 0; i ¼ j; i; j ¼ 1; 2; 3; one gets a representation
as well, if one takes Fij ¼ 0; F
of Morera type, of the form
r11 ¼ 2F23;23 ; r22 ¼ F31;31 ; r33 ¼ F12;12 ; ð5:162Þ

€ Þ;
r23 ¼ ðF23;1 þ F31;2 þ F12;3 Þ;1 þ qF 23

r ¼ ðF
31 þF 31;2 þ F Þ þ qF
12;3
€ Þ;23;1 ;2 31 ð5:1620 Þ
€ Þ;
r12 ¼ ðF12;3 þ F23;1 þ F31;2 Þ;3 þ qF 12

 12;2 þ F
u1 ¼ F  31;3 ; u2 ¼ F
 23;3 þ F
 12;1 ; u3 ¼ F
 31;2 þ F
 23;2 : ð5:16200 Þ

These two representations depend each one only on 2  3 ¼ 6 potential functions.


Making the same considerations as in Sect. 5.2.2.3, one can show that both
representations are complete for a simply connected domain.
 ij ¼ 0 for i 6¼ j or i ¼ j; one obtains
We also notice that, taking only Fij ¼ 0 or F
other complete representations.

5.3.2.4 Teodorescu’s Representation

On the way indicated in Sect. 5.2.2.4, one can obtain a representation analogue to
that given by Schaefer in the static case, starting from the representation (5.159),
5.3 Dynamical Problem. Potential Functions 233

(5.590 ) of Beltrami-Finzi type. We choose thus the components of the tensor stress
functions Fij ; F ij ; in the form (5.150) and in the form

 ij þ ðH
 ij ¼ H
F  ij ;
 ll  XÞd ð5:163Þ
 ij ¼ H
where the potential functions Hij ¼ Hij ðx1 ; x2 ; x3 ; tÞ, H  ij ðx1 ; x2 ; x3 ; tÞ are
0
given by (5.150 ) and by

H  ij þ 1 ðF
 ij ¼ F  ij ;
 ll þ XÞd ð5:1630 Þ
2
respectively; these functions are univocally determined, excepting the arbitrary
 ¼ Xðx
functions X ¼ Xðx1 ; x2 ; x3 ; tÞ and X  1 ; x2 ; x3 ; tÞ.
Introducing in the representations (5.160), (5.1590 ), we get the representation

rij ¼  ikl jmn Hkm;ln þ DðHll  XÞdij  ðHll  XÞ;ij


 qH€
 þ qðH € €
 ; i; j ¼ 1; 2; 3;
  XÞd ð5:164Þ
ij ll ij

 ij;j þ ðH
ui ¼  H  ;
 ll  XÞ ð5:1640 Þ
;i

as well as the relations

H ¼ rll ¼ Hij;ij þ DðHll  2XÞ þ qð2H €


€  3XÞ; ð5:16400 Þ
ll

 ij;ij þ DðH
H ¼ ui;i ¼ H 
 ll  XÞ; ð5:16400 Þ

if the stresses (5.164) satisfy the Eq. (5.1490 ) too, then one obtains

 ikl jmn h2 Hkm;ln þ Dh2 ðHll  XÞdij  h2 ðHll  XÞ;ij  qh2 H €


ij

€ €
 ij þ
 ll  XÞd 1 €  3qX €
þ qh2 ðH D Hll;ij  2DX;ij þ Hkl;klij þ 2qH ll;ij ;ij
1þm

m q € € þH €

€ kl;kl þ 2qH €€
 ll  3qXÞd
þ ðDHll  2DX ij ¼ 0:
2ð1  mÞ l
It results thus

 ikl jmn h2 Hkm;ln  h2 ðHll  XÞ;ij  qh2 H €


ij
1 €  3qX € Þ ¼ 0;
þ ðDHll;ij  2DX;ij þ Hkl;klij þ 2qH ll;ij ;ij ð5:165Þ
1þm

mh2 ðHll  XÞ þ qh2 ðH€ €



 ll  XÞ
m q € € þH €€
€ kl;kl þ 2qH €€
þ 2
ðDHll  2DX ll  3qXÞ ¼ 0: ð5:1650 Þ
2ð1  m Þ l
234 5 General Equations. Formulation of Problems

If the components of the stress functions tensor satisfy the equations


 ij ¼ 0;
h2 Hij ¼ h2 H ð5:166Þ

then we can write the conditions (5.165), (5.1650 ) in the form


 
1 €
€  3qXÞ
h2 X þ ðDHll  2DX þ Hkl;kl þ 2qH ll ¼ 0; ð5:167Þ
1þm ;ij

Dh2 X þ qh2 X€

m q € € þH €€
€ kl;kl þ 2qH €€
þ 2
ðDHll  2DX ll  3qXÞ ¼ 0: ð5:1670 Þ
2ð1  m Þ l
Supposing that following relations

Hll þ 2lH  ¼0
 ll ¼ 0; X þ 2lX ð5:168Þ

take place, the conditions (5.167), (5.1670 ) become


½ð1  mÞh1 X  Hkl;kl ;ij ¼ 0; ð5:169Þ
m q € H
€ kl;kl  ¼ 0:
h1 h2 X  ½ð1  mÞh1 X ð5:1690 Þ
2ð1  m2 Þ l
We notice that the condition (5.167) is satisfied if the function is chosen so as to
verify the relation

1
h1 X ¼ Hkl;kl ; ð5:170Þ
ð1  mÞ
taking into account (5.166), we notice that the function X must satisfy the double
waves equation
h1 h2 X ¼ 0; ð5:1700 Þ

so that the condition (5.1690 ) is verified too.


One must see if Hooke’s law is satisfied too; introducing the representation
(5.164), (5.1640 ) in the corresponding equations, one gets

ðH  H
 ll  XÞ  kði;jÞk ¼ 1 D Hij  2Hkði;jÞk þ X;ij  qH
€ ij
;ij
2l

1 € € d : ð5:171Þ
 ll  DXÞ  ð1  2mÞqX
þ ½Hkl;kl  mDHll þ ð1  mÞðqH
1þm ij

Taking into account the conditions (5.166), (5.168), (5.170), one may write the
conditions (5.171) in the form
5.3 Dynamical Problem. Potential Functions 235

 
1 q € €
  kði;jÞk ;
Hll;ij þ DHij  2Hkði;jÞk  Hll dij  qHij ¼ H ð5:1710 Þ
2l 2l

we notice that these conditions are fulfilled if one takes

 ij ¼ 1 ð2Hij  Hll dij Þ;


H ð5:172Þ
2l
relation which is concordance with the first condition (5.168). All the equations of
the basic system of equations of elastodynamics are thus satisfied.
Returning to the relations (5.58) between Finzi’s and Schaefer’s tensors, we
notice that for i 6¼ j we have the relation (5.64), while for i ¼ j one can write the
relation (5.640 ). The first case corresponds to Morera’s stress functions, which
must satisfy the transverse waves equation
h2 Fij ¼ 0; i 6¼ j; ð5:173Þ

while the second case corresponds to Maxwell’s stress functions. Which must
satisfy the double waves equation
h1 h2 Fij ¼ 0; i ¼ j: ð5:1730 Þ
Taking into account (5.166), (5.168), (5.170), (5.172) and (A.37), the repre-
sentation (5.164), (5.16400 ) becomes
rij ¼ mh2 Xdij þ X;ij  2Hkði;jÞk ; i; j ¼ 1; 2; 3; ð5:174Þ

2Gui ¼ X;i  2Hij;j ; i ¼ 1; 2; 3; ð5:1740 Þ

that is Teodorescu’s representation. The seven stress functions must be of class C4


with respect to space variables; the function X must be of class C4 with respect to
time, while the functions Hij must be of class C2 with respect to the same variable.
We notice that the sum of the normal stresses is given by
rll ¼ ð1 þ mÞh2 X: ð5:176Þ
Although we have put supplementary conditions in the above proof, the rep-
resentation (5.174, 5.1740 ) is complete for a simply connected domain; one can put
in evidence this fact by making the connection between this representation and the
Somigliana–Iacovache one.
Introducing the notation (5.70), as in the static case, one can write the above
representation in the form
rij ¼ mh2 Xdij þ X;ij  2Uði;jÞ ; i; j ¼ 1; 2; 3; ð5:176Þ

2Gui ¼ X;i  2Ui ; i ¼ 1; 2; 3: ð5:1760 Þ


236 5 General Equations. Formulation of Problems

The functions Ui ¼ Ui ðx1 ; x2 ; x3 ; tÞ of class C3 with respect to the space vari-


ables and of class C2 with respect to time are the components of a vector U which
satisfies the transverse waves equation
h2 U ¼ 0; ð5:177Þ

while the function X must verify the equation


1
h1 X ¼ div U; ð5:178Þ
1m
Using the notation
U ¼ ð1  mÞh1 C; ð5:179Þ
we notice that a relation of the form (5.76) takes place, where we have neglected a
function which satisfies the longitudinal waves equation, because it can be
included in the components of the vector C. If in the representation (5.1760 ),
written in a vector form
2Gu ¼ gradX  2U; ð5:17600 Þ

one takes into account (5.76) and (5.179), then one finds again the Somigliana-
Iacovache representation (5.122); the latter representation being complete for a
simply connected domain, it results that Teodorescu’s representation is complete
in the same conditions too.
In a developed form, one can write the state of displacement in the form (5.72),
the state of stress being given by

r11 ¼X;11  mh2 X  2U1;1 ;


r22 ¼X;22  mh2 X  2U2;2 ; ð5:180Þ
r33 ¼X;33  mh2 X  2U3;3 ;

r23 ¼X;23  ðU2;3 þ U3;2 Þ;


r31 ¼X;31  ðU3;1 þ U1;3 Þ; ð5:1800 Þ
r12 ¼X;12  ðU1;2 þ U2;1 Þ:
If the volume forces are non-zero, then one can use the same representations
(5.174), (5.1740 ) or (5.176), (5.1760 ). In this case, the basic system of equations of
elastodynamics is verified if the vector U is given by the equation
h2 U ¼ F; ð5:181Þ
while the function X is given by the Eq. (5.178); applying the operator h2 to this
equation, we notice that this function satisfies the equation
1
h1 h2 X ¼ divF: ð5:182Þ
1m
5.3 Dynamical Problem. Potential Functions 237

As in the static case, from the point of view of the particular solution, corre-
sponding to given volume forces, we notice that one can replace the Eq. (5.178) by
the Eq. (5.182); indeed, the last equation is equivalent to the equation
h2 ½ð1  mÞh1 X  Ui;i  ¼ 0;

which leads to
ð1  mÞh1 X  Ui;i ¼ u;

where the function u satisfies the transverse waves equation. Taking into account
(5.181), one can make a decomposition of the form (5.7300 ), where the function U 
satisfies the transverse waves equation, while U0 is a particular integral. We notice
that the function u may be included, as in the static case, in the components of the
vector U; hence, we may use the Eq. (5.181) to obtain the particular integral
corresponding to the given volume forces.
An important particular case is that of conservative forces, of the form (5.77),
where v ¼ vðx1 ; x2 ; x3 ; tÞ is a function of class C2 with respect to the space vari-
ables; one can choose a particular integral of the form (5.78), where U ¼
 1 ; x2 ; x3 ; tÞ is a function of class C with respect to the space variables and of
Uðx 4

class C2 with respect to time. The Eq. (5.181) leads to


 ¼ v:
h2 U ð5:183Þ
We notice that the Eq. (5.78) becomes
1 
h1 X ¼ DU; ð5:184Þ
1m
while the relation (5.106) leads to

ð1  2mÞh2 X ¼ DðX  2UÞ;
we introduce the scalar potential x ¼ xðx1 ; x2 ; x3 ; tÞ by means of a relation of the
form (5.81). The corresponding state of stress is given by a relation of the form
(5.83) and the state of displacements by a relation of the form (5.82).
We notice that the Eq. (5.184) becomes

ð1  mÞh1 x ¼ DU 
  2ð1  mÞh1 U; ð5:1840 Þ
it results, in this case, that the function x must be a particular integral of the
equation
1  2m
h1 x þ v ¼ 0: ð5:185Þ
1m
Thus, the particular solution can be obtained by means of only one stress function
x, which must be of class C3 with respect to space variables and of class C2 with
238 5 General Equations. Formulation of Problems

respect to time. If we wish to satisfy the Beltrami-Michell type equations too, then
the function x must be of class C4 with regard to space variables.
The considerations made in Sect. 5.2.2.4 concerning the boundary conditions
remain valid in this case too.

5.3.2.5 Teodorescu’s Representation for the Beltrami-Michell Type


Equations

As in the static case, one can obtain a representation in stresses of the dynamical
problem of the theory of elasticity starting from the Beltrami type Eq. (5.1490 ); the
condition to verify the equations of motion and Hooke’s law too is then put.
One introduces a function F by the relation
H ¼ ð1 þ mÞh2 F; ð5:186Þ

in this case, the Eq. (5.1490 ) show that the stresses must be of the form
 
m q o2
rij ¼  oi oj þ d ij F þ uij ; i; j ¼ 1; 2; 3; ð5:187Þ
2ð1  mÞ l ot2

where uij ¼ uji are functions which verify the transverse waves equation

h2 uij ¼ 0; i; j ¼ 1; 2; 3: ð5:188Þ

Taking into account (5.186), we get


ð2 þ mÞh1 F ¼ ukk ; ð5:189Þ
while the conditions (5.188) show that the function F must verify the double waves
equation
h1 h2 F ¼ 0: ð5:190Þ
By the relations
uij ¼ ð2 þ mÞh1 Fij ; ð5:191Þ

we introduce the functions Fij ¼ Fij ðx1 ; x2 ; x3 ; tÞ; i; j ¼ 1; 2; 3; too, components of


a symmetric stress tensor function; the Eq. (5.188) show that these functions verify
the double wave equation
h1 h2 Fij ¼ 0; i; j ¼ 1; 2; 3: ð5:192Þ
The condition (5.189) leads, in this case, to F ¼ Fkk ; where we neglect a
function which verifies the longitudinal waves equation and which can be intro-
duced in each of the functions F11 ; F22 ; F33 if one takes into account Boggio’s
formula (A.107).
5.3 Dynamical Problem. Potential Functions 239

We obtain thus Teodorescu’s representation


 
m q o2
rij ¼ ð2 þ mÞh1 Fij  oi oj þ d ij Fkk ; i; j ¼ 1; 2; 3; ð5:194Þ
2ð1  mÞ l ot2

which is complete for the system of equations of Beltrami type, in case of a simply
connected domain; the six stress functions Fij must be of class C4 .
In case of non-zero volume forces, one can use the same representation, the
functions Fij being particular integrals of the equations
1 h m i
h1 h2 Fij ¼  Fij þ Fk;k dij ; i; j ¼ 1; 2; 3; ð5:1920 Þ
2þm 1m
thus, the Eq. (5.149) of the Beltrami-Michell type are identically verified.
Using Boggio’s formula (A.107), we may write the functions Fij in the form

Fij ¼ Wij þ Uij ; i; j ¼ 1; 2; 3; ð5:194Þ

where
h1 Wij ¼ 0; h2 Uij ¼ 0; i; j ¼ 1; 2; 3; ð5:1940 Þ

thus, we obtain a new representation for the state of stress in the form
 
m q o2
rij ¼ ð2 þ mÞh1 Uij  oi oj þ dij ðUll þ U0 Þ; i; j ¼ 1; 2; 3; ð5:195Þ
2ð1  mÞ l ot2

where we have introduced the function


U0 ¼ Wll : ð5:19400 Þ
The functions Uij ¼ Uij ðx1 ; x2 ; x3 ; tÞ and the function U0 ¼ U0 ðx1 ; x2 ; x3 ; tÞ will
verify the equations
h2 Uij ¼ 0; h1 U0 ¼ 0; i; j ¼ 1; 2; 3; ð5:1900 Þ

both the tensor potential and the scalar one being functions of class C4 with respect
to all variables.
In a developed form, one can write the state of stress in the form
 2 
o m q o2
r11 ¼ð2 þ mÞh1 U11  þ ðU11 þ U22 þ U33 þ U0 Þ;
ox21 2ð1  mÞ l ot2
 2 
o m q o2
r22 ¼ð2 þ mÞh1 U22  þ ðU11 þ U22 þ U33 þ U0 Þ; ð5:1950 Þ
ox22 2ð1  mÞ l ot2
 2 
o m q o2
r33 ¼ð2 þ mÞh1 U33  þ ðU11 þ U22 þ U33 þ U0 Þ;
ox23 2ð1  mÞ l ot2
240 5 General Equations. Formulation of Problems

r23 ¼ð2 þ mÞh1 U23  ðU11 þ U22 þ U33 þ U0 Þ;23 ;


r31 ¼ð2 þ mÞh1 U31  ðU11 þ U22 þ U33 þ U0 Þ;31 ; ð5:19500 Þ
r12 ¼ð2 þ mÞh1 U12  ðU11 þ U22 þ U33 þ U0 Þ;12 :

To show the conditions in which the state of stress given by (5.183) corresponds
to the basic system of equations of elastodynamics, one can make a study analogue
to that in the static case (see Sect. 5.2.2.5).

References

A. Books

1. Finzi, B., Pastori, M.: Calcolo tensoriale e applicazioni, 2nd edn. Zanichelli, Bologna (1961)
2. Gurtin, M.E.: The Linear Theory of Elasticity. Encyclopedia of Physics, vol. VI a/2, Springer,
Berlin-Heidelberg-New York (1972)
3. Haimovici, M.: Teoria elasticitătßii (The Theory of Elasticity). Ed. did. ped, Bucuresßti (1969)
4. Jaunzemis, W.: Continuum Mechanics. The Macmillan Comp, New York-London (1967)
5. Kecs, W., Teodorescu, P.P.: Applications of the Theory of Distributions in Mechanics. Ed.
Academiei, Bucuresßti, Abacus Press, Tunbridge Wells, Kent (1974)
6. Lamé,G.: Leçons sur la théorie mathématique de l’élasticité des corps solides. Paris (1852)
7. Moisil, Gr. C.: Matricele asociate sistemelor de ecuatßii cu derivate partßiale. Introducere în
studiul cercetărilor lui I. N. Lopatinschi (Matrices Associated to Systems of Partial
Derivative Equations. Introduction to the Study of I. N. Lopatinski’s Researches).
Ed. Academiei, Bucuresßti (1950)
8. Nowacki, W.: Dynamics of Elastic Systems. Chapman and Hall Ltd., London (1963)
9. Nowacki, W.: Teoria spre_zistošci (Theory of Elasticity). Państ. Wyd. Nauk, Warsawa (1970)
10. Sedov, L.I.: Mehanika splošnoi sredy (Mechanics of Deformable Media). I. II., Izd. Nauka,
Moskva (1970)
11. Teodorescu, P.P.: Dynamics of Linear Elastic Bodies. Ed. Academiei, Bucuresßti, Abacus
Press, Tunbridge Wells, Kent (1975)

B. Papers

12. Beltrami, E.: Osservazioni sulla nota precedente. Atii R. Accad. dei Lincei. Rend. Cl. Sci. fis.
mat. e nat., ser. 5(3), 141 (1892)
13. Blokh, V.I.: Funktsii napryzhenii v teorii uprugosti (Stress functions in the theory of
elasticity). Prikl. mat. i mekh 14, 415 (1950)
14. Bondarenko, B.A.: Ob odnom klass reshenii dinamicheskikh uravnenii teorii uprugosti (On a
class of solutions of dynamical equations in the theory of elasticity). Akad. Nauk. Uzbek.
SSR, Trudy Inst. Mat. Mekh. 21, 41 (1957)
15. Finzi, B.: Integrazione delle equazioni indefinite della meccanica dei sistemi continui. Atti.
R. Acad. dei Lincei, Rend., Cl. Sci. fis. mat e nat., ser. 6, 19, 578, 620 (1934)
References 241

16. Galerkin, B.G.: Contribution à la solution générale du problème de la théorie de l’élasticité


dans le cas de trois dimensions. C. Rend. Hebd. des séances de l’Acad. des Sci. 190, 1047
(1930)
17. Gurtin, M.E.: A generalisation of the beltrami stress functions in continuum mechanics. Arch.
Rat. Mech. Anal. 13, 321 (1963)
18. Iacovache, M.: O extindere a metodei lui Galerkin pentru sistemul ecuatßiilor elasticitătßii (A
generalisation of Galerkin’s method for the elasticity equations system). Bul. ßst. Acad., ser. A
1, 592 (1949)
19. Iacovache, M.: Asupra unor integrale ale ecuatßiilor micilor misßcări ale corpurilor elastice (On
Certain Integrals of the Equations of Small Motions of Elastic Bodies). Acad. Rom., Lucr.
ses. gen. 299 (1951)
20. Ignaczak, J.: Direct determination of stresses from the stress equation of motion in elasticity.
Arch. Mech. Stos. 11, 671 (1959)
21. Ionescu, D.: Asupra vectorului lui Galerkin in teoria elasticitătßii ßsi în hidrodinamica fluidelor
vâscoase (On Galerkin’s Vector in the theory of elasticity and in the hydrodynamics of
viscous fluids). Bul. ßst. Acad., sec. ßst. mat. fiz., 6, 555 (1954)
22. Kröner, E.: Die Spannungsfunktionen der dreidimensionalen isotropen Elastizitätstheorie.
Z. Physik 139, 175 (1954); Korrektur. 143, 374 (1955)
23. Langhaar, H., Stippes, M.: Three-dimensional stress functions. J. Franklin Inst. 258, 371
(1954)
24. Marguerre, K.: Ansätze zur Lösung der grundgleichungen der Elastizitätstheorie. Z. A. M. M.
35, 242 (1955)
25. Michell, J.H.: On the direct determination of stress in an elastic solid with applications to
theory of plates. Proc. London Math. Soc. 31, 100 (1900)
26. Moisil, G.C.: Un analog al vectorului lui Galerkin în hidrodinamica lichidelor vâscoase (An
analogous of Galerkin’s Vector in hydrodynamics of viscous liquids). Bul. ßst. Acad., ser. A.
1, 803 (1949)
27. Moisil, G.C.: Asupra descompunerii undelor seismice în unde de condensare ßsi unde de
forfecare (On the decomposition of seismic waves in condensation and shear waves). Bul. ßst.
Acad., ser. mat. fiz. chim. 2, 235 (1950)
28. Moisil, G.C.: Asupra ecuatßiilor echilibrului corpurilor elastice (On the equilibrium equations
of elastic bodies). An. Acad., ser. mat. fiz. chim. 3, 739 (1950)
29. Morera, G.: Soluzione generale delle equazioni indefinite dell’equilibrio di un corpo
continuo. Atti. R. Accad. dei Lincei, Rend., Cl. Sci. fis., mat. e nat., ser. 5, 1, 137 (1892)
30. Morera, G.: Appendice alla nota: Sulla soluzione generale delle equazioni indefinite
dell’equilibrio di un corpo continuo. Atti. R. Accad. dei Lincei, Rend., Cl. Sci. fis., mat. e
nat., ser. 5, 1, 233 (1892)
31. Neuber, H.: Ein neuer Ansatz z} ur Lösung räumlicher Probleme der Elastizitätstheorie. Der
Hohlkegel unter Einzellast als Beispiel. Z. A. M. M., 14, 203 (1934)
32. Ornstein, W.: Stress functions of Maxwell and Morera. Quart. Appl. Math. 12, 198 (1954)
33. Papkovich, P.F.: Solution générale des équations différentielles fondamentales d’élasticité
exprimée par trois fonctions harmoniques. C. Rend. hebd. des séances de l’Acad. Sci. 195,
513 (1932)
34. Peretti, G.: Significato del tensore arbitrario che interviene nell’integrale generale delle
equazioni della statica dei continui. Atti. Sem. Mat. Fis., Univ. Modena 3, 77 (1949)
35. Rieder, G.: Topologische Fragen in der Theorie der Spannungsfunktionen. Abh.
Braunschweig. Wiss. Ges. 7, 4 (1960)
36. Rieder, G.: Die Berechnung des Spannungsfeldes von Einzelkräften mit Hilfe räumlicher
Spannungsfunktionen und ihre Anwendung zur quellenmässingen Darstellung der
Verschiebung bei Eigenspannungstäanden. Oesterr. Ing. Archiv. 18, 173 (1964)
37. Rieder, G.: Die Randbedingungen f} ur den Spannungsfunktionentensor an ebenen und
gekr}ummten belasteten Oberflächen. Oesterr. Ing. Archiv. 18, 208 (1964)
242 5 General Equations. Formulation of Problems

}
38. Rieder, G.: Uber die Spezialisierung des Schaeferschen Spannungsfunktionenansatzes in der
räumlichen Elastizitätstheorie. Z. A. M. M. 44, 329 (1964)
39. Schaefer, H.: Die Spannungsfunktionen des dreidimensionalen Kontinuums und des
elastischen Körpers. Z. A. M. M. 33, 356 (1953)
40. Schaefer, H.: Die Spannungsfunktionen einer Dyname. Abh. Braunschweig. Wiss. Ges 7, 107
(1955)
41. Schaefer, H.: Die Spannungsfunktionen des dreidimensionalen Kontinuums: statische
Deutung und Randwerte. Ing.–Aechiv, 28, 291 (1959)
42. Schuler, K.W., Fosdick, R.L.: Generalized Beltrami Stress Functions. Dept. Mech. Rept.,
Illinois Inst. Techn. (1967)
43. Somigliana, C.: Sulle espressioni analitiche generali dei movimenti oscillatori.
Atti. R. Accad. dei Lincei, Rend., Cl. Sci. fis., mat., ser. 5, 1, 111 (1892)
44. Soós, E.: The Galerkin vector for the dynamic problems of an elastic isotropic and non-
homegeneous body. Rev. Roum. Math. Pures et Appl. 10, 855 (1965)
45. Sternberg, E.: On the integration of the equations of motion in the classical theory of
elasticity. Arch. Rat. Mech. Analysis 6, 34 (1960)
46. Sternberg, E., Eubanks, R.A.: On stress functions for elastokinetics and the integration of the
repeated wave equation. Quart. Appl. Math 15, 149 (1957)
47. Teodorescu, P.P.: Sur une solution générale du problème en espace de la théorie de
l’élasticité. IXme Congrès IUTAM, Bruxelles, 1956, Actes, 5, 155 (1957)
48. Teodorescu, P.P.: Sur une représentation par potentiels dans le problème tridimensionnel de
l’élastodynamique. C. Rend. hebd. des séances de l’Acad. Sci., 250, 1972 (1960)
}
49. Teodorescu, P.P.: Uber das kinetische Problem nichthomogener elastischer K} orper. Bull.
Acad. Pol., sér. Sci. Techn., 12, 595 [867] (1964)
50. Teodorescu, P.P.: Schwingungen der elastischen Kontinua. III Konferenz f} ur nichtlineare
Schwingungen, Berlin, 1964, Abh. der deutschen Akad. Wiss., Kl. Math., Phys. u. Techn. 2,
29 (1965)
51. Teodorescu, P.P.: Sur le tenseur de Finzi et sur quelques de ses applications et
généralisations. Ann. Mat. Pura ed Appl., ser. IV, 84, 225 (1970)
52. Teodorescu, P.P.: Sur l’introduction des fonctions-potentiel en élasticité linéaire. An. Univ.
Bucuresßti, mat.–mec. 20, 131 (1971)
}
53. Teodorescu, P.P.: Uber ein Analogon der Schaeferschen Darstellung in der Elastokinetik und
einige Anwendungen. Z. A. M. M., 52, Sonderheft, T 154 (1972)
54. Teodorescu, P.P.: Stress functions in three-dimensional elastodynamics. Acta Mech. 14, 103
(1972)
55. Truesdell, C.: invariant and complete stress functions for general continua. Arch. Rat. Mech.
Anal. 4, 1 (1959/1960)
56. Weber, C.: Spannungsfunktionen des dreidimensionalen Kontinuums. Z. A. M. M. 28, 193
(1948)
57. Wo_zniak, Cz.: Introduction to dynamics of deformable bodies. Arch. Mech. Stos. 19, 647
(1967)
Chapter 6
Principles and General Theorems
of the Theory of Elasticity. Computation
Methods

We shall state in what follows some principles and general theorems of the theory
of elasticity, laying stress chiefly on the ideas connected with the notions of work
of deformation and external work.
Hereafter, we shall make a survey of some of the most important methods used
in the theory of elasticity, insisting on the method of Fourier representations, as
well as on the methods based on the theory of distributions.

6.1 Principles and General Theorems of the Theory


of Elasticity

The general (universal) theorems of mechanics in the theory of elasticity are


reflected on the equations concerning the stresses they lead to; thus, the theorem of
linear momentum allows us to obtain the equations of equilibrium or of motion for
the stresses, while the consequence of the theorem of angular momentum consist in
the relations of symmetry of the tangential stresses, hence the symmetry of the
tensor Tr . We shall not deal with these theorems, since the respective equations
have been deduced from other considerations. As to the theorem of kinetic energy,
a few of its consequences will appear later on.

6.1.1 Work

In what follows we introduce the notion of work, putting in evidence the external
work and the work of deformation.

P. P. Teodorescu, Treatise on Classical Elasticity, Mathematical and Analytical 243


Techniques with Applications to Engineering, DOI: 10.1007/978-94-007-2616-1_6,
 Springer Science+Business Media Dordrecht 2013
244 6 Principles and General Theorems. Computation Methods

6.1.1.1 External Work

The bodies submitted to the action of external loads undergo certain deformations,
while the point of application of these loads move about, by certain paths, and
yield a work, called work of external loads or external work.
We have seen in Sect. 4.1.2.2 that the elementary work, corresponding to a
force P; which imparts to the application point a displacement u; is expressed in
space co-ordinates (we consider to be in the case of infinitesimal deformations) in
the form
dW ¼ P  du ¼ P  udt
_ ¼ Pi dui ¼ Pi u_ i dt; ð6:1Þ
thus, the external work corresponding to the volume forces F and to the superficial
n
forces p is given by
ZZZ ZZ
dWe n
¼ F  udV_ þ p udS
_
dt
ZZZ V ZZ S
n
¼ Fi u_ i dV þ pi u_ i dS: ð6:2Þ
V S

The formula is valid in the dynamical case. In the statical one, we may assume
n
that, in a relatively small interval of time, the intensity of the external forces Fi ; pi
increases from zero till the final intensity; the displacements ui increase, as well,
from zero to their final magnitude. Such an action is called statical and the work
does not depend on time; in this case, the work of the force of component Pi ;
which produces the displacement ui ; which corresponds to an element of volume
or of area, is given by
Z ui
 i d
P ui ð!Þ; ð6:3Þ
0

where P i is an intermediary value of the external force, while ui is the corre-
sponding displacement. Because the phenomenon is linear, we can assume that
 i ¼ ki P0i ; Pi ¼ ki P0i ð!Þ;
P ð6:4Þ

ui ¼ ki u0i ; ui ¼ ki u0i ð!Þ;


 ð6:40 Þ

where P0i ; u0i correspond to an intermediary fixed situation, so that


Z ui Z ki
1 1
 i d
P ui ¼ P0i u0i ki dki ¼ P0i u0i ki2 ¼ Pi ui ð!Þ; ð6:30 Þ
0 0 2 2
we thus notice that appears the factor 1/2 by the product of the constant static force
with the corresponding displacement.
6.1 Principles and General Theorems of the Theory of Elasticity 245

If the force is suddenly applied, having from the very beginning its entire
intensity, while the displacement instantaneously appears, the factor 1/2 disappears
in the above relation; such an action is called dynamical, although it is a particular
case of such an action. In this case, one can write

Pi ¼ P0i hðtÞ; ui ¼ u0i hðtÞ; ð6:5Þ

introducing the distribution of Heaviside; it results

Pi dui ¼ P0i u0i hðtÞdðtÞdt ð!Þ; ð6:6Þ

which leads to
Z 1
Pi u_ i dt ¼ P0i u0i ð!Þ; ð6:60 Þ
1

relation which corresponds to the above statement.


Thus, in case of a statical action, the external work is given by the remarkable
relation
ZZZ ZZ
1 1 n
We ¼ F  udV þ p  udS
2 V 2 S
ZZZ ZZ
1 1 n
¼ Fi ui dV þ pi ui dS: ð6:7Þ
2 V 2 S

In the case of a concentrated volume force PdðrÞ; e.g., we may write


1 1
We ¼ P  u ¼ Pi u i ; ð6:8Þ
2 2
in case of a concentrated superficial force, one obtains an analogous result.

6.1.1.2 Work of Deformation

The work affected by the external loads is stored in the elastic body and given
back, by the unloading and the returning to its initial form, in the form of work of
internal forces or internal work; the latter is called the work of deformation too,
corresponding to the strain energy.
We shall denote by Ws the specific work of deformation, i.e. the work per unit
volume in the neighbourhood of a point of the body, produced by the state of stress
due to the state of strain of the body at the respective point. We shall give the name
of elementary work of deformation to the corresponding work of a volume element
dV ¼ dx1 dx2 dx3 of the body and will have
dWi ¼ Ws dV; ð6:9Þ
246 6 Principles and General Theorems. Computation Methods

it result the internal work


ZZZ
Wi ¼ Ws dV: ð6:10Þ
V

Taking into account the results in Sect. 4.1.2.2, we notice that


Ws ¼ W; ð6:11Þ

we may express the this work in the form (4.510 ), which represents the formula of
Clapeyron [19], or under one of the formulae (4.90)–(4.90000 ), (4.93)–(4.95), (4.98).
If we refer to the principal directions, then we can write
1
W ¼ kh2 þ lðe21 þ e22 þ e23 Þ
2
1
¼ ½ð1 þ mÞðr21 þ r22 þ r23 Þ  mH2 
2E
1
¼ ðr1 e1 þ r2 e2 þ r3 e3 Þ: ð6:12Þ
2
These results have the same form both in the case of statical and dynamical
actions.

6.1.1.3 Green’s and Castigliano’s Formulae

The formulae of Green [23] are given by the relations (4.9)–(4.900 ), which express
the fact that the partial derivative of the specific work of deformation with respect
to a strain is equal to the corresponding stress. As well, the formulae of Castigliano
[2, 18] (4.59)–(4.5900 ) show that the partial derivative of the specific work of
deformation with respect to a stress is equal to the corresponding strain.
One observes, from (4.9) to (4.100 ), that if Green’s formulae take place, then the
elementary work of deformation is an exact differential, wherefrom the tensor Tr
is conservative. As well Castigliano’s formulae (4.59) correspond to the fact that
the tensor Te is conservative.
We will assume now that the external work, given by (6.7), is an exact dif-
n
ferential; hence, Fi and pi must be the partial derivatives with respect to the
corresponding displacements of a potential function in these displacements. In
case of the action of a concentrated volume force F ¼ PdðrÞ or of a concentrated
n
superficial force p ¼ PdðrÞ it results
dWe ¼ P  du ¼ Pi dui ; ð6:13Þ
6.1 Principles and General Theorems of the Theory of Elasticity 247

if the force P is conservative in the sense considered above, then it results that
oWe
Pi ¼ ; i ¼ 1; 2; 3: ð6:130 Þ
oui
In a linear statical theory, we have

ui ¼ kij Pj ; Pi ¼ kij uj ; ð6:14Þ

taking into account (6.8), we have


1 1 1
We ¼ Pi ui ¼ kij Pi Pj ¼ kij ui uj ; ð6:15Þ
2 2 2
representing the external work corresponding to a concentrated force, which is a
quadratic form with respect to the components of the force or of the displacement.
It results also the relation
oWe
ui ¼ ; i ¼ 1; 2; 3: ð6:16Þ
oPi
One may thus write
oWe
¼ d; ð6:17Þ
oP
oWe
¼ P; ð6:170 Þ
od
using Clapeyron’s principle (6.200 ), which will be proved in the Sect. 6.1.2.1, one
may write, in the static case,
oWi
¼ d; ð6:18Þ
oP
oWi
¼ P; ð6:180 Þ
od
which represent the theorems of Castigliano [2, 18], corresponding to the work
stored by the whole body (Fig. 6.1). The first theorem shows that the derivative of
the internal work with respect to an external concentrated force is equal to the
displacement of its point of application on its direction. The second theorem shows
that the derivative of the internal work with respect to a displacement of a point is

Fig. 6.1 Castigliano’s


theorems d P
r

O
248 6 Principles and General Theorems. Computation Methods

equal to the external concentrated force which acts at this point on the direction of
the displacement.
The formula (6.18) allows to determine the constraint forces at the fixity points
of the elastic solids, where one has d ¼ 0:
In the case of another type of concentrated load (e.g., a rotational concentrated
moment) the theorems hold too, appearing a corresponding type of displacement
(e.g., a rotation).
In the dynamical case, the relations (6.17), (6.170 ) remain still valid at a given
moment t.

6.1.2 General Principles

Hereafter we shall present some general principles of the theory of elasticity.


Starting from these principles, we can generally obtain the equations of the stated
problems; if we start from other results, then these principles will be obtained as
consequences and will offer the aspect of theorems. Owing to their importance, we
shall maintain the name of principles. Interesting results about this subject are to
be found in the monograph of M. Gurtin [5].

6.1.2.1 Clapeyron’s Principle

If we start from the formula (6.2) and take into account the formulae (3.200 ), (3.63)
and the Gauss-Ostrogradskiı formula (A.960 ), which allows to pass from a surface
integral to a volume one, one may write
ZZ ZZ ZZZ ZZZ
n    
pi u_ i dS ¼ rji u_ i nj dS  rji u_ i ;j dV ¼ rji;j u_ i dV
S S V V
ZZZ ZZZ ZZZ
þ rji ð_eji þ x_ ji ÞdV ¼ ðq€ui  Fi )u_ i dV þ rij e_ ij dV;
V V V

we are thus led to the relation


ZZZ ZZZ
dWe
¼ ui u_ i dV þ
q€ rij e_ ij dV: ð6:20 Þ
dt V V

Taking into account Green’s formula (4.9), it results


ZZZ ZZZ
dW dWi
rij e_ ij dV ¼ dV ¼ ;
V V dt dt
on the other hand, we introduce the kinetic energy, given by
ZZZ
1
T¼ qu_ i u_ i dV: ð6:19Þ
2 V
6.1 Principles and General Theorems of the Theory of Elasticity 249

Assuming that q ¼ const with respect to time, we get


dWe ¼ dT þ dWi ;

if the natural state of stresses and the initial state of deformations correspond to the
initial moment t ¼ 0; then one obtains the generalized principle of Clapeyron in
the form
We ¼ T þ Wi : ð6:20Þ
This result constitutes, in fact, a consequence of the general theorem of energy,
representing a conservation relation of it. We have obtained this principle as a
theorem.
In the statical case, if the body is in equilibrium at any moment t; we have
T ¼ 0 and obtain the relation
We ¼ Wi ; ð6:200 Þ
which represents the principle of Clapeyron [19].

6.1.2.2 The D’Alembert-Lagrange Principle

We shall call virtual displacement dui of a system of displacements ui ; i ¼ 1; 2; 3;


of an elastic body, any system of infinitesimal variations, compatible with the
constraints (with the supporting conditions of the body), that do not depend on the
applied external loads and are not produced in time; the condition for these
variations to be infinitesimal is necessary if we do not want them involving a stress
variation. To these virtual displacements we add the corresponding virtual strains
deij ; given by relations of the form
1
deij ¼ ½ðdui Þ;j þ ðduj Þ;i ; i; j ¼ 1; 2; 3: ð6:21Þ
2
Using the principle of virtual displacements, in the static case, one can affirm
that
dWe ¼ dWi ; ð6:22Þ
hence, the external work, yielded by the external loads through the agency of some
system of virtual displacements (which drive the body out of the position of elastic
equilibrium), is equal to the internal work, produced by the state of stress of the
body through the agency of the virtual state of strain corresponding the virtual
displacements.
The external work can be expressed by
ZZZ ZZ
n
dWe ¼ Fi dui dV þ pi dui dS; ð6:23Þ
V S
250 6 Principles and General Theorems. Computation Methods

while the corresponding internal work is


ZZZ
dWi ¼ rij deij dV; ð6:230 Þ
V

we remark that the factor 1/2 does not appear here, because the external loads and
the stresses act, in this case, with their full intensity from the very start. By using
the relations (6.23), (6.230 ), one may obtain the relation (6.22) on a way analogous
to that in case of the proof of Clapeyron’s relation.
If in this principle we substitute d’Alembert’s lost forces to the volume forces,
we shall obtain the d’Alembert-Lagrange principle in the form
ZZZ ZZZ
dWe ¼ dWi þ q€ui dui dV ¼ dWi  ðq€ui Þdui dV; ð6:24Þ
V V

therefore, the external work, yielded by the external loads through the agency of a
system of virtual displacements (that drive the body out of its position of elastic
equilibrium) is equal to the difference between the internal work produced by the
state of stress of the body through the agency of the virtual state of strain corre-
sponding to the assumed system of virtual displacement (relations (6.21)) and the
work of the forces of inertia corresponding to the same system of virtual
displacements.
We have to mention that the principle of virtual displacements and the d’Al-
embert-Lagrange principle express the condition of static equilibrium and the
condition of dynamic equilibrium and the limit conditions (the equations of con-
tinuity being verified by the manner of applying the principles), respectively.
In the static case, it has been shown that this principle may be applied to all
systems which verify Betti’s reciprocity principle (see Sect. 6.1.2.7).
We remark, on the other hand, that the internal work yielded by the unloading
of the body is equal in absolute value but of an opposite sign to the work yielded
when loading it. In the statement presented above, we took into consideration the
work yielded by the unloading; if we make use of the internal work produced by
the external loads dWi0 ¼ dWi we can express the principle of virtual displace-
ments in the form

dWi þ dWi0 ¼ 0; ð6:220 Þ

which shows that the total work corresponding to an elastic solid, in case of virtual
displacements, vanishes. In the dynamic case, the d’Alembert-Lagrange principle
reads
ZZZ
0
dWe þ dWi þ ðq€ui Þdui dV ¼ 0; ð6:240 Þ
V

leading to an analogous interpretation.


6.1 Principles and General Theorems of the Theory of Elasticity 251

6.1.2.3 Principle of the Virtual Variations of the State of Stress

If, in state of the virtual displacements, we take into consideration the virtual
variations of the state of stress (defined analogically), for which the equations of
equilibrium ði ¼ 1; 2; 3Þ
 
drij ;j þdFi ¼ 0; ð6:25Þ

are satisfied, as well as the boundary conditions


n
dpi ¼ drji nj ; ð6:250 Þ
n
where dFi ; dpi are corresponding virtual variations, so that the state of stress be
statically possible, then we may enounce the principle of the virtual variations of
the state of stress, in the static case, in the same form (6.22); the external work is
given by
ZZZ ZZ
n
dWe ¼ ui dFi dV þ ui dpi dS; ð6:26Þ
V S

while the internal work reads


ZZZ
dWi ¼ eij drij dV: ð6:260 Þ
V

Thus, the external work, yielded by some system of virtual variations of the
external loads through the agency of the displacements of the body, is equal to the
internal work, yielded by the corresponding virtual variation (see the relations
(6.25), (6.250 )) of the state of stress through the agency of the state of strain of the
body.
The principle of the virtual variations of the state of stress expresses the con-
ditions of continuity of the body (the equations of equilibrium and the conditions
on the boundary being included in the manner of application of the principle). This
principle too may be written in the form (6.220 ).

6.1.2.4 Hamilton’s Principle

Let us consider an elastic body the states of which vary continuously between the
times t0 and t1 : Integrating the relation (6.24) on the assumed time interval, we
obtain
Z t1 Z t1 Z t1 ZZZ
dWe dt ¼ dWi dt þ dt q€ui dui dV;
t0 t0 t0 V
252 6 Principles and General Theorems. Computation Methods

the second integral in the right member of this relation can be related to the
variation of the kinetic energy. Indeed, we may introduce the kinetic energy in the
form (6.19), wherefrom
ZZZ ZZZ ZZZ
o
dT ¼ qu_ i du_ i dV ¼ q ð€ui dui ÞdV  q€ui dui dV;
V V ot V

if we admit that the virtual displacements dui are such that they vanish at times t0
and t1 (they are synchronous virtual displacements) it follows that
Z t1 Z t1 ZZZ
dTdt ¼  dt q€ui dui dV; ð6:27Þ
t0 t0 V

so that the relation (6.27) becomes


Z t1 Z t1
d ðWi  TÞdt ¼ dWe dt: ð6:28Þ
t0 t0

Since in (6.23) we consider that the external loads do not vary, we can remove the
operator d from under the integral (volume integral); we introduce the potential
energy
P ¼ Wi  2We ; ð6:29Þ
where We is the external work, corresponding to the static case, supplied by (6.7).
By so doing, the relation (6.28) leads to Hamilton’s principle in the form
Z t1
d ðP  T Þdt ¼ 0: ð6:30Þ
t0

The above integral takes the name of action; Hamilton’s principle affirms that
the action is steady, while the motion of the elastic body corresponds to the
extremals of this functional, in case of synchronous variations of the
displacements.

6.1.2.5 Principle of the Minimum Potential Energy

Because in (6.23) we suppose that the external loads do not vary, while in (6.26)
we consider that the displacements have the same property, we can remove the
operator d from under the integral sign, so that we may write, in the static case,
dP ¼ 0; ð6:31Þ
6.1 Principles and General Theorems of the Theory of Elasticity 253

where the potential energy is given by (6.29). This relation represents the principle
of the minimum potential energy; indeed, one obtains for P an extreme value and
one can prove that it corresponds to a minimum.
According to this principle, the real displacements ui ; corresponding to the
deformation of the elastic body, are those for which the potential energy is minimum
(Lagrange’s principle); Wi is expressed here by means of the components of the
displacement vector.
As well, from all the states of stress statically possible (which equilibrate the
external loads), takes place that state of stress for which the potential energy is
minimum (Castigliano’s principle); Wi is expressed here by means of the com-
ponents of the stress tensor.
We mention that, in case of a non-linear relation between stresses and strains,
the expressions (6.230 ) and (6.260 ) of the first variation of Wi will be different. In
this case, the energy obtained by using a virtual variation of the state of stress is
called complementary energy; thus, Castigliano’s principle will be the principle of
the minimum complementary energy. The potential energy and the complementary
energy have the same value only in the case of a linear constitutive law (see also
Sect. 4.1.2.6).
It is convenient to apply the principle of the minimum complementary energy in
case of a solution in stresses of the problems of the theory of elasticity; as well, the
principle of the minimum potential energy is useful in case of a solution in
displacements.

6.1.2.6 Principle of the Minimum Internal Work

We consider the states of strain and stress the variations of which do not affect the
external loads or their corresponding displacements; we have, in this case,
n
dFi ¼ dpi ¼ 0; i ¼ 1; 2; 3; ð6:32Þ

or
dui ¼ 0; i ¼ 1; 2; 3: ð6:320 Þ
Taking into account (6.23) or (6.26), it results
dWi ¼ 0; ð6:33Þ

relation which expresses the principle of the minimum internal work. In this case,
we can assert that among all the states of stress (or of strain) statically corre-
sponding to the given external loads, takes place only that state of stress (or of
strain) which minimizes the internal work (the deformation be produced with a
minimum expense of energy).
As a consequence of the results stated above, we notice as well that the
hyperstatic efforts, to which correspond null displacements, are determined by the
condition that the internal work be minimum for these efforts.
254 6 Principles and General Theorems. Computation Methods

6.1.2.7 Principle of Reciprocity. Betti’s Formulae

Considering two states of strain and stress of an elastic body, marked by (0 ) and (00 ),
corresponding to two distinct systems of external loads, we shall be able to express
the work yielded by one the loads through the agency of the displacements cor-
responding to the other load in the form (one observes that the factor 1/2 does not
appear, because the external loading acts with the whole intensity along the
displacements)
ZZZ ZZ n ZZZ ZZ n
0 00 0 00 00 0
Fi ui dV þ pi ui dS; Fi ui dV þ p00i u0i dS;
V S V S

we notice that one may write, e.g.,


ZZ n ZZ   ZZZ  
0 00 0 00
pi ui dS ¼ rji ui nj dS ¼ r0ji u00i dV
S0 0 ;j
ZZZ S ZZZ V
ZZZ
¼ r0ji u00i dV þ r0ij u00i;j dV ¼  Fi0 u00i dV
V V V
ZZZ   ZZZ ZZZ
00 00 00 0 00
þ rij eij  xij dV ¼  Fi ui dV þ r0ij e0ij dV;
V V V

where we have used the relations (3.200 ), (3.210 ), (3.61) and the Gauss-Ostro-
gradskiı formula (A.950 ). Thus, there result the relations
ZZZ ZZ n ZZZ
Fi0 u00i dV þ p0i u00i dS ¼ r0ij e00ij dV; ð6:34Þ
V S V
ZZZ ZZ n
ZZZ
Fi00 u0i dV þ p00i u0i dS ¼ r00ij e0ij dV: ð6:340 Þ
V S V

Taking into account Hooke’s law (4.56), valid for a linearly elastic body with
linear anisotropy, one has the relation

r0ij e00ij ¼ r00ij e0ij ; ð6:35Þ

so that the right members of the relations (6.34), (6.340 ) are equal. Thus, it results
ZZZ ZZ n ZZZ ZZ n
Fi0 u00i dV þ p0i u00i dS ¼ Fi00 u0i dV þ p00i u0i dS; ð6:36Þ
V S V S

we thus obtain the principle of reciprocity of the work (Betti’s principle [17]) in
the static case. This principle asserts the fact that the work effected by a system of
external loads by corresponding displacements of another system of external loads
is equal to the work effected by the second system of external loads by the
displacements corresponding to the first system.
6.1 Principles and General Theorems of the Theory of Elasticity 255

Replacing the volume forces by the lost forces of d’Alembert, we may write
ZZZ ZZ n
 0 
0 00
Fi  q€
ui ui dV þ p0i u00i dS
V S
ZZZ ZZ n
 00 
¼ Fi  q€ u00i u0i dV þ p00i u0i dS; ð6:37Þ
V S

we get thus the principle of reciprocity of the work in the dynamic case, where the
system of external loads includes the inertia forces too.
D. Graffi [21, 22] showed that, in the case of homogeneous initial conditions, one
can state a theorem of reciprocity in the form (the time variable was brought into
relief, by assuming that t0 ¼ 0)
ZZZ Z t ZZ Z t
0 00 n
dV Fi ðt  sÞui ðsÞds þ dS p0i ðt  sÞu00i ðsÞds
V 0 S 0
ZZZ Z t ZZ Z t
00 0 n
¼ dV Fi ðt  sÞui ðsÞds þ dS p00i ðt  sÞu0i ðsÞds: ð6:38Þ
V 0 S 0

An extension of this theorem of reciprocity to infinite domains was made by


L. T. Wheeler and E. Sternberg [48] in 1968 (the demonstration of Grafii was
confined to finite domains).

6.1.2.8 Maxwell’s Formulae. Coefficients of Influence

Particularly, we can express such a theorem of reciprocity of the work in a finite


form. We assume that the first system of external loads is formed by a concentrated
force Pi dðr  ri Þ acting at the point Ai of position vector ri and producing a
displacement dji at the point Aj of position vector rj ; while the second system of
external loads is formed by a concentrated force Pj dðr  rj Þ; acting at the point Aj
and producing a displacement dji at the point Ai (Fig. 6.2); the forces may be
volume forces or superficial ones. In this case, the relation (6.36) leads to
Pi  dij ¼ Pj  dji ð!Þ; ð6:39Þ

the first index of the displacement shows the point where it is produced, while the
second index shows the point where the external load that yields this displacement
is acting.
If dij is the component of dij along the direction of Pi , while dji is the component
of dji along the direction of Pj , one may write the relation (6.39) also in the form

Pi dij ¼ Pj dji ð!Þ: ð6:390 Þ


In particular, if we take Pi ¼ 1 and Pj ¼ 1; i.e. if we take forces equal to unity,
we get the reciprocity formulae of Maxwell [30]
256 6 Principles and General Theorems. Computation Methods

Fig. 6.2 Theorem of


reciprocity Pi
ji Pj
dij d ji
ij Aj
Ai rj
ri

dij ¼ dji : ð6:40Þ


We notice that one may establish formulae of the form (6.39)–(6.40) for other
concentrated loads too; e.g. to a concentrated rotational moment corresponds a
rotation. Hence, the above forces and displacements have the significance of
generalized quantities. Thus, the relation (6.40) should be considered in the
numerical sense (not dimensional) since one of the quantities can be a length and
the other an angle.
Using the principle of superposition of effects, we may express the displace-
ment along a direction, due to a finite number of concentrated forces Pj ; j ¼
1; 2; . . .; n; in the form
X
n
di ¼ dij Pj ; ð6:41Þ
j¼1

hence, dij are coefficients of influence for the lines of influence of the displacements
in case of movable concentrated forces.
In case of n points at which we calculate the displacements, we may solve the
system (6.41); we obtain
X
n
Pi ¼ cij dj ; ð6:410 Þ
j¼1

relation which is, as well, a consequence of the principle of superposition of


effects. We notice that cij are coefficients of influence for the lines of influence of
the forces in case of movable forces.

6.1.3 Other Considerations

We make, in what follows, some considerations concerning the theorems of


existence and uniqueness; we introduce then Saint-Venant’s principle, particularly
useful in applications.
6.1 Principles and General Theorems of the Theory of Elasticity 257

6.1.3.1 Theorems of Existence and Uniqueness

From the mathematical standpoint, the mathematical problem of the theory of


elasticity is reduced to the integration of a fundamental system of differential
equations, under given limit conditions. We have seen that, from a practical point
of view, this leads to a solution in stresses or to a solution in displacements.
An important question arises: does a solution of the problem exist? To this
question, several theorems answer showing that, under certain, rather general,
conditions, the problem always admits a solution. Such theorems make sure about
the possibility of searching a solution of an elasticity problem.
We notice that one may put certain conditions which are necessary (but not
sufficient!) in this direction. For instance, if the domain occupied by the elastic
solid is bounded and F ¼ 0; then a necessary condition of existence of the solution
is represented by the static equivalence to zero of the superficial forces, expressed
in the form
ZZ ZZ
n n
p dS ¼ 0; r  p dS ¼ 0; ð6:42Þ
S S

obviously, in the case of a non bounded domain these conditions are no more
necessary.
We mention that the impossibility to prove a theorem of existence may arise
because, for the given body and in the given conditions of loading, the considered
equations do not describe accurately the physical phenomenon.
Data about this problem are to be found in the monograph of W. Nowacki [12],
in the monograph of V. D. Kupradzhe, T. G. Gegelya, M. C. Basheleıshvili and T.
V. Burchuladzhe [8] and in the monograph of M. Gurtin [5].
Another question arises about the number of solutions that this problem admits.
In any case, if a solution is found, indifferent on what way, it means that it exists;
but it is important to know if this solution is unique or not. To this, a theorem of
uniqueness due to G. R. Kirchhoff [27], in the static case, which affirms that given
a perfectly elastic body, occupying a finite domain, isotropic, homogeneous, avoid
of given initial stresses, to which can be applied the superposition of effects, on
which acts a system of external loads, which increase continuously from zero to the
maximum value, with corresponding strains and rigid body local rotations neg-
ligible with respect to unity and with no influence whatsoever on the equations of
equilibrium, a single state of strain and stress possible. We mention that the
domain occupied by the body must be simply connected (otherwise, initial stresses
would appear).
We remark that, in the absence of the external loads (superficial loads and
volume forces), the state of strain and stress vanish if these are not initial stresses.
Indeed, from the above hypotheses it results that We ¼ 0; while Clapeyron’s
principle (6.200 ) shows that one must have Wi ¼ 0: But Wi is given by (6.10); from
the formulae (6.12) we see that the elastic potential W is a definite positive
quadratic form, both in stresses and in strains. Consequently, when Wi vanishes,
258 6 Principles and General Theorems. Computation Methods

W vanishes too, but this only occurs simultaneously with the vanishing of all the
components of the strain tensor or concomitantly with the vanishing of all the
components of the stress tensor, which had to be proved; one may have at most a
displacement and a rotation of rigid body. This result is valid also in case of null
displacements on the boundary.
We shall now suppose that to a system of given external loads correspond,
under the same boundary conditions, two states of strain and stress. By subtracting
the two loadings (volume forces and superficial loads, including eventually con-
ditions in displacements on the boundary), as well as the two states of strain and
stress (application of the principle of superposition of effect), we obtain a state of
strain and stress which corresponds to external loads equal to zero (including
boundary conditions equal to zero), hence, a null state of strain and stress. By
virtue of the above conclusions, the two states of strain and stress supposed to take
place will coincide, making abstraction of a global motion of rigid body. Hence, if
the problem of elastostatics has a solution, it will be a single one; the theorem of
uniqueness is thus proved.
We remark that the above result is valid both for the two fundamental problems
as well as for the mixed fundamental problem. The nondetermined motion of rigid
body appears only in case of the second fundamental problem (conditions in
stresses on the boundary); it can be specified by conditions of fixity of the elastic
solid. In the case of the other two problems, this motion is determined by the
conditions on the boundary where displacements does appear.
The theorem remains valid in more general conditions, e.g., in the case of
anisotropic bodies; as well, it is valid for infinite domains, if one puts certain
supplementary conditions of regularity of the solution at infinity. In case of
multiply connected domains one puts, as well, certain supplementary conditions.
The above results are valid as long as the assumptions made in establishing
them are satisfied, especially the hypothesis which states that W is a positive
definite quadratic form. But it is possible that, owing to the instability of the
material (by passing into the plastic domain, by the appearance of yield phe-
nomena etc.), W no longer assumes a positive definite quadratic form. On the other
hand, fundamental changes are liable to occur in the equations of the problem; we
can thus have finite deformations. The theorem of uniqueness is valid as long as
the elastic displacements do not affect the action of the external loads. There are
however cases when these deformations, not negligible any longer, must be taken
into account when writing the equations of equilibrium. In these cases, we cannot
prove that the solution is unique; there can be several possible forms of static
equilibrium for the same system of external loads. This leads us to problems of
elastic instability (eigenvalue problems or branching problems). We can be
equally led to this kind of phenomena in case of the action of non-conservative
external loads, as well as in the case when these loads are functionals of the body
deformation or of the story of this deformation.
We must mention that, in all the cases when the principle of superposition of
effects is applicable, the state of strain and stress, corresponding to the action of
the external loads, is not affected by the eventual existing of initial stresses and can
6.1 Principles and General Theorems of the Theory of Elasticity 259

be computed as in the absence of such stresses. The total stress is obtained, in this
case, by superposition of effects. But, if this principle is not applicable, then the
state of stress due to the external loads can no longer be determined if the initial
conditions are not known from the very beginning. These stresses can be produced
by temperature variations or by many other causes.
In the dynamic case, a theorem of uniqueness due to Fr. Neumann [10], who
states that, given a perfectly elastic body, isotropic, homogeneous, devoid of initial
stresses, to which can be applied the principle of superpositions of effects, on
which a system of external loads acts, admitting given initial conditions, with
corresponding strains and rigid body local rotations negligible with respect to
unity and with no influence whatsoever on the equations of motion, a single state of
strain and stress is possible.
We remark that, in the absence of the external loads (superficial loads and
volume forces) and under homogeneous (vanishing) initial condition, the state of
strain and stress vanish if there are no initial stresses. Indeed, from the above
hypotheses, it follows that We ¼ 0; while the generalized Clapeyron’s principle
(6.20) shows that we must have T þ Wi ¼ 0; since both these quantities are
positive, it follows that we must simultaneously have T ¼ Wi ¼ 0: As in the static
case, from Wi ¼ 0 it result that all the components of both strain and stress tensors
must vanish. Likewise, from the vanishing of the kinetic energy T; it results that
the displacement velocities must vanish. Therefore, the state of strain and stress is
reduced to a rigid body motion that, owing to the homogeneous initial conditions,
vanishes to.
We shall now suppose that to a system of given external loads correspond,
under assumed initial conditions, two states of strain and stress under the same
boundary conditions. By subtracting the two loadings (volume forces, superficial
loads, hence boundary conditions in displacements too), we shall find a state of
strain and stress, corresponding to external loads equal to zero (including boundary
conditions equal to zero) and to initial conditions equal to zero too, therefore—by
virtue of the above conclusions—to a state of strain and stress equal to zero; the
above two states of strain and stress will coincide, while if the problem of elas-
todynamics has a solution, then it will be a single one and the theorem of
uniqueness is proved.
In the dynamic case, one can make considerations analogous to those made in
the static case.
L. T. Wheeler and E. Sternberg [48] extend the theorem of uniqueness, dem-
onstrated in the case of some finite domains, to infinite domains. The demon-
stration of this result is based on the generalization of a theorem of energetical
nature, given by S. Zaremba [49] for the scalar waves equation, independently
rediscovered by A. Rabinovich [31], discussed later on by Fritz John and resumed
by K. O. Friedrichs and H. Lewy [20] and by R. Courant [3, 4]. Similar results
were supplied, in the anisotropic case, by L. T. Wheeler [47].
With the results given by M. E. Gurtin and E. Sternberg [24] in the case of finite
domains, it can be shown that the theorem of uniqueness for the second funda-
mental problem remains equally valid in the case of infinite domains, even when
260 6 Principles and General Theorems. Computation Methods

the two wave propagation velocities, which have real values, are no more related
by (5.5). We mention moreover that the results given in [24] were extended by
M. E. Gurtin and R. A. Toupin [25] to anisotropic bodies. R. J. Knops and
L. E. Payne [28] presented moreover, in the case of finite domains, a theorem of
uniqueness for the weak solutions in elastodynamics.

6.1.3.2 Saint-Venant’s Principle

To solve the chief problems raised by the practice, a particularly important prin-
ciple is usually applied: the principle of Barré de Saint-Venant [32], stated by him
in 1855.
According to Saint-Venant, this principle is stated as follows: If upon an elastic
body acts a system of external loads in static equilibrium, the state of strain and
stress within in is practically equal to zero, except in a zone of the order of
magnitude of the boundary on which these loads are acting (Fig. 6.3)
We shall bring into relief the following corollary, practically important in
practical applications: The state of strain and stress within an elastic body, that
admits the principle of superposition of the effects, are practically independent
from the manner of applying the external loads, excepting in the zone neigh-
bouring the boundary on which these loads are applied (Fig. 6.4a, b, c).
We must mention that this principle can only be applied when the strain and the
rigid body local rotations can be neglected with respect to unity; for instance, it is
inapplicable, in the above mentioned form, to bodies with thin walls.
Saint-Venant’s principle is widely applied in practice. It allows, for instance, to
schematize the actual loading mode of the elements of construction, giving a
mathematical formulation of the problems raised; this principle allows us to obtain
most of the state of strain and stress of the body. So as to find the state of strain and
stress in the zones where the external loads are applied, the study must be com-
pleted by a local one, that helps to determine the local states of strain and stress
born there. Such a problem takes the name of contact problem and involves special
methods of approximating the physical phenomenon.
The above formulation of the principle of Saint-Venant has the disadvantage of
an insufficient accuracy and of a rather empirical character. Later on, justifications
of a theoretical nature of this principle were searched for and endeavours were

Fig. 6.3 Saint-Venant’s


principle
6.1 Principles and General Theorems of the Theory of Elasticity 261

(a) P (b) P (c)


p p

Fig. 6.4 Corollary to Saint-Venant’s principle

made of more accurate formulations (with a more accentuated mathematical


character).
Let thus be an elastic body acted upon by a system of selfequilibrating external
loads (the resultant force and the resultant moment vanish), all in the interior of a
sphere S. We make a section R; in the exterior of the sphere S; the body is thus
divided into two parts. We consider now that the two parts are acted upon only by
the stresses which appear on the two faces of the section R; due to the external
loads contained in the interior of the sphere S. One assumes that a convenient
measure of the intensity of those stresses may be the internal work induced by
them in the two parts of the body (without the internal work corresponding to the
external loads which act upon the interior of the sphere S). If R0 and R00 are two
sections in the body, which have not common points, both in the exterior of the
sphere S; the first one being nearer to the sphere than the second one, then one can
state, after Zanaboni [49], that

Wi ðR00 Þ\Wi ðR0 Þ: ð6:43Þ


To Boussinesq and Mises we also owe considerations concerning the applica-
tion of Saint-Venant’s principle. Thus, let be un elastic half-space x3  0; acted by
normal or by tangential loads upon the separation plane x3 ¼ 0; assuming that
these loads act in the interior of a circle C of diameter g; it has been shown that the
greatest component of the stress tensor at a point situated at the distance R from the
centre of the circle C is:
• of the order of magnitude r0 ¼ F=R2 if the resultant force is of the order of
magnitude F;
• of the order of magnitude ðg=RÞr0 in case of a resultant couple, obtained by
means of forces of order of magnitude F;
• of the order of magnitude ðg=RÞr0 in case of a dipol of forces (selfequilibrated
system), obtained by means of forces of order of magnitude F;
• of the order of magnitude ðg=RÞ2 r0 in case of a selfequilibrated system in astatic
equilibrium, obtained be means of forces or order of magnitude F.
A selfequilibrated system in astatic equilibrium is a system which remains
selfequilibrated if all its forces are subjected to a rotation of the same arbitrary
n
angle. For the external loads p which act on an element of area of external normal
n, on a surface of area A, the conditions of astatical equilibrium are of the form
262 6 Principles and General Theorems. Computation Methods

ZZ ZZ
n n
pi dA ¼ 0; xj pi dA ¼ 0; i; j ¼ 1; 2; 3: ð6:44Þ
S S

According to Sternberg, in case of an elastic solid acted upon by external loads


contained in the interior of a sphere of diameter g; the greatest component of the
tensor Tr is of the order magnitude Oðeq Þ; where
• q  2 for a nonzero resultant force (one assumes that the external loads are finite
and that the resultant force tends to zero as g2 ; because the area upon which act
these loads tends thus to zero);
• q  3 for a system of external loads which are selfequilibrated or are reduced to
a couple;
• q  4 for a system of external loads which are selfequilibrated, being in astatic
equilibrium.

6.1.4 Simply Connected Domains. Multiply Connected


Domains

Generally, by a simply connected domain we mean a domain without internal


holes; hereafter, we shall give a rigorous definition of the notion. The domains that
do not posses this property will obviously be multiply connected domains. The
material bodies occupying these domains can or cannot have internal holes.
Consequently, a study of the connection of a domain proves to be necessary.

6.1.4.1 Properties

We shall call a simply connected domain that one for which each closed surface
R in its interior (which has not common points with the boundary) can be reduced
to a point, by a continuous deformation. In the three-dimensional case, e.g., one
may define another type of connection, replacing the closed surface R by a closed
curve C. Thus, e.g., the domain contained between two concentric spheres is
simply connected according to the second definition but is doubly connected
according to the first one; in the following discussion we shall confine ourselves to
the first of the above given definitions. In the two-dimensional case, the simply
connection is obviously defined by means of a closed curve C.
Let now be a domain D, bounded by a surface S and having n  1 closed
surfaces S1 ; S2 ; . . .; Sn1 ; within the domain and outside of one another; we
assume, obviously, that these surfaces do not posses multiple points. Let equally
be the closed surface R1 within the domain D and enclosing within it the surface
S1 ; we remark that the surface R1 can be reduced to the sole surface S1 ; by
continuous deformation, contrary to the closed surface R; that can be reduced to a
point (Fig. 6.5a; for sake of simplicity, it will correspond to the plane case). This
property of the multiply connected domains is a fundamental one and helps to
define them; in our case, the domain is n-connected.
6.1 Principles and General Theorems of the Theory of Elasticity 263

(a) (b)
S1 1
S1 D
S2 S n-1
S2 S n-1
D
S S

(c) (d)
S1
S1 S n-1 c1 S n-1

D c2 D
S2 S2 S
S

Fig. 6.5 Multiply connected domain (a). Cuts: reduction to a simply connected domain (b, c),
equivalent cuts (d)

One important problem is to transform a multiply connected domain D into a


simply connected domain. For this end, we shall link the n surfaces bounding the
domain by double surfaces, so that the boundary becomes unique and continuous.
This is possible by introducing n  1 double surfaces or cuts, as in Fig. 6.5b or as
in c. Confining to the two-dimensional case (that of Fig. 6.5), we maintain the
counter-clockwise sense of description along the boundary, which leaves the
domain to the left. Two such cuts t1 and t2 are equivalent if they coincide as a
consequence of a continuous deformation (Fig. 6.5d).
We say that a domain is multiply connected of nth order if, by n  1 cuts, one
can transform it in a simply connected domain. An n-connected domain contains
n  1 internal holes. Such domains are often to be found in practice.
Let be a function F ¼ Fðx1 ; x2 ; x3 ; tÞ single-valued and continuous within the
simply connected domain D, except in n  1 isolated points. If these points, called
singular, are surrounded by the closed surfaces S1 ; S2 ; . . .; Sn1 , we shall obtain an
n-connected domain, within which the function, considered as a function of the
space variables, is single-valued and continuous. This is, e.g., the case of the
bodies acted upon by internal concentrated loads.

6.1.4.2 Multiply Connected Bodies. Distorsions

If a body occupies a multiply connected domain, we shall say that it is a multiply


connected body. If in such a body, acted upon by external loads, we effect several
cuts that transform it into a simply connected body, the components of the dis-
placement vector (and of the rigid body local rotation vector) on one part and on
the other part of the cut will no more be the same; the two faces of the cut cðkÞ will
displace one another, the displacement being composed by an infinitesimal
translation of vector uðkÞ and an infinitesimal rotation of vector xðkÞ . The dis-
ðk Þ ðkÞ
placement vectors on the two faces of the cut cðkÞ will be uþ and u and the rigid
ðkÞ ðkÞ
body local rotation vectors will be denoted by xþ and x (Fig. 6.6).
264 6 Principles and General Theorems. Computation Methods

Fig. 6.6 Distorsions and


initial stresses on a cut
c(k)
(k)
p
(k) c (k)
u-
(k)
p

(k)
u+

The strains will be continuous and single-valued functions, since the contri-
bution of the difference of displacements between one part and the other of the cut
vanishes. Taking into account the formulae (2.71)–(2.71000 ) which emphasize the
rigid body local motion, we can write
ðkÞ ðkÞ
uþ  u ¼ uðkÞ þ xðkÞ  rðkÞ ; k ¼ 1; 2; . . .; n  1; ð6:45Þ
ðkÞ ðkÞ
xþ  x ¼ xðkÞ ; k ¼ 1; 2; . . .; n  1; ð6:46Þ

where rðkÞ is the position vector of a point of the cut; in components, one may write
ðkÞ ðkÞ ðkÞ ðkÞ
uþj  uj ¼ uj  jlm xl xðkÞ
m ; j ¼ 1; 2; 3; k ¼ 1; 2; . . .; n  1; ð6:450 Þ
ðkÞ ðkÞ ðkÞ
xþj  xj ¼ xj ; j ¼ 1; 2; 3; k ¼ 1; 2; . . .; n  1: ð6:460 Þ
ðkÞ ðkÞ
The constant vectors uðkÞ and xðkÞ and the constant scalars uj and xj ; j ¼
1; 2; 3; respectively, are characteristic for every cut ck ; k ¼ 1; 2; . . .; n  1; being
the same for each point of a cut, as well as for two equivalent cuts.
Inversely, if, in the absence of the external loads, in a multiply connected body,
in which the necessary number of cuts is performed so as to transform it into a
simply connected body, we displace the faces of each cut by a translation uðkÞ and
a rotation xðkÞ and then suppress the cuts by sticking their faces in their new
positions after the displacement, then a state of stress will be born due exclusively
to these displacements. These displacements are called distorsions, while the
stresses they yield are called initial stresses.
In this case, the theorems of uniqueness can no more be applied. A multiply
connected elastic body, under no external loads, is not necessarily in a natural
state. Besides, we can affirm that the state of strain and stress of a multiply
connected elastic body is determined by the external loads and by the six com-
ponents of the distorsion with regard to each of the cuts transforming the multiply
connected body into a simply connected one. Since the components of the dis-
torsion can be arbitrarily chosen, this leads to a multiple-valued solution.
6.1 Principles and General Theorems of the Theory of Elasticity 265

A thorough study of these problems was performed, in the static case, at the
very beginning of twentieth century century by Vito Volterra [45, 46]. He showed
that, to an arbitrary distribution of distorsions, corresponds a unique state of strain
in a multiply connected body undergoing the action of external loads. This
theorem of uniqueness corresponds to the theorems of Kirchhoff and Neumann for
simply connected bodies in the static and in the dynamic case, respectively.
Results in this direction are to be found equally in the volume [16] published by
E. Volterra in 1960.
To compute the internal work in case of a multiply connected body, to the
expression (6.10) of the internal work of a simply connected domain we must add
an additional term in the form
ZZ
1Xn1
ðkÞ
Wi ¼ ðuþ  uðkÞ ðkÞ
 Þ  p dS
ðkÞ
2 k¼1 SðkÞ
ZZ
1Xn1
ðkÞ ðkÞ ðkÞ
¼ ðuþj  uj Þ pj dSðkÞ ; ð6:47Þ
2 k¼1 SðkÞ

where SðkÞ is the area of the corresponding cut. Introducing the torsor of the stress
vectors which appear on the cut cðkÞ
ZZ
PðkÞ ¼ pðkÞ dSðkÞ ; k ¼ 1; 2; . . .; n  1; ð6:48Þ
SðkÞ
ZZ
MðkÞ ¼ rðkÞ  pðkÞ dSðkÞ ; k ¼ 1; 2; . . .; n  1; ð6:49Þ
SðkÞ

and, taking into account that in a scalar triple product one can interchange the
scalar and the vector products, we may write the additional term in the form

1Xn1
Wi ¼ uðkÞ  PðkÞ þ xðkÞ  MðkÞ ; ð6:470 Þ
2 k¼1

if
ZZ
ðkÞ ðkÞ
Pj ¼ pj dSðkÞ ; j ¼ 1; 2; 3; k ¼ 1; 2; . . .; n  1; ð6:480 Þ
SðkÞ
ZZ
ðkÞ ðkÞ
Mj ¼ jlm xl pðkÞ ðkÞ
m dS ; j ¼ 1; 2; 3; k ¼ 1; 2; . . .; n  1; ð6:490 Þ
SðkÞ

then one may also write


n1  
1X ðkÞ ðkÞ ðkÞ ðkÞ
Wi ¼ uj Pj þ xj Mj : ð6:470 Þ
2 k¼1
266 6 Principles and General Theorems. Computation Methods

This is the expression of the internal work due solely to distorsions.


To solve the problems of the theory of elasticity in case of a multiply connected
body, we must assume some additional conditions of uniqueness of the dis-
placements; in other words, we shall begin by solving the problem as in the case of
a simply connected domain and afterwords we shall introduce the influence of the
effected cuts (therefore, of the introduced distorsions).
What concerns the continuity conditions of deformations of Saint-Venant
(2.68)–(2.68000 ), these are now only necessary conditions; they must be completed
by certain numerical conditions corresponding to each internal frontier.

6.2 Computation Methods

In what follows; we shall deal with particular integrals, fundamental formulae, as


well as with the most important methods of computation, i.e.: successive
approximations, variational methods, methods of the theory of distributions etc.

6.2.1 Particular Integrals

We have seen that, both in case of a solution in displacements and of a solution in


stresses, we are led to the search of certain potential functions which must verify a
harmonic or a biharmonic equation or a simple or double waves equation. It is thus
useful to put in evidence some simple properties of these functions as well as
particular integrals of the corresponding equations.

6.2.1.1 Properties of Symmetry and Antisymmetry

In the case of a body admitting a plane of geometrical and mechanical symmetry


(from the points of view of the mechanical properties of the material and from the
supporting standpoint), considerable computation simplifications can be made; the
plane of symmetry is considered as plane of co-ordinates (the Ox1 x2 -plane), any
case of loading can be decomposed into two cases: one that is symmetrical with
respect to this plane and the other antisymmetrical with respect to the same plane;
the same break down is applied to the boundary and initial conditions. The state of
strain and stress must posses the same properties.
In the case of a symmetry with respect to the Oxy-plane, the normal stress
r11 ; r22 ; r33 as well as the tangential stress r12 ¼ r21 are even functions with
respect to x3 ; while the tangential stresses r23 ¼ r32 ; r31 ¼ r13 are odd functions
with respect to the same variable; the displacements u1 and u2 are even function
with regard to x3 , while u3 is an odd one. In the static case, e.g., the displacement
functions C1 and C2 ; which appear in Galerkin’s representation, will be even with
respect to x3 ; while C3 will be odd with respect to the same variable; in the
6.2 Computation Methods 267

Papkovich-Neuber representation, the displacement functions v1 and v2 are, ana-


logically, even with regard to x3 ; while the function v3 is odd with respect to this
variable. In case of a solution in stresses, we mention Schaefer’s representation for
the static problem; thus, the stress functions H11 ; H22 ; H33 ; H12 and X must be
even with respect to x3 ; while the stress functions H23 and H31 must be odd with
respect to this variable. In case of an antisymmetry with respect to Oxy-plane, one
has an inverse parity. If the displacement or the stress functions are chosen by
taking these properties into account, it shall be sufficient to put the boundary
conditions on a single side of the contour; on the symmetrical part with it, they will
be concomitantly fulfilled.
Besides, if the body admits two or three planes of symmetry, these planes will
be chosen as planes of co-ordinates. Any loading case can be decomposed in eight
cases, according to their properties of symmetry or antisymmetry with respect to
the three planes. The problems can be studied separately, for each of the com-
ponent loading, the final result being then obtained by superposition of the effects;
obviously, each of these particular cases is much more simple to study.
If, in a plane problem, we have an axis of geometrical and mechanical sym-
metry, then we can make similar considerations. On the other hand, in case of an
axially symmetrical problem, we can consider a plane of geometrical and
mechanical symmetry, normal to this axis, and proceed to a similar study.

6.2.1.2 Particular Integrals for the Harmonic and for the Biharmonic
Equations

The harmonic and biharmonic equations, defined in Sect. A.1.2.4, play an


important rôle in elastostatics. In Sect. A.1.2.7 one presents Almansi’s relations,
which allow to build a biharmonic function by means of two harmonic ones; thus,
we shall try—especially—to obtain particular integrals for Laplace’s equation
DUðx1 ; x2 ; x3 Þ ¼ 0; ð6:50Þ

putting some supplementary conditions.


Thus, if we admit that

Uðx1 ; x2 ; x3 Þ ¼ Uðx21 þ x22 þ x23 Þ ¼ UðuÞ; ð6:51Þ

we are led to
 
d dUðuÞ
DUðuÞ ¼ 2 UðuÞ þ 2u ¼ 0;
du du
pffiffiffi
we get the integrals 1= u and const, hence the harmonic functions
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1
; const; R ¼ x21 þ x22 þ x23 6¼ 0; ð6:52Þ
R
268 6 Principles and General Theorems. Computation Methods

where R is the polar radius. These integrals are useful in case of a spherical
symmetry of pole O. By using Almansi’s formulae, we get also the biharmonic
functions
xi
R; R2 ; ; i ¼ 1; 2; 3; R 6¼ 0: ð6:53Þ
R
The condition R 6¼ 0 is essential; indeed, if we have to do with distributions, the
results in Sect. A.3.4.3 allow to write

1
D ¼  4pdðRÞ; DDR ¼ 8pdðRÞ: ð6:54Þ
R
Assuming other relations between the three variables, one may obtain, ana-
logically, other particular integrals.
A particularly fruitful method is that of separation of variables; we may thus
choose a function of the form
Uðx1 ; x2 ; x3 Þ ¼ uðx1 ; x2 Þwðx3 Þ; ð6:55Þ
obtaining

d2 wðx3 Þ
DUðx1 ; x2 ; x3 Þ ¼ wðx3 ÞDuðx1 ; x2 Þ þ uðx1 ; x2 Þ ¼ 0;
dx23

which leads to the equations

d2 wðx3 Þ
 k2 wðx3 Þ ¼ 0;
dx23
ðD k2 Þuðx1 ; x2 Þ ¼ 0;

k being an arbitrary constant. The first equation leads to trigonometric or hyper-


bolic functions, while the second equation leads to metaharmonic functions. If we
proceed with the second equation as with the first one, we get particular integrals
of the form
eai x i ; ð6:56Þ

where the constants ai ; i ¼ 1; 2; 3; must verify the condition


ai ai ¼ 0; ð6:560 Þ
these integrals may be of one of the forms

sin a1 x1 sin a2 x2 ea3 x3 ; sin a1 x1 sin a2 x2 sinh a3 x3 ; ð6:57Þ

ea1 x1 ea2 x2 sin a3 x3 ; sinh a1 x1 sinh a2 x2 sin a3 x3 ;


ð6:570 Þ
ea1 x1 sinh a2 x2 sin a3 x3 ;
6.2 Computation Methods 269

where the constant ai ; i ¼ 1; 2; 3; are linked by the relation


qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
a3 ¼ a21 þ a22 : ð6:570 Þ

Other harmonic functions are obtained by replacing some lines ‘‘sin’’ by ‘‘cos’’
or ‘‘sinh’’ by ‘‘cosh’’ or interchanging the rôle of the variables.
By means of the formula (A.1.100) of Almansi, one may then obtain particu-
larly useful biharmonic functions. Starting from these results, one can construct
Fourier representations (Fourier series or Fourier integrals), important for many
boundary value problems; the properties of evenness with respect to certain
variables are easily put in evidence.
Another method to obtain particular harmonic or biharmonic integrals consists
in the differentiation of the already obtained integrals with respect to one of the
variables; thus, differentiating 1=R; we get the harmonic functions
xi
; i ¼ 1; 2; 3; R 6¼ 0: ð6:58Þ
R2
Taking into account (6.54), in case of a distribution, one has
x  o
i
D 2 ¼ 4p dðRÞ; i ¼ 1; 2; 3: ð6:580 Þ
R oxi
Because the harmonic and the biharmonic equations are linear partial differ-
ential equations, one may apply the principle of superposition of the effects; any
linear combination of particular integrals will be an integral of the respective
equation. Thus, one obtains, e.g. the Fourier representations mentioned above.

6.2.1.3 Harmonic and Biharmonic Polynomials

We shall pay a particular attention, in what follows, to the harmonic and bihar-
monic polynomials, due to their importance in various applications.
One can easily verify that following monomials are harmonic functions
const; x1 ; x2 ; x3 ; x2 x3 ; x3 x1 ; x1 x2 ; x1 x2 x3 ð6:59Þ
being the only monomials which have this property.
In general, let be a homogeneous polynomial of nth degree
XXX
Pn ðx1 ; x2 ; x3 Þ ¼ Ai1 i2 i3 xi11 xi22 xi33 ; i1 þ i2 þ i3 ¼ n; ð6:60Þ
i1 i2 i3

where all the indices are nonnegative. We notice that for i1 fix one has i2 þ i3 ¼
n  i1 corresponding n  i1 þ 1 different coefficients; making i1 to take all pos-
sible values, we find the total number of the coefficients to be determined, i.e.
270 6 Principles and General Theorems. Computation Methods

1
ðn þ 1Þ þ n þ ðn  1Þ þ    þ 2 þ 1 ¼ ðn þ 2Þðn þ 1Þ: ð6:61Þ
2
The condition of harmonicity leads to
XXX
Ai1 i2 i3 i1 ði1  1Þx1i1 2 xi22 xi33 þ i2 ði2  1Þxi11 xi22 2 xi33
i1 i2 i3

þi3 ði3  1Þxi11 xi22 x3i3 2 ¼ 0;

making a change of indices for the second and the third term in the square
brackets, one may write
XXX
½i1 ði1  1ÞAi1 i2 i3 þ ði2 þ 2Þði2 þ 1ÞAi1 2;i2 þ2;i3
i1 i2 i3

þði3 þ 2Þði3 þ 1ÞAi1 2;i2 ;i3 þ2 x1i1 2 xi22 xi33 ¼ 0:

This polynomial is identically null if all its coefficients are identically null. By a
change of indices, we find that between the ðn þ 2Þðn þ 1Þ=2 coefficients must
take place the relations

ði1 þ 2Þði1 þ 1ÞAi1 þ2;i2 ;i3 þ ði2 þ 2Þði2 þ 1ÞAi1 ;i2 þ2;i3
þði3 þ 2Þði3 þ 1ÞAi1 ;i2 ;i3 þ2 ¼ 0; ð6:62Þ

where i1 þ i2 þ i3 ¼ n  2; hence nðn  1Þ=2 relations.


A homogeneous harmonic polynomial of nth degree, in three variables, will
thus depend on ðn þ 2Þðn þ 1Þ=2  nðn  1Þ=2 ¼ 2n þ 1 arbitrary constants,
hence there exist 2n þ 1 linearly independent homogeneous harmonic polynomials
of nth degree, in three variables.
One can choose as arbitrary constants the coefficients
A0;0;n ; A0;1;n1 ; . . .; A0;n1;1 ; A0;n;0 ; ð6:63Þ

A1;0;n1 ; A1;1;n2 ; . . .; A0;n2;1 ; A1;n1;0 ; ð6:630 Þ

there are n þ 1 coefficients on the first line and n coefficients on the second line.
The other coefficients Ai1 i2 i3 (for the other values of the index i1 ) may be obtained
as functions of the above ones by means of the recurrence relation (6.62).
We find thus that, for an even index i1 , the coefficients are given by
X
q
Clq 2ðq1Þ
A2q;j;k ¼ ð1Þq Clþ2ðq1Þ C2l
kþ2l A0;jþ2ðqlÞ;kþ2l ;
l¼0 C2l
2q

2q þ j þ k ¼ n; ð6:64Þ

which may be verified by complete induction; here Crs is the symbol of the
combination of s things, r at a time. Analogically, for an odd index i1 one has
6.2 Computation Methods 271

ð1Þq X Cq 2ðqlÞ 2l
q l
A2qþ1;j;k ¼ Cjþ2ðqlÞ Ckþ2l A1;jþ2ðqlÞ;kþ2l ;
2q þ 1 l¼0 C2l
2q

2q þ j þ k ¼ n  1; ð6:640 Þ
If n is an even number, then the polynomial Pn can be even with respect to all
three variables, where one may take the arbitrary constants A0;2r;n2r ; r ¼
0; 1; . . .; n=2; which are n=2 þ 1 ones, or may be even with respect to one variable
and odd with respect to the other variables, choosing thus the arbitrary constants
A0;2rþ1;n2r1 ; A1;2r;n2r1 ; A1;2rþ1;n2r2 ; r ¼ 0; 1; . . .; n=2  1; being n=2 of
each one. Analogically, if n is an odd number, then the polynomial Pn can be odd
with respect to all three variables and one can choose the arbitrary constants
A1;2rþ1;n2r2 ; r ¼ 0; 1; . . .; ðn  3Þ=2; which are ðn  1Þ=2 ones, or may be odd
only with respect to one variable and even with respect to the other variables,
choosing the arbitrary constants A1;2r;n2r1 ; A0;2rþ1;n2r1 ; A0;2r;n2r ; r ¼
0; 1; . . .; ðn  1Þ=2; being ðn þ 1Þ=2 of each one.
One can construct the 2n þ 1 linearly independent harmonic polynomials by
means of the formulae (6.64), (6.640 ). Thus, an even polynomial with respect to the
three variables, corresponding to the coefficient A0;2r;n2r ; reads
n=2 X
X n=2
Csþqr
q 2ðrsÞ 2ðsþqrÞ n2ðqþsÞ
Pn ðx1 ; x2 ; x3 Þ ¼ ð1Þq 2ðsþqrÞ
C2r Cn2ðqþsÞ x2q 2s
1 x2 x3 ; ð6:65Þ
q s C2q

where q þ s
n=2; one may have r ¼ 0; 1; . . .; n=2. The other polynomials may be
expressed analogically.
Starting from the above results, one may easily obtain homogeneous bihar-
monic polynomials.
Thus, starting from the harmonic polynomials (6.59) and using Almansi’s
representation, we obtain the biharmonic monomials

x21 ; x22 ; x23 ; x31 ; x32 ; x33 ; x1 x22 ; x1 x23 ; x2 x23 ; x2 x21 ; x3 x21 ; x3 x22 ;
x1 x32 ; x1 x33 ; x2 x33 ; x2 x31 ; x3 x31 ; x3 x32 ; x21 x2 x3 ; x22 x3 x1 ; x23 x1 x2 ;
x31 x2 x3 ; x32 x3 x1 x33 x1 x2 : ð6:66Þ
Using the formula (A.1.100), one may express a homogeneous biharmonic
polynomial of nth degree in the form

Qn ðx1 ; x2 ; x3 Þ ¼ Pn ðx1 ; x2 ; x3 Þ þ ðx21 þ x22 þ x23 ÞPn2 ðx1 ; x2 ; x3 Þ; ð6:67Þ

where Pn ; Pn2 are homogeneous harmonic polynomials; thus, one can state that a
homogeneous biharmonic polynomial of degree n [ 1; in three variables, depends
on ð2n þ 1Þ þ 2ðn  2Þ þ 1 ¼ 2ð2n  1Þ arbitrary constants, which one may
choose on the same considerations as in case of harmonic polynomials.
If the homogeneous biharmonic polynomial Qn has a certain evenness with
respect to the variables x1 ; x2 ; x3 ; then the constituent harmonic polynomials have
272 6 Principles and General Theorems. Computation Methods

the same evenness. Based on this observation and proceeding as above, one may
state following results: If n is an even number, then the polynomial Qn may be
decomposed in an even polynomial with respect to all three variables, which
depends on n þ 1 arbitrary constants, and three polynomials even only with respect
of one variable and odd with respect to the other variables, which depend each one
on n  1 arbitrary constants. As well, if n is an odd number, then the polynomial Q
is decomposed in an odd polynomial with respect to all three variables, which
depends on n  2 arbitrary constants, and three polynomials odd with respect to
one variable and even with respect to the other variables, which depend each one
on n arbitrary constants.
The properties of evenness or oddness with respect to the variables which occur
are particularly important in the case in which the body admits planes with
properties of geometrical and mechanical symmetry.

6.2.1.4 Fourier Representations

In the case of many computation methods, it is useful to can approximate the


external loads (superficial or volume ones) by means of certain Fourier repre-
sentations (series or integrals); this makes easier the putting of boundary condi-
tions. In case of superficial loads, one uses double series or integrals, while in case
of volume loads one uses triple series or integrals.
In the case of some superficial loads, there are useful the double Fourier series.
We assume to have a plane contour of rectangular form, of dimensions L1 and L2
and of equation x3 ¼ const, on which acts a normal load p ¼ pðx1 ; x2 Þ. The
following considerations can be used for a curvilinear contour (in curvilinear
co-ordinates) too, or for a tangential load; as well, they may be used for dis-
placements imposed on the contour.
A function pðx1 ; x2 Þ may be represented by a double Fourier series if the suf-
ficient conditions of Lejeune-Dirichlet are fulfilled, i.e.: the function must be
piecewise continuous, have a finite number of maxima and minima on the rect-
angular interval mentioned above and be periodical
pðx1 þ L1 ; x2 Þ ¼ pðx1 ; x2 þ L2 Þ ¼ pðx1 ; x2 Þ: ð6:68Þ
The condition of piecewise continuity is fulfilled by usual loads, which appear
in practice. But in case of concentrated loads, this condition is no more fulfilled;
introducing the methods of the distributions theory, one may use the results given
in Sect. A.3.3.1 for the Fourier transforms. Formally, one can make a computation
for a uniform distributed load p on a rectangle of dimensions 2c and 2d, obtaining
then, by a formal process passing to the limit,
P ¼ p!1
lim 4pcd; ð6:69Þ
c!0
d!0
6.2 Computation Methods 273

the representation corresponding to a concentrated force P. This representation is


not convergent. But one can make formal operations with it, the results obtained
for the state of strain and stress in the interior of the body being correct; this may
be explained because the formal passing to the limit can be made on the final
results. From the point of view of the practical computation it is more convenient
to use the formal representations considered above.
In the mentioned conditions, one may represent the function pðx1 ; x2 Þ in the
form
X X
pðx1 ; x2 Þ ¼ d00 þ bl0 sin a1l x1 þ dl0 cos a1l x1
l l
X X
þ c0m sin a2m x2 þ d0m cos a2m x2
m m
XX
þ alm sin a1l x1 sin a2m x2
l m
XX
þ blm sin a1l x1 cos a2m x2
l m
XX
þ clm cos a1l x1 sin a2m x2
l m
XX
þ dlm cos a1l x1 cos a2m x2 ; ð6:70Þ
l m

where
2pl 2pm
a1l ¼ ; a2m ¼ ; l:m ¼ 1; 2; 3; . . .; ð6:71Þ
L1 L2
L1 and L2 being the lengths of the periods on the two directions. The unknown
parameters are given by
Z Z
1
d00 ¼ pðn1 ; n2 Þdn1 dn2 ; ð6:72Þ
L1 L2 L1 L2
Z Z
2
bl0 ¼ pðn1 ; n2 Þ sin a1l n1 dn1 dn2 ;
L1 L2 L1 L2
Z Z
2
dl0 ¼ pðn1 ; n2 Þ cos a1l n1 dn1 dn2 ;
L1 L2 L1 L2
Z Z ð6:720 Þ
2
c0m ¼ pðn1 ; n2 Þ sin a2m n2 dn1 dn2 ;
L1 L2 L1 L2
Z Z
2
d0m ¼ pðn1 ; n2 Þ cos a2m n2 dn1 dn2 ;
L1 L2 L1 L2
274 6 Principles and General Theorems. Computation Methods

Z Z
4
alm ¼ pðn1 ; n2 Þ sin a1l n1 sin a2m n2 dn1 dn2 ;
L1 L2 L1 L2
Z Z
4
blm ¼ pðn1 ; n2 Þ sin a1l n1 cos a2m n2 dn1 dn2 ;
L1 L2 L1 L2
Z Z ð6:720 Þ
4
clm ¼ pðn1 ; n2 Þ cos a1l n1 sin a2m n2 dn1 dn2 ;
L1 L2 L1 L2
Z Z
4
dlm ¼ pðn1 ; n2 Þ cos a1l n1 cos a2m n2 dn1 dn2 :
L1 L2 L1 L2
The free term d00 represents the mean loading on a rectangular surface of
dimensions equal to the periods L1 and L2 .
At a point of discontinuity, the sum Sðx1 ; x2 Þof the series represents the
arithmetic mean of the two limits at the right and at the left, corresponding to the
two variables,
1
Sðx1 ; x2 Þ ¼ ½pðx1  0; x2  0Þ þ pðx1  0; x2 þ 0Þ
4
þ pðx1 þ 0; x2  0Þ þ pðx1 þ 0; x2 þ 0Þ: ð6:73Þ

In the case of loadings with properties of symmetry or antisymmetry with


respect to the co-ordinate axes, the above representations have important simpli-
fications. For instance, in case of a loading skewsymmetric with respect to both
axes of co-ordinates (pðx1 ; x2 Þ ¼ pðx1 ; x2 Þ ¼ pðx1 ; x2 Þ) we use an expansion
into series odd with respect to both variables
XX
pðx1 ; x2 Þ ¼ alm sin a1l x1 sin a2m x2 ; ð6:74Þ
l m

with
Z L1 =2 Z L2 =2
16
alm ¼ pðn1 ; n2 Þ sin a1l n1 sin a2m n2 dn1 dn2 ; ð6:740 Þ
L1 L2 0 0

Analogically, one may use triple Fourier series, useful for periodic volume
loads (of periods L1 ; L2 ; L3 along the three directions). A volume loading F ¼
Fðx1 ; x2 ; x3 Þ; with properties of antisymmetry with the three planes of co-ordinates,
e.g., may be represented in the form
XXX
Fðx1 ; x2 ; x3 Þ ¼ almn sin a1l x1 sin a2m x2 sin a3m x3 ; ð6:75Þ
l m n

where
6.2 Computation Methods 275

Z L1 =2 Z L2 =2 Z L3 =2
64
alm ¼ Fðn1 ; n2 ; n3 Þ
L1 L2 L3 0 0 0
 sin a1l n1 sin a2m n2 sin a3n n3 dn1 dn2 dn3 ; ð6:750 Þ

using the notations (6.71), as well as


2np
a3n ¼ ; n ¼ 1; 2; 3; . . . ð6:710 Þ
L3
On can write the representation (6.70) in a complex form
X
1 X
1
pðx1 ; x2 Þ ¼ klm eiða1l x1 þa2m x2 Þ ð6:76Þ
l¼1 m¼1

too, where
Z Z
1
klm ¼ pðn1 ; n2 Þeiða1l n1 þa2m n2 Þ dn1 dn2 ; ð6:760 Þ
L1 L2 L1 L2

the indices l and m may take any positive or negative entire value.
Analogically, for local loadings one uses double Fourier integrals
Z 1Z 1Z 1Z 1
1
pðx1 ; x2 Þ ¼ 2 pðn1 ; n2 Þ
4p 1 1 1 1 ð6:77Þ
 ei½ða1 ðn1 x1 Þþa2 ðn2 x2 Þ da1 da2 dn1 dn2 :

Passing to trigonometric lines and noting that an integral between symmetric limits
of an odd function vanishes, one may represent the loading in the form
Z Z
1 1 1
pðx1 ; x2 Þ ¼ 2 pðn1 ; n2 Þ
p 0 0
 cos a1 ðn1  x1 Þ cos a2 ðn2  x2 Þda1 da2 dn1 dn2 ð6:770 Þ

too; these results may be applied in the case in which the function pðx1 ; x2 Þ
accomplishes the conditions of Lejeune-Dirichlet and is absolutely integrable in
the whole plane, i.e. the integral
Z 1Z 1
pðn1 ; n2 Þdn1 dn2 ð6:78Þ
1 1

has sense. At the points of discontinuity, the above integral represents the arith-
metic mean of the four limits at the right and at the left too.
In the case of a symmetric loading with respect to both axes of co-ordinates, we
use an even function with respect to both variables
276 6 Principles and General Theorems. Computation Methods

Z 1 Z 1
pðx1 ; x2 Þ ¼ dða1 ; a2 Þ cos a1 x1 cos a2 x2 da1 da2 ; ð6:79Þ
0 0

where
Z 1 Z 1
1
dða1 ; a2 Þ ¼ pðn1 ; n2 Þ cos a1 n1 cos a2 n2 dn1 dn2 ; ð6:790 Þ
4p2 0 0

other cases of symmetry or antisymmetry lead to similar results.


Analogically, one can use triple Fourier integrals; e.g., a volume load
Fðx1 ; x2 ; x3 Þ, with some properties of symmetry or of antisymmetry with respect to
the three planes of co-ordinates, may be represented in the form
Z 1Z 1Z 1
Fðx1 ; x2 ; x3 Þ ¼ dða1 ; a2 ; a3 Þ
0 0 0
 cos a1 x1 cos a2 x2 cos a3 x3 da1 da2 da3 ; ð6:80Þ

with
Z Z Z
8 1 1 1
dða1 ; a2 ; a3 Þ ¼ Fðn1 ; n2 ; n3 Þ
p2 0 0 0
 cos a1 n1 cos a2 n2 cos a3 n3 dn1 dn2 dn3 : ð6:800 Þ

Sometimes it is convenient to use mixed representations too, which may be


Fourier series on one direction and Fourier integrals on the other one. Let thus be
the representation
Z 1 X Z 1
pðx1 ; x2 Þ ¼ a0 ða1 Þ cos a1 x1 da1 þ cos a2m x2 am ða1 Þ cos a1 x1 da1 ;
0 m 0

ð6:81Þ
corresponding to a loading local along the direction x1 and periodic along the
direction x2 ; the variable coefficients of this representation are
Z L1 =2 Z 1
4
a0 ða1 Þ ¼ pðn1 ; n2 Þ cos a1 n1 dn1 dn2 ; ð6:810 Þ
pL1 0 0
Z L1 =2 Z 1
8
am ða1 Þ ¼ pðn1 ; n2 Þ cos a1 n1 cos a2 n2 dn1 dn2 : ð6:8100 Þ
pL1 0 0

6.2.1.5 Integrals for the Simple and for the Double Wave Equations

We saw that, in the solving in displacements, as well as in the case of solving in


stresses, the problems of electrodynamics are reduced, from the mathematical
standpoint, to the integration of wave equations, defined in Sect. A.1.2.4, with
6.2 Computation Methods 277

certain given limit conditions. To this end, it is useful to give some results con-
cerning the integration of these equations of hyperbolic type. In Sect. A.1.2.7 have
been presented Boggio’s relations which allow to construct the solution of the
double wave equation by means of the solutions of two simple wave equations. We
have thus to deal with the equation
h1 h2 Uðx1 ; x2 ; x3 ; tÞ ¼ 0: ð6:82Þ
If we choose a function U of the form (the method of separation of variables)
Uðx1 ; x2 ; x3 ; tÞ ¼ uðx1 ; x2 ; x3 Þ#ðtÞ; ð6:83Þ

then the equation (6.82) shows that the temporal function must satisfy an equation
of the form
€  x2 #ðtÞ ¼ 0:
#ðtÞ ð6:84Þ
Taking the sign -, it results an aperiodic motion; if we have x ¼ 0; then one
obtains a critical aperiodic motion. Both cases are less interesting in practice.
Taking the sign +, one obtains periodic vibrations. The temporal function will be
thus of the form
#ðtÞ ¼ A cos xt þ B sin xt: ð6:840 Þ
In general, the potential function will be of the form
X
1
Uðx1 ; x2 ; x3 ; tÞ ¼ Ui cosðxi t  ui Þ; ð6:85Þ
i¼1

where Ui ¼ Ui ðx1 ; x2 ; x3 Þ; hence a superposition of harmonic vibrations of period


T (one assumes a periodic motion), where ui is the phase shift, while
2pi
xi ¼ ð6:86Þ
T
is the pulsation (circular frequency). The functions Ui are given by the bimeta-
harmonic equation
2 2
ðD þ k1i ÞðD þ k2i ÞUðx1 ; x2 ; x3 Þ ¼ 0; ð6:87Þ

where we have introduced the notations


ð1 þ mÞð1  2mÞ q 2 2 q
2
k1i ¼ xi ; k2i ¼ 2ð1 þ mÞ x2i : ð6:870 Þ
1m E E
Taking into account Boggio’s theorem, we may write
Ui ¼ U1i þ U2i ; ð6:88Þ
278 6 Principles and General Theorems. Computation Methods

where the metaharmonic functions Uji ¼ Uji ðx1 ; x2 ; x3 Þ; j ¼ 1; 2; verify the


equations

ðD þ kji2 ÞUji ¼ 0; j ¼ 1; 2: ð6:880 Þ

These equations lead to a state of strain and stress corresponding to steady-state


vibrations.
In the case of small non-steady vibrations, one can make a Laplace transform
with respect to time, admitting homogeneous initial conditions (vanishing initial
displacements and displacement velocities); thus, the problem is formally reduced
to a steady-state one (a quasi-static problem).
As to the boundary conditions, we shall mention two important cases, where no
initial conditions appear (which constitutes a considerable simplification of the
computation):
(i) The body is subjected to the action of a periodic load (steady-state case). In
this case, one can use expansions into Fourier series with respect to time. i.e.
functions of the form (6.85). We shall express the external loads in the same
form, which will facilitate the setting of the boundary conditions.
(ii) The body is subjected to the action of an aperiodic load (e.g., a shock) with
respect to time. In this case, to the Fourier series, a Fourier integral with
respect to time will be submitted; therefore we shall use functions of the form

Z 1
Uðx1 ; x2 ; x3 ; tÞ ¼ Us0 ðx1 ; x2 ; x3 Þ cos stds
1
Z 1
þ Us00 ðx1 ; x2 ; x3 Þ sin stds: ð6:89Þ
1

To simplify the putting of the boundary conditions, the external loads will be
similarly expressed. The limit
lim PDt; ð6:890 Þ
P!1
Dt!0

will occur, thus introducing the notion of impulse.


We can obtain integrals of the Eq. (6.880 ) by the separation of variables. i.e. in
the form

ea1i x1 þa2i x2 þa3i x3 ; ð6:90Þ


where

a21i þ a22i þ a23i þ kji2 ¼ 0; j ¼ 1; 2; ð6:900 Þ

we find thus integrals of the form


6.2 Computation Methods 279

sin a1i x1 sin a2i x2 ea3i x3 ; sin a1i x1 sin a2i x2 sinh a3i x3 ; ð6:91Þ
where

a21i þ a22i ¼ a13i þ kji2 ¼ 0; j ¼ 1; 2; ð6:910 Þ

or integrals of the form

ea1i x1 ea2i x2 sin a3i x3 ; sinh a1i x1 sinh a2i x2 sin a3i x3 ;
ea1i x1 sinh a2i x2 sin a3i x3 ; ð6:92Þ

where

a21i þ a22i þ kji2 ¼ a23i ; j ¼ 1; 2: ð6:920 Þ

Other particular integrals are obtained by substituting some functions such as


‘‘cos’’ to ‘‘sin’’ or ‘‘cosh’’ to ‘‘sinh’’ or by changing with one another the space
variables.
We remark that, in general, the simple waves equation admits particular inte-
grals in the form

ea1 x1 þa2 x2 þa3 x3 þst ; ð6:93Þ


where
1 2
a21 þ a22 þ a23  s ¼ 0; j ¼ 1; 2: ð6:930 Þ
c2j

cj being the wave propagation velocities.


By the superposition of the effects we can build, in this way, Fourier represen-
tations (series or integrals), useful in solving limiting problems. If the boundary
conditions are non-homogeneous, then we have to deal with forced vibrations, the
external loads playing the rôle of perturbing loads. In the case of homogeneous
boundary conditions (absence of the external loads), a characteristic equation results,
corresponding to the free vibrations (i.e. giving the eigenvalues) of the elastic body.

6.2.2 General Methods of Computation

So as to find the potential functions and to solve the problems of the theory of
elasticity, two general computation methods are used: the inverse method and the
direct method.
280 6 Principles and General Theorems. Computation Methods

6.2.2.1 The Inverse Computation Method

The inverse computation method consists of admitting a certain state of stress or a


certain state of strain within the body, such as to fulfill the boundary conditions
and to verify whether all the equations of the theory of elasticity are satisfied.
Contingently, these can be results supplied by elementary computation methods.
It may happen that some elements, do not be completely specified, i.e.: the form
of the body or the mechanical properties of the material and the external loads; in
this case, the respective elements remain to be determined, so as to correspond to
the given state strain and stress. Such a problem is, in fact, a design problem. One
has thus to deal with an optimal theory of the theory of elasticity, in which it is
asked that the elastic system be stressed as uniform as possible, so as not to have
any parts of the body superstressed. However, in the case of a sudden change of the
geometric or mechanical configuration of the body, there appear stress concen-
trations, for which special methods of computation have to be used.
If the equations of the theory of elasticity are not verified, then one must make
another hypothesis concerning the state of strain and stress and repeat the com-
putation; eventually, by simplifying hypotheses, compatible with the complete
system of equations of the theory of elasticity and with the limit conditions, the
number of the unknown functions, as well as of the equations, may be reduced.
Thus, one has to integrate a reduced system of equations with certain limit con-
ditions; taking into account the theorem of uniqueness, the solution of this system
of equations will correspond to the searched solution. One can add some correction
functions too, which have to be determined. This is a semi-inverse computation
method, due to Saint-Venant [32]; in principle, it is different from the simplifying
methods of strength of materials, leading to a solution which satisfies all the
differential equations, as well as the limit conditions.

6.2.2.2 The Direct Computation Method

The direct computation method consists in choosing certain potential functions


(eventually fulfilling some conditions of evenness or oddness with respect to the
three space variables or other conditions depending on the particular conditions of
the problem), such that they comprise certain arbitrary parameters, which have to
be determined by means of the limit conditions. If we can univocally determine
these parameters, while fulfilling all the limit conditions, the problem is solved.
Contrarily, we must choose other potential functions, allowing us to effect the
computation (or to add correction functions to the initially chosen ones).
Another aspect of the direct computation methods consists in the possibility of
tackling the problems by a general method, liable to lead to a systematic com-
putation. Obviously, each of the computation methods mentioned above is suitable
to a particular case; hereafter, we shall chiefly use direct computation methods.
6.2 Computation Methods 281

6.2.2.3 Exact and Approximate Computations

From the mathematical standpoint, it is possible to effect either an exact compu-


tation or an approximate computation. The cases when an exact computation is
possible are rather rare, because the physical phenomena are, in general, intricate
and can be only approximately expressed in various mathematical forms. That is
why most of the methods in use are approximate ones.
The approximations arise either from choosing the potential functions in a form
that gives as good an approximation of their real form, as we require it (e.g., a
Fourier representation) and that fully verifies the differential equation of the
problem as well as the limit conditions, or from choosing a potential function that
fulfils approximately some of these conditions (the method of finite differences, the
method of collocations etc.); besides—and this occurs often enough—both these
kinds of approximations can be jointly used.
Among the general computation methods which we shall mention: the ele-
mentary methods (using, in general, some particular integrals), the variational
methods (frequently using Fourier representations energetical methods etc.),
operational methods (where the Fourier or Laplace transforms are used), the
method of finite differences (comprising, e.g., the relaxation method), the method of
collocations (the point matching method), iteration methods (methods of succes-
sive approximations), methods of integral equations (by reducing the problems to
the integration of such equations), methods of fundamental solutions (formulating
the problems in distributions and searching solutions in the space of distributions),
method of ‘‘spline’’ functions, method of functions of complex variables (useful
especially for one- or two-dimensional problems), method of functions of hyper-
complex variables, method of p-analytic functions (useful for axisymmetrical
problems), experimental methods (methods by which one may determine the state
of strain and stress in the interior of an elastic body and not check up the
hypotheses which have been made, i.e.: analogies with other problems, photo-
elastic methods, experiments on models) etc.
In connection with these various computation methods, details can be found in
the monographs [14, 15]. Hereafter, we shall confine ourselves to give some details
about the variational methods, the methods of successive approximations and the
point matching method; as well, we shall deal with the method of fundamental
solutions both in the static and in the dynamic case, using results of the distri-
butions theory.

6.2.3 Variational Methods

The variational methods are approximate ones, based on the extremal of a func-
tional. To do this, we choose certain expressions which depend on a number
(theoretically infinite, practically finite) of arbitrary parameters for the functions
282 6 Principles and General Theorems. Computation Methods

we search and the condition that these expressions do approximate as good as


possible these functions.

6.2.3.1 General Considerations

Let be a function of the form


X
n
F ¼ f0 þ f ¼ f0 þ ai fi ; ð6:94Þ
i¼1

where f0 and fi are given functions, while ai ; i ¼ 1; 2; . . .; n; are parameters to be


determined. The functions f0 and fi must be conveniently chosen, suitable for the
problem we have to solve. From this point of view, we distinguish three methods
to choose these functions, i.e.:
(i) The functions f0 and fi do not satisfy the equation which must be verified by
the function F, but satisfy each one some of the boundary conditions; the
parameters ai are determined from the condition that the expression (6.94)
does approximate as good as possible the function F, both in the interior of
the domain and on its boundary.
(ii) The functions f0 and fi do not satisfy the equation which must be verified by
the function F, but verify each one the boundary conditions; the parameters ai
are obtained from the condition that the expression chosen for the function
F does approximate it the best possible in the interior of the domain.
(iii) The functions f0 and fi are particular integrals of the equation which must be
satisfied by the function F, but do not satisfy the boundary conditions; the
parameters ai are determined from the condition that the expression (6.94)
does approximate the best possible the function F on the frontier.
The first of these methods is the most simple from the point of view of the
construction of the approximate expression of the function F, but it is more dif-
ficult to determine the parameters ai ; the other methods (when we may choose
easily the functions f0 and fi ) lead to a quicker computation. If it is possible, then it
is convenient to choose the function f0 so as to correspond, with a certain
approximation, to the solution of the problem, even of another nature than the
other functions fi ; in this case, the function f is a correction function.
The correction function f may be determined by various methods of approxi-
mation. To do this, we choose a measure of the error which we make, replacing the
real value of the function by its approximate one. Such a measure may be, e.g., the
maximum of the absolute value of the deviation of the approximation function
from the real value
A0 ¼ maxjF  f0  f j: ð6:95Þ
6.2 Computation Methods 283

Chebyshev’s uniform approximation method consists in putting the condition


that a0 does have a minimum value. This method is seldom used because of the
difficulties of computation.
In this case, one calculates the mean value of order s of a deviations
A ¼ jF  f0  f j

in the form
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ZZZ
s 1
As ¼ As dV ; ð6:96Þ
V V

the integral being extended to the domain of volume V.


For s ¼ 1 one obtains the mean arithmetic error. Putting the condition that A1
be minimum, it is possible not to obtain a unique solution. Therefore, one takes,
usually, as measure of the error the mean quadratic error ðs ¼ 2Þ.
Thus we are led, e.g., to an integral of the form
Z x1
IðyÞ  Fðx; y; y0 Þ dx; ð6:97Þ
x0

0
where Fðx; y; y Þ represents a known real function of arguments x; y and
y0  dy=dx; the value of this integral depends on the choice of the function y ¼
yðxÞ; wherefrom it results the notation used and the denomination of functional.
We assume that the admissible arguments yðxÞ are of class C2 and that at the ends
of the interval ½x0 ; x1  they take the values y0 ; y1 ; in this case, the set fyðxÞg of
admissible arguments yðxÞ may be seen as a family of smooth curves, passing
through the points ðx0 ; y0 Þ and ðx1 ; y1 Þ; from which we must choose one which
minimizes the functional IðyÞ. A necessary condition which may determine this
curve is the Euler-Poisson equation
dFy0
Fy  ¼ 0; ð6:970 Þ
dx
associated to the variational problem IðyÞ ¼ min, where Fy ; Fy0 represent the
partial derivatives with regard to the respective arguments; one obtains thus

d2 y dy
F y0 y0 2
þ Fy0 y þ Fy0 x  Fy ¼ 0: ð6:970 Þ
dx dx
But this condition is only necessary; one must then verify if the solution of the
differential equation effective minimizes the functional IðyÞ.
Let be now a functional of the form
ZZ
IðuÞ  Fðx1 ; x2 ; u; u1 ; u2 Þ dx1 dx2 ð6:98Þ
D
284 6 Principles and General Theorems. Computation Methods

on a set fuðx1 ; x2 Þg of functions of class C2 ; which take given continuous values


u ¼ uðsÞ on the frontier of the domain D; F is a given function of class C2 ; with
respect to the arguments x1 ; x2 ; u; u1  ou=dx1 ; u2  ou=dx2 in the domain of
definition of those arguments. The Euler-Ostrogradskiı̆ equation corresponding to
the problem of minimum is written in the form
o o
Fu  Fu1  Fu ¼ 0; ð6:980 Þ
ox1 ox2 2
being a necessary condition too. In particular, in case of the functional
ZZ
IðuÞ  ½u21 þ u22 þ 2f ðx1 ; x2 Þu dx1 dx2 ; ð6:99Þ
D

we find an effective minimum, given by Poisson’s equation


Duðx1 ; x2 Þ ¼ f ðx1 ; x2 Þ; ð6:990 Þ

in the domain D.
Analogically, the functional
ZZ
IðuÞ  Fðx1 ; x2 ; u; u1 ; u2 ; u11 ; u12 ; u22 Þ dx1 dx2 ð6:100Þ
D

leads to the Euler-Ostrogradskiı̆ equation

o o o2 o2 o2
Fu  Fu 1  Fu2 þ 2 Fu11 þ Fu12 þ 2 Fu22 ¼ 0: ð6:1000 Þ
ox1 ox2 ox1 ox1 ox2 ox2
As well, for the functional
ZZ
IðuÞ  Fðx1 ; x2 ; x3 ; u; u1 ; u2 ; u3 Þ dx1 dx2 dx3 ; ð6:101Þ
D

defined on a three-dimensional domain D, one obtains the Euler-Ostrogradskiı̆


equation
o o o
Fu  Fu  Fu  Fu ¼ 0: ð6:1010 Þ
ox1 1 ox2 2 ox3 3
We make now some considerations on the least squares method and on the Ritz
method.

6.2.3.2 The Least Squares Method

If the potential function F is biharmonic, e.g., then we introduce the functional


6.2 Computation Methods 285

ZZZ
1
A02
2 ¼ ½DDðf0 þ f Þ2 dV; ð6:102Þ
V V

we must put the condition that these functions be approximate the best possible in
the interior of the given domain.
n n
As well, let be given the normal component p and the tangential component q of
the external load on the element of area dS, of external normal n; the mean
quadratic error may be calculated in the form
ZZ ZZ
002 1 n n 2 1 n n
A2 ¼ ðp  rÞ dS þ ðq  sÞ2 dS; ð6:1020 Þ
S S S S

n n
where r and s are the normal and the tangential stress, respectively, corresponding
to the same element of area, given by potential functions of the form (6.94).
In the first method mentioned at the previous Subsection (the most general one),
we calculate the mean quadratic error in the form (we observe that one can have
several components of the form (6.102) of the mean quadratic error, taking into
account all the potential functions used)

A22 ¼ A02 002


2 þ A2 : ð6:103Þ
The mean quadratic error depends on the n arbitrary parameters ai . By putting
conditions of minimum for A2 or for A22 (it is the same, because a2  0), we get the
equations

oA22
¼ 0; i ¼ 1; 2; . . .; n; ð6:104Þ
oai
which form a linear system
X
n
cij aj ¼ di ; i ¼ 1; 2; . . .; n; ð6:1040 Þ
j¼1

with regard to the parameters ai .


In general, we have cij ¼ cji ; the system being symmetric with respect to the
principal diagonal. If one uses an orthogonal system of n functions fi ; then cij ¼
0; i 6¼ j; which leads to important simplifications of computation, each equation
remaining with only one unknown parameter.
In the computation method presented above, known under the name of least
squares method, one passed directly to the minimizing of the functional A2 ;
without calculating its first variation; just this is the simplification made in the
computation.
286 6 Principles and General Theorems. Computation Methods

6.2.3.3 Ritz’s Method

One may obtain the correction function by using the principle of the minimum
potential energy or the principle of the minimum complementary energy; the
respective method is called Ritz’s method.
If the boundary conditions are put in displacements and one solves the problem
in displacements, then it is convenient to use the principle of minimum potential
energy (Lagrange); we express, in this case, Wi by means of the components ui of
the displacement, using the formula (4.90). We choose (in case of linear
elastostatics)
nj
X
uj ðx1 ; x2 ; x3 Þ ¼ u0j ðx1 ; x2 ; x3 Þ þ akj ukj ðx1 ; x2 ; x3 Þ; ð6:105Þ
k¼1

where j ¼ 1; 2; 3 and k ¼ 1; 2; . . .; nj being possible to use a different number of


parameters for each component of the displacement vector.
There result the variations
nj
X
duj ¼ ukj dakj ; ð6:106Þ
k¼1

while for the internal work one has


nj
3 X
X oWi
dWi ¼ dakj : ð6:107Þ
j¼1 k¼1
oakj

Taking into account the principle of virtual displacements (from which results
Lagrange’s principle), the variations dakj of the displacement are arbitrary. We
obtain thus the equations
ZZZ ZZ
oWi k n
¼ u F
j j dV  ukj pj dS ¼ 0 ð!Þ; j ¼ 1; 2; 3; k ¼ 1; 2; . . .; nj ;
oakj V S

ð6:108Þ
which represents the canonical form of the Lagrange-Ritz equations for linear
elastostatics. We obtain thus a system of n1 þ n2 þ n3 linear algebraic equations,
which determine the unknown parameters akj .
If we put the boundary conditions in stresses and solve the problem in stresses,
then we use the principle of minimum of complementary energy; the internal work
will be expressed by means of stresses in the form (4.9000 ). We write (in the frame
of linear elastostatics too)
njk
X
rjk ðx1 ; x2 ; x3 Þ ¼ r0jk ðx1 ; x2 ; x3 Þ þ aljk rljk ðx1 ; x2 ; x3 Þ; ð6:109Þ
l¼1
6.2 Computation Methods 287

where j; k ¼ 1; 2; 3 and l ¼ 1; 2; . . .; njk ; being possible to use a different number


of parameters for each component of the stress tensor.
Thus, there result the variations
njk
X
drjk ¼ rljk daljk ; ð6:110Þ
l¼1

while for the internal work we may write


njk
3 X
X oWi
dWi ¼ daljk : ð6:111Þ
j;k¼1 l¼1
oaljk

Taking into account the principle of virtual variations of the state of stress (from
which Castigliano’s principle is obtained), one can consider the variations daljk as
arbitrary; if the relations (6.25), (6.250 ) are verified, then we get the equations
ZZZ
oWi
ð1 þ djk Þ l þ ðuk rljk;j þ uj rlkj;k Þ dV
oajk V
ZZ
 ðuk nj þuj nk Þ dS ¼ 0 ð!Þ; j; k ¼ 1; 2; 3; l ¼ 1; 2; . . .; njk ; ð6:112Þ
S

which represent the canonical form of the Castigliano-Ritz equations for the linear
elastostatics. The parameters aljk are thus given by a system of n11 þ n22 þ n33 þ
n23 þ n31 þ n12 linear algebraic equations.
If the functions r0jk verify the equilibrium equations, while the functions rljk are
particular integrals of these equations in the absence of the volume forces, the
latter ones do not intervene any more in the statement of the principle of minimum
complementary energy; hence, the Eq. (6.112) take a simpler form, without
volume integrals. If the functions r0jk verify the boundary conditions too and the
functions rljk verify homogeneous (zero) conditions on the boundary, then the
Eq. (6.112) take a simpler form, neither these conditions intervening. We obtain
thus the equations
oWi
¼ 0; j; k ¼ 1; 2; 3; l ¼ 1; 2; . . .; njk ; ð6:113Þ
oaljk

which express the principle of the minimum internal work.


The Eq. (6.112) may be used even on the contour are given conditions in
displacements. We must mention that, in the case of the representations (6.105)
and (6.109) one can use the least squares method too.
Another computation method which we wish to emphasize is the Bubnov-
Galerkin method, based on the property of orthogonality which may have two
functions. An analogous computation method is Trefftz’s method, where one
calculates the square of the gradient of the difference between the function which
288 6 Principles and General Theorems. Computation Methods

is searched and the approximation one; by putting conditions of extremum for this
functional (least squares method), one can determine the unknown parameters.
One can also apply the Bubnov-Galerkin method, putting the condition that the
gradient of the difference mentioned above be orthogonal with certain functions,
e.g. with the functions generating the expression of approximation. In practice we
shall use some methods or other ones, as the functions f0 and fi do satisfy or not all
the limit conditions or are or not particular integrals of the potential differential
equation of the problem. As well, if it is necessary for simplifying the computa-
tion, then we may use for the same problem ideas from different methods.
It is interesting to mention that Ritz’s method approximates the value of the
function in excess, while the Trefftz one leads just to a contrary approximation;
hence, the value of the searched function may be bordered between two limits.
Besides the computation methods mentioned above as the most important ones,
many other methods have been developed, which may be interesting in various
particular cases. Thus, the method of the least squares has an interesting extension
in the method of the least products; thus, the integral
ZZZ
½DDðf0 þ f ÞðF  f0  f Þ dV ð6:114Þ
V

must have a minimum, for which the constructed function represents a better
approximation.

6.2.4 Method of Fundamental Solutions

The introduction of the fundamental solutions allows a systematic study of the


problems of the theory of elasticity, using the methods of the theory of distribu-
tions, presented in Sect. A.3. After a study concerning the representation of the
concentrated forces, one deals with the fundamental solutions of the static and
dynamic problems of the elastic bodies, which allow to express the solutions in a
general case of loading.

6.2.4.1 Representation of the Concentrated Forces

Let be a bound vector V of components Vi ; i ¼ 1; 2; 3; with respect to an ort-


honormed frame of reference Ox1 x2 x3 applied at the point O. Let us consider a
deformable domain Xe , which contains the origin O, so that for e ! þ0; Xe tends
to the point O; we define on this domain the field of vectors Qe ðx1 ; x2 ; x3 Þ, of
components Vi fie ðx1 ; x2 ; x3 Þ ð!Þ; i ¼ 1; 2; 3, so that fie be representative sequences d,
which have bean introduced in Sect. A.3.1.1, i.e.
lim fie ðx1 ; x2 ; x3 Þ ¼ dðx1 ; x2 ; x3 Þ; i ¼ 1; 2; 3: ð6:115Þ
e!0
6.2 Computation Methods 289

Passing to the limit in the sense of the theory of distributions, we obtain


QðrÞ ¼ Qðx1 ; x2 ; x3 Þ ¼ lim Qe ðx1 ; x2 ; x3 Þ ¼ Vj ij dðx1 ; x2 ; x3 Þ;
e!þ0

hence, the field of vectors equivalent to the bound vector V may be expressed in
the form
QðrÞ ¼ VdðrÞ: ð6:116Þ

This field of vectors may be considered as volume density of the bound vector V.
If the bound vector is applied at a point A of position vector r0 ; then the
equivalent vector field is given by
QðrÞ ¼ Vdðr  r0 Þ: ð6:1160 Þ
To represent the bound vector V, we can use an arbitrary representative
sequence d; because that one does not intervene in the final result by its concrete
form, but only by its properties.
Let us consider now the field of parallel vectors
Qe ðx1 ; x2 ; x3 Þ ¼ Vfe ðx1 ; x2 ; x3 Þ;
defined on the sphere Xe with the centre at the origin and of radius e; passing to the
limit, in the sense of the theory of distributions, we obtain

QðrÞ ¼ Qðx1 ; x2 ; x3 Þ ¼ lim Qe ðx1 ; x2 ; x3 Þ


e!þ0
¼ V lim fe ðx1 ; x2 ; x3 Þ ¼ Vdðx1 ; x2 ; x3 Þ ¼ VdðrÞ:
e!þ0

To put into evidence the correctness of this representation, we shall calculate


the torsor at the point O of the considered field of vectors. We obtain thus the
resultant
ZZZ ZZZ ZZZ
Re ¼ Qe dX ¼ Vfe dX ¼ V fe dX; ð6:117Þ
Xe Xe Xe

wherefrom, passing to the limit for e ! þ0; we get


R ¼ V: ð6:1170 Þ
The resultant moment is given by
ZZZ ZZZ ZZZ
Me ¼ r  Qe dX ¼ fe r  V dX ¼ fe V  rdX ¼ 0; ð6:118Þ
Xe Xe Xe

where it has been assumed that fe depends only on e for jrj


e and where it has
been taken into account that
290 6 Principles and General Theorems. Computation Methods

ZZZ
r dX ¼ 0; ð6:119Þ
Xe

the integral corresponding to the static moment of the homogeneous sphere Xe


with respect to its centre; it is obvious that
MO ¼ 0; ð6:1180 Þ

the representation (6.116) being thus justified from the mechanical point of view.
Let be V1 ; V2 ; . . .; Vn a system of n bound vectors, applied at the point Aðr0 Þ;
the equivalent vector field will be given by
Qi ðrÞ ¼ Vi dðr  r0 Þ; i ¼ 1; 2; . . .; n: ð6:120Þ

Introducing the resultant vector


X
n
R¼ Vi ; ð6:121Þ
i¼1

applied at the same point Ai ; one may write the equivalent vector field in the form
X
n
QðrÞ ¼ Qi ðrÞ ¼ Rdðr  r0 Þ: ð6:122Þ
i¼1

Hence, the properties of the bound vectors, applied at the very same point, are
preserved in their representation by means of distributions.
The concentrated loads, applied to a deformable solid, may be mathematically
modelled by means of bound vectors. Thus, taking into account the above con-
siderations, one may represent the equivalent load of a concentrated force F,
applied at a point A of position vector r0 , in the form
QðrÞ ¼ Fdðr  r0 Þ; ð6:123Þ
the support of the equivalent load QðrÞ is the point A.

6.2.4.2 The Static Problem of the Theory of Elasticity

Let us consider the elastic space 1\x1 ; x2 ; x3 \1 acted upon by a volume


force Fi ; i ¼ 1; 2; 3; expressed in distributions. The state of displacements is
given by Lamé’s equations, where the operations are considered in the sense of the
theory of the distributions too. The solutions of this system must satisfy certain
regularity conditions at the infinite; thus, if
pffiffiffiffiffiffiffi
R ¼ xi xi ð6:124Þ

is the vector radius, with the notation used in case of the spherical co-ordinates,
then the components of the displacement vector must tend to zero for R ! 1.
6.2 Computation Methods 291

The tensor of components uij ¼ uij ðx1 ; x2 ; x3 Þ; i; j ¼ 1; 2; 3; will be called


fundamental solution tensor, corresponding to the elastic space, if these compo-
nents satisfy the system of equations
lDuij þ ðk þ lÞukj;ki þ dij dðx1 ; x2 ; x3 Þ ¼ 0; i; j ¼ 1; 2; 3; ð6:125Þ

where dðx1 ; x2 ; x3 Þ is Dirac’s distribution concentrated at the origin. As well, one


must verify the boundary conditions
lim uij ¼ 0; i; j ¼ 1; 2; 3: ð6:126Þ
R!1

By the application of the Fourier transform in distributions and by taking into


account the properties mentioned in Sect. A.3.3.1, we get the linear algebraic
system of equations
lak ak F½uij  þ ðk þ lÞai ak F½ukj  ¼ dij ;

where ai ; i ¼ 1; 2; 3; are complex variables; multiplying by ai and summing, it


results
1 ai
aj F½uij  ¼ :
k þ 2l ak ak
We obtain thus the Fourier transforms in the form
kþl ai aj 1 dij
F½uij  ¼  þ ; i; j ¼ 1; 2; 3: ð6:127Þ
lðk þ 2lÞ ðak ak Þ2 l ak ak

Taking into account the formulae (A.237) and (A.2370 ), we may write

1
D ¼ 4pdðx1 ; x2 ; x3 Þ; ð6:128Þ
R

applying the Fourier transformation, it results


 
1 4p
F ¼ :
R ak ak
By means of a formula of the form (A.3.40) for distributions of several vari-
ables and choosing the polynomial P of the form ixj , we get
hx i aj
j
F ¼ 8pi ; j ¼ 1; 2; 3; ð6:129Þ
R ðak ak Þ2

analogically, using a formula of the form (A.3.41) for the differentiation of the
function under the Fourier transform operator with respect to the variables xi , we
get
292 6 Principles and General Theorems. Computation Methods

  hx x i
1 i j ai aj
dij F F ¼ 8p ; i; j ¼ 1; 2; 3: ð6:1290 Þ
R R ðak ak Þ2
We notice that between the elastic constants of the material take place the
relations
kþl 1 k þ 3l 3  4m
¼ ; ¼ :
k þ 2l 2ð1  mÞ k þ 2l 3ð1  mÞ
Starting from the formulae (6.127) and using the above results, we get
1 1h xi xj i
uij ðx1 ; x2 ; x3 Þ ¼ ð3  4mÞdij þ 2 ; i; j ¼ 1; 2; 3; ð6:130Þ
16pð1  mÞG R R

this is the Kelvin-Somigliana fundamental solution tensor, which corresponds to a


formulation in displacements of the problem. We notice that the components of
this tensor may also be written in the form
"
#
1 1 1 2 1
uij ðx1 ; x2 ; x3 Þ ¼ ð5  6mÞdij þ R ; i; j ¼ 1; 2; 3:
24pð1  mÞG R 2 R ;ij
ð6:1300 Þ
The regularity conditions at infinite (6.126) are automatically satisfied, because of
the Fourier transformations.
In the case of some arbitrary volume forces Fi ; i ¼ 1; 2; 3; expressed in distri-
butions, the state of displacements is given by
ui ðx1 ; x2 ; x3 Þ ¼ uij ðx1 ; x2 ; x3 Þ Fj ðx1 ; x2 ; x3 Þ; i ¼ 1; 2; 3; ð6:131Þ

where we have introduced the convolution product with respect to the space
variables.
Using the matrices
2 3 2 3 2 3
u11 u12 u13 u1 F1
U  4 u21 u22 u23 5; u  4 u2 5; Q  4 F2 5; ð6:132Þ
u31 u32 u33 u3 F3

i.e. the fundamental solution matrix U, the displacement state matrix u and the
volume forces field matrix Q, we also may write
u ¼ U Q: ð6:1310 Þ
What concerns the numerical computation, the values corresponding to regular
distributions are calculated as usual, while the singular distributions are approxi-
mated by means of representative sequences. Obviously, for the numerical cal-
culation of various convolution products, the electronic computer may be very
useful.
6.2 Computation Methods 293

In the case of the elastic space subjected to the action of a concentrated force P
at the point Aðx01 ; x02 ; x03 Þ; we introduce the equivalent field of components

Fi ðx1 ; x2 ; x3 Þ ¼ Pi dðx1  x01 ; x2  x02 ; x3  x03 Þ; i ¼ 1; 2; 3; ð6:133Þ

the formula (6.131) leads to the state of displacement

ui ðx1 ; x2 ; x3 Þ ¼ Pi uij ðx1  x01 ; x2  x02 ; x3  x03 Þ; i ¼ 1; 2; 3: ð6:134Þ


In particular, if the force P is applied at the origin, it results
1 1h xi xj i
ui ðx1 ; x2 ; x3 Þ ¼ ð3  4mÞPi þ Pj 2 ; i ¼ 1; 2; 3; ð6:135Þ
16pð1  mÞG R R
in a vector form, we get the Kelvin-Somigliana displacement vector
 
1 1 RP
uðRÞ ¼ ð3  4mÞP þ 2 R ; ð6:1350 Þ
16pð1  mÞG R R

R being the position vector applied at the point at which acts the force P. We can
get this result on another way too, using a proof due to N. S ßandru [36]; thus, on
behalf of the formula (A.3.13), we can write the Eqs. (5.34), (5.35), which give the
vector potential and the scalar potential, respectively, in the Papkovich-Neuber
representation (4.4.22), in the form
Dv þ 2Pdðx1 ; x2 ; x3 Þ ¼ 0; Dv0 ¼ 0:
Assuming that the function v is regular at infinite and taking into account (6.128),
we get
1 P
v¼ ; v ¼ 0; ð6:136Þ
2p R 0
so that the representation (5.30) leads to the same result.
If Fj ; j ¼ 1; 2; 3; are locally integrable functions, we can write
Z 1Z 1Z 1
ui ðx1 ; x2 ; x3 Þ ¼ uij ðx1  n1 ; x2  n2 ; x3  n3 Þ
1 1 1
 Fj ðn1 ; n2 ; n3 Þdn1 dn2 dn3 ; i ¼ 1; 2; 3; ð6:137Þ

introducing thus the volume elastic potential; as a matter of fact, we made a


translation at a point of position vector n and co-ordinates ni ; i ¼ 1; 2; 3; which
corresponds to the introduction of Green’s functions of the elastostatics problem,
in case of a solution in displacements.
Starting from the Kelvin-Somigliana fundamental solution (6.130), calculating
the state of strain by means of Cauchy’s formulae and then the state of stress by
means of Hooke’s law, we obtain the fundamental solution tensor, which corre-
sponds to a formulation in stresses of the problem, in the form
294 6 Principles and General Theorems. Computation Methods

2m
sijk ¼ G uik;j þ ujk;i þ ulk;l dij ; ð6:138Þ
1  2m

taking into account the derivatives


1 xi
¼ 2; ð6:139Þ
R ;i R
x x  1  xi xj xk 
i j
¼ x d
i jk þ x d
j ik  3 ; ð6:1390 Þ
R3 ;k R3 R2
we obtain, finally
1
sijk ðx1 ; x2 ; x3 Þ ¼
2pð1  mÞ
1 h xi xj xk i
 3 ð1  2mÞðxk dij  xi djk  xj dij Þ  3 2 ; i; j; k ¼ 1; 2; 3:
R R
ð6:1380 Þ
The state of stress corresponding to the arbitrary volume forces
Fk ðx1 ; x2 ; x3 Þ; k ¼ 1; 2; 3; will be given by
rij ðx1 ; x2 ; x3 Þ ¼ sijk ðx1 ; x2 ; x3 Þ Fk ðx1 ; x2 ; x3 Þ; i; j ¼ 1; 2; 3; ð6:140Þ

where we have introduced, as well, the convolution product with respect to the
space variables.
We notice that
sijk ¼ sjik ; i; j; k ¼ 1; 2; 3; ð6:1400 Þ

due to the symmetry of the stress tensor.


Introducing the matrices
2 3 2 3
s111 s112 s113 r11
6 s221 s222 s223 7 6 r22 7
6 7 6 7
6 s331 s332 s333 7 6 r33 7
S6 6 7 6 7 ð6:1320 Þ
7; r  6 r23 7;
6 s231 s232 s233 7 6 7
4 s311 s312 s313 5 4 r31 5
s121 s122 s123 r12

i.e. the fundamental solution matrix S and the stress tensor matrix r; respectively,
and taking into account the notation (6.132), we may write
r¼S Q ð6:141Þ
If the case of the elastic space subjected to the action of a concentrated force P,
expressed by the equivalent load (6.133), we obtain the state of stress
6.2 Computation Methods 295

rij ðx1 ; x2 ; x3 Þ ¼ Pk sijk ðx1  x01 ; x2  x02 ; x3  x03 Þ; i; j ¼ 1; 2; 3: ð6:1340 Þ


In particular, if the force P is applied at the origin, it results
1 1 h xi xj xk i
rij ðx1 ; x2 ; x3 Þ ¼ 3
ð1  2mÞðxk Pk dij  xi Pj  xj Pi Þ3Pk 2
8pð1  mÞ R R
1 1 n h xi xj i o
¼ 3
xk Pk ð1  2mÞdij  3 2 2ð1  2mÞxði PjÞ ; i; j ¼ 1; 2; 3:
8pð1  mÞ R R
ð6:142Þ
Taking into account the formula (5.5), one may write
n 1 1 n xi o
p¼ ð1  2mÞ½ðR  PÞn i  ðR  nÞP i Þ  3ðR  PÞðR  nÞ ;
i 8pð1  mÞ R3 R2

wherefrom we get the stress vector corresponding to an element of area of external


normal n in the form

n 1 1 3
p¼ ð1  2mÞ½ðR  PÞn  ðP  nÞR  ðR  nÞP  ðR  PÞðR  nÞR ;
8pð1  mÞ R3 R2
ð6:143Þ
starting from (5.500 ) and using the formula (6.1350 ), we get the same result.
We mention that the above results may be put in connection with Green’s matrix,
which allows a formal solution of the three fundamental problems of the linear
elastostatics. One can make the connection also with Somigliana’s fundamental
formula, based on Betti’s reciprocity theorems, where the Kelvin-Somigliana dis-
placement vector (6.1350 ) plays an essential rôle.
Starting from the fundamental formulae stated above, we may obtain also other
results for various types of concentrated loads; we will deal in Chap. 7 with these
problems.

6.2.4.3 The Dynamic Problem of the Theory of Elasticity

We start, for the dynamic problem, from the considerations made in Sect. 5.2 and
in Sect. 4.3. We will use Lamé’s equations to study the problem of the elastic
space 1\x1 ; x2 ; x3 \1; acted upon by arbitrary volume forces, in a solution
in displacements, or the Eqs. (5.182), (5.183) in a solution in stresses, for
arbitrary volume forces too; in the case of conservative volume forces, one may
use the Eq. (5.94).
Using the Laplace transform in distributions with respect to the time t, one may
express a particular solution of the Eq. (5.182) in the form
 
1 1 1 pR=c2
U¼ L e L½F ; ð6:144Þ
4p R
296 6 Principles and General Theorems. Computation Methods

as well, a particular solution of the equation (5.183) is given by


(  )
1 2 1 1 1  pR=c1 
X¼ c L e  epR=c2 L½Fi  : ð6:145Þ
2p 2 p2 R ;i

In both cases, the product of convolution operates with regard to the space
variables.
Analogically, a particular solution of the equation (5.94) is of the form
 
1  2m 1 1 pR=c1
x¼ L e L½v ; ð6:146Þ
4ð1  mÞp R

where the convolution product operates with space variables too.


We consider now the elastic space subjected to the action of certain dynamic
loads Fi ðx1 ; x2 ; x3 ; tÞ; i ¼ 1; 2; 3; using a prolongation with null values at the left,
we introduce the distributions defined by the functions

F i ðx1 ; x2 ; x3 ; tÞ ¼ hðtÞFi ðx1 ; x2 ; x3 ; tÞ; i ¼ 1; 2; 3: ð6:147Þ


As well, we introduce the generalized displacements
ui ðx1 ; x2 ; x3 ; tÞ ¼ hðtÞui ðx1 ; x2 ; x3 ; tÞ; i ¼ 1; 2; 3: ð6:148Þ
Putting initial conditions of the form (5.7) at the moment t0 ¼ 0; we may write

o ~
o
ui ðx1 ; x2 ; x3 ; tÞ ¼ ui ðx1 ; x2 ; x3 ; tÞ þ u0i ðx1 ; x2 ; x3 ; tÞdðtÞ;
ot ot
o2 ~
o2
u i ðx 1 ; x 2 ; x 3 ; tÞ ¼ ui ðx1 ; x2 ; x3 ; tÞ þ u_ 0i ðx1 ; x2 ; x3 ; tÞdðtÞ
ot2 ot2
_
þ u0i ðx1 ; x2 ; x3 ; tÞdðtÞ;
o2 ~
o2
ui ðx1 ; x2 ; x3 ; tÞ ¼ ui ðx1 ; x2 ; x3 ; tÞ;
oxj oxk oxj oxk

hence Lamé’s equations in distributions become


_
lh2 ui þ ðk þ lÞuj;ji þ F i þ q½u_ 0i dðtÞ þ u0i dðtÞ ¼ 0; i ¼ 1; 2; 3; ð6:149Þ

which include the initial conditions too.


Applying the Laplace transform with respect to the time and the Fourier
transform with respect to the space variables, we get

ðlak ak þ qp2 ÞF½L½ui  þ ðk þ lÞai aj F L½uj 
 
¼ F L½F i  þ q pF½u0i  þ F½u_ 0i  ; i ¼ 1; 2; 3;
6.2 Computation Methods 297

where p is a complex variable, corresponding to the Laplace transform, while


ai ; i ¼ 1; 2; 3; are complex variables corresponding to the Fourier transform.
Multiplying the equations by ai and summing, we may write

f ða1 ; a2 ; a3 ; pÞ ¼ aj F L½uj 
1 h  i
0 0
¼ a j F L½F j  þ q pF½u j  þ F[ _
u j  ;
ðk þ 2lÞak ak þ qp2
ð6:150Þ
therefore, the integral transforms will be
1
F½L½ui ðx1 ; x2 ; x3 ; tÞ ¼ 2
F L½F i 
lak ak þ qp
 
þ q pF½u0i  þ F[u_ 0i   ðk þ lÞai f ða1 ; a2 ; a3 ; pÞ; i ¼ 1; 2; 3:
ð6:151Þ

Observing that
  
kþl 1 1 1
F1 ¼ F
½ðk þ 2lÞak ak þ qp2 ðlak ak þ qp2 Þ qp2 ak ak þ p2 =c21

1 1  pR=c1 pR=c2

 ¼ e  e ;
ak ak þ p2 =c22 4pqp2 R

where we have introduced the wave propagation velocities, given by the formulae
(5.105), (5.1050 ), and using the properties of the convolution product concerning
the differentiation, we may express the Laplace transforms of the generalized
displacements in the form

1  
L½ui ðx1 ; x2 ; x3 ; tÞ ¼ ½L½F i  þ q pu0i þ u_ 0i 
4pG
1 h i
epR=c2 þ c22 ðL½F j Þ;j þ qðpu0j þ u_ 0j Þ;j
R
1 1  pR=c1  
pR=c2
2 e e ; ð6:152Þ
p R ;i
where the convolution product concerns the space variables.
The inverse Laplace transformation leads to the generalized state of displace-
ment in the form
298 6 Principles and General Theorems. Computation Methods

 
1 1
ui ðx1 ; x2 ; x3 ; tÞ ¼ L1 L½F i  epR=c2
4pG R



0 1 _ R 0 1 R
þ q ui d t  þ u_ j d t 
R c2 R c2
"   #
1 1 
þ c22 L1 L½F j  2 epR=c1  epR=c2
p R ;ij



1 R R
þ q u0j;j h t h t
R c1 c2 ;i


 
1 R R
þu_ 0j;j t  t ; i ¼ 1; 2; 3:
R c1 þ c2 þ ;i
ð6:153Þ
In the case of homogeneous (null) initial conditions, we get
(  
1 1 1 pR=c
ui ðx1 ; x2 ; x3 ; tÞ ¼ L L½F i  e 2
4pG R
"  #)
1 1  pR=c1 
2 1 pR=c2
þ c2 L L½F j  2 e e ; i ¼ 1; 2; 3;
p R ;ij

ð6:154Þ
while, in the absence of the volume forces, we have
( 


1 1 0 1_ R 0 1 R
ui ðx1 ; x2 ; x3 ; tÞ ¼ u d t  þ _
u d t 
4p c22 i R c2 i
R c2



1 R R
þ u0j;j h t h t
R c1 c2 ;i


 )
1 R R
þu_ 0j;j t  t ; i ¼ 1; 2; 3: ð6:155Þ
R c1 þ c2 þ ;i

For homogeneous initial conditions, we may successfully use a representation


by potentials; in this case, we will assume that the respective results are expressed
in distributions, the derivatives in the sense of the theory of distributions being
equal to those in the usual sense. Let be a concentrated force PðtÞ; acting at the
origin of the co-ordinate axes; we can write

F i ¼ cos bi PðtÞdðx1 ; x2 ; x3 Þ; i ¼ 1; 2; 3; ð6:156Þ


where cos bi ; i ¼ 1; 2; 3; are the direction cosines of the vector PðtÞ.
Using the formulae (6.144) and (6.145), we get

1 cos bi R
Ui ¼  P t ; i ¼ 1; 2; 3; ð6:157Þ
4p R c2
6.2 Computation Methods 299




1 2 1 R R
X¼ c2 cos bj PðtÞ t  t ; ð6:158Þ
2p ðtÞ R c1 þ c2 þ ; j

where the convolution product is performed with respect to the time variable. We
obtain thus the generalized state of displacement
(

cos bj 1 R
ui ðx1 ; x2 ; x3 ; tÞ ¼ P t dij
4pG R c2


 )
2 1 R R
þc2 PðtÞ t  t ; i ¼ 1; 2; 3;
ðtÞ R c1 þ c2 þ ; ij
ð6:159Þ
where

PðtÞ ¼ hðtÞPðtÞ: ð6:160Þ


In particular, if

PðtÞ ¼ P0 dðtÞ; ð6:161Þ


hence in the case of a concentrated force acting as a shock at the initial moment,
we get
(

P0 cos bj 1 R
ui ðx1 ; x2 ; x3 ; tÞ ¼ d t d
4pG R c2 ij


 )
2 1 R R
þc2 t  t ; i ¼ 1; 2; 3: ð6:162Þ
R c1 þ c2 þ ; ij

For P0 ¼ 1; we get the fundamental solution tensor in the sense of the theory
distributions, of components
(

1 1 R
uij ðx1 ; x2 ; x3 ; tÞ ¼ d t d
4pG R c2 ij


 )
2 1 R R
þc2 t  t ; i; j ¼ 1; 2; 3; ð6:163Þ
R c1 þ c2 þ ; ij

The state of generalized displacement, corresponding to arbitrary volume forces


and tot non-homogeneous initial conditions, is given, in this case, by
n o
_
ui ðx1 ; x2 ; x3 ; tÞ ¼ uij F j þ q½u_ 0j dðtÞ þ u0j dðtÞ ; i; j ¼ 1; 2; 3; ð6:164Þ

where the convolution product is performed with respect both to the space vari-
ables and the temporal one.
300 6 Principles and General Theorems. Computation Methods

In the absence of the volume forces, e.g. we find again the formulae (6.155).
It is interesting to notice that, for a usual function PðtÞ; corresponding to a force
which acts at the origin along the Ox1 -axis, G. C. Stokes [35] established, in 1849,
the formula (6.159), in the form
"

1 1 x21 R
u1 ðx1 ; x2 ; x3 ; tÞ ¼ P t 
4pqR c21 R2 c1



Z 1=c2 #
1 x21 R x21
þ 2 1 2 P t  13 2 kPðt  kRÞdk ;
c2 R c2 R 1=c1
"

x1 x2 1 R 1 R
u2 ðx1 ; x2 ; x3 ; tÞ ¼ P t   P t 
4pqR c21 c1 c22 c2
Z #
1=c2
þ3 kPðt  kRÞdk ; ð6:165Þ
1=c1
"

x1 x3 1 R 1 R
u3 ðx1 ; x2 ; x3 ; tÞ ¼ P t   P t 
4pqR c21 c1 c22 c2
Z 1=c2 #
þ3 kPðt  kRÞdk :
1=c1

A. E. H. Love [29] demonstrated this result by a process of passing to the limit


of a field of volume forces, dependent on the time variable. Extending a formula of
G. R. Kirchhoff [27] for the scalar waves equation, Wheeler and E. Sternberg [48]
gave an interesting integral identity, based on the formulae (6.165). This identity
was mentioned in an approximate form by Somigliana [34] in 1906, extending an
identity given earlier by himself [33], in 1894. O. Tedone [37] gave, in 1896,
similar results for the cubical dilatation and for the rigid body local rotation vector.
Finally, if we sudden apply a concentrated force at the origin of the co-ordinate
axes at the initial moment and then we maintain it constant in time, we may write
PðtÞ ¼ P0 hðtÞ; ð6:166Þ

there result the generalized displacements


(

P0 cos bj 2 R
ui ðx1 ; x2 ; x3 ; tÞ ¼ d t d
8pG R c2 ij
( "
2
2 #) )
1 R R
þ c22 t  t ; i ¼ 1; 2; 3: ð6:167Þ
R c1 þ c2 þ
;ij

If the domain occupied by the elastic body is infinite, but has at least a part of
the frontier at a finite distance or it is finite, then one can no more use the
computation method indicated above. In this case, we must use a procedure
6.2 Computation Methods 301

analogous to that given by the formulae (6.147), (6.148) in what concerns the
space variables too; the derivatives in the sense of the theory of distributions, with
respect to the space variables, will no more be equal to the corresponding ones, in
the usual sense; one must introduce distributions concentrated on surfaces.

6.2.5 Other Computation Methods

Among the methods of approximate computation reviewed in Sect. 6.2.2.3, we


will deal with some aspects of the successive approximations method as well as of
the point matching method. Other methods will be put into evidence by studying
some particular problems.

6.2.5.1 Successive Approximation Method

The iterative computation methods can be very useful in some cases. Starting from
a first approximate form of the solution, various degrees of approximation are then
obtained. The two important mathematical problems about it are: the demon-
stration of the convergence of the method and the estimation of the error (the
remainder computation).
Hereafter, we shall indicate such a method of successive approximations in case
of the non-homogeneous bodies, by using the results given by P. P. Teodorescu
[38, 39] in the dynamic case, that generalize those supplied by P. P. Teodorescu
and M. Predeleanu [43] in the static case.
Using the considerations made in Sect. 4.1.3.8, one can write the equilibrium
equations (in the static case) in the form
rij;j þ f;j rij þ Fi ¼ 0; i ¼ 1; 2; 3; ð6:168Þ

the boundary conditions in stresses become


n
pi ¼ rij nj ; i ¼ 1; 2; 3; ð6:169Þ

where we have introduced the reduced external loads


n 1 n
pi ¼ pi ; i ¼ 1; 2; 3: ð6:170Þ
E
The reduced stresses are expressed in the form
X
n
ðkÞ
rij ¼ r0ij þ rij : ð6:171Þ
k¼1

If we neglect the terms with variable coefficients from the equilibrium equa-
tions (6.168), then the latter ones will take a form similar to that of the equilibrium
equations in the case of the homogeneous bodies (3.610 ); if, in the frame of a
302 6 Principles and General Theorems. Computation Methods

solution in stresses of the problem, we add Hooke’s law (4.101) and Saint-
Venant’s equations (2.68), then we shall find the approximation of zeroth order,
which solves the problem in the case of a homogeneous body, undergoing given
volume forces and boundary conditions (6.169).
So as to find the approximation of kth order if the approximation of ðk  1Þth
order, that supplies the reduced conventional volume forces
ðkÞ ðk1Þ
F i ¼ f;j rij ; i ¼ 1; 2; 3; ð6:172Þ

is supposed known, the problem is solved now in the case of a homogeneous body
ðkÞ
acted upon by the volume forces F i ; i ¼ 1; 2; 3; with homogeneous boundary
conditions (vanishing surface loads).
Thus, the solution in stresses of the elastostatic problem of a non–homogeneous
body can be approximated by a succession of solutions in stresses of elastostatic
problems of the very same body, considered as homogeneous and acted upon by
certain conventional loads, being submitted to certain conventional boundary
conditions (zero ones, excepting the zeroth approximation).
The convergence of the above mentioned iterative process can be practically
ðnÞ
estimated by its physical interpretation—the conventional volume forces F i ; i ¼
1; 2; 3; must yield a state of stress, negligible with respect to the looked real state
of stress.
An analogous solution in displacements of the problem can be supplied by
starting from the Lamé equations (5.13).
In the dynamic case one may use an analogous method of successive approx-
imations if one admits that the reduced density and the reduced damping coefficient
are constant coefficients (q ¼ const and k ¼ const), as defined in Sect. 4.1.3.8;
to this end, the non-homogeneity corresponding to q and k must be of the same
type as the non-homogeneity of the modulus of longitudinal elasticity.

6.2.5.2 The Point Matching Method

Some of the computation methods presented above (e.g., the variational methods)
allow to obtain, with an as well as we wish approximation the potential functions
or the state of strain and stress at any point in the interior of the body. Other
methods (e.g., the method of finite differences) allow to obtain approximate values
of those functions at a finite number of points in the interior of the body, satisfying
the boundary conditions at a finite number of points too.
One can imagine a computation method which has common parts with each of
the methods considered above. The method consists in searching a function, of the
most simple form, which does verify the boundary conditions at a finite number of
points; this function, which satisfies a certain partial differential equation, will give
the state of strain and stress at any point in the interior of the body.
6.2 Computation Methods 303

For the sake of simplicity, we will consider a potential function represented by


the polynomial (e.g., one of the biharmonic functions)
X
n
Fðx1 ; x2 ; x3 Þ ¼ Pi ðx1 ; x2 ; x3 Þ; ð6:173Þ
i¼2

which contains 6 þ 10 þ 14 þ    þ 2ð2n  1Þ ¼ 2ðn2  1Þ arbitrary constants.


At a point of the boundary we put conditions in stresses or in displacements,
hence at such a point (not an angular one) we get three relations between the
constants to be determined. If the function Fðx1 ; x2 ; x3 Þ fulfils the boundary con-
ditions at m ¼ E½2ðn2  1Þ=3 points, where E½p is the greatest entire contained in
p, we get a system of 3m linear algebraic equations with 2ðn2  1Þ unknowns
(eventually, one or two of the unknowns may be taken arbitrarily).
Thus, the biharmonic function (6.173), which one determines will correspond to
a state of stress in the body given by a distribution D1 of stresses or of dis-
placements on the boundary, very close to the distribution D2 of the problem
which we study. Let be D ¼ D1  D2 the difference between the two distributions
of boundary conditions (e.g., conditions in stresses). If the number of the points on
the boundary where we have put exact conditions is sufficiently great, then the
parasitary stresses D give, in the interior of the body, a state of stress negligible
with respect to the state of stress given by the loading D2 . This approximation is as
much better as the loading D on the boundary is pieceswise self-equilibrated, so
that one can apply Saint-Venant’s principle.
The number of points at which one must put boundary conditions, as well as the
approximation due to the computation method may be determined practically,
from case to case, calculating the difference D on the boundary.
In the case of a contour with properties of symmetry and of a symmetric or
antisymmetric loading one obtains important simplifications in computation; one
may thus use the even or odd, with respect to the variables, parts of the biharmonic
polynomials, while the boundary conditions are put only on a part of the contour.
If there are necessary two or three potential functions to determined, e.g., the
state of stress, then—obviously—the computation becomes more intricated.
This computation method is called the point matching method or the collocation
method and has been widely developed under this denomination. Among its
advantages, we mention that one can study the problem of an elastic body with any
contour with a desired exactitude (the form of the contour does not complicate the
problem from the point of view of the effective computation); as well, one can put,
without any difficulty, boundary conditions either in stresses or in displacements.
In addition to the elementary representations (e.g., biharmonic polynomials), one
can use also other functions which may be more adequate for some particular
problems.
As in the case of the finite differences method, one must make a separate
calculation for each ratio between the dimensions of the considered body.
304 6 Principles and General Theorems. Computation Methods

Sometimes, one may introduce parameters of the loading to indicate, e.g., an


arbitrary position of it; this is possible also because, in the linear system of
equations to be solved, the loading has an influence only on the free terms.

References

A. Books

1. Bers, L., John, F., Schechter, M.: Partial Differential Equations. Interscience Publisher, New
York (1964)
2. Castigliano, A.: Théorie de l’équilibre des systèmes élastiques. Turin (1879)
3. Courant, R.: Partial Differential Equations. Interscience Publisher, New York (1962)
4. Courant, R., Friedrichs, K.O.: Supersonic Flow and Shock Waves. Interscience Publisher,
New York (1948)
5. Gurtin, M.E.: The Linear Theory of Elasticity. In: Encyclopedia of Physics, VIa/2. Springer,
New York (1972)
6. Kecs, W., Teodorescu, P.P.: Application of the Theory of Distributions in Mechanics. Ed.
Academiei, Bucuresßti, Abacus Press, Tunbridge Wells, Kent (1974)
7. Kirchhoff, G.R.: Gesammelte Abhandlungen. Leipzig (1882)
8. Kupradzhe, V.D., Gegelya, T.G., Basheishvili, M.O., Burchuladzhe, T.V.: Trehmernye
zadachi matematicheskoı teorii uprugosti (Three Dimensional Problems of the Mathematical
Theory of Elasticity). Izd. Tbilisskogo Univ, Tbilissi (1968)
9. Love, A.E.H.: A Teatrise on the Mathematical Theory of Elasticity, 2nd edn. University
Press, Cambridge (1934)
10. Neumann, F.: Vorlesungen } uber die Theorie der Elasticität der festen Körper und der
Lichtäthers. B. G. Teubner, Leipzig (1885)
11. Nowacki, W.: Dynamics of Elastic Systems. Chapman and Hall Ltd., London (1963)
12. Nowacki, W.: Teoria spre_zystości (Theory of Elasticity). Państ. Wydawn. Naukowe,
Warszawa (1970)
13. Sneddon, I.N.: Fourier Transforms. McGraw Hill Book Co., Inc., New York (1951)
14. Teodorescu, P.P.: Probleme spatßiale în teoria elasticitătßii (Space Problems in the Theory of
Elasticity). Ed. Academiei, Bucuresßti (1970)
15. Teodorescu, P.P.: Dynamics of Linear Elastic Bodies. Ed. Academiei, Bucuresßti, Abacus
Press, Tunbridge Wells, Kent (1975)
16. Volterra, V., Volterra, E.: Sur les distorsions des corps élastiques (théorie et applications).
Mém. Sci. Math. CXLVII. Gauthier Villars, Paris (1960)

B. Papers

17. Betti. E.: Teoria dell’elasticità. Nuovo Cimento, 7–8, 5, 69, 158; 9, 34 (1872); 10, 58 (1873)
18. Castigliano, A.: Atti R. Accad. Sci. Torino (1875)
19. Clapeyron, B.P.E.: Mémoire sur le travail des forces élastiques dans un corps solide déformé
par l’action des forces extérieures. C. R. des séances de l’Acad. des Sci. 46, 208 (1858)
20. Friedrichs, K.O., Lewy, H.: Uber } die Eindentigkeit und das Abhängigkeitsgebiet der
Lösungen beim Anfangswertproblem linearer hyperbolischer Differentialgleichungen. Math.
Annalen 98, 192 (1928)
References 305

21. Graffi, D.: Sul teorema di reciprocità nella dinamica dei corpi elastici. Mem. della Accad.
delle Sci., Bologna, ser. 10, 4, 103 (1946/47)
22. Graffi, D.: Sui teoremi di reciprocità nei fenomeni non stazionari. Atti della Accad. delle Sci.,
Bologna, ser. 11, 10, 33 (1963)
23. Green, G.: On the propagation of light in crystallized media. Trans. Cambridge Phil. Soc. 7,
121 (1841)
24. Gurtin, M.E., Sternberg, E.: A note on uniqueness in classical elastodynamics. Quart. Appl.
Math. 19, 169 (1961)
25. Gurtin, M.E., Toupin, R.A.: A uniqueness theorem for the displacement boundary—value
problem of linear elastodynamics. Quart. Appl. Math. 23, 29 (1965)
26. Iesßan, D.: On the Reciprocity Theorem in the Linear Elastodynamics. An. ßst. Univ. ‘‘Al.
I. Cuza’’, Iasßi (ser. nouă), sectß. I, Mat., 18, 193 (1972)
27. Kirchhoff, G.R.: Zur Theorie der Lichtstrahlen. Sitzungsberichte der k. Akad. der Wiss.,
Berlin, 2, 641 (1882)
28. Knops, R.J., Payne, L.E.: Uniqueness in classical elastodynamics. Arch. Rat. Mech. Anal. 27,
349 (1968)
29. Love, A.E.H.: The propagations of wave motion in an isotropic elastic solid medium. In:
Proceedings of the London Mathematical Society, ser. 2, 1, 291 (1904)
30. Maxwell, J.C.: On the reciprocal figures, frames and diagrams of forces. Trans. Roy. Soc.
Edinburgh 26, 1 (1870)
31. Rubinovicz, A.L.: Herstellung von Lösungen gemischter Randwertprobleme bei
Hyperbolischen Differenzialgleichungen zweiter Ordnung durch Zusammenst} uckelung aus
Lösungen einfacher gemischter Randwertaufgaben. Monatshefte f} ur Math. u. Phys. 30, 65
(1920)
32. Saint Venant, B. de: Mémoire sur la torsion des prismes. Mém. Acad. savants étrangers
(1885)
33. Somigliana, C.: Sulle equazioni della elasticità. Ann. di Matem., ser. 2, 17, 37 (1889)
34. Somigliana, C.: Sopra alcune formule fondamentali della dinamica dei mezzi isotropi. Atti.
R. Accad. delle Sci., Torino, 41, 869, 1071 (1905/6); 42, 387 (1906/7)
35. Stokes, G.G.: On the dynamical theory of diffraction. Trans. Cambridge Phil. Soc. 9, 1 (1849)
ßandru, N.: O deıstvii peremenyh sil v neogranichennom prostransve (On the Action of
36. S
Variable Forces in the Infinite Space). Bull. Acad. Pol. Sci., ser. Sci. Techn. 12, 45 (1964)
37. Tedone, O.: Sulle vibrazioni dei corpi solidi, omogenei ed isotropi. Mem. R. Accad. delle
Sci., Torino, 47, 181 (1896/97)
38. Teodorescu, P.P.: Uber} das kinetische Problem nichthomogener elastischer Körper. Bull. de
l’Acad. Pol. des Sci., sér. Sci. Techn. 12, 867 (1964)
39. Teodorescu, P.P.: Schwingungen der elastischen Kontinua. III. Konferenz } uber nichtlineare
Schwingungen, Berlin, 1964, Abh. der deutschen Akad. der Wiss. zu Berlin, Kl. f} ur Math.,
Phys. u. Techn., Akad. Verlag, Berlin, 29 (1965)
40. Teodorescu, P.P.: Uber} das dreidimensionale Problem der Elastokinetik. Z.A.M.M. 45, 513
(1965)
41. Teodorescu, P.P.: Sur quelques problèmes dynamiques de la théorie de l’élasticité. Rev.
Roum. Math. Pures Appl. 11, 773 (1966)
42. Teodorescu, P.P.: Stress functions in three dimensional elastodynamics. Acta Mech. 14, 103
(1972)
43. Teodorescu, P.P., Predeleanu, M.: Quelques considérations sur le problèmw des corps
élastiques hétérogènes. IUTAM Symp., Non Homogenity in Elasticity and Plasticity, 1958,
Bull Acad. Pol. Sci., sér. Sci. Techn., 7, 81 (1959)
44. Volterra, V.: Sur les vibrations des corps élastiques isotropes. Acta Mathematica 16, 161
(1894)
45. Volterra, V.: Sull’equilibrio dei corpi elastici più volte connessi. Atti R. Accad. dei Lincei,
Rendiconti, Cl. Sci. fis., mat. e nat., ser. 5, 14, 193 (1905)
306 6 Principles and General Theorems. Computation Methods

46. Volterra, V.: Sur l’quilibre des corps élastiques multiplement connexes. Ann. École Norm.
Sup., 2 me sér., 24, 401 (1907)
47. Wheeler, L.T.: Some results in the linear dynamical theory of anisotropic elastic solids.
Quart. Appl. Math. 28, 91 (1970)
48. Wheeler, L.T., Sternberg, E.: Some theorems in classical elastodynamics. Arch. Rat. Mech.
Anal. 31, 51 (1968)
49. Zaremba, S.: Sopra un teorema d’unicità relativo alla equazione delle onde sferiche. Atti R.
Accad. dei Linceu, ser. 5, Rendicanti, Cl. Sci. fis., mat. e nat., 24, 904 (1915)
Chapter 7
Introduction to the Theory of Cosserat
Type Bodies

We deal hereafter with a synthesis study on the Cosserat type bodies, insisting on
the linear elastic bodies in the case of small deformations and rotations. After some
hystorical considerations, the fundamental equations are presented and formula-
tions in displacements and stresses of the problems are given. Some important
results are then given, e.g., theorems of reciprocity, fundamental solutions corre-
sponding to concentrated loads etc.

7.1 General Equations

After a short review of the problems and of the development of the Cosserat type
bodies, the state of deformation and the states of stress and couple-stress are
considered; then, the constitutive laws are put in evidence.

7.1.1 Introduction to Cosserat Type Bodies

On the way to construct mathematical models corresponding better to the real


bodies, one may put in evidence two aspects: (i) the generalization of kinematics
by considering particles with more then three degrees of freedom or by considering
displacement gradients of higher order; (ii) the taking into consideration of the
asymmetry of the stress tensor, as well as of the couple-stresses (eventually,
hyperstresses). However, the two aspects are interdependent.
Thus appear the Cosserat type bodies with free rotations, as well as the bodies
of second order (bipolar bodies). In particular, starting from the two types of
bodies, one may obtain the Cosserat type bodies with constraint rotations.

P. P. Teodorescu, Treatise on Classical Elasticity, Mathematical and Analytical 307


Techniques with Applications to Engineering, DOI: 10.1007/978-94-007-2616-1_7,
Ó Springer Science+Business Media Dordrecht 2013
308 7 Introduction to the Theory of Cosserat Type Bodies

Mounting the scale of generalization, the Cosserat type bodiea with dislocations
and the bodies of second degree with initial stresses and deformations [196–198]
are obtained. Another generalization of the bipolar bodies is constituted by the
multipolar ones (of Green-Rivlin type [64, 67, 148]).

7.1.1.1 Introductory Considerations. Short History

Such problems as the asymmetry of the stress tensor have been mentioned in the
19th century by A. L. Cauchy [33, 34], in 1850 and 1851, B. de Saint-Venant
[150], in 1869, S.-D. Poisson [144], in 1842, Lord Kelvin [8] and W. Voigt [206],
in 1887. But the first systematic works in this direction have been published in
1909 by the brothers E. Cosserat and F. Cosserat [2], the name of which has been
later put in connection with these bodies; as a matter of fact, the Cosserat brothers
began their studies earlier, in 1896 [41, 42]. But their studies remain unnoticed for
many years, when—after 1950—the problem became a new development.
In the meantime, one can mention the studies of E. Hellinger [5], et K. Heun
[6], in 1914, as well as those of F. Klein [94] and E. Noether [131], in 1918, which
bring not hing new. In his thesis, in 1929, T. J. Jaramillo [7] puts in evidence the
fact that the action density depends on the gradients of second order too.
The possibility of asymmetry of the stress tensor has also been mentioned by E.
Reissner [146], in 1944, but on a wrong way. One must also remember that in his
lectures, in 1899, at Clark University, L. Boltzmann [1] considers the property of
symmetry of this tensor as an axion; G. Hamel [4] gives the name of Boltzmann
axiom to this property, the bodies which have properties of mechanical asymmetry
being non-Boltzmannian ones. The bodies of Cosserat type are just such bodies.
After 1950, many various researches on the energetic state of the crystalline net
have been made, having—as a consequence—the asymmetry of the stress tensor
rij ; indeed, if the work of deformation reads
1
W ¼ Aijkl ui; j uk;l ; ð7:1Þ
2
then it results
rij ¼ Aijkl uk;l : ð7:2Þ
One can mention the studies of Y. Le Corre [37–40] and of J. Laval [104] in this
direction. R. Tiffen and A. C. Stevenson [200] make somewhat different consid-
erations. N. Joel and W. A. Wooster [83, 84], E. S. Rajagopal [145] and
E. S. Krishnan and E. S. Rajagopal [97] make the criticism of these results. One
may state that the majority of these authors are not aware of the work of the
Cosserat brothers.
S. Bodaszewski [26], based on the studies of W. Burzynski [29] of 1949,
introduced, in 1953, the inertia of rotation. Other initial studies on the asymmetry
of the stress tensor are due to N. Oshima [137, 138].
7.1 General Equations 309

If one attaches to any point of the deformable body a rigid orthogonal trihedron,
then one obtains a polar body (with orientation) and one can speak about the
rotation of a point. A body having such properties is called a Cosserat type body,
where any point of which is an infinitesimal rigid; the apparition of couple-stresses
is thus explained.
In 1958, J. L. Ericksen and C. Truesdell [49] have introduced the deformable
director trihedrons, being thus led to the generalized bodies of Cosserat type. As a
continuation of a paper of C. Truesdell [204], in 1952, and of another one of
W. Noll [132], in 1958, these general theories are presented in two volumes of the
Encyclopedia of Physics by C. Truesdell and R. A. Toupin [18], in 1960, and by C.
Truesdell and W. Noll [17], in 1965. Some following studies of F. A. McClintock,
P. A. André, K. R. Schwedt and R. E. Stoeckly [36] or of F. A. McClintock [35]
do not bring anything new.
In 1958 appears an important study of W. Günther [71] on the statics and on the
kinematics of a Cosserat type body; one makes the connection with the theory of
dislocations too. Other interesting researches in this direction, after 1958, are due
to E. Kröner [9, 98–100], F. Hehl and E. Kröner [76] and R. Stojanović [172].
C. Teodosiu [194, 195] deals with the determination of stresses and couple-stresses
within bodies with dislocations. E. Reissner and F. Y. M. Wan [147] deal
with Günther’s theories in a variational form, generalizing thus some results of
P. M. Naghdi [121].
We mention also the studies of fluids of Cosserat type made by J. L. Ericksen
[45–48] and by E. L. Aero, A. N. Bulugin and E. V. Kuvshinskiı̆ [21]. Studies
concerning the plastic, elastic-plastic or viscoelastic bodies with asymmetric
mechanical properties have been made by M. Misßicu [114–119], V. A. Lomakin
and L. N. Savov [105], Savov [156] and V. D. Kubenko and N. A. Shul’ga [101].
A study on a geometrical basis has been made by Y. Yamamoto [212].
The problem of the constitutive laws, especially linear ones, has been the object
of many researches. After the paper of W. Günther [71], C. Truesdell and
R. A. Toupin [18], G. Grioli [3, 68, 69] and E. L. Aero and E. V. Kuvshinskiı̆ [22]
dealt with bodies with constraint rotations. H. Schaefer [160] considers the bodies
with free rotations in the plane case; E. V. Kuvshinskiı̆ and E. L. Aero [102], in
1963, and V. A. Palmov [139] and A. C. Eringen and E. S. Suhubi [52] (dealing
with bodies with microstructure), in 1964, consider bodies with free rotations in
the three-dimensional case. Interesting studies on the constitutive laws are due
to R. Stojanović and L. Vujośević [176] and to R. Stojanović, S. Djurić and
L. Vujośević [174].
In 1966, J. Schijve [165] makes experimental researches to determine an elastic
constant l, having the dimension of a length; he obtains, for certain metals, l ¼ 0:1
mm. G. Adomeit [20] and E. Sooś and P. P. Teodorescu [167] deal with the
determination of elastic constants in bodies of generalized Cosserat type.
S. Kaliski [85, 86] deals with a model where may appear asymmetric stresses
rij (a spatial frame). W. H. Hoppman and F. O. F. Shahman [79] give a model of
isotropic elastic body with three constants.
310 7 Introduction to the Theory of Cosserat Type Bodies

Other important problems studied are those of coupled fields. We mention thus
the work of S. Kaliski, Z. Plochocki and D. Rogula [87], of P. D. Kelly [88] and of
R. C. Dixon and A. C. Eringen [43], who deal with the coupling of the electro-
magnetic field and the field of deformations; works in this direction are due also to
J. B. Alblas [23], who considers Cosserat type bodies with electronic spin. With
the coupling of the thermic field and the field of deformations dealt chiefly
W. Nowacki [133–136].
In 1963, R. D. Mindlin [197] generalizes the classic conception of elastic body,
introducing the model of an linearly elastic body with microstructure. An year
later, R. A. Toupin [192] considers, in the same manner, the theories where appear
the couple-stresses; similar ideas can be found in the papers of A. C. Eringen and
E. S. Suhubi [52].
Detailed studies, where appear dislocations too, are due to C. Teodosiu [186–
189]. We mention, in this direction, the researches of I. A. Kunin [103] and of R.
A. Toupin [203]; R. D. Mindlin [109] comes back later to this problem.
More sophisticated models are conceived by A. E. Green and R. S. Rivlin [64,
67, 148]; there are multipolar bodies, where one takes into account gradients of
higher order of the displacements. Cz. Woźniak [209] develops a theory of fibre
bodies and A. C. Eringen [50] deals with microfluids.
In Italy, G. Grioli [3, 68–70] considers especially the problem of thermody-
namic potentials for bodies of Cosserat type. Various other studies are due to
A. Bressan [28] and to D. Galletto [54–57], the latter one dealing also with
incompressible systems with reversible transformations.
Books on linearly elastic Cosserat type bodies have been written by H. Schaefer
[13] (a cycle of lectures presented at the University of Trieste, dealing especially
with the contribution of the German school to the problem) and by G. N. Savin
[12] (lectures given at the University of Kiev on the plane problem of asymmetric
elasticity, treated by means of functions of complex variables). In Romania we
mention the syntheses of D. Iesßan [16] and N. S ßandru [16], as well the books of
P. P. Teodorescu [14, 15].
We remark also an excellent synthesis due to H. Schaefer [163], in which the
hystorical process of the development of the problem is put into evidence. Another
synthesis paper, dealing with the general problems of polar bodies has been written
by W. Barański, K. Wilmański and Cz. Woźniak [25].
Various studies on problems of a general character are due to W. T. Koiter [95]
(asymmetric elasticity with constraint rotations), Cz. Woźniak and M. Jukowski
[210] (model of an elastic subsoil with couple-stresses), Z. Weselowski [208], A.
E. Green, P. M. Naghdi and W. L. Wainright [66], P. M. Naghdi [122] (a static-
geometric analogy for bodies with couple-stresses), V. A. Gordon and L. A. Tol-
okonnikov [61], B. V. Gorskiı̆ [62] and R. D. Mindlin [111] (relation between the
second gradient of displacements and couple-stresses).
In 1967, an IUTAM-Symposium on the theory of dislocations and on the
Cosserat type bodies took place in Freudenstadt and Stuttgart in Germany; we
7.1 General Equations 311

mention the papers of R. D. Mindlin [112], A. E. Green and P. M. Naghdi [65]


(about Cosserat surfaces), G. Hermann and J. D. Achenbach [77] (dynamics of
composite materials in the frame of generalized bodies), K. Anthony, U. Essmann,
A. Seeger and H. Träuble [24] (relations between Volterra’s dislocations of second
species and Cosserat type bodies with incompatible rotations), A. C. Eringen [51]
(on micromorphic bodies), K. Kondo [96] (general considerations on bodies with
imperfections) and P. P. Teodorescu [184] (basic concentrated loads in Cosserat
type bodies).

7.1.1.2 Other Problems

We have presented shortly only some problems in this direction, but their number
is greater. We will remember other studies concerning bodies with asymmetric
elasticity.
L. P. Vinokurov and N. I. Derevianko [205] dealt with straight bars which are
not subjected to torsion; M. Sokolovski [166] considers the torsion of these bars.
The thin plates have been studied by O. Hoffman [78]. The first studies on thin
shells have been made by W. G} unther [72]; a thesis in this direction has been
written by K. D. Schade [157].
Kirichenco [93] deals with spherical bodies with axial symmetry; analogous
studies have been made by Ju. N. Podil’chuk and A. M. Kirichenko [143].
Anisotropic bodies have been considered by E. V. Kuvchinskiı̆ and E. L. Aero
[102] and by Kessel [89].
Another group of researches deal with the concentration of stresses and couple-
stresses. Let us mention a first study of R. D. Mindlin [106]. H. Neuber [128–130]
continues his previous studies [10] for bodies with symmetric elasticity.
R. Muki and E. Sternberg [120] deal with analogous problems; as well,
E. Sternberg and R. Muki [170] consider concentrations of stresses around cracks.
Concentrations of stresses around cylindrical inclusions have been studied by
Y. Weitsmann [207].
The chief results in asymmetric thermoelasticity are due to W. Nowacki [133–
136], who generalizes the reciprocity theorems of V. Ionescu-Cazimir [81, 82]; the
Green’s functions, corresponding to Nowacki’s formulation have been found by
J. Wyrwiński [211]. F. Pietras and J. Wyrwiński [142] considered the plane
thermoplastic problem in the anisotropic case.
The interest of the researches has been observed in the direction of the dynamical
problems too. D. C. Gazis and R. F. Wallis [59, 60] deal with the propagation of
weaves in crystals. G. Rymarz [149], G. Adomeit [19] and K. F. Graff and Yih-Hsing
Pao [63] consider the surface waves of Rayleigh type. Other results have been given
by R. Stojanović, S. Djurić and L. Vujośević [175] and by S. Djurić [44]. G. N. Savin
and N. A. Shul’ga [155] study the plane dynamic problem.
312 7 Introduction to the Theory of Cosserat Type Bodies

7.1.1.3 Conclusion

As a conclusions, we observe that the theory of Cosserat type bodies has been
much developed in the second part of the twentieth century; one can put in evi-
dence, firstly, the researches in the United States then the studies in Germany
(chiefly in Braunschweig, Karlsruhe and Clausthal-Zellerfeld), the Russian results
(especially in Kiev and Moskow), the Italian research (in Padova), the Romanian
studies (in Bucharest), the Polish studies (in Warsaw) etc.
An important direction of research is the theory of the most general possible
bodies (multipolar bodies, bodies of generalized Cosserat type, with or without
internal defects, with or without initial deformations); another direction of study
consists in the theory of linearly elastic body (in the case of great or small
deformations or rotations).
Nevertheless, this last direction of study generalizes the classical theory of
elasticity; thus, the methods which are generally used may by applied in this case too.
One can easily study the two-dimensional problems, using the results obtained by us
[182, 183] in the symmetric case, various plane problems (using the methods given
by us [179], previously), as well as ideas given by us [178] in a synthesis paper.
But a problem which remains to be considered consists in the experimental
determination of the elastic constants; one must obtain numerical values for the
mechanical properties of the material. It has not been made much in this direction.
It is true that one can imagine plane or spatial structures (e.g., frames of various
constructions), which may be modelled as a Cosserat type body. But it is not
sufficient. To make arise the confidence in such a theory, one must show that it can
be applied at least to a small number of bodies, in certain circumstances, being
absolutely necessary at least for a small number of cases. Thus, experimental
researches very thoroughly made, very difficult too, are to be expected for the
consolidation of a theory the beauty of which influenced its actual development.

7.1.2 State of Deformation

In the case of Cosserat type bodies, besides the displacements, the strains and the
local rotations of rigid body, appear the free rotations and the characteristics of the
deformation; the state of deformation is thus characterized by the displacement
vector u and by the free rotation vector U.

7.1.2.1 Free Rotations. Constrained Rotations. Characteristics


of Deformation

Let us consider, in a more general case, that—beside the local rotation of rigid
body, represented by the vector x—we assume the existence of a rotation of vector
U, of components Ui ¼ Ui ðx1 ; x2 ; x3 Þ; i ¼ 1; 2; 3: This rotation is called free
7.1 General Equations 313

rotation or simply rotation; it contributes to the state of deformation of the body,


which is characterized by the vectors u and U. A particle of the body has thus six
degrees of freedom. We will consider the case of infinitesimal deformations.
As in the classical case, we introduce the antisymmetric tensor of second order
(the free rotation tensor)
2 3
0 U12 U31
TU  ½Uij   4 U12 0 U23 5; ð7:3Þ
U31 U23 0

the components of which are given by


Uij ¼ Uij ðx1 ; x2 ; x3 Þ ¼ ijk Uk ; ð7:4Þ

where Ui ¼ Ui ðx1 ; x2 ; x3 Þ are the components of the free rotation vector U; con-
versely, we have
1
Ui ¼ ijk Ujk : ð7:40 Þ
2
Let us introduce the local rotation vector
X¼xU ð7:5Þ
which represents the local rotation, without the contribution of the free rotation U;
its components are Xi ¼ Xi ðx1 ; x2 ; x3 Þ; i ¼ 1; 2; 3. The antisymmetric local rota-
tion tensor of second order
2 3
0 X12 X31
TX  ½Xij   4 X12 0 X23 5; ð7:6Þ
X31 X23 0

of components
Xij ¼ Xij ðx1 ; x2 ; x3 Þ ¼ ijk Xk ; ð7:7Þ

may be considered too; conversely, we have


1
Xi ¼ ijk Xjk : ð7:70 Þ
2
Obviously, the relation (7.5) may be written also in the form
x ¼ X þ U; ð7:50 Þ
which shows that the local rotation of rigid body x is the sum of the local rotation
X and the free rotation U.
The definition relation (7.5) leads to the equivalent relation
TX ¼ Tx  TU : ð7:8Þ
314 7 Introduction to the Theory of Cosserat Type Bodies

Fig. 7.1 Deformation of an x3


angular element
P*
P**
32

32

1
N*
P

1
N **
23
23

M* N
M ( x2, x3)
O x2

Let us take again the deformation of the angular element NMP, considered in
Sect. 2.2.1.3, the sides of which are parallel to the co-ordinate axes in the plane
x1 ¼ const (Fig. 7.1); analogically, we admit that the deformation does not depend
on x1 and takes place only in this plane, being in a linear case.
After deformation, without the contribution of the free rotation U, this element
becomes N  M  P , the angles b23 and b32 , made with the co-ordinate axes and
measured from these axes, being put into evidence; if we take into consideration
the influence of the free rotation U1 (corresponding to a positive rotation in the
x2 Ox3 -plane) too, then the angular element becomes N  M  P , the sides of which
make the angles a23 and a32 with the co-ordinate axes. The angle a23 represents,
e.g., the unit sliding of two elements of line parallel to the axis Ox3 , one with
regard to the other.
One may easily write
a23 ¼ b23 þ U1 ¼ u3;2 ; a32 ¼ b32  U1 ¼ u2;3 ;

where we used also the formulae (1.60). We get thus the relations
bij ¼ uj;i  Uk ; bji ¼ ui;j þ Uk ; i 6¼ j 6¼ k 6¼ i;

if we add the relations


bii ¼ eii ¼ ui;i ð!Þ;

we have, finally,
bij ¼ uj;i  ijk Uk ; i; j; k ¼ 1; 2; 3: ð7:9Þ
7.1 General Equations 315

We can introduce thus the asymmetric tensor of second order


2 3
b11 b12 b13
Tb  ½bij   4 b21 b22 b23 5: ð7:10Þ
b31 b32 b33
We notice that
cij ¼ bij þ bji ¼ 2bðijÞ ; ð7:11Þ

1  1 
Xij ¼ bij  bji ¼ b½ij ¼ uj;i  ui;j  Uij : ð7:110 Þ
2 2
It results that the symmetric part of the tensor Tb is the tensor Te , while its
antisymmetric part is the tensor TX ; hence
Tb ¼ Te þ TX ; ð7:12Þ

generalizing thus the relation (2.60).


Analogically, one may introduce the gradient of the free rotation vector U
2 3
u11 u12 u13
Tu ¼ GradU ¼ ½uij   4 u21 u22 u23 5; ð7:13Þ
u31 u32 u33

which is an asymmetric tensor of second order; the components uij ¼


uij ðx1 ; x2 ; x3 Þ are given by

uij ¼ Uj;i ; i; j ¼ 1; 2; 3: ð7:130 Þ

The asymmetric tensors Tb and Tu will characterize the deformation of the


body; their 2  9 ¼ 18 components are called the characteristics of the deformation.
The symmetric and antisymmetric components of the tensor Tb have interesting
interpretations but for those of the tensor Tu we cannot say the same thing. One
can only mention the first invariant of the tensor Tu , i.e.

w ¼ uii ¼ divU: ð7:14Þ


It is interesting to notice that, in the classical case, the position of a particle of
the body in the actual state is characterized only by the displacement u (hence, by
three parameters). But in the general case may appear privileged directions, the
position of a particle being characterized by the rotation U too, hence by six
parameters. In other words, we attach to each particle a rigid frame of reference,
which may undergo certain rotations during the deformation.
By state of deformation we understand now the totality of displacements and
rotations (as well as the totality of the characteristics of deformation) in the
neighbourhood of a point or in the domain occupied by the deformable body.
Hence, each point of the body is an infinitesimal rigid one.
316 7 Introduction to the Theory of Cosserat Type Bodies

We discussed the above problems in the static case; in the dynamic case, all the
quantities which are involved depend also on the temporal variable t.

7.1.2.2 Equations of Continuity. Cesàro Type Formulae


for Displacements and Rotations

In the general case, the relations between characteristics of the deformation, dis-
placements and rotations can be assumed as a system of 18 equations (7.9) and
(7.130 ) with six unknowns ui ; Ui ; i ¼ 1; 2; 3. So as this system be compatible, the
characteristics of the deformation must verify certain conditions of compatibility.
Supposing that Ui are functions of class C2 , one may write (Schwarz’s theorem)
Uj;ik ¼ Uj;ki ;

wherefrom
uij;k  ukj;i ¼ 0; i; j; k ¼ 1; 2; 3; ð7:15Þ

these relations may be written also in one of the following forms


ikl uij;k ¼ 0; j; l ¼ 1; 2; 3; ð7:150 Þ

o½k ui j ¼ 0; i; j; k ¼ 1; 2; 3: ð7:1500 Þ

Analogically, assuming that ui are functions of class C2 , the relations (7.9) lead
to
bij;l  blj;i ¼ jkl uik  ijk ulk ; i; j; l ¼ 1; 2; 3; ð7:16Þ

where the relations (7.130 ) have been taken into consideration; one may write these
relations in one of the forms
ilm ðbij;l  jkl uik Þ ¼ 0; j; m ¼ 1; 2; 3; ð7:160 Þ

o½l bi j ¼ jk½l uik ; i; j; l ¼ 1; 2; 3: ð7:1600 Þ

These 18 equations represent the necessary and sufficient conditions of com-


patibility of the displacements and of the rotations in case of a simply connected
domain; they have been obtained independently by S. Kessel [88] and N. S ßandru
[153]. P. P. Teodorescu [172, 173] dealt with the problem too.
Taking into account the relation (A.38), the relations (7.16) may be written in
the form
ilm bij;l  dmj ukk þ ujm ¼ 0; j; m ¼ 1; 2; 3: ð7:17Þ

Making m ¼ j and summing, one obtains the relation


ijl bij;l þ 2ukk ¼ 0: ð7:18Þ
7.1 General Equations 317

Taking into account the decomposition (7.12) and that the product of a
symmetric tensor and an antisymmetric one, with respect to the same indices,
vanishes, one also obtains
ijl Xij;l þ 2ukk ¼ 0 ð7:180 Þ

or, in a developed form


X23;1 þ X31;2 þ X12;3 þ u11 þ u22 þ u33 ¼ 0: ð7:1800 Þ
The relations (7.17, 7.18) lead to
1
uij ¼  klm bkl;m dij þ jlm bmi;l ; ð7:19Þ
2
relation equivalent with each of the relations (7.16)–(7.1600 ); it can be written also
in the form
1
uij ¼  klm Xkl;m dij þ jlm bmi;l : ð7:190 Þ
2
Taking into account (7.150 ), the relation (7.19) leads to
1
jlm inp bmi;ln  jnp klm bkl;mn ¼ 0;
2
but
klm ekl ¼ 0
and
1
jlm inp Xmi;ln  jnp klm Xkl;mn ¼ 0;
2
for each j and p (as one can prove easily assuming j ¼ p or j 6¼ p and giving all
possible values to the indices), so that one obtains the classical conditions of
compatibility of B. de Saint-Venant (2.6800 ), which are now only necessary. We
observe that the relations (2.680 ) and (7.15) do not represent sufficient conditions
of compatibility; we must take into account the relations (7.16) or (7.19) too.
The 18 conditions of compatibility offer a particular important physical sig-
nificance. Indeed, let be two adjacent elements of a body, initially undeformed
(Fig. 7.2a); if the conditions of compatibility are not fulfilled, then the two ele-
ments (the possibility of a free rotation is also taken into account) would separate
from each other after the deformation (Fig. 7.2b) or would interpenetrate
(Fig. 7.2c). If these conditions were fulfilled, then the elements would stick to each
other (Fig. 7.2d). Therefore, it follows that, from a physical standpoint, the 18
conditions of compatibility are necessary and sufficient conditions of continuity of
the deformation (rotation included), in the case of simply connected domains.
318 7 Introduction to the Theory of Cosserat Type Bodies

* *
N1 N2
P N R P *
R*
** **
N1 N2
P ** Q* R **
* S*
M
Q M S
**
(a) Q (b) S **
P*
P* R* **
P
P** R ** R **
*
Q* ** R*
* N2* ** **N1 N2
Q N2 N1 S* ** *
**
N1 N S **
Q S ** Q**
S*
M* M *

(c) (d)

Fig. 7.2 Deformation of two elements: before deformation (a), separation (b), interpenetration
(c), continuity (d)

We observe that the relation (7.16) can be written also in the form
eij;l  elj;i þ Xij;l þ Xjl;i ¼ jkl uik  ijk ulk ; ð7:16000 Þ

by circular permutation of indices i; j; l and summing the obtained relations, we get


Xij;l þ Xjl;i þ Xli;j þ jlk uik þ lik ujk þ ijk ulk ¼ 0: ð7:18000 Þ

We notice that this relation is equivalent to the relation (7.180 ). From (7.16000 ) and
(7.18000 ), one obtains
eij;l  elj;i þ Xil;j ¼ ikl ujk ;

multiplying by iml and taking into account the relation (A.39), we have, finally,
1  
uij ¼ jkl 2o½k eli  Xkl;i ; ð7:1900 Þ
2
a relation equivalent to the relation (7.190 ), as it can be easily proved. This relation
may be written also in the form
 
1
uij ¼ jkl eli;k  Xkl;i : ð7:19000 Þ
2
Observing that (see Fig. 2.8)
Z Z Z
Ui ¼ U0i þ _ dUi ¼ U0i þ _ Ui;j dxj ¼ _ uji dxj ; ð7:20Þ
P0 P P0 P P0 P
7.1 General Equations 319

one can calculate the free rotations at an arbitrary point P, if one knows these
quantities at a point P0 , in case of a simply connected domain. Taking into account
(7.190 ), one becomes, finally.
Z Z
1
Ui ¼ U0i  jkl _ Xjk;l dxi þ ikl _ blj;k dxj : ð7:200 Þ
2 P0 P P0 P

The displacements are given by


Z Z
0 0
ui ¼ ui þ _ dui ¼ ui þ _ ui;j dxj
0 P0 P
ZP P
 
¼ u0i þ _ bji  ijk Uk dxj
0
ZP P Z
 
¼ u0i þ _ bli dxl þ ijk _ Uk d xj  nj dnj ;
P0 P P0 P

where Uk ¼ Uk ðn1 ; n2 ; n3 Þ. Integrating by parts and taking into account the relation
(7.130 ), we obtain the Cesàro type formulae
Z h
    i
ui ¼ u0i  ijk xj  x0j U0k þ _ bli  ijk xj  x0j ulk dxl ; ð7:21Þ
P0 P
0
by the aid of (7.19 ), the relation (7.21) takes the form
 
ui ¼ u0i  ijk xj  x0j U0k
Z   
 0
 1
þ _ bki  xj  xj 2o½i bjk  ijk lmn Xlm;n dxk : ð7:210 Þ
P0 P 2
Both the rotations and the displacements are thus expressed by means of the
components of the tensor Tb and of the displacements and rotations of a given
point P0 . The integrals do not depend on the path if the equations of continuity of
the deformations are verified.
Using the relations (7.19000 ), we may write the Cesàro type formulae in the form
Z Z
0 1
Ui ¼ Ui þ ikl _ elj;k dxj  ikl _ Xkl;j dxj ; ð7:2000 Þ
P0 P 2 P0 P
 
ui ¼ u0i  ijk xj  x0j U0k
Z
   
þ _ bki  xj  x0j 2o½i ejk  Xij;k dxk ; ð7:2100 Þ
P0 P
equivalent with those obtained above.
Observing that
Z
_ Xkl;j dxj ¼ Xkl  X0kl ;
P0 P
Z
  0
  0
 0
_ Xki þ xj  xj Xij;k dxk ¼  xj  xj Xij ;
P0 P
320 7 Introduction to the Theory of Cosserat Type Bodies

where we used an integration by parts, these formulae become


Z
0 1  0

Ui ¼ Ui  ijk Xjk  Xjk þ ikl _ elj;k dxj ; ð7:20000 Þ
2 P0 P
  
ui ¼ u0i  xj  x0j ijk U0k þ X0ij
Z
  
þ _ eki  2 xj  x0j o½i ejk dxk : ð7:21000 Þ
P0 P

The antisymmetric part of the tensor Tb could be integrated; hence, the sym-
metric part leads to the classical formulae, as it has been shown by P. P. Teodo-
rescu [189].
If
Tu ¼ 0; Tb ¼ 0; ð7:22Þ

then one obtains the motion of rigid body. The formula (2.71000 ) remains valid; the
rotation is given by

U ¼ x0 : ð7:23Þ

7.1.2.3 Case of Constrained Rotations

If the antisymmetric part of the tensor Tb vanishes, i.e. if

TX ¼ 0; ð7:24Þ

then we say that we have to do with constrained rotations; in this case, the vector
U represents the constrained rotation vector while the tensor TU is the constrained
rotation tensor.
There does not exist a relation between the vectors u and U in the case of free
rotations; but, in case of constrained rotations, one has
1
U¼ curl u; ð7:25Þ
2
which is equivalent to
1
ijk Uk ¼ ðuj;i  ui;j Þ ð7:250 Þ
2
or to
b½ij ¼ 0; ð7:2500 Þ

so that
U ¼ x: ð7:25000 Þ
7.1 General Equations 321

The classical Cesàro formulae (2.70), (2.700 ) are thus valid; in case of a simply
connected domain, the integrals do not depend on the path if the Saint-Venant
conditions are fulfilled. These conditions which—in general—are only necessary
conditions of compatibility, become now also sufficient ones.

7.1.2.4 Kinematics of Deformation

We can, similarly, introduce the angular velocity of the free rotation


dU oU dxi oU
¼  þ ; ð7:26Þ
dt oxi dt ot

corresponding to the free rotation /; in the case of infinitesimal deformations,


there remains only
dU oU _
¼ ¼ U; ð7:27Þ
dt ot
of components U_ i ¼ U_ i ðx1 ; x2 ; x3 ; tÞ; i ¼ 1; 2; 3, where material co-ordinates are
used equally.
As well, we define the angular acceleration of the free rotation

d2 U o2 U €
¼ 2 ¼ U; ð7:28Þ
dt2 ot
of components U €i ¼ U€ i ðx1 ; x2 ; x3 ; tÞ; i ¼ 1; 2; 3, in the case of infinitesimal
deformations, in material co-ordinates too.
The vector relation (7.5) allows to introduce the rate of local rotations vector
_ ¼ x_  U;
X _ ð7:29Þ

of components X_ i ¼ X_ i ðx1 ; x2 ; x3 ; tÞ; i ¼ 1; 2; 3. By introducing the rate of local


rotations tensor
TX_ ¼ Tx_  TU_ ; ð7:290 Þ

of components X_ ij ¼ X_ ij ðx1 ; x2 ; x3 ; tÞ; i; j ¼ 1; 2; 3, we get the antisymmetric part


of the tensor Tb_ .
As well, one may introduce the gradient of the rate of the free rotation vector U _
in the form
_
Tu_ ¼ GradU: ð7:30Þ
The tensors Tb_ and Tu_ will characterize the rate of deformation of the body;
their 18 components are called characteristics of the rate of deformation.
322 7 Introduction to the Theory of Cosserat Type Bodies

We have thus seen how the symmetrical part and the antisymmetrical part of the
tensor Tb_ were thrown into relief; we can equally emphasize the symmetrical part
and the antisymmetrical one of the tensor Tu_ , although offering a remarkable
physical interpretation.
In the case of constrained rotations ðTX_ ¼ 0Þ one has

_ ¼ 1 curl u_ ¼ 1 curl v:
U ð7:31Þ
2 2

7.1.3 State of Stress and Couple-Stress

In the following, we shall consider problems of mechanics of stresses, including


couple-stresses. Obviously, the study made in the classical case (see Chap. 3) is
thus generalized.

7.1.3.1 General Considerations

In general, after effecting a section S through a body (Fig. 3.1a), on one of the parts
(on the left one), on the area element DA in the tangent plane at point M, beside the
n n n
stress D P, there appears D M, of components DMn (along the external normal to
n
the section at the respective point) and DMt (in the plane tangent to the section at
the point M), that makes the relation with the remote part (the right one)
n
(Fig. 7.3a). We admit that the effort D M is of the nature of a moment. We obtain
thus the mean couple-stress vector in the neighbourghood of the point M for the
area element of external normal n, in the form

x3 t
t n n
n mt m
n
Mt M
M
I
M n
I n
r dA mn n
A Mn n

x1 O x2
(a) (b)

Fig. 7.3 Couple-stresses: on the element DA (a), on the element dA (b)


7.1 General Equations 323

n
n DM
mmean ¼ ; ð7:32Þ
DA
at the limit, admitting that this exists, we find (Fig. 7.3b)
n
n n dM
m ¼ lim mmean ¼ ; ð7:33Þ
D A!0 dA
i.e. the couple-stress vector at the point M of an area element of external normal n.
n
The magnitude m of this vector represents the couple-stress at the point M. This
notion should be introduced with the same precautions, from a mathematical
standpoint, as the stress notion.
n
The component along the direction m of the couple-stress vector m shall be
n
noted with mm . By introducing the components along the three axes of co-ordi-
n n
nates mi ¼ mi ðx1 ; x2 ; x3 Þ; i ¼ 1; 2; 3, we shall be able to write
n n n n
n n
m2 ¼ mi mi ¼ m21 þ m22 þ m23 : ð7:34Þ
n n
The component of m along the external normal n shall be denoted by mn and
n
called normal (torsion) couple-stress; on the other hand, the component of m in
n
the plane tangent at M to the section S shall be noted by mt and shall be called
n n
tangential (bending) couple-stress. These components result from DMn and DMt
by a process of tending to the limit, analogous to the above indicated one
(Fig. 7.3b).
n
We shall equally introduce the notions of normal couple-stress vector mn and
n
tangential couple-stress vector mt . We shall have
n n n
m ¼ mn þ mt ; ð7:340 Þ

as well as
n n n
m2 ¼ m2n þ m2t : ð7:3400 Þ
The normal and tangential couple-stresses lead to the torsion and bending,
respectively. The normal couple-stress shall be considered positive if it corre-
n
sponds to a positive rotation (the vector mt in the same direction as n; an observer
along n will see the rotation from right to left) and negative in the opposite case.
As to the tangential couple-stress, its sign shall be specified in connection with a
system of axes of co-ordinates.
n
Now, by state of stress of the body, beside the totality of the stresses p in all
n
directions n, we shall equally mean the totality of the couple-stresses m in the
same direction. So as to understand the manner in which the couple-stresses
324 7 Introduction to the Theory of Cosserat Type Bodies

appear, we shall admit that, on an area element DA, of external normal n, the
internal forces are not uniformly distributed; that is why their resultant is not
n n
reduced to the sole stress D P but consists also in the couple-stress D M, which
leads to the couple-streses. We can even admit the existence of moments of higher
order (evidently with weaker and weaker effects), which leads to higher order
couple-stresses (hyperstresses); we shall confine ourselves to couple-stresses of the
first order (stresses of second order).
n
From a dimensional standpoint, the couple-stresses are ½m ¼ MT2 .
n n
As in the case of the stress vector p, the couple-stress vector m depends on the
n n
direction n. When p and m depend only on the direction n, we say that we have to
deal with a uniform state of stress, that will be determined by simple algebraic
equations of equilibrium; if in this case we do not take into account the volume
forces and the volume moments, we may confine ourselves to a global study of the
body. In the contrary case, we have a non-uniform state of stress and couple-stress
for which certain differential equations must be established; a local study of the
considered body should be carried out so as to complete the study made in the
previous section.
As in the classical case, the lost moments can be substituted to the volume
moments (introduced in the same manner as d’Alembert’s lost forces). Thus, we
introduce the volume inertia couple dðI dU=dtÞ=dt (analogous to the volume
inertia force); as a matter of fact, the substantial derivative of the angular
momentum is thus used. Here I is a rotation inertia moment per unit volume that
characterizes the body mass in a rotation motion.
In the case of homogeneous body, we have I ¼ const or I ¼ I ðtÞ; in the first
of that cases, the volume inertia couple is I d2 U=dt2 . But, in general, we have
I ¼ I ðx1 ; x2 ; x3 ; tÞ.
In the particular case of the infinitesimal deformations, for the substantial
derivatives, time derivatives are substituted, that supply the volume inertia couple
_
dðI dUÞ=dt; finally, when I ¼ I ðx1 ; x2 ; x3 Þ, therefore independent of time
(eventually a non-homogeneous body), the volume inertia couple assumes the
simple form I U. €
We can equally introduce the volume damping couple proportional to the
angular velocity, in the form k0 dU=dt, where the sign—appears owing to the fact
that this couple opposes the motion; here k0 is a damping coefficient per unit
volume, that—in general—can take the form k0 ¼ k0 ðx1 ; x2 ; x3 ; tÞ. In the case of
infinitesimal deformations the volume damping couple will be k0 U. _
In the case of infinitesimal, deformations that we shall consider later, the lost
moment will be supplied by

M €
 ¼ M  I U; ð7:35Þ

in the case of small undamped motions, and by


7.1 General Equations 325

€  k0 U;
M ¼ M  IU _ ð7:350 Þ
in the case of small damped motions. In general, we shall consider the first of these
case.

7.1.3.2 Stress Tensor

In the case of Cosserat type bodies, the components of the stress tensor are
introduced in the same way as in the classic case, by the formula (2.1.17). The
theorem of Cauchy, expressed by the relation (3.160 ). as well as the relation (3.20),
which specifies the variation of the stresses around a point, remain valid too.
The stress tensor Tr is expressed by (3.26) and has the components (3.27) in the
dynamic case. The normal and the tangential components of the stress vector are
still given by the relation (3.25) and (3.28), respectively.
The quadric of the normal stresses is given by the Eq. (3.30), as in the classic
case. It depends solely on the symmetric part of the stress tensor; its antisymmetric
part brings no contribution to the determination of the principal normal stresses or
to the principal directions either. Therefore, in the results from Sect. 3.2.1.1, we
must substitute the symmetrical parts ðrij þ rji Þ=2 to the tangential stresses
rij ; i 6¼ j; obviously, here also we shall obtain three triorthogonal directions, since
the equation of the third degree supplying the values of the principal normal
stresses has always real roots. About this quadric, a discussion can be made,
similar to that in the above-mentioned subsection. As to the tangential stresses
corresponding to the principal directions, they are reduced to the antisymmetric
part of the tensor Tr (they do not vanish any more). For this reason, the corre-
sponding stress vectors are no more normal to the quadric of the normal stresses at
the points under consideration and we cannot any longer build Lamé’s stress
ellipsoid.
We shall consider again the equilibrium of a small tetrahedron MM1 M2 M3 cut
out of a body (as in Sect. 3.1.2.2); by taking also into account the lost moments
M i ; i ¼ 1; 2; 3, acting at the centre of gravity G of the tetrahedron, we remark that

MG M
i ¼ Mi þ gi ; i ¼ 1; 2; 3;

with gi ! 0 for h ! 0; we can write the moment equation with regard to an axis
parallel to the axis Oxi , passing through G0 . At the limit, with h ! 0, the volume
forces vanish from the computation and one obtains

rij  rji þ ijk ðMk  I U


€ k Þ ¼ 0; i; j ¼ 1; 2; 3: ð7:36Þ

Hence, the antisymmetric part of the stress tensor does not vanish any more,
so that the tensor Tr is an asymmetric one. This result is only valid when the
couple-stresses are not taken into account.
326 7 Introduction to the Theory of Cosserat Type Bodies

M3

x3
1
m

dx3
11

12 G1
21
M3 mn 3
13
Gn n
G2 m n
22
n
M (x1,x2, x3)G M2 m2 dx2
2 n
m r m1 M2
23 M1 31

O 32 x2
G3
x1 dx1

3
m 33

M1

Fig. 7.4 Equilibrium of an elementary trihedron acted upon by couple-stresses

7.1.3.3 Couple-Stress Tensor

i
Let us consider now the couple-stress vectors m, corresponding to the co-ordinate
axis Oxi ; we denote by
i
lij ¼ m  ij ð7:37Þ

its component along the axis Oxj of unit vector ij . We obtain thus the components
lij ¼ lij ðx1 ; x2 ; x3 ; tÞ; i; j ¼ 1; 2; 3 of the couple-stress tensor
2 3
l11 l12 l13
Tl  ½lij   4 l21 l22 l23 5; ð7:38Þ
l31 l32 l33

which is an asymmetric one. The couple-stresses are called micromoments too.


7.1 General Equations 327

Let now be the tetrahedron MM1 M2 M3 on the faces of which we admit that
couple-stresses are acting (Fig. 7.4). Proceeding as in the classical case for
stresses, we get the relations
n
mi ¼ lij nj ; i ¼ 1; 2; 3; ð7:39Þ

which put in evidence the variation of the couple-stresses around a point.


The normal component of the couple-stress vector is supplied by
n
ln ¼ mn ¼ lij ni mj ; ð7:40Þ

while the tangential component reads


n
lnm ¼ mm ¼ lij ni mj : ð7:400 Þ

As in the case of the tensor Tr , we can draw a quadric of the normal couple-
stresses, supplied by

lij x0i x0j ¼ 1; ð7:41Þ

the sign will be ? or - according to whether the rotation due to the couple-stress
is positive or negative, so that the resulting surface be real. We remark that in this
case only the symmetric part of the tensor Tl occurs, its antisymmetric part having
no influence on the results.
We find thus three orthogonal principal directions, since the equation of the
third degree that supplies the values of the principal couple-stresses always has
real roots l1  l2  l3 .
As concerns the appearing invariants, we shall only emphasize the first
invariant, which is the most important,
W ¼ lii ¼ l1 þ l2 þ l3 : ð7:42Þ

7.1.3.4 Equations of Equilibrium and Motion

As in the classical case, we must see what occurs when we pass from an area
element going through the point M to an area element going through an adjacent
point. We shall thus start from a parallelepipedical element cut out of the body,
acted upon not only by stresses but by couple-stresses too.
Taking into account the variation of stresses and couple-stresses when passing
from a parallelipiped face to a neighbouring one, by formulae of Taylor type, and
considering also the volume forces and the volume moments, the inertia forces
and the inertia moments (i.e. the lost forces and the lost moments) (Fig. 7.5), and
remarking moreover that we must introduce relations permitting to pass from the
328 7 Introduction to the Theory of Cosserat Type Bodies

x3
33 + 33,3 dx3
32 + 32,3 dx3
31 + 31,3 dx3
12 11
dx3
23 + 23,2 dx2
M3
13

21 M1 M2 22 + 22,2 dx2
13 + 13,1 dx1 G
22 +
21 21,2 dx2
23
M (x1,x2, x3)
11 + 11,1 dx1 12 + 12,1 dx1
31
r dx1
32

O 33 x2

x1
dx2

Fig. 7.5 Equilibrium of an elementary parallelepiped acted upon by couple-stresses

centres of gravity of every adjacent face to the point Mðx1 ; x2 ; x3 Þ, we obtain the
equations of motion for stresses and couple-stresses in the form
rji;j þ Fi ¼ q€
ui ; ð7:43Þ

lji;j þ ikl rkl þ Mi ¼ I U


€ i; ð7:430 Þ

for i ¼ 1; 2; 3. These equations of motion are valid assuming that the stress tensor
is an asymmetric one.
These equations can also be obtained by using the flux-divergence formula
(A.95).
In the case of damped motions, the above equations become
rji;j þ Fi ¼ q€
ui þ ku_ i ; ð7:44Þ

lji;j þ ikl rkl þ Mi ¼ I U


€ i þ k0 U_ i ; ð7:440 Þ

for i ¼ 1; 2; 3.
In the static case, one obtains the equilibrium equations
rji;j þ Fi ¼ 0; ð7:45Þ

lji;j þ ikl rkl þ Mi ¼ 0; ð7:450 Þ

with i ¼ 1; 2; 3.
7.1 General Equations 329

In the absence of volume forces and volume moments, one obtains


rji;j ¼ q€
ui ; ð7:46Þ

€ i;
lji;j þ ikl rkl ¼ I U ð7:460 Þ

with i ¼ 1; 2; 3, in the classical case of motion.


Obviously, concomitantly with the disappearance of the couple-stresses and of
the volume moments, the stress tensor becomes symmetrical; we remark, however,
that the tensor Tr can be symmetrical too if, simultaneously, the following sup-
plementary conditions are fulfilled
€ i ; i ¼ 1; 2; 3
lji;j þ Mi ¼ I U

which have the same form as the relations (7.43). If the contrary occurs, then the
stress tensor is asymmetrical.

7.1.4 Constitutive Laws

In endeavouring to build mathematical models liable to correspond better to the


real bodies, two trends can be observed: of the kinematics generalization, by
considering particles with more then three degrees of freedom or of displacements
gradient of higher order, and the introducing of the asymmetry of the stress tensor.
Besides, these two tendencies are independent.
Thus were born, on one hand the Cosserat type bodies with free rotations and,
on the other hand, the bipolar bodies. In particular, starting from these two types of
bodies, we can obtain the Cosserat type bodies with constrained rotations.

7.1.4.1 The Linearly Elastic Constitutive Law

We shall admit that we have to deal with centrosymmetrical homogeneous bodies


(every point is a centre of elastic symmetry with properties of linear elasticity); we
introduce thus the constitutive law (corresponding to the general case of the free
rotations)
rij ¼ khdij þ 2lbðijÞ þ 2ab½ij ; ð7:47Þ

lij ¼ k0 wdij þ 2l0 uðijÞ þ 2a0 u½ij ; ð7:470 Þ

where the components of the tensor Tr are only related to the components of the
tensor Tb , while the components of the tensor Tl depend only on the components
of the tensor Tu .
330 7 Introduction to the Theory of Cosserat Type Bodies

Here h and w are given by (2.64) and (7.14); k and l are the elastic constants of
Lamé, while a is a third constant with the same dimensions ½a ¼ ML1 T2 . The
elastic constants k0 ; l0 and a0 correspond to the elastic constant k; l and a and play
a similar rôle; from the dimensional standpoint, all these constants have the same
dimensions as the above ones, but multiplied by a length squared ½k0  ¼ ½l0  ¼ ½a0 
¼ MLT2 . Generally, a linearly elastic body of the Cosserat type with free rota-
tions depends on six independent elastic constants.
If k0 ¼ l0 ¼ a0 ¼ 0, while a 6¼ 0, we are in the case of asymmetrical elasticity
without couple-stresses, with free rotations, in presence of a volume moment M.
We remark thus that the presence of the volume moments involves the appearance
of free rotations in the case of bodies with three elastic constants ðk; l; aÞ. In the
dynamic case, the relations (7.36) are verified.
Summing up the symmetric parts of the constitutive Eqs. (7.47), (7.470 ), we get
the relations
H ¼ ð3k þ 2lÞh; ð7:48Þ

W ¼ ð3k0 þ 2l0 Þw: ð7:480 Þ


Solving the relations (7.47, 7.470 ) with respect to the characteristics of the
deformations, we get
1 1
bij ¼ ½ð1 þ mÞrðijÞ  mHdij  þ r½ij ; ð7:49Þ
E 2a
1 1
uij ¼ 0
½ð1 þ m0 ÞlðijÞ  m0 Wdij  þ 0 l½ij ; ð7:490 Þ
E 2a
where the modulus of longitudinal elasticity E and Poisson’s ratio m are supplied
by the relations (4.80); analogically, the elastic constants

l0 ð3k0 þ 2l0 Þ 0 k0
E0 ¼ ; m ¼ ð7:50Þ
k0 þ l0 2ðk0 þ l0 Þ

are introduced, the first one of them having the dimension ½E0  ¼ LMT2 , while
the second one is dimensionless.
These relations can be equally expressed inversely in the form
m0 E 0 E0
k0 ¼ ; l0
¼ ¼ G0 : ð7:500 Þ
ð1 þ m0 Þð1  2m0 Þ 2ð1 þ m0 Þ
It is often useful to introduce an elastic constant l, with the dimension of a
length ð½l ¼ LÞ, expressed by
ðl þ aÞðl0 þ a0 Þ
l2 ¼ ; ð7:51Þ
4la
7.1 General Equations 331

this constant characterizes, in the case of certain problems, the mechanical


asymmetry of the material.
On the other hand, the elastic constant h is used, the dimension of which is also
of a length ð½h ¼ LÞ and which is expressed by

k0 þ 2l0
h2 ¼ : ð7:52Þ
4a
In the case of asymmetrical elasticity without couple-stresses ðk0 ¼ l0 ¼ a0 ¼ 0;
a 6¼ 0Þ, we have l ¼ h ¼ 0; and the same occurs in the case of classical elasticity.

7.1.4.2 Case of Constrained Rotations

We can pass to the case of constrained rotations (for which TX ¼ 0) by assuming


that a ! 1 (which involves the indetermination of the antisymmetrical part of the
tensor Tr ) and k0 ! 1 (which involves the indetermination of the spherical part
of the tensor Tl ); indeed, from (7.14) and (7.25) it follows, in this case, that
1
w ¼ div curl u ¼ 0: ð7:53Þ
2
Consequently, the constitutive law will depend, in case of the constrained rota-
tions, on only four distinct elastic constants k; l; l0 ; a0 .
The additional conditions l0 ¼ a0 ¼ 0 lead to the case of classical (symmetri-
cal) elasticity. It is to be remarked that we cannot pass to the case of classical
elasticity simply by making a ¼ k0 ¼ 0; in this case we obtain a theory of sym-
metric elasticity with free rotations.
For a ! 1, we obtain l ! l0 ðl  l0 Þ with
l0 þ a0
l02 ¼ ; ð7:510 Þ
4l
a relation corresponding to constrained rotations. Likewise, one has h ¼ 0, in this
case.

7.2 Formulations of the Static and Dynamic Problems.


General Theorems

The Eqs. (7.45), (7.450 ) and the Eqs. (7.44), (7.440 ) form a complete system of 18
equations for the linearly elastic bodies of the Cosserat type, in the static and in the
dynamic case, respectively; to these equations, we must add the limit conditions
(boundary conditions and initial conditions).
332 7 Introduction to the Theory of Cosserat Type Bodies

x3 x3
n
p n
n u n
n n
m
r M r M
dA dA

x1 x2 x1 x2
O O
(a) (b)

Fig. 7.6 Fundamental problems. Boundary conditions: first (a), second (b)

In the case of the first fundamental problems, the boundary conditions are put in
displacements and rotations, in the form (Fig. 7.6a)

ui ¼   i ; i ¼ 1; 2; 3;
ui ; U i ¼ U ð7:54Þ
for a solution in displacements and rotations. The second fundamental problem
requires a solution in stresses and couple-stresses, so that the boundary conditions
are put consequently in the form (Fig. 7.6b)
n n
pi ¼ rji nj ; mi ¼ lji nj ; i ¼ 1; 2; 3: ð7:55Þ

In the case of the mixed problem, one puts conditions of the form (7.54) on one
part of the contour and conditions of the form (7.55) an another part of it.
W. T. Koiter [95] has shown that, in case of the Cosserat type bodies with
constrained rotations, the number of boundary conditions at a point of the contour
is reduced from 6 to 5.
In case of a dynamic problem, initial conditions of the form

ui ðx1 ; x2 ; x3 ; t0 Þ ¼ u0i ðx1 ; x2 ; x3 Þ; u_ i ðx1 ; x2 ; x3 ; t0 Þ ¼ u_ 0i ðx1 ; x2 ; x3 Þ; ð7:56Þ

Ui ðx1 ; x2 ; x3 ; t0 Þ ¼ U0i ðx1 ; x2 ; x3 Þ; U_ i ðx1 ; x2 ; x3 ; t0 Þ ¼ U_ 0i ðx1 ; x2 ; x3 Þ; ð7:560 Þ

where u0i ; U0i ; u_ 0i ; U_ 0i are the displacements and the rotations and their velocities,
respectively, at an arbitrary point of the body.

7.2.1 Formulation of the Static Problem

Replacing the constitutive law (7.47), (7.470 ) in the equations of equilibrium


(7.45), (7.450 ) and taking into account the relations (7.9), (7.130 ) between the
characteristics of deformation and the displacements and rotations, we get the
equations in displacements and rotations of Lamé’s type, in the static case
7.2 Formulations of the Static and Dynamic Problems. General Theorems 333

ðl þ aÞDui þ ðk þ l  aÞuj;ji þ 2aijk Uj;k þ Fi ¼ 0; ð7:57Þ

ðl0 þ a0 ÞDUi þ ðk0 þ l0  a0 ÞUj;ji þ 2aijk uj;k  4aUi þ Mi ¼ 0; ð7:570 Þ

where i ¼ 1; 2; 3. In a vector form, we may write


ðl þ aÞDu þ ðk þ l  aÞgrad divu þ 2a curlU þ F ¼ 0; ð7:58Þ

ðl0 þ a0 ÞDU þ ðk0 þ l0  a0 Þgrad divU þ 2a curlu  4aU þ M ¼ 0; ð7:580 Þ

7.2.1.1 Galerkin’s Type Representation

In 1966, N. S
ßandru [153] gives a representation of Galerkin’s type for the vector
unknowns of the problem

u ¼ ðk þ 2lÞD½ðl0 þ a0 ÞD  4aC1  ½ðl0 þ a0 Þðk þ l  aÞD


 4aðk þ lÞ grad divC1  2a½ðk0 þ 2l0 ÞD  4a curlC2 ; ð7:59Þ

U ¼ ðl þ aÞD½ðk0 þ 2l0 ÞD  4aC2  ½ðl þ aÞðk0 þ l0  a0 ÞD


 4a2  grad divC2  2aðk þ 2lÞD curlC1 ; ð7:590 Þ

where the vector potentials Ck ¼ Ck ðx1 ; x2 ; x3 Þ; k ¼ 1; 2, satisfy the equations

ðk þ 2lÞD2 ½ðl þ aÞðl0 þ a0 ÞD  4laC1 þ F ¼ 0; ð7:60Þ

D½ðk0 þ 2l0 ÞD  4a½ðl þ aÞðl0 þ a0 ÞD  4laC2 þ M ¼ 0: ð7:600 Þ


Let us assume that the volume moment vanishes ðM ¼ 0Þ. In the case of an
irrotational field of volume forces, one has
F ¼ gradP0 ; ð7:61Þ

in this case, the state of displacement and rotation is specified by only one scalar
potential
u ¼ gradK0 ; U ¼ 0; ð7:610 Þ
where
ðk þ 2lÞDK0 þ P0 ¼ 0: ð7:6100 Þ
In the case of a solenoidal field of volume forces, hence if
F ¼ curlP; ð7:62Þ
334 7 Introduction to the Theory of Cosserat Type Bodies

then the state of displacement and rotation reads

u ¼ ½ðl0 þ a0 ÞD  4a curlK; U ¼ 2a curlcurlK; ð7:620 Þ


where the vector potential K verifies the potential differential equations

D½ðl þ aÞðl0 þ a0 ÞD  4laK þ P ¼ 0: ð7:6200 Þ


A second hypothesis is that of a vanishing volume force ðF ¼ 0Þ. If the field of
volume moments is irrotational, i.e.

M ¼ gradP00 ; ð7:63Þ

then the state of displacement and rotation is given by

u ¼ 0; U ¼ gradK00 ; ð7:630 Þ

the scalar potential satisfying the equation

½ðk0 þ 2l0 ÞD  4aK00 þ P00 ¼ 0: ð7:6300 Þ

In the case of a solenoidal field of volume forces

M ¼ curlP0 ð7:64Þ
one introduces a vector potential, which leads to the state of displacement and
rotation

u ¼ 2a curlcurlK0 ; U ¼ ðl þ aÞD curlK0 ð7:640 Þ


and verifies the partial differential equation

D½ðl þ aÞðl0 þ a0 ÞD  4laK0 þ P0 ¼ 0: ð7:6400 Þ


By superposition of effects, one obtains the results for a general case of volume
forces and moments.

7.2.1.2 Papkovich-Neuber Type Representation

One may obtain also a representation of Papkovich-Neuber type of the form

u ¼ ½ðl0 þ a0 ÞD  4av
ðk þ l  aÞðl0 þ a0 ÞD  4aðk þ lÞ
 gradðr  v þ v0 Þ  2a curlv0 ; ð7:65Þ
2ðk þ 2lÞ
7.2 Formulations of the Static and Dynamic Problems. General Theorems 335

U ¼ ½ðl þ aÞD  4av0


ðk0 þ l0  a0 Þðl þ aÞD  4aðk0 þ l0 Þ
 gradðr  v0 þ v00 Þ  2a curlv;
2ðk0 þ 2l0 Þ
ð7:650 Þ

where the vector potentials v ¼ vðx1 ; x2 ; x3 Þ and v0 ¼ v0 ðx1 ; x2 ; x3 Þ satisfy the


equations

Dðl2 D  1Þv ¼ 0; Dðl2 D  1Þv0 ¼ 0; ð7:6500 Þ

and the scalar potentials are given by


1 0 
Dv0 þ r  Dv ¼ 0; Dv00 þ r  Dv0 ¼ 2
v0 þ r  Dv0 : ð7:65000 Þ
h
We mention that all the potentials are of the class C5 . These representations are
complete for a simply connected domain.

7.2.1.3 Case of Constrained Rotations

In the case of constrained rotations ða ¼ k0 ¼ 0Þ, the Lamé type Eqs. (7.58),
(7.580 ) become
lDu þ ðk þ lÞ grad divu þ F ¼ 0; ð7:66Þ

ðl0 þ a0 ÞDU þ ðl0  a0 Þ grad divU þ M ¼ 0; ð7:660 Þ


the Eq. (7.66) is identical with that in the classical case.
The representation of Sßandru (7.59), (7.590 ) becomes
u ¼ ðk þ 2lÞDC1  ðk þ lÞ grad divC1 ; ð7:67Þ

U ¼ 2DC2  grad divC2 ; ð7:670 Þ

where the vector potentials Ck ¼ Ck ðx1 ; x2 ; x3 Þ; k ¼ 1; 2, verify the equations

lðk þ 2lÞD2 C1 þ F ¼ 0; ð7:68Þ

2ðl0 þ a0 ÞD2 C2 þ M ¼ 0: ð7:680 Þ


We mention that we have introduced some elastic constants in the potential
functions, as well as the operator D. The representation (7.67) corresponds to the
classical one.
One can make analogous considerations for the Papkovich-Neuber type
representation.
336 7 Introduction to the Theory of Cosserat Type Bodies

7.2.2 Formulation of the Dynamic Problem

Hereafter we shall deal with certain formulations of the problems of the Cosserat
type bodies with the help of potential functions; we shall thus consider both the
cases of free and constrained rotations.

7.2.2.1 Equations of Lamé’s Type

The displacements and rotations are the unknowns in a solution in displacements


and rotations of the problems. After eliminating the stresses, the couple-stresses
and the characteristics of the deformation, we find the Lamé type equations in the
form
ðl þ aÞh2 u þ ðk þ l  aÞ grad divu þ 2a curlU þ F ¼ 0; ð7:69Þ

ðl0 þ a0 Þh02 U þ ðk0 þ l0  a0 Þ grad divU þ 2a curlu þ M ¼ 0; ð7:690 Þ

we applied the differential operators

1 o2 0 1 o2 1a
h2 ¼ D  ; h 2 ¼ D   02 ; ð7:70Þ
c22 ot2 c02
2 ot 2 l l

where the wave propagation velocities are supplied by


l þ a 02 l0 þ a0
c22 ¼ ; c2 ¼ : ð7:700 Þ
q I
When applying the differential operators

1 o2 1 o2 1
h1 ¼ D  2 2
; h01 ¼ D  02 2  2 ; ð7:71Þ
c1 ot c1 ot h

the wave propagation velocities will be supplied by

k þ 2l 02 k0 þ 2l0
c21 ¼ ; c1 ¼ : ð7:710 Þ
q I
We remark that these operators are related to one another by the relations
ðk þ 2lÞh1  ðl þ aÞh2 ¼ ðk þ l  aÞD; ð7:72Þ

ðk0 þ 2l0 Þh01  ðl0 þ a0 Þh02 ¼ ðk0 þ l0  a0 ÞD: ð7:720 Þ


Applying the divergence operator to the Eqs. (7.69), (7.690 ), we find
ðk þ 2lÞh1 h þ divF ¼ 0; ð7:73Þ
7.2 Formulations of the Static and Dynamic Problems. General Theorems 337

ðk0 þ 2l0 Þh01 w þ divM ¼ 0; ð7:730 Þ

i.e. the equation that should be verified by first invariants of the tensors Te and Tu ;
the first equation coincides with the corresponding one in the classical case.
In the absence of the volume forces and moments, the Lamé type equations
become
ðl þ aÞh2 u þ ðk þ l  aÞ grad div u þ 2a curlU ¼ 0; ð7:74Þ

ðl0 þ a0 Þh02 U þ ðk0 þ l0  a0 Þ grad divU þ 2a curlu ¼ 0 ð7:740 Þ

and the two invariants verify the equations

h1 h ¼ 0; h01 w ¼ 0: ð7:75Þ
0 0
Taking into account the relation (A.91 ), the Eqs. (7.74), (7.74 ) can be equally
expressed in the form
ðk þ 2lÞh1 u þ ðk þ l  aÞ curl curlu þ 2a curlU ¼ 0; ð7:76Þ

ðk0 þ 2l0 Þh01 U þ ðk0 þ l0  a0 Þ curl curlU þ 2a curlu ¼ 0: ð7:760 Þ


Applying the operator div, one obtains the same Eq. (7.75).

7.2.2.2 S
ßandru’s Representation

N. S
ßandru [153] gave a representation of Iacovache [80] type of the solutions of
the Eqs. (7.69), (7.690 ).
Eliminating the rotation vector and the displacement vector, respectively,
between the Eqs. (7.74), (7.740 ), we obtain

½ðl þ aÞh2 þ ðk þ l  aÞ grad div][ðl0 þ a0 Þh02 þ ðk0 þ l0  a; Þ grad div]u


 4a2 curlu ¼ 0;
ð7:77Þ

½ðl þ aÞh2 þ ðk þ l  aÞ grad div][ðl0 þ a0 Þh02 þ ðk0 þ l0  a; Þ grad div]U


 4a2 curlU ¼ 0;
ð7:770 Þ
We notice that the first of these equations, e.g., may be written in the form

Du þ ðD1 þ D02 Þ grad divu ¼ 0 ð7:78Þ

or in the form

Du þ ðD01 þ D2 Þ grad divu ¼ 0; ð7:780 Þ


338 7 Introduction to the Theory of Cosserat Type Bodies

with

D ¼ ðl þ aÞðl0 þ a0 Þh2 h02 þ 4a2 M; ð7:79Þ

D1 ¼ ðk0 þ l0  a0 Þðk þ 2lÞh1 ;


ð7:790 Þ
D01 ¼ ðk þ l  aÞðk0 þ 2l0 Þh01 ;

D2 ¼ ðk0 þ l0  a0 Þðl þ aÞh2  4a2 ;


ð7:7900 Þ
D02 ¼ ðk þ l  aÞðl0 þ a0 Þh02  4a2 :
We notice that

D1 þ D02 ¼ D01 þ D2 ¼ ðk þ 2lÞðk0 þ l0  a0 Þh1


þ ðk0 þ 2l0 Þðk þ l  aÞh01 þ ðk þ l  aÞðk0 þ l0  a0 ÞD  4a2
¼ ðl þ aÞðk0 þ l0  a0 Þh2 þ ðl0 þ a0 Þðk þ l  aÞh02
þ ðk þ l  aÞðk0 þ l0  a0 ÞD  4a2 : ð7:79000 Þ
Introducing the notation
divu ¼ Du; ð7:80Þ

from (7.78) it follows that

u ¼ ,  ðD1 + D02 Þgradu; ð7:81Þ

the function , ¼ ,ðx1 ; x2 ; x3 ; tÞ being given by


D, ¼ 0: ð7:82Þ
Introducing the relation (7.81) in (7.80), we find the condition

ðk þ 2lÞðk0 þ 2l0 Þh1 h01 u ¼ div,; ð7:83Þ

we adopt the notation

, ¼ ðk þ 2lÞðk0 þ 2l0 Þh1 h01 C; ð7:84Þ

where the potential vector C must verify the equation

h1 h01 DC ¼ 0: ð7:85Þ
In this case, it follows that
u ¼ divC; ð7:86Þ
where we neglected a function C0 which verifies the equation
7.2 Formulations of the Static and Dynamic Problems. General Theorems 339

h1 h01 C0 ¼ 0 ð7:87Þ

and which can be included in the vector potential C.


We obtain thus the representation

u ¼ ðk þ 2lÞðk0 þ 2l0 Þh1 h01 C  ðD1 + D02 Þ grad divC; ð7:88Þ

that is complete for the Eq. (7.78) in the case of a simply connected domain (this
result has been obtained step by step, starting from the Eq. (7.78)); the vector
potential C ¼ Cðx1 ; x2 ; x3 ; tÞ should verify the Eq. (7.85).
Similarly, for the function U, which verifies an equation analogous to the Eq.
(7.78), we get the complete representation

U ¼ ðk þ 2lÞðk0 þ 2l0 Þh1 h01 C0  ðD01 þ D2 Þ grad divC0 ; ð7:880 Þ

where the potential function C0 ¼ C0 ðx1 ; x2 ; x3 ; tÞ satisfies the equation

h1 h01 DC0 ¼ 0: ð7:850 Þ


These representations are equivalent and can, eventually, be used for each
other.
Taking into account the relations (7.730 ) and (A.910 ), the representation (7.88)
reads

u ¼ ðk þ 2lÞðl0 þ a0 Þh1 h02 C  D02 grad divC  D1 curl curlC; ð7:89Þ

analogically, we obtain

U ¼ ðk0 þ 2l0 Þðl þ aÞh01 h2 C0  D2 grad divC0  D01 curl curlC0 : ð7:890 Þ
We remark, however, that the representation (7.88), (7.880 ) should verify the
Eqs. (7.74), (7.740 ), from which we started (the Eqs. (7.77), (7.770 ) are only their
consequences and represent necessary conditions).
Taking into account the relations

divu ¼ DdivC; divU ¼ DdivC0 ; ð7:90Þ

the Eqs. (7.75) show that the potential functions C and C0 must verify the equations

h1 DC ¼ 0; h01 DC0 ¼ 0; ð7:900 Þ

which represent more restricted conditions than (7.85) and (7.850 ); as a matter of
fact, these conditions will nevertheless be fulfilled. We remark moreover that we
have neglected a component in the form of a curl of the vectors C and C0 ,
respectively, that vanishes in the relations (7.90).
In this case, we remark that, in the representations (7.89), (7.890 ), we can neglect
the terms in curl, verifying the Eqs. (7.77), (7.770 ) by means of the conditions
(7.900 ). Moreover, we remark that we can introduce additional terms such as
340 7 Introduction to the Theory of Cosserat Type Bodies

u ¼ curlW; ð7:91Þ

U ¼ curlW0 ; ð7:910 Þ

with the condition that the potentials W and W0 are supplied by

DW ¼ 0; DW0 ¼ 0; ð7:92Þ
so that the Eqs. (7.77), (7.770 ) should be verified. In point of fact, these additional
terms correspond to the passage from the Eqs. (7.74), (7.740 ) to the Eqs. (7.77),
(7.770 ).
Now, laying the condition that the Eqs. (7.74), (7.740 ) should be verified by the
representation (7.89), (7.890 ) with the addition of (7.91), (7.910 ), respectively, we
obtain

ðk þ 2lÞh1 DC þ ðl þ aÞh2 curl½W þ 2aðk0 þ 2l0 Þh01 C0 


þ 2a curl curl½W0 þ 2aðk þ 2lÞh1 C ¼ 0; ð7:93Þ

ðk0 þ 2l0 Þh01 DC0 þ ðl0 þ a0 Þh02 curl½W0 þ 2aðk þ 2lÞh1 C


þ 2a curl curl½W þ 2aðk0 þ 2l0 Þh01 C0  ¼ 0: ð7:930 Þ

Taking into account (7.900 ), we see that these conditions are fulfilled when

W ¼ 2aðk0 þ 2l0 Þh01 C0 ; ð7:94Þ

W0 ¼ 2aðk þ 2lÞh1 C; ð7:940 Þ

the conditions (7.92) will also be fulfilled, since they agree with the conditions
(7.900 ).
In this case, we obtain the representation of N. ßSandru

u ¼ ðk þ 2lÞðl0 þ a0 Þh1 h02 C


 D02 grad divC  2aðk0 þ 2l0 Þh01 curlC0 ; ð7:95Þ

U ¼ ðk0 þ 2l0 Þðl þ aÞh01 h2 C0


 D2 grad divC0  2aðk þ 2lÞh1 curlC; ð7:950 Þ

where the vector potentials C and C0 are functions of class C6 that verify the
Eq. (7.900 ) in the absence of the volume forces and moments; this representation is
complete for a simply connected domain.
If the volume forces and moments are not vanishing, i.e. in the case of the
system of Eqs. (7.69), (7.690 ), we shall use the same representation, and the
potential functions must verify the equations

h1 DC þ F ¼ 0; h01 DC0 þ M ¼ 0: ð7:96Þ


7.2 Formulations of the Static and Dynamic Problems. General Theorems 341

From (7.900 ), (7.95), (7.950 ) we see that, in the absence of the volume forces and
moments, the displacement and the rotation vectors verify the equations

h1 Du ¼ 0; h01 DU ¼ 0: ð7:97Þ
Since the operators h1 ; h01 and D are prime with one another, we can use
Boggio’s theorem; it follows that
u ¼ u 1 þ u2 ; U ¼ U 1 þ U 2 ; ð7:98Þ

with

h1 u1 ¼ 0; h01 U1 ¼ 0: ð7:99Þ

Du2 ¼ 0; DU2 ¼ 0; ð7:990 Þ

which corresponds to a decomposition into two types of waves. We remark that the
displacements verifying the first Eq. (7.99) correspond to the classical case (lon-
gitudinal waves).

7.2.2.3 Particular Cases of Volume Loads

We shall admit that the volume moment vanishes


M ¼ 0: ð7:100Þ
Let us consider an irrotational field of volume forces, therefore of the form
F ¼ gradP0 ; ð7:101Þ
in this case, the state of displacement and rotation will be specified by a single
scalar potential K0 ¼ K0 ðx1 ; x2 ; x3 ; tÞ in the form
u ¼ gradK0 ; U ¼ 0; ð7:102Þ

where
ðk þ 2lÞh1 K0 þ P0 ¼ 0: ð7:1020 Þ
In the case of a solenoidal field of volume forces, therefore of the form
F ¼ curlP; ð7:103Þ

the state of displacement and rotation is supplied by

u ¼ ðl0 þ a0 Þh02 curlU; U ¼ 2a curl curlK; ð7:104Þ


342 7 Introduction to the Theory of Cosserat Type Bodies

where the vector potential K ¼ Kðx1 ; x2 ; x3 ; tÞ satisfies the partial differential


equation
DK þ P ¼ 0: ð7:1040 Þ
We shall also consider a second hypothesis, namely that the volume force
vanishes
F ¼ 0; ð7:105Þ

in the case of an irrotational field of volume moments

M ¼ gradP00 ; ð7:106Þ

the state of displacement and rotation is expressed by

u ¼ 0; U ¼ gradK00 ; ð7:107Þ

the scalar potential K00 ¼ K00 ðx1 ; x2 ; x3 ; tÞ being supplied by the equation

ðk0 þ 2l0 Þh01 K00 þ P00 ¼ 0: ð7:1070 Þ


In the case of a solenoidal field of volume forces

M ¼ curlP0 ; ð7:108Þ

a potential vector K0 ¼ K0 ðx1 ; x2 ; x3 ; tÞ is introduced, expressing the state of dis-


placement and rotation

u ¼ 2a curl curlK0 ; U ¼ ðl þ aÞh2 curlK0 ; ð7:109Þ


and verifying the equation

DK0 þ P0 ¼ 0: ð7:1090 Þ
By the method of superposition of effects, we can obtain the results corre-
sponding to a general loading case with volume forces and moments.

7.2.2.4 Representation of Lamé-Clebsch Type

Starting from the S


ßandru representation, we shall offer—hereafter—a representa-
tion of the Lamé-Clebsch type. For instance, in the absence of volume forces and
moments, we can, by means of Boggio’s theorem, express the potential vectors C
and C0 in the form

C ¼ , þ ,1 ; C0 ¼ ,0 þ ,01 ; ð7:110Þ

where
7.2 Formulations of the Static and Dynamic Problems. General Theorems 343

D, ¼ 0; h1 ,1 ¼ 0; ð7:111Þ

D,0 ¼ 0; h01 ,01 ¼ 0; ð7:1110 Þ

in this case, the state of displacement and rotation will be supplied by

u ¼ ðk þ 2lÞðl0 þ a0 Þh1 h02 ,  D02 grad ðdiv, þ ,0 Þ


 2aðk0 þ 2l0 Þh01 curl,0 ; ð7:112Þ
0
U ¼ ðk0 þ 2l0 Þðl þ aÞh01 h2 ,0  D2 grad ðdiv,0 þ ,0 Þ
 2aðk þ 2lÞh1 curl,: ð7:1120 Þ

We introduced here the scalar potentials

,0 ¼ div,1 ; ,00 ¼ div,01 ; ð7:113Þ

that verify the equations

h1 ,0 ¼ 0; h01 ,00 ¼ 0; ð7:1130 Þ

by so doing, the state of displacement and rotations is expressed by means of two


vector potentials , ¼ ,ðx1 ; x2 ; x3 ; tÞ, ,0 ¼ ,0 ðx1 ; x2 ; x3 ; tÞ of class C6 and of two
scalar potentials ,0 ¼ ,0 ðx1 ; x2 ; x3 ; tÞ, ,00 ¼ ,00 ðx1 ; x2 ; x3 ; tÞ of class C5 .
Let us remark that the scalar potentials express the state of displacement and
rotation

u ¼ D02 grad,0 ; U ¼ D02 grad,00 ; ð7:114Þ

corresponding to irrotational waves that verify the Eqs. (7.99); on the other hand,
the state of displacement and rotation expressed solely by the vector potentials ,
and ,0 is solenoidal (h ¼ 0; w ¼ 0, therefore incompressible) and verifies the Eqs.
(7.990 ). This can constitute an extension of the decomposition into two wave types,
from the classical case, a fact realized in the previous subsection too.
Taking into account the relations (7.72), (7.720 ), (7.79), (7.7900 ), the represen-
tation (7.112), (7.1120 ) can be written as follows

u ¼ D02 ½D,  grad ðdiv, þ ,0 Þ  2aðk0 þ 2l0 Þh01 curl,0 ; ð7:115Þ

U ¼ D2 ½D,0  grad ðdiv,0 þ ,00 Þ  2aðk þ 2lÞh1 curl,; ð7:1150 Þ

by introducing the scalar potentials

,0 ¼ D02 ,0 ; ,00 ¼ D02 ,00 ; ð7:116Þ

that verify the equations


344 7 Introduction to the Theory of Cosserat Type Bodies

h1 ,0 ¼ 0; h01 ,00 ¼ 0; ð7:1160 Þ

and making use of the differential relations (A.910 ), we can moreover express the
state of displacement and rotation in the form

u ¼ grad,0  curl½D02 curl, þ 2aðk0 þ 2l0 Þh01 ,0 ; ð7:117Þ

U ¼ grad,00  curl[D2 curl,0 þ2aðk þ 2lÞh1 ,; ð7:1170 Þ

which may constitute a generalization, for the Cosserat type bodies, of the classical
Lamé-Clebsch representation. We introduce thus two vector potentials , ¼
,ðx1 ; x2 ; x3 ; tÞ and ,0 ¼ ,0 ðx1 ; x2 ; x3 ; tÞ of class C6 and two scalar potentials ,0 ¼
,0 ðx1 ; x2 ; x3 ; tÞ and ,00 ¼ ,00 ðx1 ; x2 ; x3 ; tÞ of class C3 .
With the help of the results of previous subsection, we can write a represen-
tation of the same type in the form

u ¼ gradK0 þ curl½ðl0 þ a0 Þh02 K  2a curlK0 ; ð7:118Þ

U ¼ gradK00 þ curl[(lþaÞh2 K0  2a curlK; ð7:1180 Þ

where the scalar potentials should be functions of class C3 , verifying the equations

h1 K0 ¼ 0; h01 K0 0 ¼ 0 ð7:119Þ

and the vector potentials should be of class C5 and satisfy the equations

DK ¼ 0; DK0 ¼ 0: ð7:1190 Þ
Obviously, the representations (7.117), (7.1170 ) and (7.118), (7.1180 ) are
equivalent.

7.2.2.5 Representation of Sternberg-Eubanks Type

Starting from S ßandru’s representation (7.95), (7.950 ), we can give a representation


liable to extended the Papkovich-Neuber representation from the static case (the
classical problem). For this end, let us introduce the vector potentials , ¼
,ðx1 ; x2 ; x3 ; tÞ and ,0 ¼ ,0 ðx1 ; x2 ; x3 ; tÞ as well as the scalar potentials ,0 ¼
,0 ðx1 ; x2 ; x3 ; tÞ and ,00 ¼ ,00 ðx1 ; x2 ; x3 ; tÞ with the help of the relations

v ¼ ðk þ 2lÞh1 C; v0 ¼ ðk0 þ 2l0 Þh01 C0 ; ð7:120Þ

v0 ¼ 2ðk þ 2lÞdivC  r  v;
ð7:1200 Þ
v00 ¼ 2ðk0 þ 2l0 ÞdivC0  r  v0 :
Taking into account the Eqs. (7.900 ), we see that the vector potentials should verify
the equations
7.2 Formulations of the Static and Dynamic Problems. General Theorems 345

Dv ¼ 0; Dv0 ¼ 0; ð7:121Þ
likewise, with the help of the notations (7.120), the relations (7.1200 ) lead to the
conditions

h1 v0 ¼ 2divv  h1 ðr  vÞ;
ð7:122Þ
h01 v00 ¼ 2divv0  h01 ðr  v0 Þ;

that can also be written as follows

h1 v0 þ r  h1 v ¼ 0;
ð7:1220 Þ
h01 v00 þ r  h01 v0 ¼ 0:
The state of displacement and rotation becomes

u ¼ ðl0 þ a0 Þh02 v
1
 D0 gradðr  v þ v0 Þ  2a curlv0 ;
2ðk þ 2lÞ 2
ð7:123Þ
U ¼ ðl þ aÞh2 v0
1
 D2 gradðr  v0 þ v00 Þ  2a curlv;
2ðk0 þ 2l0 Þ

where both the vector and scalar potentials must be functions of class C5 . One
obtains thus the Sternberg-Eubanks type representation.

7.2.3 General Theorems

Hereafter, we shall give some theorems concerning the uniqueness of the solution of
the equations of linearly elastic bodies of Cosserat type, using the monograph of W.
Nowacki [11]; as well, we shall enounce, without demonstration, some theorems
of reciprocity.

7.2.3.1 Theorems of Uniqueness

First of all, we shall state a theorem of non-negativity of the work of deformation


per unit volume, i.e.
k
W ¼ bkk bll þ lbðijÞ bðijÞ þ ab½ij b½ij
2
k0
þ ukk ull þ l0 uðijÞ uðijÞ þ a0 u½ij u½ij : ð7:124Þ
2
346 7 Introduction to the Theory of Cosserat Type Bodies

On the basis of this theorem, the necessary and sufficient conditions in which W
is positive definite are
3k þ 2l [ 0; l [ 0; a [ 0; ð7:125Þ

3k0 þ 2l0 [ 0; l0 [ 0; a0 [ 0: ð7:1250 Þ


The theorem of uniqueness in the dynamic case, states that if the conditions
I [ 0 and X non-identical zero are satisfied in a bounded, regular domain D of the
space, the frontier of which is S, then there exists at most a solution uðx1 ; x2 ; x3 ; tÞ
and Uðx1 ; x2 ; x3 ; tÞ of class C2 (D ½0; 1Þ), which satisfies the Eqs. (7.43), (7.430 ),
the initial conditions (7.56), (7.560 ), as well as the boundary conditions (7.54) or
(7.55) on S.
Analogically, the theorem of uniqueness in the static case, states that if the
conditions u_  0; U _  0 and X non-identical zero are fulfilled in a bounded,
regular domain D of the space, the frontier of which is S, then there exists at most a
field of stresses rij and a field of couple-stresses lij , which result from the vector
fields uðx1 ; x2 ; x3 ; tÞ, Uðx1 ; x2 ; x3 ; tÞ of class C2 , which satisfy the Eqs. (7.45, 7.450 )
and the boundary conditions (7.54) or (7.55) on S.

7.2.3.2 Theorems of Reciprocity

In 1965, N. Sßandru [151, 152] stated theorems of reciprocity of the work both for
the static case (of Betti type) and for the dynamic case (of Graffi type).
Thus, considering two states of displacement and rotation and of stress and
couple-stress of an linearly elastic body of Cosserat type, marked by (0 ) and (00 ),
corresponding to two distinct systems of external loads, one may write, in the static
case,
ZZZ ZZ  n 
 0 00 0 00
 0 00
n
0 00
Fi ui þ Mi Ui dV þ pi ui þ mi Ui dS
V S
ZZZ ZZ  n 
 00 0 00 0
 00 0
n
00 0
¼ Fi ui þ Mi Ui dV þ pi ui þ mi Ui dS: ð7:126Þ
V S

In the dynamic case, we introduce the lost forces and the lost moments, so that
the corresponding relation of reciprocity reads
ZZZ ZZ  n 
 0 
0 00
 0 0
 00 0 00
n
0 00
Fi  q€ €
ui ui þ Mi  I Ui Ui dV þ pi ui þ mi Ui dS
V S
ZZZ ZZ  n 
 00 
00 0
 00 00
 0 00 0
n
00 0
¼ Fi  q€ €
ui ui þ Mi  I Ui Ui dV þ pi ui þ mi Ui dS:
V S
ð7:127Þ
7.2 Formulations of the Static and Dynamic Problems. General Theorems 347

The theorem of reciprocity of the Graffi type takes the form


ZZZ Z t
 0
dV Fi ðt  sÞu00i ðsÞ þ Mi0 ðt  sÞU00i ðsÞ ds
V 0
ZZ Z tn n

0 00 0 00
þ dS pi ðt  sÞui ðsÞ þ mi ðt  sÞUi ðsÞ ds
S 0
ZZZ Z t
 00
¼ dV Fi ðt  sÞu0i ðsÞ þ Mi00 ðt  sÞU0i ðsÞ ds
V 0
ZZ Z t n n

00 0 00 0
þ dS pi ðt  sÞui ðsÞ þ mi ðt  sÞUi ðsÞ ds: ð7:128Þ
S 0

References

A. Books

1. Boltzmann, L.: Populäre Schriften, 3rd edn. Barth Verlag, Leipzig (1925)
2. Cosserat, E., Cosserat, F.: Théorie des corps déformables. A. Herman et Fils, Paris (1909)
3. Grioli, G.: Mathematical Theory of Elastic Equilibrium (Recent Results) (Erg. Angew.
Math.). Springer, Berlin (1962)
4. Hamel, G.: Die Axiome der Mechanik. Handbuch der Physik, vol. 5. Springer, Berlin (1927)
5. Hellinger, E.: Die allgemeinen Ansätze der Mechanik der Kontinua. In: Klein, F., Wagner,
K. (eds.) Enz. Math. Wiss., vol. 4. Springer, Berlin (1914)
6. Heun, K.: Ansätze und allgemeine Methoden der Systemmechanik. Enz. Math. Wiss., vol. 4
(1914)
7. Jaramillo, T.J.: A Generalization of the Energy Function of Elasticity Theory. Dissertation,
University of Chicago (1929)
8. Kelvin, L.: Mathematical and Physical Papers, vol. 1–3. 1882, 1884, 1890
9. Kr}oner, E.: Kontinuumstheorie der Versetzungen und Eigenspannungen (Erg. Angew.
Math.). Springer, Berlin (1958)
10. Neuber, H.: Kerbspannungslehre. Grundlagen f} ur genaue Festigkeitsberechnung mit
Ber}ucksichtigung von Konstruktionsform und Werkstoff, 2nd edn. Springer, Berlin (1958)
11. Nowacki, W.: Teoria niesymetrycznej spre_zistości (Theory of Asymmetric Elasticity).
Państ. Wydawn. Naukowe, Warszawa (1971)
12. Savin, G.N.: Osnovy ploskoı̆ padachi momentnoi teorii uprugosti (Fundamentals of the
Problems of the Moment Theory of Elasticity). T. G. Shevchenko University, Kiev (1965)
13. Schaefer, H.: Continui di Cosserat. Funzioni potenziali. Calcolo numerico delle piastre.
Univ. di Trieste. Ist. di Mecc., Lezione e conferenzes, vol. 7, Trieste (1965)
14. Teodorescu, P.P.: Probleme spatßiale în teoria elasticitătßii (Space Problems in the Theory of
Elasticity). Ed. Acad., Bucuresßti (1970)
15. Teodorescu, P.P.: Dynamics of Linear Elastic Bodies. Ed. Acad., Bucuresßti. Abacus Press,
Tunbridge Wells (1975)
16. Teodorescu, P.P. (ed.): Probleme actuale în mecanica solidelor (Present-day Problems in
Solid Mechanics), vol. 1. Ed. Acad., Bucuresßti (1975)
348 7 Introduction to the Theory of Cosserat Type Bodies

17. Truesdell, C., Noll, W.: The Non-Linear Field Theories of Mechanics. Encycl. Phys. vol.
III(3). Springer, Berlin (1965)
18. Truesdell, C., Toupin, R.A.: The Classical Field Theories. Encycl. Phys. vol. III(1).
Springer, Berlin (1960)

B. Papers

19. Adomeit, G.: Ausbreitung elastischer Wellen in einen Cosserat Kontinuum mit freien
Oberfläche. ZAMM 46, 158 (1966)
20. Adomeit, G.: Determination of elastic constants of a structured material. IUTAM Symp.
1967, Mech. Generalized Continua, p. 80 (1968)
21. Aero, E.L., Buligin, A.N., Kuvshinskii, E.V.: Asymmetrical hydrodynamics. Prikl. mat. e
mekh. 2, 297 (1965)
22. Aero, E.L., Kuvshinskii, E.V.: Fundamental equations of the theory of elastic media with
rotational interaction of the particles. Fiz. tviordogo tiela 2, 1399 (1960)
23. Alblas, J.B.: The Cosserat continuum with electronic Spin. Mechanics of Generalized
Continua, IUTAM Symp. 1967, p. 350 (1968)
24. Anthony, K., Essmann, U., Seeger, A., Träuble, H.: Disclinations and the Cosserat
continuum with incompatible rotations. IUTAM Symp. 1967, Mechanics of Generalized
Continua, p. 355 (1968)
25. Barański, K., Wilmański, K., Wo_zniak, Cz.: Mechanika ośrodków ciaglich typa Cosseratów
(mechanics of bodies of Cosserat type). Mech. teoret. i stos. 5, 215 (1967)
26. Bodaśzewski, S.: O niesymetrycznym stanie niapecia i o zastowaniach w mechanike
ósrodków ciagluch (On the asymmetric state of stress and couple stress in the mechanics of
continuous bodies). Arch. Mech. Stos. 5, 351 (1953)
27. Boggio, T.: Sull’integrazione di alcune equazioni lineari alle derivate parziali. Ann. Mat.
Pura Appl. 8(Ser. 3), 181 (1903)
28. Bressan, A.: Sui sistemi continui nel caso asimmetrico. Ann. Mat. 512(Ser. 4), 169 (1963)
29. Burzyński, W.: Skrecenie bez skrecania. Przegl. Mech. (1949)
30. Carlson, D.E.: The general solution of the stress equations of equilibrium of a Cosserat
continuum. Proc. Fifth U. S. Natl. Congr. Appl. Mech., p. 249 (1966)
31. Carlson, D.E.: Stress functions for couple and dipolar stresses. Appl. Math. 24, 29 (1966)
32. Carlson, D.E.: On G} unther’s stress functions for couple stresses. Appl. Math. 25, 139 (1967)
33. Cauchy, A.L.: Mémoire sur les systèmes isotropes de points matériels. Mém. Acad. Sci.
Paris 22, 605 (1850)
34. Cauchy, A.L.: Note sur l’équilibre et les mouvements vibratoires des corps solides. C. Rend.
hebd. des séances Acad. Sci. Paris 32, 326 (1851)
35. Clintock, F.A.Mc: Contribution of interface couples to the energy of dislocation. Acta
metallurgica 8, 127 (1960)
36. Clintock, F.A.Mc., André, P.A., Schwedt, K.R., Stoeckly, R.E.: Interface couples in
crystals. Nature 182, 652 (1958)
37. Corre, Y.Le.: Constantes élastiques et piézoélectriques cristallines. Bull. Soc. Franc. Minér.
Crist. 76, 464 (1953)
38. Corre, Y.Le: Étude de l’élasticité. Bull. Soc. Franc. Minér. Crist. I, II, 77, 1363, 1393
(1954); III, IV, 28, 33, 54 (1955)
39. Corre, Y.Le.: La dissymétrie du tenseur des efforts et ses conséquences. J. Phys. Radium 17,
934 (1956)
40. Corre, Y.Le.: Les densités de couple et les pseudo rotations dans la théorie de l’élasticité de
Lavel. J. Phys. Radium 19, 541 (1958)
41. Cosserat, E., Cosserat, F.: Sur la théorie de l’élasticité. Ann. Éc. norm. Toulouse 10, 1
(1896)
References 349

42. Cosserat, E., Cosserat, F.: Sur la mécanique générale. C. Rend. hebd. Acad. Sci., Paris 145,
1139 (1907)
43. Dixon, R.C., Eringen, A.C.: A dynamical theory of polar elastic dielectrics. Int. J. Eng. Sci.
3, 359 (1965)
44. Djurić, S.: Dynamique et petites vibrations du continuum de Cosserat. Thèse Droit.,
Belgrade (1964)
45. Ericksen, J.L.: Conservation laws for liquid crystals. Trans. Soc. Rheol. 5, 23 (1961)
46. Ericksen, J.L.: Hydrostatic theory of liquid crystals. Arch. Rat. Mech. Anal. 9, 374 (1962)
47. Ericksen, J.L.: Nilpotent energies in liquid crystals theory. Rat. Mech. Anal. 10, 189 (1962)
48. Ericksen, J.L.: Orientation induced by flow. Trans. Soc. Rheol. 6, 275 (1962)
49. Ericksen, J.L., Truesdell, C.: Exact theory of stress and strain in rods and shells. Arch. Rat.
Mech. Anal. 1, 295 (1958)
50. Eringen, A.C.: Simple microfluids. Int. J. Eng. Sci. 2, 205 (1964)
51. Eringen, A.C.: Mechanics of micromorphic continua. IUTAM Symp. 1967, Mechanics
Generalized Continua, p. 18 (1968)
52. Eringen, A.C., Suhubi, E.S.: Nonlinear theory of simple microelastic solids. Int. J. Eng. Sci.
2(I, II), 189, 389 (1964)
53. Galerkin, B.G.: On the problem of stresses and deformations in an isotropic elastic body.
Dokl. Akad. Nauk SSSR 14(Ser. A), 353 (1930)
54. Galletto, D.: Sulle equazioni in coordinate generali della statica dei continui con
caratteristiche di tensione asimmetriche. St. Anna University Hospital, Ferrara 10(Ser. 7),
33 (1962)
55. Galletto, D.: Nuove forme per le equazioni in coordinati generali della statica dei continui
con caratteristiche di tensione asimmetriche. Ann. Scuola Norm. Sup. Pisa, Sci. Fis. Mat.
17(Ser. 3), 297 (1963)
56. Galletto, D.: Contributo allo studio dei sistemi continui a transformazioni reversibili con
caratteristiche di tensione asimmetriche. Rend. Sem. Mat. Univ. Padova, 35, 299 (1965)
57. Galletto, D.: Sull’unicità in presenza di vincoli interni di una condizione cinematica
fondamentale nella teoria delle deformazione finite. Atti Ist. Veneto Sci. Lett. Arti, Cl. Sci.
Mat. Nat. 123, 197 (1965)
58. Galletto, D.: Sistemi incomprimibili a trasformazioni reversibili nel caso asimmetrico.
Rend. Sem. Mat. Univ. Padova I, II, 36, 243 (1966); 37, 18 (1967)
59. Gazis, D.C., Wallis, R.F.: Extensional waves in cubic crystal plates. Proceedings of 4th U.
S. National Congress of Applied Mechanics, p. 161 (1962)
60. Gazis, D.C., Wallis, R.F.: Surface vibrational waves in crystal lattices with complex
interatomic interactions. J. Math. Phys. 3, 190 (1962)
61. Gordon, V.A., Tolokonnikov, L.A.: Equations of equilibrium in the moment theory of
elasticity. Prikl. Mekh., Kiev 4, 118 (1968)
62. Gorskii, B.V.: On some problems of the theory of asymmetric elasticity. Prikl. mekh., Kiev
3, 74 (1967)
63. Graff, K.F., Pao, Y.H.: The effects of couple stresses on the propagation and reflection of
plane waves in an elastic half space. J. Sound Vibr. 6, 217 (1967)
64. Green, A.E.: Micro materials and multipolar continuum mechanics. Int. J. Eng. Sci. 3, 533
(1965)
65. Green, A.E., Naghdi, P.M.: The Cosserat surface. IUTAM Symp. 1967, Mechanics of
Generalized Continua, p. 36 (1968)
66. Green, A.E., Naghdi, P.M., Wainrght, W.L.: A general theory of Cosserat surface. Arch.
Rat. Mech. Anal. 20, 287 (1965)
67. Green, A.E., Rivlin, R.S.: Multipolar continuum mechanics. Arch. Rat. Mech. Anal. 17, 113
(1964)
68. Grioli, G.: Elasticità asimmetrica. Ann. Mat. Pura Appl. 50(Ser. 4), 389 (1960)
69. Grioli, G.: Sulla meccanica dei continui a transformazioni reversibili con caratteristiche di
tensioni asimmetriche. Sem. Ist. Alta Mat. 535 (1962–1963)
350 7 Introduction to the Theory of Cosserat Type Bodies

70. Grioli, G.: On the thermodynamic potential of Cosserat continua. IUTAM Symp. 1967,
Mechanics of Generalized Continua, p. 63 (1968)
71. G}unther, W.: Zur Statik und Kinematik des Cosseratschen Kontinuums. Abh. der
Braunschw. Wiss. Gesellschaft, 10, 195 (1958)
72. G}unther, W.: Analoge systeme von schalengleichungen. Ing. Archiv. 30, 160 (1961)
73. Gurtin, M.E.: A generalization of the beltrami stress functions in continuum mechanics.
Arch. Rat. Mech. Anal. 13, 321 (1963)
74. Hartranft, R.J., Sih, G.C.: The effect of couple stresses on the stress concentrations of a
circular inclusion. J. Appl. Mech. 32, 429 (1965)
75. Hehl, F.: Space time as generalized Cosserat continuum. IUTAM Symp. 1967, Mechanics
of Generalized Continua, p. 347 (1968)
76. Hehl, F., Kr}oner, E.: Zum Materialgesetz eines elastischen Mediums mit Moment-
spannungen. Z. Naturforschung 20, 336 (1965)
77. Hermann, G., Achenbach, J.D.: Applications of theories of generalized Cosserat continua to
the dynamics of composite materials. IUTAM Symp. 1967, Mech. Generalized Continua,
p. 69 (1968)
78. Hoffman, W.H.: On the bending of thin elastic plates in the presence of couple stresses.
J. Appl. Mech. 31, 706 (1964)
79. Hoffman, W.H., Shahman, F.O.F.: Physical model of a 3 constant isotropic elastic material.
J. Appl. Mech. 4, 837 (1965)
80. Iacovache, M.: O extindere a metodei lui Galerkin pentru sistemul ecuatßiilor elasticitătßii (A
generalization of galerkin’s method for the system of elasticity equations). Bul. Acad. I(Ser.
A), 592 (1949)
81. Ionescu Cazimir, V.: Problem of linear coupled thermoelasticity. I. Theorems on reciprocity
for the dynamic problem of coupled thermoelasticity. Bull. Acad. Pol. Sci. Sér. Sci. Techn
12, 473 (1964)
82. Ionescu Cazimir, V.: Problem of linear coupled thermoelasticity. II. Some applications of
the theorems of reciprocity for the dynamic problem of coupled thermoelasticity. Bull.
Acad. Pol. Sci. Sér. Sci. Techn12, 481 (1964)
83. Joel, N., Wooster, W.A.: Theories of crystal elasticity. Nature 180, 430 (1957)
84. Joel, N., Wooster, W.A.: Number of elastic constants requested in crystal elasticity. Nature
182, 1078 (1958)
85. Kaliski, S.: O pewym modelu ošrodka ciaglego z istotnie niesymetricznym tensorem napié
(on the model of bodies with asymmetric stress tensor). Biul. WAT 4, 104 (1962)
86. Kaliski, S.: On a model of the continuum with essentially non symmetric tensor of
mechanical stress. Arch. Mech. Stos. 15, 33 (1963)
87. Kaliski, S., Plochocki, D., Rogula, D.: Asymmetric stress tensor and angular momentum
conservation law in the equations of combined mechanical and electromagnetic field in a
continuous medium. Proc. Vibr. Probl. 3, 253 (1962)
88. Kelly, P.D.: A reacting continuum. Int. J. Eng. Sci. 2, 129 (1964)
89. Kessel, S.: Lineare elastizitätstheorie des anisotropen Cosserat kontinuums. Abh. der
Braunschw. Wiss. Gesellschaft 16, 1 (1964)
}
90. Kessel, S.: Uber die Verträglichkeitsbedingungen in einem Cosseratschen Kontinuums.
Abh. der Braunschw. Wiss. Gesellschaft 17, 51 (1965)
91. Kessel, S.: Die spannungsfunktionen des Cosserat kontinuums. ZAMM 47, 329 (1967)
92. Kessel, S.: Stress functions and loading singularities for the infinitely extended, linear,
elastic isotropic Cosserat continuum. IUTAM Symp. 1967, Mechanics of Generalized
Continua, p. 114 (1968)
93. Kirichenko, A.M.: Couple stresses in the Study of the axisymmetrical states of stress of an
elastic sphere. Prikl. mekh. Kiev 4, 47 (1968)
94. Klein, F.: Nachr. kgl. Ges. Wiss, Göttingen, 235 (1918)
95. Koiter, W.T.: Couple stresses in the theory of elasticity. I, II. Koninkl. Nederl. Akad. van
Wettenschappen, Proc. 67(Ser. B), 17, 30 (1964)
References 351

96. Kondo, K.: On the two main currents of the geometrical theory of imperfect continua.
IUTAM Symp. 1967, Mechanics of Generalized Continua, p. 200 (1968)
97. Krishnan, R.S., Rajagopal, E.S.: The atomistic and the continuum theories of crystal
elasticity. Ann. der Physik 8, 121 (1961)
98. Kr}oner, E.: On the physical reality of torque stresses in continuum mechanics. Int. J. Eng.
Sci. 1, 261 (1963)
99. Kr}oner, E.: Das physikalische problem der antisymmetrischen Spannungen und der
sogennanten Momentspannungen. Proceedings of the 10th International Congress of
Applied Mechanics, München, p. 143,1964 (1966)
100. Kr}oner, E.: Interrelations between variuos branches of continuum mechanics. IUTAM
Symp. 1967, Mechanics of Generalized Continua, p. 330 (1968)
101. Kubenko, V.D., Shul’ga, N.A.: Plane dynamic problem of the moment theory of elasticity
and viscoelasticity. Prikl. mekh. Kiev 3, 82 (1967)
102. Kuvshinskii, E.V., Aero, E.L.: Theory of continuum in the asymmetric elasticity,
considerations of the ‘‘internal’’ rotations. Fiz. tviord. tiela 5, 2591 (1963)
103. Kunin, I.A.: The theory of elastic media with microstructure and theory of dislocations.
IUTAM Symp. 1967, Mechanics of Generalized Continua, p. 321 (1968)
104. Laval, J.: L’élasticité du milieu cristallin. J. Phys. Radium 18(1–3), 217, 289, 369 (1957)
105. Lomakin, V.A., Savov, L.N.: Deformation problems of viscoelastic micro non
homogeneous bodies and the theory of moments in viscoelasticity. Mekh. polimerov 2
(1967)
106. Mindlin, R.D.: Influence of couple stresses on stress concentrations. Exper. Mech. 3, 1
(1963)
107. Mindlin, R.D.: Microstructure in linear elasticity. Arch. Rat. Mech. Anal. 16, 51 (1963)
108. Mindlin, R.D.: Representation of displacements and stresses in plane strain with couple
stresses. IUTAM Symp. Tbilisi, 1963, p. 256 (1964)
109. Mindlin, R.D.: On the equations of elastic materials with microstructure. Int. J. Solid Struct.
1, 73 (1965)
110. Mindlin, R.D.: Stress functions for a Cosserat continuum. Int. J. Solid Struct. 1, 265 (1965)
111. Mindlin, R.D.: Second gradient of strain and surface tension in linear elasticity. Int. J. Solid
Struct. 1, 417 (1965)
112. Mindlin, R.D.: Theories of elastic continua and crystal lattice theories. IUTAM Symp. 1967,
Mechanics of Generalized Continua, p. 312 (1968)
113. Mindlin, R.D., Tiersten, H.F.: Effect of couple stresses in linear elasticity. Arch. Rat. Mech.
Anal. 11, 415 (1962)
114. Misßicu, M.: Theory of viscoelasticity with couple stresses and some reductions to the two
dimensional problems. Rev. Roum. Sci. Techn., Méc. Appl. I, II, 8, 921 (1963); 9, 3 (1964)
115. Misßicu, M.: On a theory of asymmetric plastic and viscoelasticplastic solids. Rev. Roum.
Sci. Techn., Méc. Appl. 9, 477 (1964)
116. Misßicu, M.: A generalization of the Cosserat equations of the motion of deformable bodies
(with internal degrees of freedom). Rev. Roum. Sci. Techn. Méc. Appl. 9, 1351 (1964)
117. Misßicu, M.: On a general solution of the theory of singular dislocation of media with couple
stresses. Rev. Roum. Sci. Techn. Méc. Appl. 10, 34 (1965)
118. Misßicu, M.: A generalization of the motion of deformable bodies. Arch. Mech. Stos. 17, 183
(1965)
119. Misßicu, M.: The generalized dual continuum in elasticity and dislocation theory. IUTAM
Symp. 1967, Mechanics Generalized Continua, p. 141 (1968)
120. Muki, R., Sternberg, E.: The influence of the couple stresses on singular stress concentration
in elastic bodies. ZAMP 16, 611 (1965)
121. Naghdi, P.M.: On a variational theorem in elasticity and its applications to shell theory.
J. Appl. Mech. 31, 647 (1964)
122. Naghdi, P.M.: A static geometric analogue in the theory of couple stresses. Proc. Konicl.
Nederl. Akad. Wet. 68(Ser. B), 29 (1965)
352 7 Introduction to the Theory of Cosserat Type Bodies

123. Nemish, Iu.N.: Plane problem of the moment theory of elasticity in a domain with a circular
hole. Prikl. Mekh. Kiev I, 52 (1965)
124. Nemish, Iu.N.: Stress concentration in the vicinity of curvilinear holes in asymmetric theory
of elasticity. Prikl. Mekh. Kiev II, 85 (1966)
125. Nemish, Iu.N.: Reinforced circular hole in an elastic domain with an asymmetric stress
tensor. Prikl. Mekh. Kiev II, 43 (1966)
126. Neuber, H.: Ein neuer Ansatz zur Lösung rämlicher Probleme der Elastizitätstheorie. Der
Hohlkegel unter Einzellart als Beispiel. ZAMM 14, 203 (1934)
127. Neuber, H.: On the general solution of linear elastic problems in anisotropic and isotropic
Cosserat continua. Proceedings of the 11th International Congress on Applied Mechanics,
M}unchen, p. 153 (1964)
}
128. Neuber, H.: Uber Probleme der Spannungskonzentrationen in Cosserat Körper. Acta Mech.
II, 48 (1966)
129. Neuber, H.: Die schubbeanspruchte Kerbe in Cosserat Körper. ZAMM 47, 313 (1967)
130. Neuber, H.: On the effect of stress concentration in Cosserat continuum. IUTAM Symp.
1967, Mechanics of Generalized Continua, p. 109 (1968)
131. Noether, E.: Nachr, kgl. Ges. Wiss., Göttingen, p. 171 (1918)
132. Noll, W.: A mathematical theory of mechanical behaviour of continuous media. Arch. Rat.
Mech. Anal. 2, 195 (1958)
133. Nowacki, W.: Couple stresses in the theory of thermoelasticity. Bull. Acad. Pol. Sci. Sér.
Sci. Techn. 14(1–3), 97, 203, 505 (1966)
134. Nowacki, W.: Some theorems of asymmetric thermoelasticity. Bull. Acad. Pol. Sci. Sér. Sci.
Techn. 15, 289 (1967)
135. Nowacki, W.: Couple stresses in thermoelasticity. Prikl. mekh., Kiev 3, 3 (1967)
136. Nowacki, W.: Couple stresses in the theory of thermoelasticity. IUTAM Symp. 1966,
Irreversible Aspects of Continuum Mechanics (1968)
137. Oshima, N.: Asymmetrical stress tensor and its application to a granular medium. In:
Proceedings of 3rd Japanese National Congress of Applied Mechanics, 1953, p. 77 (1954)
138. Oshima, N.: Dynamics of Granular Media, vol. I, p. 563. Memory RAAG, Tokyo (1955)
139. Pal’mov, V.A.: Fundamental equations of the theory of asymmetric elasticity. Prikl. mat.
mekh. 28, 401 (1964)
140. Pal’mov, V.A.: Plane problem of the theory of asymmetric elasticity. Prikl. mat. mekh. 28,
1117 (1964)
141. Papkovich, P.F.: Solution générale des équations differentielles fondamentales d’élasticité
exprimée par trois fonctions harmoniques. C. Rend. hebd. séances Acad. Sci., Paris 195, 513
(1932)
142. Pietras, F., Wyrwiński, J.: Thermal stresses in a plane anisotropic Cosserat continuum.
Arch. Mech. Stos. 19, 627 (1967)
143. Podil’chuk, N., Kirichenko, A.M.: Influence of a spherical hole on the state of stress of an
elastic medium, taking into account couple stresses. Prikl. mekh. Kiev 3, 69 (1967)
144. Poisson, S.D.: Mém. Acad. Paris 18 (1842)
145. Rajagopal, E.S.: The existence of interfacial couples in infinitesimal elasticity. Ann. Physik
6, 192 (1960)
146. Reissner, E.: Note on the theorem of the symmetry of the stress tensor. J. Math. Phys. 23,
192 (1944)
147. Reissner, E., Wan, F.Y.M.: A note on G} unther’s analysis of couple stress. IUTAM Symp.
1967, Mechanics of Generalized Continua, p. 83 (1968)
148. Rivlin, R.S.: Generalized mechanics of continuous media. IUTAM Symp. 1967, Mechanics
Generalized Continua, p. 1 (1968)
149. Rymarz, G.: Surface waves in Cosserat’s medium. Bull. Acad. Pol. Sci. Sér. Sci. Techn. 15,
177 (1967)
150. Saint Venant, A.J.C.B.: Note sur les valeurs que prennent les pressions dans un solide
élastique isotrpe lorsque l’on tient compte des dériveés d’ordre supérieur des déplacements
References 353

trés petits que leurs points ont éprouvés. C. Rend. hebd. séances Acad. Sci., Paris 68, 569
(1869)
151. S
ßandru, N.: Le théorème de réciprocité du type de Betti dans l’élasticité asymétrique.
C. Rend. hebd. séances Acad. Sci. Paris 260, 3565 (1965)
152. S
ßandru, N.: Le théorème de réciprocité dans l’élasticité asymétrique (cas dynamique). Atti
Accad. Naz. Lincei, Ser. VIII, Rend., Cl. Sci. Fis., Mat. e Nat., 38, 78, (1965)
153. S
ßandru, N.: On some problems of the linear theory of the asymmetric elasticity. Int. J. Eng.
Sci. 4, 81 (1966)
154. Savin, G.N., Guz’, A.N.: On a method of solution of the plane problems of the moment
theory of elasticity for multiply connected domains. Prikl. mekh. Kiev 2, 3 (1966)
155. Savin, G.N., Shul’ga, N.A.: Plane dynamic problem of the moment theory in elasticity.
Prikl. mekh. Kiev. 3, 1 (1967)
156. Savov, L.N.: Plane problem of the moment theory of viscoelasticity on the stress
concentration in the proximity of a circular hole. Prikl. mekh. Kiev. 4, 6 (1968)
157. Schade, K.D.: Cosserat Fläche und Schalentheorie. Doktorthesis, Techn. Hochschule,
Darmstadt (1966)
158. Schaefer, H.: Die Spannungsfunktionen des dreidimensionalen Kontinuums und des
elastischen Körpers. ZAMM 33, 356 (1953)
159. Schaefer, H.: Die Spannungsfunktionen des dreidimensionalen Kontinuums; statische
Dentung und Randwerte. Ing. Archiv 28, 291 (1959)
160. Schaefer, H.: Versuch einer Elastizitätstheorie des zweidimensionalen ebenen Cosserat
Kontinuums. Misz. Angew. Mech. Festschrift W. Tollmien, p. 277 (1962)
161. Schaefer, H.: Analysis der Motorfelder in Cosserat Kontinuum. ZAMM 47, 319 (1967)
162. Schaefer, H.: Die Spannungsfunktionen eines Kontinuums mit Momentspannungen. Bull.
Acad. Pol. Sci. Sér. Sci. Techn. I, II. 15, 63, 69 (1967)
163. Schaefer, H.: Das Cosserat kontinuum. ZAMM 47, 485 (1967)
164. Schaefer, H.: The basic affine connection in a Cosserat continuum. IUTAM Symp. 1967,
Mechanics of Generalized Continua, p. 57 (1968)
165. Schijve, J.: Note on couple stresses. J. Mech. Solid 14, 113 (1966)
166. Sokolowski, M.: Couple stresses in problems of torsion of prismatical bars. Bull. Acad. Pol.
Sci. Sér. Sci. Techn. 13, 419 (1965)
167. Soós, E., Teodorescu, P.P.: Les constantes de matérieau dans le cas des modèles du type de
Cosserat des cristaux élastiques. Lett. Appl. Eng. Sci. 1, 209 (1973)
168. Stefaniak, J.: Concentrated loads as body forces. Rev. Roum. Math. Pures Appl. 14, 119
(1969)
169. Sternberg, E.: Couple stresses and singular stress concentrations in elastic solids. IUTAM
Symp. 1967, Mechanics of Generalized Continua, p. 95 (1968)
170. Sternberg, E., Eubanks, R.A.: On stress functions for elastokinetics and the integration of
the repeated wave equation. Appl. Math. 15, 149 (1957)
171. Sternberg, E., Muki, R.: The effect of couple stresses on the stress concentration around a
crack. Int. J. Solid Struct. 3, 69 (1967)
172. Stoianović, R.: Dislocation in the generalized elastic Cosserat continuum. IUTAM Symp.
1967, Mechanics of Generalized Continua, p. 152 (1968)
173. Stoianović, R., Djurić, S.: Le problème plan de la théorie des corps élastiques orientés. 8th
Congress on Young Mechanics, Split (1966)
174. Stoianović, R., Djurić, S., Vujośević, L.: Stress strain relations for the elastic Cosserat
continuum. Conference on Department Continuing Medical, P.A.N., Zakopane (1961)
175. Stoianović, R., Djurić, S., Vujośević, L.: Contribution à la dynamique des continuum de
Cosserat. Mat. Vesnik 1, 127 (1964)
176. Stoianović, R., Vujośević, L.: Couple stress in non euclidean continua. Publ. Inst. Math.
New Ser. 2, 71 (1962)
177. Teodorescu, P.P.: Schwingungen der elastischen Kontinua. III. Konf. Nichtlin. Schwing.,
1964, Abh. deut. Akad. Wiss., Berlin 29 (1965)
354 7 Introduction to the Theory of Cosserat Type Bodies

178. Teodorescu, P.P.: One hundred years of investigations in the plane problem of the theory of
elasticity. Appl. Mech. Surv. p. 245, Spartan Books, Washington, DC (1966)
179. Teodorescu, P.P.: Sur un certain caractère tensoriel des charges concentrées. Rend. Naz.
Lincei, Ser. 8, Cl. Sci. Fis. Mat. Nat. 150, 251 (1966)
180. Teodorescu, P.P.: Sur l’action des charges concentrées dans le problème plan de la
mécanique des solides déformables. Arch. Mech. Stos. 18, 567 (1966)
181. Teodorescu, P.P.: Sur la notion de moment massique dans le cas des corps du type de
Cosserat. Bull. Acad. Pol. Sci., Sér. Sci.Tech. 20, 57 (1967)
182. Teodorescu, P.P.: Problèmes bidimensionnels de la théorie de l’élasticité. I. Une tension
normale nulle. Atti Acad. Naz. Lincei, Ser. III, Rend. Cl. Sci. Fis. Mat. Nat. 154, 141 (1968)
183. Teodorescu, P.P.: Problémes bidimensionels de la théorie de l’élasticité. II. Deux tensions
tangentielles nulles. Atti. Accad. Naz. Lincei, Ser. VIII, Rend. Cl. Sci. Fis. Mat. Nat. 154,
232 (1968)
184. Teodorescu, P.P.: On the action of concentrated loads in the case of a Cosserat continuum.
IUTAM Symp. 1967. Mechanics of Generalized Continua, p. 120 (1968)
185. Teodorescu, P.P.: Sur les corps du type de Cosserat à l’élasticité linéaire. Ist. Naz. Alta Mat.,
Symp. Mat. 1, 375 (1969)
186. Teodorescu, P.P.: Sur l’introduction des fonction potentiels en élasticité linéaire. An. Univ.
Bucuresßti, Mat. Mec. 20, 131 (1971)
187. Teodorescu, P.P.: Sur une représentation du type de Papkovich Neuber dans le cas des corps
du type de Cosserat. An. Univ. Bucuresßti, Mat. Mec. 21, 115 (1972)
188. Teodorescu, P.P.: Sur le problème dynamique des corps du type de Cosserat. Rev. Roum.
Math. Pures Appl. 17, 1097 (1972)
189. Teodorescu, P.P.: On continuity equations and cesàro type formulae in case of Cosserat type
solids. Mech. Res. Comm. 4, 63 (1977)
190. Teodorescu, P.P.: Sur l’état de déplacement et de rotation dans le cas des corps du type de
Cosserat. Bull. Math. Soc. Sci. Math. Roum. 17(65), 309 (1973)
191. Teodorescu, P.P.: Sur la représentation de S ßandru dans le cas des corps du type de Cosserat.
An. Univ. Bucuresßti, Mat. Mec. 22, 155 (1973)
192. Teodorescu, P.P.: Sur le calcul des récipients soumis à des charges concentrées dans le cas
de l’élasticité symétrique ou asymetrique. Cons. naz. ric., Quaderni de ‘‘la ric. sci.’’ 78, 24
(1973)
193. Teodorescu, P.P., S ßandru, N.: Sur l’action des charges concentrée en élasticité asymétrique
plane. Rev. Roum. Math. Pures Appl. 12, 1399 (1967)
194. Teodosiu, C.: On the determination of internal stresses and couple stresses in the continuum
theory of dislocations. Bull. Acad. Pol. Sci., Sér. Sci. Tech. 12, 605 (1964)
195. Teodosiu, C.: The determination of stresses and couple stresses generated by dislocations in
isotropic media. Rev. Roum. Sci. Tech. Sér. Mec. Appl. 10, 1461 (1965)
196. Teodosiu, C.: Non linear theory of the materials of grad 2 with initial stresses and
hyperstresses. I. Basic geometric and static equations. Bull. Acad. Pol. Sci., Sér. Sci. Techn.
15, 95 (1967)
197. Teodosiu, C.: Non linear theory of the materials of grad 2 with initial stresses and
hyperstresses. II. Constitutive equations. Bull. Acad. Pol. Sci. Sér. Sci. Tech. 15, 193 (1967)
198. Teodosiu, C.: Contributions to the continuum theory of dislocations and initial stresses. Rev.
Roum. Sci. Techn., Sér. Méc. Appl. 12(1–3), 961, 1061, 1291 (1967)
199. Teodosiu, C.: Continuous distribution of dislocations in hyperelastic material of grade 2.
IUTAM Symp. 1967, Mechanics of Generalized Continua, p. 279 (1967)
200. Tiffen, R., Stevenson, A.C.: Elastic isotropy with body force couple. Q. J. Mech. Appl.
Math. 9, 306 (1956)
201. Toupin, R.A.: Elastic materials with couple stresses. Arch. Rat. Mech. Anal. 11, 385 (1962)
202. Toupin, R.A.: Theories of elasticity with couple stress. Arch. Rat. Mech. Anal. 17, 85
(1964)
203. Toupin, R.A.: Dislocated and oriented media. IUTAM Symp. 1967, Mechanics of
Generalized Continua, p. 126 (1968)
References 355

204. Truesdell, C.: The mechanical foundations of elasticity and fluid dynamics. Rat. Mech.
Anal. 5, 125 (1952)
205. Vinokurov, L.P., Derevianko, N.I.: Construction of fundamental equations for calculation
without torsion with consideration of couple stresses. Prikl. mekh., Kiev 2, 72 (1966)
206. Voigt, W.: Theoretische Studien über die Elastizitätsverhältnisse der Krystalle. I, II. Abh.
Königl. Ges. Wiss., Göttingen, p. 34 (1887)
207. Weistmann, Y.: Couple stresses effects on stress concentrations around a cylindrical
inclusion in a field of uniaxial tension. J. Appl. Mech. 32, 424 (1965)
208. Weselowski, Z.: On the couple stresses in an elastic continuum. Arch. Mech. Stos. 17, 219
(1965)
209. Woźniak, Cz.: Theory of fibrous media, I, II. Arch. Mech. Stos. 17, 651, 777 (1965)
_
210. Woźniak, Cz., Zukowski, M.: On a model of elastic subsoil carrying surface moments. Bull.
Acad. Pol. Sci., Sér. Sci. Techn. 14 (1966)
211. Wirwiński, J.: Green functions for a thermoelastic Cosserat medium. Bull. Acad. Pol. Sci.
Sér. Sci. Tech. 14, 145 (1966)
212. Yamamoto, Y.: An intrinsic theory of a Cosserat continuum. RAAG, 3rd ser
Chapter 8
Theory of Concentrated Loads

A complete mathematical model of Newtonian mechanics includes the represen-


tation of loads acting on an arbitrary continuous body. In the mechanics of rigid
solids it is sufficient to consider that the loads are just forces which may be rep-
resented by sliding vectors, while in the case of a deformable continuum the loads
must be represented by bound vectors. Owing to the deformability of the body, the
loads have at the same time a local and an overall effect; the manner of representing
the vector fields which correspond to various loads is particularly important. In this
problem, the theory of distributions plays a very important rôle [1, 2, 8].
In the following discussion we shall treat of the representation of arbitrary loads
in general and especially of the representation of concentrated loads. Also, we
shall give a general classification of concentrated loads for classical as well as for
Cosserat type bodies [5–7, 9].

8.1 Case of Linearly Elastic Bodies

We shall first treat of the representation of concentrated forces in case of linearly


elastic bodies; these are concentrated loads which will be considered as funda-
mental since starting from them we may construct any other concentrated loads.

8.1.1 Construction of Concentrated Loads

With the aid of concentrated forces we may construct an arbitrary distributed load,
the representation of which will be considered in the following, i.e.: systems of
concentrated forces, directed moments and dipoles of forces; then one may obtain
centres of rotation and centres of dilatation.

P. P. Teodorescu, Treatise on Classical Elasticity, Mathematical and Analytical 357


Techniques with Applications to Engineering, DOI: 10.1007/978-94-007-2616-1_8,
Ó Springer Science+Business Media Dordrecht 2013
358 8 Theory of Concentrated Loads

8.1.1.1 Concentrated Force

A category of loads the mathematical representation of which is simple and


includes all its characteristics is that of concentrated forces. We shall first consider
a single concentrated force and then a system of concentrated forces.
As we have shown in Sect. 6.2.4.1, the density of a concentrated force F,
applied at the point of position vector r0 is given by
QðrÞ ¼ Qðx1 ; x2 ; x3 Þ ¼ Fdðr  r0 Þ; ð8:1Þ

this representation includes in a synthetic form all the characteristics of a con-


centrated force F, considered as a bound vector, namely: direction, modulus and
location (point of application).
For example, the representative d sequence
8 
>
> 1 x
> 1 þ ; e\x\0;
>
<e e
qe ðxÞ ¼ 1  x  0\x  e; ð8:2Þ
>
> 1  ;
>
>
:e e
0; jxj  e;

allows the representation of a concentrated force as the limit of a sequence of loads


distributed in a triangular form (Fig. 8.1) [11]; we have
Qe ðxÞ ¼ Fie qe ðxÞ; ð8:3Þ

so that
Re ðxÞ ¼ Fie ; Me ¼ 0; ð8:30 Þ
where ie is the unit vector of the Ox2 -axis (we have assumed that the concentrated
force F acts at the point O in the direction of the Ox2 -axis and has been obtained as
the limit of loads acting in the Ox1 x2 -plane).

Fig. 8.1 Concentrated force x2


represented by a
representative d sequence (0, F)
F
F
(0, 2 )

F
(0, 3 )
F
(0 , ε )
Qε (x1)

( -ε,0) (-3,0) (-2,0) (-1,0) O (1,0) (2,0) (3,0) (ε,0) x1


8.1 Case of Linearly Elastic Bodies 359

8.1.1.2 Composition of Concentrated Forces

By the composition of concentrated forces of the type defined above we obtain forces
of the same type; thus, the forces F1 ; F2 ; . . .; Fn ; applied at the point Aðr0 Þ; are
equivalent to the force F1 þ F2 þ    þ Fn ; applied at the very same point (Fig. 8.2).
Indeed, the n forces are equivalent to the loads
Qi ðrÞ ¼ Fi dðr  r0 Þ; i ¼ 1; 2; . . .; n; ð8:4Þ

by the composition of which we obtain


X
n
QðrÞ ¼ Qi ðrÞ ¼ Rdðr  r0 Þ; ð8:40 Þ
i¼1

where we have introduced the resultant force


X
n
R¼ Fi : ð8:400 Þ
i¼1

Also, by multiplying a bound force by a real number, the point of application is


not altered, a fact which is maintained in the mathematical representation (8.40 );
this shows that the representation of a concentrated force with the aid of the Dirac
distribution is correct [14–21, 22].
Although the composition of bound forces has a meaning only when these ones
have the same point of application, we may perform, in a certain sense, the
composition of concentrated forces with different points of application. Such is the
case of a system of n  2 forces, applied at different points, if they are parallel
ones; otherwise, one may decompose every force along three directions, obtaining
thus three systems of parallel forces (the vector components of the forces initially
given).

Fig. 8.2 Composition of Fi


concentrated forces F2 R
F1

Fn
A ( x10, x 02 , x 30 )
r0

O
360 8 Theory of Concentrated Loads

8.1.1.3 Parallel Concentrated Forces

Let us consider, e.g., two parallel, equal, bound forces ðF1 ¼ F2 ¼ F; in the sense
of free vectors), applied at the points A1 ða; 0Þ and A2 ða; 0Þ; a [ 0 (Fig. 8.3a). To
these parallel forces there correspond the equivalent vector fields
Q1 ðx1 ; x2 Þ ¼ F1 dðx1 þ a; x2 Þ ¼ Fdðx1 þ a; x2 Þ;
ð8:5Þ
Q2 ðx1 ; x2 Þ ¼ F2 dðx1  a; x2 Þ ¼ Fdðx1  a; x2 Þ:
The vector field equivalent to the system formed by the two bound forces reads
Qðx1 ; x2 Þ ¼ Q1 ðx1 ; x2 Þ þ Q2 ðx1 ; x2 Þ ¼ F½dðx1 þ a; x2 Þ þ dðx1  a; x2 Þ;

taking into account a rule of composition of distributions, one gets

Qðx1 ; x2 Þ ¼ 2aFdðx21  a2 ; x2 Þ; a [ 0; ð8:6Þ

which is a rule of composition of two bound equipolent forces (equal in the sense
of free forces).
The fact of considering an Ox1 x2 -plane and points of application on the Ox1 -
axis is not essential to particularize the problem.
Let us consider now two bound forces F1 and F2 ; having the same modulus, the
same direction, but opposite sense, for which we may write F1 ¼ F and F2 ¼ F
(in the sense of free vectors), applied at the points A1 ða; 0Þ and A2 ða; 0Þ; a [ 0
respectively (Fig. 8.3b). The equivalent vector field is given by
Qðx1 ; x2 Þ ¼ F½dðx1 þ a; x2 Þ  dðx1  a; x2 Þ;
wherefrom

Qðx1 ; x2 Þ ¼ 2Fx1 dðx21  a2 ; x2 Þ; a [ 0; ð8:7Þ

obtaining thus the corresponding composition formula.


The above representation has no more sense in the case of sliding forces; in the
latter case we obtain a couple to which corresponds a free vector.

Fig. 8.3 Parallel x2 x2


concentrated forces: a of the
same sense; b of opposite F F F
sense
A1(-a,0) O A2 (a,0) x1 A1(-a,0) O A2 (a,0) x1
-F

(a) (b)
8.1 Case of Linearly Elastic Bodies 361

8.1.1.4 Directed Concentrated Moments

Starting from the notion of concentrated force, we can construct directed


concentrated moments of the first order as well as directed concentrated moments
of a higher order.
Usually, in the theory of sliding forces, by a couple we mean a set of two forces
of the same modulus but of opposite direction; the lines of action of the two forces
must not coincide (if they do, the couple vanishes). The couple is characterized by
its moment, which is a vector normal to the plane determined by the lines of action
of the two component forces; its magnitude is equal to the area of the parallelo-
gram formed by one of the force and the arm of the couple, which connects the
points of application of the two forces (irrespective of their location on the lines of
action); its direction is such that it corresponds to a positive rotation in the plane of
the two forces. Such a couple is characteristic of a rigid body [13].
In the case of a continuously deformable body, however, we have to consider
sets of bound forces; in this case, a couple which results from a process involving a
passage to the limit (the arm of the couple tends to zero) vanishes no longer. We
are thus led to introduce the notion of directed concentrated moment.
Let F of components Fi ; i ¼ 1; 2; 3 be a concentrated force acting at the
fixed point A of position vector r0 ðx01 ; x02 ; x03 Þ and F1 ¼ F (equality in the sense of
free vectors) of components Fi ; i ¼ 1; 2; 3 a concentrated force acting at the var-
!
iable point B of position vector qðn1 ; n2 ; n3 Þ. We denote by d ¼ AB the arm of the
corresponding couple, by F0 the unit vector of F and by u the unit vector of the
arm of the couple (Fig. 8.4).
By definition, the directed concentrated moment at the point A is the limit, in
the sense of the theory of distributions, of the set of concentrated forces fF; F1 g;
when the arm d of the couple tends to zero. The point A is considered fixed and the
point B variable; we admit that the unit vectors F0 and u are constant and that the
magnitude M of the moment is constant.
The distance between the two lines of action is
 
d 0 ¼ du  F0  ¼ d sinðu; F0 Þ ð8:8Þ

and the magnitude M of the moment is

M ¼ Fd0 : ð8:9Þ

Fig. 8.4 Directed


concentrated moment 0
n F d' F1= F

A ( x10, x02 , x30 ) ud


B ( ξ1, ξ 2 , ξ 3 )

-F
362 8 Theory of Concentrated Loads

The loads Q0 and Q00 equivalent to the concentrated forces applied at the points
A and B, respectively, are

Q0d ðrÞ ¼ Fdðr  r0 Þ;


ð8:10Þ
Q00d ðrÞ ¼ Fdðr  qÞ;

hence, the set fF; F1 g is equivalent to

Qd ðrÞ ¼ Q0d ðrÞ þ Q00d ðrÞ


¼ F½dðr  qÞ  dðr  r0 Þ
MF0
¼   ½dðr  r0 Þ  dðr  qÞ; ð8:100 Þ
d u  F0 

where account has been taken of relations (8.8) and (8.9).


The load Q equivalent to the directed concentrated moment will therefore be
QðrÞ ¼ lim Qd ðrÞ ð8:11Þ
d!0

We may introduce the directional derivative in the direction of u of the Dirac


distribution by the relation
1 o
lim ½dðr  r0 Þ  dðr  qÞ ¼ dðr  r0 Þ: ð8:12Þ
d!0 d ou
In that case, the load corresponding to the directional moment is expressed by

MF0 o
QðrÞ ¼    dðr  r0 Þ: ð8:110 Þ
u  F0  ou

It is assumed that u and F0 are not collinear, hence

u  F0 6¼ 0 ð8:120 Þ
From the relation (8.110 ) we see that the directed concentrated moment is
characterized by: (i) the location Aðx01 ; x02 ; x03 Þ; (ii) the unit vector F0 of the forces
which generate the moment; (iii) the unit vector u of the direction in which the
passage to the limit is effected (one observes that different results are obtained,
depending on the choice of the location B on the line of action of the force F1);
(iv) the magnitude M of the moment.
The direction of the moment results from the above data and is specified by the
unit vector n of the normal to the plane in which the directed concentrated moment
is acting; the unit vector is chosen such that the scalar triple product ðu; F0 ; nÞ [ 0:
The directed concentrated moment thus defined is positive since the rotation, as
seen from the unit vector n, occurs in the positive direction.
An important particular case is that where the unit vectors u and F0 are normal
to each other ðu  F0 ¼ 0Þ; the corresponding equivalent load is
8.1 Case of Linearly Elastic Bodies 363

Fig. 8.5 Directed x3


concentrated moment
i3 F1= F
d
i1
O i2 B (a , d, 0 ) x2

-F
x1

o
QðrÞ ¼ MF0 dðr  r0 Þ ¼ MF0 ½u  graddðr  r0 Þ: ð8:13Þ
ou
For example, if we take the origin of the co-ordinate axes as point of application
and F0 ¼ i1 ; u ¼ i2 : Where ij ; j ¼ 1; 2; 3; are the unit vectors of the co-ordinate
axes; we obtain a direct concentrated moment, which induces a positive rotation in
the Ox1 x2 -plane (Fig. 8.5) and is expressed by the equivalent load
QðrÞ ¼ Mi1 d;2 ðrÞ: ð8:130 Þ
The directed concentrated moment introduced above is a moment of first order.
Starting from a directed concentrated moment of (n-1)th order, the vectors
ðn1Þ
Qðn1Þ and Q1 of which are applied at the points Aðr0 Þ and BðqÞ; respec-
tively, one may obtain, analogically, a directed concentrated moment of nth order.
The directed concentrated moments of higher order may be useful for certain
generalized deformable media.

8.1.1.5 Centre of Rotation

Proceeding from the notion of directed concentrated moment, we can define


another type of concentrated moment: the rotational concentrated moment.
Let thus be a plane P passing through the point Aðr0 Þ and n be the unit vector of
its normal (Fig. 8.6a). We denote by Q the set of directed concentrated moments
Qi ðrÞ; applied at the point A, having the unit vectors ui and F0i ; i ¼ 1; 2; . . .; n and
belonging to the plane P.
By definition, the sum of two or more elements of the set Q is called rotational
concentrated moment (centre of rotation), corresponding to the point A and the
plane P, when the sum does not depend on the unit vectors ui and F0i ;
i ¼ 1; 2; . . .; n:
The load Q is expressed by (the equivalent load of a moment of magnitude M)
(Fig. 8.6b)
X
n
QðrÞ ¼ Qi ðrÞ; n  2; ð8:14Þ
i¼1
364 8 Theory of Concentrated Loads

n n

Q2 Q1 P P
Qn
M
A ( x10, x02 , x30 )
A ( x10, x02 , x30 )

(a) (b)

Fig. 8.6 Rotational concentrated moment: a component directed moments; b centre of rotation

where

Mi F0i o
Qi ðrÞ ¼    dðr  r0 Þ; i ¼ 1; 2; . . .; n: ð8:15Þ
ui  F0i  oui

Replacing (8.15) in (8.14) and complying with the condition that the element
Q should not depend on the unit vectors ui and F0i ; i ¼ 1; 2; . . .; n one obtains

X
3
o
QðrÞ ¼ aj dðr  r0 Þ; ð8:16Þ
j¼1
oxj

where a1 ; a2 ; a3 are constant vectors depending only on the magnitudes Mi ; we


shall now show that the representation of the rotational concentrated moment in
the form (8.16) is unique.
Let us assume that this moment can be written also in the form
X
3
o
QðrÞ ¼ a0j dðr  r0 Þ; ð8:160 Þ
j¼1
oxj

then, by comparing the relations (8.16), (8.160 ) it follows that


X
3
o
ðaj  a0j Þ dðr  r0 Þ ¼ 0: ð8:17Þ
j¼1
oxj

If uðrÞ is an arbitrary fundamental function, we may write


!
X3
0 o
ðaj  aj Þ dðr  r0 Þ; uðrÞ ¼ 0; ð8:170 Þ
j¼1
oxj

whence
X
3
o
ðaj  a0j Þ uðr0 Þ ¼ 0; ð8:1700 Þ
j¼1
oxj
8.1 Case of Linearly Elastic Bodies 365

a relation which, owing to its linearity with respect to the derivatives of the
arbitrary fundamental function uðrÞ; can occur only if

aj ¼ a0j ; j ¼ 1; 2; 3: ð8:18Þ

Hence, the representation (8.16) of the rotational concentrated moment is unique,


which justifies the definition given above.
We shall now show that the set Q is non-void; this and the fact that the
representation (8.16) is unique allow the computation of the coefficients a1 ; a2 ; a3 :
Let thus the directed concentrated moments, represented by the loads Q1 and
Q2, be such that the following conditions are fulfilled:
M
M1 ¼ M2 ¼ ; ð8:19Þ
2
F01 ¼ u2 ; F02 ¼ u1 ; ð8:190 Þ

u1  u2 ¼ 0; u1  u2 ¼ n: ð8:1900 Þ
The equivalent load QðrÞ will be given, in this case, by
M 0 o M o
QðrÞ ¼  F1 dðr  r0 Þ  F02 dðr  r0 Þ
2 ou1 2 ou2
M
¼  fu2 ½u1  graddðr  r0 Þu1 ½u2  graddðr  r0 Þg
2
M
¼  ðu1  u2 Þ  graddðr  r0 Þ;
2
where we took into account the formula (A.41) of the vector triple product; thus,
the equivalent load of the centre of rotation is expressed in the form
1
QðrÞ ¼  Mn  graddðr  r0 Þ: ð8:20Þ
2
We notice that the vectors
1
aj ¼ Mij  n; j ¼ 1; 2; 3; ð8:21Þ
2
are coplanar and located in the plane P.
Thus we see that the rotational concentrated moment is characterized by: (i) the
location Aðr0 Þ; (ii) the plane P determined by the unit vector n of the normal to it
at the point A; (iii) the magnitude M. This moment thus defined is positive since
the rotation in the plane P, seen from n (Fig. 8.6b), is counter-clockwise. In this
way, the direction of the moment results from the data specified above and is
indicated by the unit vector n.
In particular, choosing the Ox1 x2 -plane as plane Pðn ¼ i3 Þ; one obtains the
equivalent load
366 8 Theory of Concentrated Loads

1
Qðx1 ; x2 Þ ¼  Mi3  graddðx1  x01 ; x2  x02 Þ; ð8:22Þ
2
of components [25]
1 o
Q1 ðx1 ; x2 Þ ¼ Mi1 dðx1  x01 ; x2  x02 Þ;
2 ox2
ð8:220 Þ
1 o
Q2 ðx1 ; x2 Þ ¼  Mi2 dðx1  x01 ; x2  x02 Þ:
2 ox1
Taking into account the above considerations, we may use as canonical rep-
resentation of a rotational concentrated moment the representation obtained by the
superposition of two directed concentrated moments of the same direction, same
magnitude, and the component forces of which are normal to each other.
Thus, the system of concentrated forces acting on the sides of a square
(Fig. 8.7a) leads to the superposition of two orthogonal directed concentrated
moments of the same magnitude M/2 and sign (Fig. 8.7b), hence to a canonical
representation of the rotational concentrated moment (Fig. 8.7c). We remark that
the result is the same for any position of a square of which centre is O. The fact
that we have considered the Ox1 x2 -plane particularizes in no way the problem
from a physical point of view. It should be mentioned too that this points out a
method of constructing (modelling) experimentally a rotational concentrated
moment.
It can be shown that starting from loads which are tangential to and uniformly
distributed over the sides of a regular polygon or from loads tangential to and
uniformly distributed over a circle, we obtain again a rotational concentrated
moment.

x2 x2 x2

M
M1 = 2 M

O x1 x1 O x1
O

M
M2= 2

(a) (b) (c)

Fig. 8.7 Rotational concentrated moment: a on the sides of a square; b two orthogonal directed
concentrated moments; c centre of rotation
8.1 Case of Linearly Elastic Bodies 367

8.1.1.6 Concentrated Moments of Linear Dipole Type

In Sect. 8.1.1.4 we have defined a concentrated moment assuming that the unit
vectors F0 and u are not collinear ðu  F0 6¼ 0Þ: For the case where they are such,
we introduce a new type of concentrated moment: the concentrated moment of
linear dipole type [12].
Let –F, of components Fi ; i ¼ 1; 2; 3; be a concentrated force applied at the
fixed point A of position vector rðx01 ; x02 ; x03 Þ and let F1 ¼ F (the equality is con-
sidered in the sense of free vectors) be another concentrated force applied at the
variable point B of position vector qðn1 ; n2 ; n3 Þ: Let u be the unit vector of
the force F (Fig. 8.8). We assume that the two concentrated forces have the same
!
line of action; hence, the unit vector of the vector AB will be again u. We set the
arm AB ¼ d and introduce the magnitude
D ¼ Fd: ð8:23Þ
By definition, the concentrated moment of dipole type (dipole of concentrated
forces) at the point A is the limit, in the sense of the theory of distributions, of the
set of concentrated forces ðF; F1 Þ; when the arm d of the couple tends to zero.
The point A is considered fixed and the point B variable; we assume that the two
concentrated forces have the same line of action and that the unit vector u and the
magnitude D of the dipole moment are constant.
The common line of action of the two concentrated forces is the line of action
of the dipole, which is considered to be positive if it tends to further separate the
points of application of the forces (Fig. 8.8) and negative in the opposite case.
The equivalent load of the set of two concentrated forces may be
Du
Qd ðrÞ ¼ Fu½dðr  qÞ  dðr  r0 Þ ¼  ½dðr  r0 Þ  dðr  qÞ; ð8:24Þ
d
passing to the limit for d ! þ0 in the sense of the theory of distributions, we
obtain the equivalent load of a concentrated moment of dipole type
o
QðrÞ ¼ Du dðr  r0 Þ ¼ D½u  graddðr  r0 Þu; ð8:25Þ
ou
where we have introduced the directional derivative in the direction defined by the
unit vector u.
It results that the concentrated moment of dipole type is characterized by:
(i) the location Aðr0 Þ; (ii) the unit vector u of the forces which generate the
moment; (iii) the magnitude D of the dipole moment.

Fig. 8.8 Concentrated -F u F1 = F


d
moment of linear dipole type
A ( x10, x02 , x30 ) B ( ξ1, ξ 2 , ξ 3 )
368 8 Theory of Concentrated Loads

The direction of the dipole is specified by the unit vector F0 of the force F; if
F0 ¼ u; which is the case considered above, then the dipole is positive, while if
F0 ¼ u; then the dipole is negative and the sign of the relation (8.25) must be
changed.
An important particular case is that where the point of application is the origin
of the co-ordinate axes and u ¼ i1 ; where i1 is the unit vector of the Ox1 -axis; we
thus obtain a dipole of concentrated forces, which is expressed by the equivalent
load

QðrÞ ¼ Di1 d01 ðrÞ: ð8:26Þ


The concentrated moment of linear dipole type introduced above is a moment of
first order; analogically, starting from a moment of dipole type of (n - 1)th order,
one may obtain a moment of nth order, hence of a higher order.

8.1.1.7 Centres of Dilatation

Starting from a concentrated moment of linear dipole type, one can construct
concentrated moments of plane or spatial dipole type. Let thus be a plane
P passing through the point Aðr0 Þ and n be the unit vector of the normal to the
plane. We denote by Q the set of moments of linear dipole type Qi ðrÞ; applied at
the point A; with the unit vectors ui ; i ¼ 1; 2; . . .; n belonging to the plane
P (Fig. 8.9a).
By definition, the sum of two or more elements of the set Q is called a con-
centrated moment of plane dipole type (centre of plane dilatation), corresponding
to the point A and to the plane P, in the case where the sum does not depend on the
unit vectors ui :
Such a sum, corresponding to a concentrated moment of plane dipole type is
(Fig. 8.9b)
X
n
QðrÞ ¼ Qi ðrÞ; n  2; ð8:27Þ
i¼1

n n

Q2 P P
Q1
Qn A ( x10, x02 , x30 )
Dp
A ( x10, x02 , x30 )

(a) (b)

Fig. 8.9 Concentrated moment of plane dipole type: a component linear dipoles; b cen-
tre of plane dilatation
8.1 Case of Linearly Elastic Bodies 369

where
o
Qi ðrÞ ¼ Di ui dðr  r0 Þ: ð8:270 Þ
oui
A study analogous to that in Sect. 8.1.1.5 concerning centres of rotation shows
that the representation is unique, the set being non-void; the corresponding vectors
aj are given by
1
aj ¼ Dp n  ðn  ij Þ; j ¼ 1; 2; 3: ð8:28Þ
2
Let thus be two positive dipoles of concentrated forces normal to each other,
applied at the point Aðr0 Þ and contained in the plane P. We may take
1
D1 ¼ D2 ¼ Dp ; ð8:29Þ
2
the unit vectors u1 and u2 will satisfy the relations
u1  u2 ¼ 0; u1  u2 ¼ n: ð8:290 Þ
The equivalent load QðrÞ is, in this case, given by
Dp
QðrÞ¼  fu1 ½u1  graddðr  r0 Þ þ u2 ½u2  graddðr  r0 Þg
2
Dp
¼  n  fu1 ½u2  graddðr  r0 Þ  u2 ½u1  graddðr  r0 Þg
2
Dp
¼ n  ½ðu1  u2 Þ  graddðr  r0 Þ;
2
where we took into account the second relation (8.280 ) and the formula (A.1.41) of
the triple vector product; thus, the equivalent load of the plane centre of dilatation
is given by
1
QðrÞ ¼ Dp n  ½n  dðr  r0 Þ: ð8:30Þ
2
The concentrated moment of plane dipole type is considered to be positive if it
is obtained from positive linear dipoles; in this case, the concentrated moment of
plane dipole type is a centre of plane dilatation. Conversely, if we use negative
linear dipoles, then we obtain a negative concentrated moment of plane dipole
type, i.e., a centre of plane contraction and the relation (8.30) takes a changed sign
(supposing that Dp [ 0Þ:
The concentrated moment of plane dipole type is characterized by: (1) the
location Aðr0 Þ; (2) the unit vector n of the normal to the plane whereon the plane
dipole is acting; (3) the magnitude Dp of the dipole.
370 8 Theory of Concentrated Loads

Taking into account the formula (8.20) which expresses the rotational con-
centrated moment and assuming that M ¼ Dp ; we may write

QDp ðrÞ þ n  QM ðrÞ ¼ 0: ð8:31Þ

In particular, if the plane P is just the Ox1 x2 -plane, then


1
QðrÞ ¼  Dp graddðx1  x01 ; x2  x02 Þ: ð8:32Þ
2
The system of concentrated forces acting on the sides of a square (Fig. 8.10a)
leads to the superposition of two orthogonal linear dipoles of the same magnitude
Dp =2 and sign (Fig. 8.10b), hence to a canonical representation of the plane centre
of dilatation (Fig. 8.10c). A method of constructing (modelling) experimentally
such a centre of dilatation is thus pointed out.
We mention that we obtain the same result if we consider uniformly distributed
normal loads acting on the circumference of a circle.
The above considerations may be extended to a three-dimensional concentrated
load. By definition, the sum of three or more elements of the spatial set Q is called
a concentrated moment of spatial dipole type (centre of spatial dilatation) cor-
responding to the point A, in the case where the sum does not depend on the unit
vectors ui :
An analogous study to that above leads to an equivalent load of the form (8.27)
but with n C 3. One can thus show that this set is unique and non-void, the vectors
ai being given by
1
aj ¼  Ds ij ; j ¼ 1; 2; 3; ð8:33Þ
3
where

x2 x2 x2

Dp
D1 = 2

O x1 O x1 x1
O Dp
Dp
D2 = 2

(a) (b) (c)

Fig. 8.10 Concentrated moment of plane dipole type: a on the sides of a square; b two orthog-
onal linear dipoles; c centre of plane dilatation
8.1 Case of Linearly Elastic Bodies 371

Fig. 8.11 Centre of spatial


dilatation x3

Ds
3
Ds
3
O x2

Ds
x1 3

1
D1 ¼ D2 ¼ D3 ¼ Ds : ð8:34Þ
3
Let now be three linear dipoles of concentrated forces of the same sign, their
lines of action being specified by the unit vectors u1 ; u2 ; u3 ; orthogonal to each
other, which form a dextrorsum trihedron (Fig. 8.11).
The equivalent load is given by
 
Ds o o o
QðrÞ ¼  u1 dðr  r0 Þ þ u2 dðr  r0 Þ þ u3 dðr  r0 Þ
3 ou1 ou2 ou3
Ds
¼  fu1 ½u1  graddðr  r0 Þ þ u2 ½u2  graddðr  r0 Þ
3
þu3 ½u3  graddðr  r0 Þg;

wherefrom
1
QðrÞ ¼  Ds graddðr  r0 Þ: ð8:35Þ
3
A concentrated moment of spatial dipole type is considered to be positive if it
has been obtained from positive linear dipoles; in this case, it is a centre of spatial
dilatation. If the component linear dipoles are negative, then we obtain a negative
concentrated moment of spatial dipole type; it represents a centre of spatial
contraction and the relation (8.35) takes a changed sign (supposing that Ds [ 0Þ:
The concentrated moment of spatial dipole type is characterized by: (i) the
location Aðr0 Þ; (ii) the magnitude Ds of the dipole.
A canonical representation of a concentrated moment of spatial dipole type
may be obtained by three concentrated moments of linear dipole type, orthogonal
to each other, of the same magnitude Ds =3 and sign.
We observe that one obtains the same result if one considers uniform distrib-
uted normal loads acting over the surface of a sphere.
It is interesting to remark that a spatial centre of dilatation introduces a
singularity of a different type than that of the plane centre of dilatation; indeed,
372 8 Theory of Concentrated Loads

the latter concentrated load may be obtained with the aid of d representative
sequences, while the first one necessities d0 representative sequences.
We mention that one can construct concentrated moments of sectorial dipole
type (of a circle or of a sphere).

8.1.2 Tensor Properties of Concentrated Loads

We have seen in the previous section how one may obtain directed concentrated
moments and concentrated moments of linear dipole type, starting from the notion
of concentrated force. By the superposition of two directed concentrated moments,
which are orthogonal and of the same magnitude and sign (and for which the arm
of the couple is normal to the direction of the component concentrated forces), we
obtain a rotational concentrated moment (centre of rotation), which has lost its
directional effect and leads to a mechanical phenomenon with axial antisymmetry;
in fact, this is the essential property of the concentrated moment.
In the same way, we can introduce a quadripole of concentrated forces,
obtained by superposing the effect of two orthogonal concentrated moments of
linear dipole type, of the same magnitude and sign; thus, we obtain a concentrated
moment of plane dipole type (centre of plane dilatation) which, likewise, has no
directional effect any longer and leads to a mechanical phenomenon with axial
symmetry.
By the superposition of effects of three orthogonal concentrated moments of
linear dipole type of the same magnitude and sign, we obtain a concentrated
moment of spatial dipole type (centre of spatial dilatation), which leads to a
mechanical phenomenon with central symmetry.
Results of this kind have been pointed out as early as the beginning of the
twentieth century, but without any further and deeper investigation of the matter.
In the sequel we shall justify the use of these concentrated loads and emphasize
their tensorial aspect; this will aid in classifying the concentrated loads. Using also
previous considerations of E. Kröner [3], we shall use the results given by us [28,
29].
In order to emphasize the different tensorial properties of concentrated loads,
we shall consider a deformable solid on which these loads are acting, thus
inducting in the body a state of strain and stress. We shall refer to a body on which
we may apply the principle of local superposition of effect, e.g., to a linearly elastic
body.
In the case of an elastic space (or an elastic plane), the state of strain and the
state of stress do not show any particular features depending on the direction of the
line of action of an internal concentrated load (concentrated force, directed con-
centrated moment, concentrated moment of linear dipole type etc.). But in the case
of a domain whose boundary is, in some part at least, at a finite distance we have to
deal with favourite directions which depend on the boundary.
8.1 Case of Linearly Elastic Bodies 373

Fig. 8.12 Tensor properties


of concentrated loads fi
( x10, x02 , x30 )
f ( fi )
( x 1, x 2 , x3 )

8.1.2.1 Concentrated Loads of First Order

Let fi ; i ¼ 1; 2; 3; be concentrated forces equal to unity and acting at the point of


position vector r0 ðx01 ; x02 ; x03 Þ in the direction of the co-ordinate axes Oxi ; i ¼ 1; 2; 3
(Fig. 8.12); it is easy to see that they are components of a vector (tensor of the first
order). Denoting by f ðfi Þ a component of the displacement vector, strain tensor
orstress tensor at the point of position vector rðx1 ; x2 ; x3 Þ; due to the action of the
force f1 ; f2 ; f3 ; respectively, we remark that f ðfi Þ; i ¼ 1; 2; 3; are also the compo-
nents of a tensor of first order.
By definition, concentrated loads which behave like a tensor of first order from
the point of view of the state of strain and stress they induce in a body are called
concentrated loads of first order.
Therefore, the concentrated forces are concentrated loads of first order.
We can also construct other concentrated loads of first order, e.g., dipoles of
centres of spatial dilatation. Such a concentrated load is obtained by a process of
passing to the limit, in the sense of the theory of distributions, when a variable
point BðqÞ approaches a fixed point Aðr0 Þ; it is assumed that the dipole of centres
of spatial contraction acts at the fixed point and a centre of spatial dilatation at the
variable point B. An analogous concentrated plane load of first order may be
obtained, starting from a centre of rotation, i.e. a dipole of centres of rotation.

8.1.2.2 Concentrated Loads of Second Order

Let Mij ; i 6¼ j; i; j ¼ 1; 2; 3; be the directed concentrated moments corresponding to


a rectangular trihedron Ox1 x2 x3 (we assume that the arm of the couple of each
directed concentrated moment is at a right angle to the common direction of the
component concentrated forces); let Dii ; i ¼ 1; 2; 3; be the concentrated moments
of linear dipole type corresponding to the co-ordinate axes. The directed con-
centrated moments are positive if they correspond to a positive rotation in the
plane wherein they act. In the above notation, the first index refers to the direction
of the component forces and the second index to the direction of the arm of the
couple.
Taking into account the conventions regarding the signs and the fact that the
magnitude of these concentrated moments is obtained as a product of the
374 8 Theory of Concentrated Loads

components of a vector (two tensors of the first order), we obtain a tensor of


second order whose components are
2 3
D11 M12 M13
4 M21 D22 M23 5: ð8:36Þ
M31 M32 D33
Let mij ; i 6¼ j; i; j ¼ 1; 2; 3; be the corresponding directed concentrated moments
of magnitude equal to unity, and dii ; i ¼ 1; 2; 3; the corresponding concentrated
moments of linear dipole type, equal to unity, applied at the point of position
vector r0 ðx01 ; x02 ; x03 Þ: Denoting by f ðmij Þ; f ðdii Þ a geometrical or a mechanical
magnitude induced at the point of position vector rðx1 ; x2 ; x3 Þ by a single con-
centrated load, and admitting the principle of local superposition of effects, we
obtain a tensor of second order of the form
2 3
f ðd11 Þ f ðm12 Þ f ðm13 Þ
Tf  4 f ðm21 Þ f ðd22 Þ f ðm23 Þ 5: ð8:37Þ
f ðm31 Þ f ðm32 Þ f ðd33 Þ
Hence, by definition, concentrated loads which are behaving like a tensor of the
second order from the point of view of the state of strain and stress they induce are
called concentrated loads of second order.
Directed concentrated moments and concentrated moments of linear dipole type
are concentrated loads of second order.
The antisymmetric part of the tensor Tf leads to a vector whose components are
1
f ðm1 Þ ¼ ½ f ðm23 Þ þ f ðm32 Þ;
2
1
f ðm2 Þ ¼ ½ f ðm31 Þ þ f ðm13 Þ; ð8:38Þ
2
1
f ðm3 Þ ¼ ½ f ðm12 Þ þ f ðm21 Þ;
2
thus, we obtain a rotational concentrated moment, which is a concentrated load of
first order.
The determination of the principal directions of the tensor Tf allows a correct
representation of a concentrated moment acting on the surface of an elastic body
(the concentrated moments of linear dipole type vanishing).

8.1.2.3 Concentrated Loads of Zero Order

By a similar procedure we state by definition that concentrated loads which behave


like a scalar (tensor of zero order) from the point of view of the state of strain and
stress they induce are called concentrated loads of zero order.
The first invariant of the tensor Tf ; which is a scalar, leads to a centre of spatial
dilatation, expressed by
8.1 Case of Linearly Elastic Bodies 375

1
f ðds Þ ¼ ½f ðd11 Þ þ f ðd22 Þ þ f ðd33 Þ; ð8:39Þ
3
hence, the centres of spatial dilatation are concentrated loads of the zero order.

8.1.2.4 Plane Case

In the Ox1 x2 -plane, the tensor Tf has the form


 
f ðd11 Þ f ðm12 Þ
Tf ¼ ; ð8:40Þ
f ðm21 Þ f ðd22 Þ

which leads to the centre of plane dilatation


1
f ðdp Þ ¼ ½f ðd11 Þ þ f ðd22 Þ ð8:41Þ
2
and to the rotational concentrated moment
1
f ðmÞ ¼ ½f ðm12 Þ þ f ðm21 Þ: ð8:42Þ
2
The last relation is also an invariant in the plane case; therefore, the rotational
concentrated moment behaves like a load of the zero order.
In various classical textbooks the influence of a rotational concentrated moment
is often introduced with the aid of a single concentrated moment; this is possible
only in very particular cases. The problem may be associated with that of the
principal directions of the tensor Tf :
For example, in the plane case one can show that the principal directions are
expressed by
f ðm12 Þ  f ðm21 Þ
tan 2a1 ¼  ; ð8:43Þ
f ðd11 Þ  f ðd22 Þ
where a1 is the angle made with the Ox1 -axis; this leads to
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1
fextr ðda1 Þ ¼ f ðdp Þ ½f ðd11 Þ  f ðd22 Þ2 þ½f ðm12 Þ  f ðm21 Þ2 ð8:44Þ
2
and to
f ðma1 Þ ¼ f ðmÞ: ð8:440 Þ
Similarly, the principal direction expressed by
f ðd11 Þ  f ðd22 Þ
tan 2a2 ¼ ð8:45Þ
f ðm12 Þ  f ðm21 Þ
376 8 Theory of Concentrated Loads

bisect the preceding principal directions and lead to


qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1
fextr ðma2 Þ ¼ f ðmÞ ½f ðd11 Þ  f ðd22 Þ2 þ½f ðm12 Þ  f ðm21 Þ2 ð8:46Þ
2
and to
f ðda2 Þ ¼ f ðdp Þ: ð8:460 Þ
The formulae (8.440 ) and (8.460 ) show that the effect of a rotational concen-
trated moment or of a centre of plane dilatation cannot be correctly introduced by
means of a directed concentrated moment or of a concentrated moment of linear
dipole type, respectively, unless the direction of the latter loads is a principal
direction for the concentrated moments of linear dipole type or for the directed
concentrated moments, respectively.
We observe that we can also write the relation
fextr ðma2 Þ  fextr ðda1 Þ ¼ f ðmÞ  f ðdp Þ ð8:47Þ
The above results may be also used when the concentrated loads are acting on
the boundary of a plane domain. An interesting problem is the passage from a
point inside the domain to a point on the boundary; in fact, this is the only way of
obtaining a concentrated load on the boundary.

8.1.2.5 Concentrated Loads of nth Order

We have seen how, by proceeding from a concentrated force (concentrated load of


first order) and passing to the limit, in the sense of the theory of distributions, to a
neighbouring point, we obtain a concentrated load of second order. By a similar
procedure, starting from a concentrated load of (n–1)th order, one obtains a
concentrated load of nth order.
We remark further that, by a tensor contraction, of a concentrated load of nth
order one may obtain a concentrated load of (n–2)th order. That is, e.g., the case a
centre of spatial dilatation.

8.1.3 Solutions for Concentrated Loads

We give, in the following, solutions for the concentrated loads considered above
which act in an elastic space (infinite domain); we deal thus with the static case, as
well as with the dynamic one.
8.1 Case of Linearly Elastic Bodies 377

8.1.3.1 Static Case

We have considered in Sect. 6.2.4.2 the problem of the fundamental solution in


case of an elastic space 1\x1 ; x2 ; x3 \1; subjected to the action of volume
forces Fi ; i ¼ 1; 2; 3; expressed by means of distributions. We obtained thus the
Kelvin-Somigliana fundamental solution tensor (6.130), which leads to the state of
displacements (6.135), corresponding to a concentrated force applied at the origin;
using the Kelvin-Somigliana displacement vector, the displacement vector corre-
sponding to a concentrated force F acting (for the sake of simplicity) at the origin
O is given by [4]
 
1 1 rF
u¼ ð3  4mÞF þ 2 r : ð8:48Þ
16pð1  mÞG r r
If the force F acts along the Ox1 -axis ðF ¼ Fi1 Þ; then one obtains

F 1 x21
u1 ¼ 3  4m þ 2 ;
16pð1  mÞG r r
F x1 x2
u2 ¼ ; ð8:49Þ
16pð1  mÞG r 3
F x1 x3
u3 ¼ :
16pð1  mÞG r 3
Corresponding to a formulation in stresses (eventually starting from the Kelvin-
Somigliana fundamental solution tensor), one obtains the fundamental solution
tensor (6.1380 ); for the concentrated force F acting at the origin one gets
1 1h xi xj xk i
rij ¼ 3
ð1  2mÞðxk Fk dij  xi Fj  xj Fi Þ  3Fk 2 ;
8pð1  mÞ r r
i; j ¼ 1; 2; 3; ð8:50Þ

and the stress vector becomes



n 1 1 3
p¼ ð1  2mÞ½ðr  FÞn  ðF  nÞr  ðr  nÞF 2 ðr  FÞðr  nÞr :
8pð1  mÞ r 3 r
ð8:500 Þ
If the force F acts along the Ox1 -axis, then the state of stress reads

F x1 x21
r11 ¼ 1  2m þ 3 2 ;
8pð1  mÞ r 3 r

F x1 x22
r22 ¼ 1  2m  3 2 ; ð8:51Þ
8pð1  mÞ r 3 r

F x1 x23
r33 ¼ 1  2m  3 2 ;
8pð1  mÞ r 3 r
378 8 Theory of Concentrated Loads

3F x1 x2 x3
r23 ¼  ;
8pð1  mÞ r 5

F x3 x21
r31 ¼ 1  2m þ 3 2 ; ð8:510 Þ
8pð1  mÞ r 3 r

F x2 x21
r12 ¼ 1  2m þ 3 2 ;
8pð1  mÞ r 3 r

the radius r is given by


pffiffiffiffiffiffiffiffi
r¼ xk xk : ð8:52Þ
Interesting results may be obtained in cylindrical and spherical co-ordinates.
This solution has been given by W. Thomson [19].
In the case of arbitrary forces, one uses the convolution product, obtaining the
formulae (6.131) and (6.140).
Let now consider the elastic space acted upon, at the point Aðr0 Þ; by a directed
concentrated moment of magnitude M and specified by the unit vectors u and F0 :
The state of displacement will be given by
M M o F0
ui ðrÞ ¼  0
Fj0 uij ðr  r0 Þ ¼  u ;
0 ou i
i ¼ 1; 2; 3; ð8:53Þ
ju  F j ju  F j
0
where uFi are the displacements corresponding to the action of a concentrated
force F0 of magnitude equal to unity (unit vector), expressed by formulae of the
form (6.134), where the directional derivative has been introduced.
In particular, if the directed concentrated moment is applied at the origin O, in
the Ox1 x2 -plane, and if we take F0 ¼ i1 ; u ¼ i2 ; it results the equivalent load
o
Q1 ðrÞ ¼ M dðrÞ; Q2 ¼ Q3 ¼ 0; ð8:54Þ
ox2
the state of displacement takes the form

o M x2 x21
u1 ðrÞ ¼ M u11 ðrÞ ¼  3  4m þ 2 ;
ox2 16pð1  mÞG r 3 r
2

o M x1 x
u2 ðrÞ ¼ M u21 ðrÞ ¼ 1  3 22 ; ð8:55Þ
ox2 16pð1  mÞG r 3 r
o 3M x1 x2 x3
u3 ðrÞ ¼ M u31 ðrÞ ¼  :
ox2 16pð1  mÞG r 5
We remark that the derivative occuring in the above expressions may be
considered in the ordinary sense since the formal derivative of the components of
the tensor uij are locally integrable functions.
8.1 Case of Linearly Elastic Bodies 379

Let us consider now the elastic space acted upon at the point Aðr0 Þ by a
rotational concentrated moment (centre of rotation) of magnitude M, specified by
the unit vector n, normal to the plane of action. We introduce the equivalent load
1
Qi ðrÞ ¼  Mijk nj d;k ðr  r0 Þ; i ¼ 1; 2; 3: ð8:56Þ
2
It results the state of displacement
1
ui ðrÞ ¼  Mjkl nk uij;l ðr  r0 Þ; i ¼ 1; 2; 3: ð8:57Þ
2
In particular, if n ¼ i3 and r0 ¼ 0; one obtains a centre of rotation applied at the
origin O, which corresponds to a positive rotation in the Ox1 x2 -plane; there cor-
responds the equivalent load
1 1
Q1 ðrÞ ¼  Md;2 ðrÞ; Q2 ðrÞ ¼  Md;1 ðrÞ; Q2 ðrÞ ¼ 0 ð8:560 Þ
2 2
and the state of displacement reads
 
1 o o M x2
u1 ðrÞ ¼ M u11 ðrÞ  u12 ðrÞ ¼  ;
2 ox2 ox1 8pG r 3
 
1 o o M x1
u2 ðrÞ ¼ M u21 ðrÞ  u22 ðrÞ ¼ ; ð8:58Þ
2 ox2 ox1 8pG r 3
 
1 o o
u3 ðrÞ ¼ M u31 ðrÞ  u32 ðrÞ ¼ 0:
2 ox2 ox1
In the case of a concentrated moment of linear dipole type acting at the origin
O along the Ox1 -axis, the magnitude of which is D, one obtains

D x1 x21
u1 ðrÞ ¼ 1  4m þ 3 2 ;
16pð1  mÞG r 3 r

D x2 x2
u2 ðrÞ ¼  3
1  3 21 ; ð8:59Þ
16pð1  mÞG r r

D x3 x21
u3 ðrÞ ¼  13 2 :
16pð1  mÞG r 3 r
Calculating, analogically, the state of displacement corresponding to a con-
centrated moment of linear dipole type acting at the origin O along the Ox2 -axis, of
magnitude D and of the same sign, and superposing the actions of both linear
dipoles, one obtains a concentrated moment of plane dipole type (centre of plane
dilatation) of magnitude D=2 (taking into account the magnitudes of the two linear
dipoles). Thus, the centre of dilatation which acts in the Ox1 x2 -plane leads to the
state of displacement
380 8 Theory of Concentrated Loads

D x1 x23
u1 ðrÞ ¼ 3  4m þ 3 ;
32pð1  mÞG r 3 r2

D x2 x23
u2 ðrÞ ¼ 1  4m  3 2 ; ð8:60Þ
32pð1  mÞG r 3 r

D x3 x2
u3 ðrÞ ¼ 3
1  3 23 :
32pð1  mÞG r r
Superposing, analogically, a directed moment of linear dipole type, acting at
O along the Ox3 -axis, of the same magnitude D and of the same sign, one obtains a
concentrated moment of spatial dipole type (centre of spatial dilatation) of mag-
nitude D=3 (with respect to the magnitudes of the three linear dipoles). The cor-
responding state of displacement reads
ð1  2mÞD x1 ð1  2mÞD x2
u1 ðrÞ ¼ ; u2 ðrÞ ¼ ;
24pð1  mÞG r 3 24pð1  mÞG r 3
ð8:61Þ
ð1  2mÞD x3
u3 ðrÞ ¼ :
24pð1  mÞG r 3

8.1.3.2 Dynamic Case

We dealt in Sect. 6.2.4.3 with problems of linear elastodynamics; considering the


elastic space 1\x1 ; x2 ; x3 \1; the fundamental solution tensor uij in the sense
of the theory of distributions has thus been obtained, being given by (6.163).
Consequently, the generalized displacements (of the form (6.148)), corresponding
to an arbitrary generalized volume force of components

Fi ðr; tÞ ¼ Fi ðr; tÞhðtÞ; i ¼ 1; 2; 3: ð8:62Þ


are given by

ui ðr; tÞ ¼ uij
Fj ; i ¼ 1; 2; 3; ð8:63Þ

where the product of convolution refers both to the space and time variables. We
also assume that the initial conditions are homogeneous (null).
In the case of a concentrated force

Fi ðtÞ ¼ Fi ðtÞhðtÞ; i ¼ 1; 2; 3; ð8:64Þ


8.1 Case of Linearly Elastic Bodies 381

which is acting at the origin O, one obtains the generalized displacements


1 1 r
ui ðr; tÞ ¼ Fi t 
4pG r c2



2 1 r r
þ c2 Fj ðtÞ
t  t ; i ¼ 1; 2; 3: ð8:640 Þ
r c1 þ c2 þ ;ij

If we take

Fi ðtÞ ¼ Fi dðtÞ; i ¼ 1; 2; 3; ð8:65Þ


that is in the case of a shock at the initial moment, acting at the origin O, we obtain
the state of generalized displacement

Fj0 1 r
ui ðr; tÞ ¼ d t dij
4pG r c2



1 r r
þ c22 t  t ; i ¼ 1; 2; 3: ð8:650 Þ
r c1 þ c2 þ ;ij

Finally, if we apply suddenly a concentrated force

Fi ðtÞ ¼ Fi0 hðtÞ; i ¼ 1; 2; 3; ð8:66Þ

at the origin O at the initial moment, and then maintain it constant in time, the
state of generalized displacement reads
(

Fj0 2 r
ui ðr; tÞ ¼ h t dij
8pG r c2
( "

#) )
2 1 r 2 r 2
þ c2 t  t ; i ¼ 1; 2; 3: ð8:660 Þ
r c1 þ c2 þ
;ij

One observes that


8 r
>
> 0; t ;
>
> c1


>
< r r r
r r t  ;  t ;
t  t ¼ c c c ð8:67Þ
c1 þ c2 þ >>
1 1 2
>
> c  c r
>
: 1 2
r; t  ;
c1 c2 c2
382 8 Theory of Concentrated Loads

and
8 r
>
> 0; t ;
>
> c
>
2 1

2
2 >>
< r r r
r r t ; t ;
t  t ¼ c c c ð8:670 Þ
c1 þ c2 þ >> 1 1 2
>
>

>
> c 1  c 2 c 1 þ c 2 r
>
: 2t  r r; t  :
c1 c2 c1 c2 c2
We shall now consider the elastic space acted on by a rotational concentrated
moment of magnitude MðtÞ; applied at the origin O in a plane of normal n; taking
into account the equivalent load (8.20), the generalized displacement vector reads


1 1 r
uðr; tÞ ¼  n  grad M t  ; ð8:68Þ
8pG r c2

where

MðtÞ ¼ MðtÞhðtÞ: ð8:69Þ


In particular, if the centre of rotation is acting in the Ox1 x2 -plane ðn ¼ i3 Þ; we
obtain the generalized displacements


1 1 r
u1 ðr; tÞ ¼ M t ;
8pG r c2 ;2

 ð8:70Þ
1 1 r
u2 ðr; tÞ ¼  M t ; u3 ðr; tÞ ¼ 0:
8pG r c2 ;1

If

MðtÞ ¼ M0 dðtÞ; ð8:71Þ


which corresponds to a centre of rotation applied as an initial shock, we may write


M0 1 r
uðr; tÞ ¼  n  grad d t  ; ð8:710 Þ
8pG r c2

while if

MðtÞ ¼ M0 hðtÞ; ð8:72Þ


which corresponds to a centre of rotation applied suddenly at the initial moment
and then maintained constant in time, we obtain


M0 1 r
uðr; tÞ ¼  n  grad h t  : ð8:720 Þ
8pG r c2
8.1 Case of Linearly Elastic Bodies 383

Let us now consider the elastic space acted on at the origin O by a centre of
spatial dilatation (concentrated moment of spatial dipole type) of magnitude Ds ðtÞ;
taking into account the equivalent load (8.35), the generalized displacement vector
reads


1 c21 1 r
uðr; tÞ ¼  grad Ds t 
12pG c22 r c1


ð1  2mÞ 1 r
¼ grad Ds t  ; ð8:73Þ
24pð1  mÞG r c1

Ds ðtÞ ¼ Ds hðtÞ: ð8:74Þ


In particular, if

Ds ðtÞ ¼ D0s dðtÞ; ð8:75Þ

then we have to deal with an initial shock and there follows the generalized
displacement vector


ð1  2mÞD0s 1 r
uðr; tÞ ¼  grad d t  ; ð8:750 Þ
24pð1  mÞG r c1

while if

Ds ðtÞ ¼ D0s hðtÞ; ð8:76Þ

then the centre of spatial dilatation corresponds to a load applied suddenly and
maintained constant in time and the generalized displacement vector reads


ð1  2mÞD0s 1 r
uðr; tÞ ¼  grad h t  : ð8:760 Þ
24pð1  mÞG r c1

8.2 Case of Linearly Elastic Cosserat Type Bodies

In addition to the concentrated loads considered in the previous paragraph, in the


case of Cosserat type bodies appear volume moments too; the problem of the
relations between these moments and the centres of rotations studied in the clas-
sical case is thus put.
The cases of the concentrated force, of the concentrated volume moment and of
the centre of rotation have been considered by R. D. Mindlin and N. F. Tiersten
[20] for Cosserat type bodies with constraint rotations. In the case of Cosserat type
bodies with free rotations, acted upon by volume loads (concentrated forces and
moments), R. D. Mindlin [19] has given results in a condensed form for the
potential functions of Papkovich-Neuber type, while N. S ßandru [22–24] obtained
384 8 Theory of Concentrated Loads

the displacement and rotation vectors. Other concentrated loads have been con-
sidered by us [30–32] and then, independently, by S. Kessel [18].

8.2.1 Solutions for Concentrated Loads

We deal firstly with the case of a concentrated force; then we shall consider other
concentrated loads of second and zero order.

8.2.1.1 Concentrated Force

The volume force F may be expressed in the form


F ¼ gradP0 þ curlP; ð8:77Þ

where
1 1
P0 ðrÞ ¼  F  grad ; ð8:78Þ
4p r
1 1
PðrÞ ¼  F  grad ; ð8:780 Þ
4p r
in case of a concentrated force FdðrÞ; the Eqs. (7.6100 ) and (7.6200 ) lead thus to the
equations
1 1
ðk þ 2lÞDK0 ¼ F  grad ; ð8:79Þ
4p r
1 1
D½ðl þ aÞðl0 þ a0 ÞD  4laK ¼ F  grad ; ð8:790 Þ
4p r
respectively. One obtains thus
1 rF
K0 ¼ ; ð8:80Þ
8pðk þ 2lÞ r
 
l 1 1
K¼ curl v þ ð1  ev Þ F ; ð8:800 Þ
16pla 2 v

where l is an elastic constant of the nature of a length, given by (7.51), while v is


the non-dimensional quantity
r
v¼ : ð8:81Þ
l
8.2 Case of Linearly Elastic Cosserat Type Bodies 385

Taking into account the representations (7.610 ), (7.620 ), the state of displace-
ment and rotation is given by
   
1 1 rF k 0 þ l0 1 v
u¼ ð3  4mÞF þ 2 r  curl curl ð1  e ÞF ;
16pð1  mÞl r r 16pl r
ð8:82Þ
 
1 1
U¼ curl ð1  ev ÞF : ð8:820 Þ
8pl r
In the case of a body of Cosserat type with constraint rotations, the state of
displacement and rotation is given by the same formulae (8.82), (8.820 ), but the
non-dimensional constant v is replaced by
r
v0 ¼ 0 ; ð8:810 Þ
l
where l0 is given by (7.510 ).

8.2.1.2 Directed Concentrated Moments

For a force F ¼ F1 i1 acting at the origin O along the Ox1 -axis it results
F1 1h x1 i
u¼ ð3  4mÞi1 þ 2 r
16pð1  mÞl r r
ð8:83Þ
l02 F1 1 n 
2 v
 x1    o
2 v
þ 1  1 þ v þ v e i1  2 3  3 þ 3v þ v e r ;
4pl r 2 r
F1 1
U¼ ½1  ð1 þ vÞev ðx3 i2  x2 i3 Þ: ð8:830 Þ
8pl r 3
Taking into consideration also the force F1 i1 which acts at the point ð0; d; 0Þ
and passing to the limit in the sense of the theory of distributions, one obtains thus
the directed concentrated moment M12 ; the corresponding state of displacement
and rotation reads
M12 1h x1 x2 i
u¼ ð 3  4m Þx 2 i1  x 1 i 2 þ 3 r
16pð1  mÞl r 3 r2

l02 M12 1   
 5
3  3 þ 3v þ 2v2 þ v3 ev x2 i1
4pl r
  
þ 3  3 þ 3v þ v2 ev x1 i2

x1 x2  2
 
3 v
 2 15  15 þ 15v þ 6v þ v e r ; ð8:84Þ
r
386 8 Theory of Concentrated Loads


M12 1
U¼ ½1  ð1 þ vÞev i3
8pl r 3

x2   
2 v
þ 2 3  3 þ 3v þ v e ðx3 i2  x2 i3 Þ : ð8:840 Þ
r
In the case of constraint rotations, the above formulae remain valid if one
replaces v by v0 :
If we suppose that the force F1 i1 acts at the point ðd; 0; 0Þ and if we pass to
the limit in the sense of the theory of distributions, we obtain the directed con-
centrated moment of linear dipole type D11 ; in this case, the state of displacement
and rotation is given by


D11 1 x21
u¼ 2 ð 1  2m Þx i
1 1  1  3 r
16pð1  mÞl r 3 r2

l02 D11 1   
þ 5
6  6 þ 6v þ 3v2 þ v3 ev x1 i1
4pl r

  x2   
þ 3  3 þ 3v þ v2 ev  21 15  15 þ 15v þ 6v2 þ v3 ev r ;
r
ð8:90Þ
D11 x1   
U¼ 5
3  3 þ 3v þ v2 ev ðx3 i2  x2 i3 Þ: ð8:900 Þ
8pl r

8.2.2 Centres of Dilatation. Centre of Rotation

In the following, we deal with concentrated loads obtained starting from the loads
considered in Sect. 8.2.1.2, i.e. with concentrated moments of plane and spatial
dipole type as well as with rotational concentrated moments. As well, we introduce
the influence of a concentrated volume moment.

8.2.2.1 Centres of Dilatation

Calculating, analogically to the dipole D11 ; the state of displacement and rotation
corresponding to the dipole D22 and superposing the effects, one obtains the state
of displacement and rotation corresponding to a concentrated moment of plane
dipole type of magnitude D3 =2 ¼ D11 ¼ D22 and direction i3 (centre of plane
dilatation)
8.2 Case of Linearly Elastic Cosserat Type Bodies 387



D3 1 x23
u¼  2 ð 1  2m Þx i
3 3  3  4m  3 r
32pð1  mÞl r 3 r2

l02 D3 1   
 5
6  6 þ 6v þ 3v2 þ v3 ev x3 i3
8pl r

 x2   
þ 3  3 þ 3v þ v2 ev  23 15  15 þ 15v þ 6v2 þ v3 ev r ;
r
ð8:91Þ
D 3 x3   
U¼ 5
3  3 þ v þ v2 ev ðx2 i1  x1 i2 Þ: ð8:910 Þ
16pl r
In the case of constrained rotations one replaces the constant v by the constant
v0 :
Superposing also the action of the linear dipole D33 and assuming that D=3 ¼
D11 ¼ D22 ¼ D33 ; one obtains a state of displacement and rotation with central
symmetry, corresponding to a centre of spatial dilatation, i.e.
ð1  2mÞD 1
u¼ r; ð8:92Þ
24pð1  mÞl r 3

U ¼ 0: ð8:920 Þ
One observes that the results thus obtained are the same as those in the classical
elasticity, the effect of mechanical asymmetry being lost.

8.2.2.2 Centre of Rotation. Concentrated Volume Moment.


Fundamental Concentrated Loads

We obtained, in Sect. 8.2.1.2, the state of displacement and rotation for a directed
concentrated moment M12 ; analogically, in the case of a directed concentrated
moment M21 ; it results
M21 1h x1 x2 i
u¼ x i
2 1  ð 3  4m Þx i
1 2  3 r
16pð1  mÞl r 3 r2
l02 M21 1 n     
þ 5
3  3 þ 3v þ v2 ev x2 i1 þ 3  3 þ 3v þ 2v2 þ v3 ev x1 i2
4pl r
x1 x2    o
 2 15  15 þ 15v þ 6v2 þ v3 ev r ; ð8:93Þ
r
M21 1 n
U¼ ½1  ð1 þ vÞev i3
8pl r 3
x1    o
þ 2 3  3 þ 3v þ v2 ev ðx3 i1  x1 i3 Þ : ð8:930 Þ
r
388 8 Theory of Concentrated Loads

Superposing the effects of the directed concentrated moments M12 and M21 ; one
obtains a rotational concentrated moment (centre of rotation) of magnitude
M3 =2 ¼ M12 ¼ M21 and direction i3 ; i.e.
 
M3 1 a v
u¼ 1 ð1 þ vÞe r3 ih ; ð8:94Þ
8pl r 3 lþa

M3 1   
U¼  3
1  1 þ v þ v2 ev i3
16pl r

x3   
 2 3  3 þ 3v þ v2 ev r ; ð8:940 Þ
r

where
qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r3 ¼ x21 þ x22 ; ð8:95Þ

the unit vector ih (polar co-ordinates in the Ox1 x2 -plane) being given by
r3 ih ¼ x2 i1  x1 i2 : ð8:950 Þ
If we consider a concentrated volume moment M, acting at the origin O and put
MdðrÞ ¼ gradP0 þ curlP; ð8:96Þ
then, taking into account (7.630 ) and (7.640 ), we get
u ¼ 2acurl curlK ð8:97Þ

U ¼ gradK0 þ ðl þ aÞD curlK; ð8:970 Þ


where the potentials K0 and K are given by the equations
½ðk0 þ 2l0 ÞD  4aK0 þ P0 ¼ 0; ð8:98Þ

D½ðl þ aÞðl0 þ a0 ÞD  4laK þ P ¼ 0: ð8:980 Þ


By integration, we obtain  
1 1
K0 ¼  M  grad ð1  ej Þ ; ð8:99Þ
16pa r
 
1 r l2
K¼ curl þ ð1  ej Þ M ; ð8:990 Þ
16pla 2 r

where we have introduced the non-dimensional quantity


r
,¼ ; ð8:100Þ
h
where h is an elastic constant of the nature of a length, given by (7.52).
8.2 Case of Linearly Elastic Cosserat Type Bodies 389

Taking into account the representation (8.97), (8.970 ), the state of displacement
and rotation is given by
 
1 1 ,
u¼ curl ð1  e ÞM ; ð8:101Þ
8pl r
 
1 1
U¼ grad M  grad ð1  e, Þ
16pa r
 
lþa 1
þ curlcurl ð1  e, Þ M : ð8:1010 Þ
16pal r

In the case of the Cosserat type body with constraint rotations, the displacement
conserves the form (8.101), but the rotation becomes

1 1
U¼ curl curl M ; ð8:102Þ
16p r

because a ! 1; h ! 0; so that , ! 1:
The concentrated volume moment M ¼ Mi3 leads to
M 1
u¼ ½1  ð1 þ vÞev r3 ih ; ð8:103Þ
8pl r 3
M 1n , x3    o
2 ,
U¼ ½ 1  ð 1 þ , Þe i 3  3  3 þ 3, þ , e r
16pa r 3 r2
ðl þ aÞM 1  n  
 3
1  1 þ v þ v2 ev i3
16pla r
x3    o
 2 3  3 þ 3v þ v2 ev r : ð8:1030 Þ
r
Comparing the axial antisymmetric states of displacement and rotation (8.94),
(8.940 ) and (8.103), (8.1030 ), one observes that the centre of rotation M3 and the
concentrated volume moment M lead to different results for M3 ¼ M: It is thus to
be noticed that in the case of Cosserat type bodies with free rotations the con-
centrated volume moment M represents a concentrated load which cannot be
constructed starting from the notion of concentrated force as in the case of the
centre of rotation M3 : Thus, from a quantitative point of view, one can state that, in
a Cosserat continuum, the concentrated volume moment M is a fundamental load
as well as the concentrated volume force F [32–34].
Starting from the concentrated volume moment, one can thus construct many
other concentrated loads. Evidently, one can construct also loads of a mixed
nature, using simultaneously concentrated volume forces and moments.
For M3 ¼ M in the formulae (8.94), (8.940 ), (8.103), (8.1030 ) one observes that
the differences uðMÞ  uðM3 Þ and UðMÞ  UðM3 Þ tend to zero only for a !
1ðl ¼ l0 ; h ¼ 0; , ! 1Þ: Thus, only in the case of constrained rotations, the
centre of rotation and the concentrated volume moment give the same results, i.e.
390 8 Theory of Concentrated Loads

M 1h 0 v0
i
u¼ 1  ð1 þ v Þe r3 ih ; ð8:104Þ
8pl r 3
(
M 1 h 0 02 v0
i
U¼ 1  ð1 þ v þ v Þe i3
16pl r 3
)
x3 h 0 02 v0
i
 2 3  ð3 þ 3v þ v Þe r ; ð8:1040 Þ
r

where v0 is given by (8.810 ).

References
A. Books

1. Kecs, W., Teodorescu, P.P.: Applications of the Theory of Distributions in Mechanics. Ed.
Acad., Bucuresßti; Abacus Press, Tunbridge Wells, Kent (1974)
2. Kecs, W., Teodorescu, P.P.: Vvedenie v teoriyu obobshchennykh funktsiı̆ s prilozheniyami v
tekhnike (Introduction to the Theory of Distributions with Applications in Technics). Izd.
‘‘Mir’’, Moskva (1978)
3. Kr}oner, E.: Kontinuumstheorie der Versetzungen und Eigenspannungen. Springer, Berlin
(1958)
4. Love, A.E.H.: A Treatise on the Mathematical Theory of Elasticity, IVth edn. Cambridge
University Press, London (1934)
5. Nowacki, W.: Teoria niesymetrycznej spre_zystości (Theory of Asymmetrical Elasticity).
Państ. Wydawn. Nauk., Warszawa (1971)
6. Teodorescu, P.P.: Probleme spatßiale în teoria elasticitătßii (Spatial Problems in the Theory of
Elasticity). Ed. Acad., Bucuresßti (1970)
7. Teodorescu, P.P.: Dynamics of Linear Elastic Bodies. Ed. Acad. Bucuresßti; Abacus Press,
Tunbridge Wells, Kent (1975)
8. Teodorescu, P.P. (ed.): Probleme actuale în mecanica solidelor (Present-day Problems in
Solid Mechanics). Ed. Acad., Bucuresßti (1975)
9. Teodorescu, P.P., Ille, V.: Teoria elasticitătßii ßsi introducere în mecanica solidelor deformabile
(Theory of Elasticity and Introduction in Mechanics of Deformable Solids). I. Dacia, Cluj
Napoca (1976)

B. Papers

10. Khan, S.M., Dhaliwal, R.S., Chowdhury, K.L.: Singular solutions and green’s method in
micropolar theory of elasticity. Appl. Sci. Res. 25, 65 (1971)
11. Kecs, W.: Reprezentarea câmpurilor vectoriale cu ajutorul functßiilor generalizate
(Representation of the vector fields with the aid of generalized functions). St. Cerc. Mat.
15, 265 (1964)
12. Kecs, W.: Représentation mathématique des moments du type dipôle. Rev. Roum. Math.
Pures Appl. 10, 1391 (1965)
13. Kecs, W.: Les solutions généralisées des problèmes concernant l’espace élastique infini. Rev.
Roum. Math. Pures Appl. 11, 27 (1966)
References 391

14. Kecs, W.: Représentantion mathématique de certaines limites de groupes de charges et leurs
applications dans la mécanique des corps déformables. I. Bul. Inst. Polit. Iasßi, ser. nouă,
14(18), 443 (1968)
15. Kecs, W.: Les solutions généralisées de l’espace élastique sous l’action de certaines charges
variables. Rev. Roum. Math. Pures Appl. 15, 7 (1970)
16. Kecs, W., Teodorescu, P.P.: On the plane problem of an elastic body acted upon by dynamic
loads. I. General Results. II. Applications. Bull. Acad. Pol. Sci., sér. Sci. Techn. XIX, 47; 57
(1971)
17. Kecs, W., Teodorescu, P.P.: On the action of dynamic concentrated loads in case of linear
elastic bodies. Lucr. II Conf. Vibr. 49 (1978)
18. Kessel, S.: Stress functions and loading singularities for the infinitely extended linear elastic
isotropic cosserat continuum. IUTAM-Symposium 1967, Mechanics of Generalized
Continua, p. 114 (1968)
19. Mindlin, R.D.: On the equations of elastic materials with microstructure. Int. J. Solids Struct.
1, 73 (1965)
20. Mindlin, R.D., Tiersten, H.F.: Effects of couple stresses in linear elasticity. Arch. Rat. Mech.
Anal. 11, 415 (1962)
21. Stefaniak, J.: Concentrated loads as body forces. Rev. Roum. Math. Pures Appl. XIV, 119
(1969)
22. S
ßandru, N.: Asupra actßiunii unei fortße concentrate în spatßiul elastic nemărginit (On the action
of a concentrated load in the infinite elastic space). Com. Acad. Rom. XIII, 1019 (1963)
23. S
ßandru, N.: O deistviı̆ peremenykh sil v neogranichennom prostransve (On the action of
variable forces in the infinite space). Bull. Acad. Pol. Sci., sér. Sci. Technol. 12, 45 (1964)
24. S
ßandru, N.: On some problems of the linear theory of the asymmetric elasticity. Int. J. Engng.
Sci. 4, 81 (1966)
25. Teodorescu, P.P.: Sur l’action des charges concentrées dans le problème plan de la théorie de
l’élasticité. Bull. Math. Soc. Sci. Math. Roum. 8(56), 243 (1964)
26. Teodorescu, P.P.: Consideratßii asupra modului de a defini sarcinile concentrate în problema
plană a teoriei elasticitătßii (considerations on the definition of concentrated loads in the plane
problem of the theory of elasticity). An Univ. Timisßoara, ser. ßst. mat. fiz. III, 287 (1965)
27. Teodorescu, P.P.: Asupra actßiunii sarcinilor concentrate în teoria elasticitătßii (on the action of
concentrated loads in the theory of elasticity). An. Univ. Bucuresßti, ser. ßst. nat., mat. mec.
XV, 15 (1966)
28. Teodorescu, P.P.: Sur un certain caractère tensoriel des charges concentrées. Atti Accad. Naz.
Lincei, ser. VIII, Rend. Cl. Sci. fis., mat. e nat. XL, 251 (1966)
29. Teodorescu, P.P.: Sur l’action des charges concentrées dans le problème plan de la théorie de
l’élasticité. Arch. Mech. Stos. 18, 567 (1966)
30. Teodorescu, P.P.: Sur la notion de moment massique dans le cas des corps du type de
Cosserat. Bull. Acad. Pol. Sci., Sér. Sci. Techn. XV, 57 (1967)
31. Teodorescu, P.P.: On the action of concentrated loads in the case of a cosserat continuum.
IUTAM Symposium 1967, Mechanics of Generalized Continua, p. 120 (1968)
32. Teodorescu, P.P.: Sur les corps du type de Cosserat à éleasticité linéaire. Ist. Naz. Alta Mat.,
Symp. Math. I, 375 (1969)
33. Teodorescu, P.P.: Sur le calcul des récipients soumis à des charges concentrées dans le cas de
l’élasticité symétrique ou asymétrique. Quaderni de ‘‘la ricerca sci.’’, Napoli 24 (1973)
34. Teodorescu, P.P., S ßandru, N.: Sur l’action des charges concentrées en élasticité asymétrique
plane. Rev. Roum. Math. Pures Appl. XII, 1399 (1967)
35. Thomson, W. (Lord Kelvin): Cambridge and Dublin Math. J. (1848)
36. Toupin, R.A.: Elastic materials with couple stresses. Arch. Rat. Mech. Anal. 11, 385 (1962)
37. Vodička, V.: Ein durch allgemeine Massenkräfte beanspruchtes unendliches Medium.
ZAMM 39, 2 (1959)
Chapter 9
Elastic Space. Elastic Half-Space

In the present chapter and in the two following ones we deal with the study of
elastic bodies which occupy a domain in the form of a parallelepiped, as well as
with the study of the bodies which occupy infinite domains, obtained from the
above mentioned one, by throwing to infinite one or several of its faces. Thus, we
shall consider all spatial domains bounded by planes parallel to the co-ordinate
planes, hence by planes parallel or orthogonal to each other.
If all the faces, of the parallelepiped are thrown to infinite, then we have to do
with the elastic space. The elastic half-space is obtained if five or the faces of the
parallelepiped are thrown to infinite. If four of the faces of the parallelepiped are
thrown to infinite, then there are two possibilities: the two faces at a finite distance
are parallel to each other, obtaining thus an elastic layer, or are orthogonal to each
other, corresponding thus the elastic quarter-space. If three of the faces of the
parallelepiped are at a finite distance, then one obtains the elastic eighth-space,
when all these faces are orthogonal one to each other, or the elastic half-layer,
when two of these faces are parallel and orthogonal to a third one. If only two of
the faces of the parallelepiped are thrown to infinite, then there are still two
possibilities: if the faces at a finite distance are parallel two by two, then one gets
the elastic strip, while if only two faces are parallel, the other two ones being
orthogonal to the later faces and orthogonal to each other, then one obtains the
elastic quarter-layer. The elastic half-strip corresponds to the case when only one
of the faces of the parallelepiped is thrown to infinite.

9.1 Elastic Space

In the following we shall consider the elastic space acted by volume loads (forces
and moments), in particular concentrated loads; we deal with the static case.

P. P. Teodorescu, Treatise on Classical Elasticity, Mathematical and Analytical 393


Techniques with Applications to Engineering, DOI: 10.1007/978-94-007-2616-1_9,
Ó Springer Science+Business Media Dordrecht 2013
394 9 Elastic Space. Elastic Half-Space

9.1.1 Volume Loads

The elastic half-space acted by arbitrary volume forces may be studied by the
method of Fourier transforms after I. N. Sneddon and D. S. Berry [10] and after V.
Vodička [44]. Another method (in fact, another aspect of the method mentioned
above) is that of using, for the volume loads, Fourier representations: triple Fourier
series, if the volume loads are periodic, or triple Fourier integrals, if the volume
loads act locally. W. Kecs [14–16] dealt with concentrated loads, by means of the
theory of distributions.

9.1.1.1 Local Volume Forces. Solution in Displacements

We observe that any case of loading may be decomposed in eight cases of loading
after the properties of symmetry or antisymmetry with respect to the three co-
ordinate axes. We shall consider a case antisymmetric with respect to the Ox2 x3 -
plane and symmetric with respect to the Ox3 x1 -plane and to the Ox1 x2 -plane (case
useful for the study of the action of a concentrated force along the Ox1 -axis); any
other case of loading may be studied analogically.
The components of the volume load can be expressed, in this case, in the form
Z 1Z 1Z 1
F1 ¼ a1 cos a1 x1 cos a2 x2 cos a3 x3 da1 da2 da3 ;
Z0 1 Z0 1 Z0 1
F2 ¼ a2 sin a1 x1 sin a2 x2 cos a3 x3 da1 da2 da3 ; ð9:1Þ
Z0 1 Z0 1 Z0 1
F3 ¼ a3 sin a1 x1 sin a2 x2 sin a3 x3 da1 da2 da3 ;
0 0 0

where ai ¼ ai ða1 ; a2 ; a3 Þ; i ¼ 1; 2; 3, are given functions; the mentioned properties


of symmetry and antisymmetry are thus verified. Corresponding to these volume
loads, one may choose the components of the displacement vector in the form
Z 1Z 1Z 1
u1 ¼ A1 cos a1 x1 cos a2 x2 cos a3 x3 da1 da2 da3 ;
0 0 0
Z 1Z 1Z 1
u2 ¼ A2 sin a1 x1 sin a2 x2 cos a3 x3 da1 da2 da3 ; ð9:2Þ
Z0 1 Z0 1 Z0 1
u3 ¼ A3 sin a1 x1 sin a2 x2 sin a3 x3 da1 da2 da3 ;
0 0 0

where the functions Ai ða1 ; a2 ; a3 Þ; i ¼ 1; 2; 3, must be determined. We mention


that the integrals (9.1) can represent distributions too, while the integrals (9.2) are
considered in the sense used in the theory of the regularization of the Fourier
transforms in the theory of distributions [4, 5].
9.1 Elastic Space 395

Replacing in the Eq. (5.1000 ) of Lamé, we get the equations which must be
verified by the unknown functions Ai ; i ¼ 1; 2; 3,
ai  2a1 di1 1
ak a k A i þ ðak Ak  2a1 A1 Þ ¼ ai ; i ¼ 1; 2; 3; ð9:3Þ
1  2m l
by solving the system (9.3), one obtains
 
1 ðai  2a1 di1 Þðak ak  2a1 a1 Þ
Ai ¼ ai  ; i ¼ 1; 2; 3; ð9:4Þ
lak ak 2ð1  mÞak ak

the state of displacement being thus specified. The components (9.2) of the dis-
placement vector, being expressed by Fourier transforms, are regular functions at
infinite; hence, the particular integrals thus obtained are just the solutions of the
problem.

9.1.1.2 Local Volume Forces. Solution in Stresses

Considering the same volume forces (9.1) as in the precedent subsection, we may
choose the components of the stress tensor in the form.
Z 1Z 1Z 1
r11 ¼ B1 sin a1 x1 cos a2 x2 cos a3 x3 da1 da2 da3 ;
Z0 1 Z0 1 Z0 1
r22 ¼ B2 sin a1 x1 cos a2 x2 cos a3 x3 da1 da2 da3 ; ð9:5Þ
Z0 1 Z0 1 Z0 1
r33 ¼ B3 sin a1 x1 cos a2 x2 cos a3 x3 da1 da2 da3 ;
0 0 0
Z 1 Z 1 Z 1
r23 ¼ C1 sin a1 x1 sin a2 x2 sin a3 x3 da1 da2 da3 ;
Z 01 Z 01 Z 01
r31 ¼ C2 cos a1 x1 cos a2 x2 sin a3 x3 da1 da2 da3 ; ð9:50 Þ
0 0 0
Z 1Z 1Z 1
r12 ¼ C3 cos a1 x1 sin a2 x2 cos a3 x3 da1 da2 da3 ;
0 0 0

where the functions Bi ða1 ; a2 ; a3 Þ; Ci ða1 ; a2 ; a3 Þ; i ¼ 1; 2; 3, must be determined.


If the equations of equilibrium (5.1) and the Beltrami-Michell equations (5.42) are
verified, then we obtain the conditions.

a1 B1 þ a2 C3 þ a3 C2 ¼ a1 ;
a1 C 3 þ a2 B 2  a3 C 1 ¼ a2 ; ð9:6Þ
a1 C 2  a2 C 1 þ a3 B 3 ¼ a3 ;
396 9 Elastic Space. Elastic Half-Space

  a2
a21 þ a22 þ a23 B1 þ 1 ðB1 þ B2 þ B3 Þ
1þm
m
¼ ða1 a1 þ a2 a2 þ a3 a3 Þ þ 2a1 a1 ;
1m
 2  a2
a1 þ a22 þ a23 B2 þ 2 ðB1 þ B2 þ B3 Þ
1þm ð9:7Þ
m
¼ ða1 a1 þ a2 a2 þ a3 a3 Þ  2a2 a2 ;
1m
 2  a2
a1 þ a22 þ a23 B3 þ 3 ðB1 þ B2 þ B3 Þ
1þm
m
¼ ða1 a1 þ a2 a2 þ a3 a3 Þ  2a3 a3 ;
1m
 2  a2 a3
a1 þ a22 þ a23 C1  ðB1 þ B2 þ B3 Þ ¼ ða3 a2 þ a2 a3 Þ;
1þm
 2  a3 a1
a1 þ a22 þ a23 C2 þ ð B1 þ B2 þ B3 Þ ¼ a 1 a3  a 3 a 1 ; ð9:70 Þ
1þm
 2  a1 a2
a1 þ a22 þ a23 C3 þ ð B1 þ B2 þ B3 Þ ¼ a 1 a2  a 2 a 1 :
1þm
This system of 9 equations, which implies 6 unknowns is compatible, because the
Beltrami-Michell equations (5.42) are just conditions of compatibility in stresses;
by solving it, we get
 2 
1 a1 þ mða21 þ a22 þ a23 Þ ða1 a1 þ a2 a2 þ a3 a3 Þ
B1 ¼ 2  2a1 a1 ;
a1 þ a22 þ a23 ð1  mÞða21 þ a22 þ a23 Þ
  
1 a22 þ mða21 þ a22 þ a23 Þ ða1 a1 þ a2 a2 þ a3 a3 Þ
B2 ¼ 2 þ 2a 2 a 2 ;
a1 þ a22 þ a23 ð1  mÞða21 þ a22 þ a23 Þ
 2 
1 a3 þ mða21 þ a22 þ a23 Þ ða1 a1 þ a2 a2 þ a3 a3 Þ
B3 ¼ 2 þ 2a 3 3 ;
a
a1 þ a22 þ a23 ð1  mÞða21 þ a22 þ a23 Þ
ð9:8Þ
 
1 a2 a3 ða1 a1 þ a2 a2 þ a3 a3 Þ
C1 ¼  ða3 a2 þ a2 a3 Þ ;
a21 2
þ a2 þ a3 2 ð1  mÞða21 þ a22 þ a23 Þ
 
1 a3 a1 ða1 a1 þ a2 a2 þ a3 a3 Þ
C2 ¼ 2  þ a1 a3  a3 a1 ; ð9:80 Þ
a1 þ a22 þ a23 ð1  mÞða21 þ a22 þ a23 Þ
 
1 a1 a2 ða1 a1 þ a2 a2 þ a3 a3 Þ
C3 ¼ 2  þ a 1 a 2  a 2 a 1 ;
a1 þ a22 þ a23 ð1  mÞða21 þ a22 þ a23 Þ

the state of stress being thus specified.


It is obvious that, using Cauchy’s relations (5.2) and Hooke’s law (5.3), one
obtains again these results, starting from the formulae in the previous subsection.
9.1 Elastic Space 397

9.1.2 Concentrated Loads

In what follows, we introduce first a concentrated force, dealing then with other
concentrated loads; a new concentrated load (the centre of torsion) will be intro-
duced too.

9.1.2.1 Concentrated Forces

To study the case of a concentrated force, we start from the results obtained for the
volume loads. We take thus F2 ¼ F3 ¼ 0 and F1 ¼ p, the load p being uniformly
distributed in the interior of a parallelepiped of dimensions 2d1 , 2d2 and 2d3 ; the
volume load will be thus represented in the form (9.1), the function a1 ða1 ; a2 ; a3 Þ
being given by the formula (6.800 ) in the form
Z Z Z
8 d1 d2 d3
a1 ¼ 3 p cos a1 n1 cos a2 n2 cos a3 n3 dn1 dn2 dn3
p 0 0 0
8p sin a1 d1 sin a2 d2 sin a3 d3
¼ 3 : ð9:9Þ
p a1 a2 a3
Denoting F1 ¼ F ¼ 8pd1 d2 d3 and passing to the limit for p ! 1 and
d1 d2 d3 ! 0, we may represent the concentrated force, directed in the positive
sense of the Ox1 -axis, by the Fourier series
Z Z Z
F d1 d2 d3
cos a1 x1 cos a2 x2 cos a3 x3 da1 da2 da3 : ð9:90 Þ
p3 0 0 0

The formulae (9.2) and (9.4) allow to express the state of displacement in the
form
Z 1Z 1Z 1  
F ð1  2mÞa21 þ 2ð1  mÞ a22 þ a23
u1 ¼ 3  2 2
2p ð1  mÞl 0 0 0 a þ a2 þ a2 1 2 3
 cos a1 x1 cos a2 x2 cos a3 x3 da1 da2 da3 ;
Z 1Z 1Z 1
F a1 a2
u2 ¼ 3  2 2
2p ð1  mÞl 0 0 0 a þ a2 þ a2 ð9:10Þ
1 2 3
 sin a1 x1 sin a2 x2 cos a3 x3 da1 da2 da3 ;
Z 1Z 1Z 1
F a1 a3
u3 ¼  3  2 2
2p ð1  mÞl 0 0 0 a þ a2 þ a2
1 2 3
 sin a1 x1 cos a2 x2 sin a3 x3 da1 da2 da3 ;

while the formulae (9.5), (9.50 ), (9.8), (9.80 ) lead to the state of stress
398 9 Elastic Space. Elastic Half-Space

Z 1 Z 1 Z 1
  
F a1 ð1  mÞa21 þ ð2  mÞ a22 þ a23
r11 ¼ 3  2 2
p ð1  mÞ 0 0 0 a1 þ a22 þ a23
 sin a1 x1 cos a2 x2 cos a3 x3 da1 da2 da3 ;
Z 1Z 1Z 1   
F a1 ð1  mÞa22  m a23 þ a21
r22 ¼ 3  2 2
p ð1  mÞ 0 0 0 a þ a2 þ a2 ð9:11Þ
1 2 3
 sin a1 x1 cos a2 x2 cos a3 x3 da1 da2 da3 ;
Z 1Z 1Z 1   
F a1 ð1  mÞa23  m a21 þ a22
r33 ¼ 3  2 2
p ð1  mÞ 0 0 0 a þ a2 þ a2 1 2 3
 sin a1 x1 cos a2 x2 cos a3 x3 da1 da2 da3 ;
Z 1Z 1Z 1
F a1 a2 a3
r23 ¼  3 2 2
p ð1  mÞ 0 0
2 2
0 ða1 þ a2 þ a3 Þ

 sin a1 x1 sin a2 x2 sin a3 x3 da1 da2 da3 ;


Z 1Z 1Z 1  2 
F a3 ma1  ð1  mÞða22 þ a23 Þ
r31 ¼ 3
p ð1  mÞ 0 0 0 ða21 þ a22 þ a23 Þ2 ð9:110 Þ
 cos a1 x1 cos a2 x2 sin a3 x3 da1 da2 da3 ;
Z 1Z 1Z 1  2 
F a2 ma1  ð1  mÞða22 þ a23 Þ
r12 ¼ 3
p ð1  mÞ 0 0 0 ða21 þ a22 þ a23 Þ2
 cos a1 x1 sin a2 x2 cos a3 x3 da1 da2 da3 :
By computing these integrals, one obtains the state of displacement (8.49) and
the state of stress (8.51), (8.510 ).

9.1.2.2 Other Concentrated Loads

Starting from the results obtained in the above subsection, one may obtain the
states of displacement for other concentrated loads applied at the origin O.
Thus, for a directed concentrated moment along the Ox1 -axis, the state of
displacement is given by the formulae (8.55), while for a centre of rotation applied
in the Ox1 x2 -plane, the state of displacement is given by (8.58), the displacement
u3 vanishing.
A concentrated moment of linear dipole type along the Ox1 -axis leads to the
state of displacement (8.59). Starting from this load, one obtains the state of
displacement (8.60) for a plane centre of dilatation in the Ox1 x2 -plane and the
state of displacement (8.61) for a spatial centre of dilatation.
Analogically, starting from two centres of rotation of inverse sense, one may
obtain, on the same way, by a process of passing to the limit, a linear dipole of
9.1 Elastic Space 399

torsion of magnitude T1 (given by the product T1 = Mc for M ! 1 and c ! 0),


along the Ox1 -axis; the corresponding state of displacement is given by
3T1 x1 x3 3T1 x1 x2
u1 ¼ 0; u2 ¼  ; u3 ¼ : ð9:12Þ
8pl r 5 8pl r 5
Superposing the states of displacement corresponding to the linear dipoles of
torsion T1 and T2 , orthogonal to each other, one obtains the state of displacement
3Tp x2 x3 3Tp x1 x3
u1 ¼ ; u2 ¼ ; u3 ¼ 0; ð9:13Þ
16pl r 5 16pl r 5
for a plane centre of torsion in the Ox1 x2 -plane, the magnitude of which is given
by T1 ¼ T2 ¼ Tp =2.
Adding the action of a linear dipole of torsion T3 acting along the Ox3 -axis, one
obtains a spatial centre of torsion of magnitude given by T1 ¼ T2 ¼ T3 ¼ Ts =3,
which has not any effect; indeed, the displacement vector vanishes (u ¼ 0).
Starting from the centres of dilatation (contraction), one can obtain analogous
results.

9.2 Elastic Half-Space

The elastic half-space x3  0 corresponds to an infinite domain bounded at a finite


distance by the separation plane x3 ¼ 0. The study of this infinite domain is
important, because it can approximate the ground for geotechnical studies or for
foundation design; as well, such a study may be interesting in many other cases,
e. g., the elastic half-space acted upon by a concentrated load may give indications
for local concentration of stresses in case of a finite body.
This problem has been firstly studied by J. Boussinesq [1]; a demonstration of
these results, by means of the Papkovich-Neuber representation, has been given by
A. I. Lur’e [20]. However, a rich bibliography in this direction can be found in the
monograph of the same author [9]. The cases of a distributed load, normal or
tangential, as well as the case of a concentrated tangential force, have been con-
sidered by V. Cerruti [12]; the case of a normal load can be found in the famous
study of H. Hertz [13] on contact problems too. The case of a normal load dis-
tributed in the interior of a circle has been considered by J. Boussinesq [1],
H. Lamb [18], K. Terezawa [42], F. Schleicher [22, 23] and A.-E.-H. Love [19];
however, the latter one dealt also with a uniform distributed load in the interior of
a rectangle. The action of internal concentrated loads has been considered by
R. D. Mindlin [21] in the case of a free separation plane and by N. S ßandru [24]
(later, independently, by Lai Phani The and J. Mandel [17]), in the case of a built-
in separation plane. In 1959, we [31] studied a case of periodic normal loading,
while B. Tanimoto [27] considered the loading with local normal loads; in 1958,
we [31, 32] found again, independently, these results, using the stress functions
400 9 Elastic Space. Elastic Half-Space

introduced in Sect. 5.3.2.5. In [34] and [36] we considered a particular case of


loading, which leads to results in a finite form. The methods of the theory of
distributions have been introduced by W. Kecs [14].

9.2.1 Action of a Periodic Load

Let be the elastic half-space x3  0 acted upon by a normal periodic (in two
directions) load pðx1 ; x2 Þ, symmetric with respect to the co-ordinate axes Ox1 and
Ox2 ; to fix the ideas, we represent this load by means of a double Fourier series,
even with respect to x1 and x2 , of the form
XX
p3 ðx1 ; x2 Þ ¼ b0 þ bnm cos an x1 cos bm x2 ; ð9:14Þ
n m

with the notations


np mp
an ¼ ;b ¼ ; n; m ¼ 1; 2; 3; . . .; ð9:140 Þ
a1 m a2
and the periods L1 ¼ 2a1 ; L2 ¼ 2a2 .

9.2.1.1 Boundary Conditions. Stress Functions

The boundary conditions are of the form


x3 ¼ 0; r33 ¼ p3 ðx1 ; x2 Þ; r31 ¼ r32 ¼ 0 ð9:15Þ
for the plane at a finite distance, the stresses vanishing at infinite.
Using the representation given in Sect. 5.2.2.5, we may introduce the bihar-
monic stress functions.
XX
F11 ¼ K1 x21 þ ðA0nm þ B0nm cnm x3 Þecnm x3 cos an x1 cos bm x2 ;
n m
XX
F22 ¼ K2 x22 þ ðA00nm þ B00nm cnm x3 Þecnm x3 cos an x1 cos bm x2 ; ð9:16Þ
n m
XX
F33 ¼ K3 x23 þ ðA000 000
nm þ Bnm cnm x3 Þe
cnm x3
cos an x1 cos bm x2 ;
n m

where we denoted
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
cnm ¼ a2n þ b2m ð9:17Þ

and where the functions in two variables have been taken equal to zero.
9.2 Elastic Half-Space 401

One obtains thus.


XX
r33 ¼ 2ð1 þ mÞK3  c2nm Anm  2½Bnm  ð2 þ mÞB000
nm 
n m
cnm x3
þ Bnm cnm x3 e cos an x1 cos bm x2 ; ð9:18Þ
XXc
r31 ¼  nm
a2n Anm  ½ð2 þ mÞa2n þ c2nm B0nm þ ðmb2m  a2n ÞB00nm
n m
an
c x
þða2n þ mc2nm ÞB000 2
nm þ an Bnm cnm x3 e
nm 3
sin an x1 cos bm x2 ;
XXc
r23 ¼  nm
b2m Anm þ ðma2n  b2m ÞB0nm  ½ð2 þ mÞb2m þ c2nm B00nm
n m
bm
c x
þðb2m þ mc2nm ÞB000 2
nm þ an Bnm cnm x3 e
nm 3
sin an x1 cos bm x2 ;
ð9:180 Þ
where one has introduced the notations.

Anm ¼ A0nm þ A00nm þ A000 0 00 000


nm ; Bnm ¼ Bnm þ Bnm þ Bnm : ð9:19Þ
Putting the boundary conditions (9.14), (9.15), one obtains the algebraic
equations
bnm
Anm  2ðB0nm þ B00nm Þ þ 2ð1 þ mÞB000
nm ¼  ;
c2nm
a2n Anm  ½ð1 þ mÞa2n þ 2c2nm B0nm  ða2n  mb2m ÞB00nm
ð9:20Þ
þ ða2n þ mc2nm ÞB000
nm ¼ 0;
b2m Anm þ ðma2n  b2m ÞB0nm  ½ð1 þ mÞb2m þ 2c2nm B00nm
þ ðb2m þ mc2mn ÞB000
nm ¼ 0;

which give the sequences of coefficients to be determined. We notice that, using


three biharmonic functions, we get less equations then we need to determine these
constants; but we know that the state of strain and stress is unique.
We choose these constants so as to have a solution as simple as possible; we
take
bnm
A0nm ¼ A00nm ¼ 0; Anm ¼ A000
nm ¼ ð1 þ 2mÞ 2 ;
cnm
2 2 ð9:21Þ
0 00 m a n  bm 000 bnm
Bnm ¼ Bnm ¼ bnm ; Bnm ¼ Bnm ¼  2 :
2 þ m c4nm cnm

As well, it results
1
K3 ¼ b0 : ð9:22Þ
2ð1 þ mÞ

Due to the terms which are not under the sign sum in the expressions of the stress
functions, one obtains
402 9 Elastic Space. Elastic Half-Space

1 1
u1 ¼ ½K1  mðK2 þ K3 Þ; u2 ¼ ½K2  mðK1 þ K3 Þ; ð9:200 Þ
l l
putting the conditions that the displacements u1 and u2 do not increase indefinitely
together with the variables x1 and x2 (the lateral deformation is stopped by the
medium) and taking into account (9.22), one gets
m
K1 ¼ K2 ¼ b0 : ð9:220 Þ
2ð1  m2 Þ
Thus, the stress functions read
m m
F1  b0 x21 ¼ F2 þ b0 x22
2ð1  m2 Þ 2ð1  m2 Þ
m X X a2  b 2
¼ x3 n m
bnm ecnm x3 cos an x1 cos bm x2 ; ð9:23Þ
2þm n m
c3nm
XX 1
F3 ¼ 2
ð1 þ 2m  cnm x3 Þbnm ecnm x3 cos an x1 cos bm x2 :
n m
c nm

9.2.1.2 State of Strain and Stress

The state of stress is obtained in the form


" 
m XX an 2
r11 ¼ b0 þ bnm U4 ðcnm x3 Þ
1m n m
cnm
2 #
bm
þm U6 ðcnm x3 Þ cos an x1 cos bm x2 ;
cnm
" 
m XX bm 2
r22 ¼ b0 þ bnm U4 ðcnm x3 Þ ð9:24Þ
1m n m
cnm
2 #
an
þm U6 ðcnm x3 Þ cos an x1 cos bm x2 ;
cnm
XX
r33 ¼ b0 þ bnm U2 ðcnm x3 Þ cos an x1 cos bm x2 ;
n m
9.2 Elastic Half-Space 403

XX bm
r23 ¼ bnm U3 ðcnm x3 Þ cos an x1 sin bm x2 ;
n m
c nm
XX an
r31 ¼ bnm U3 ðcnm x3 Þ sin an x1 cos bm x2 ; ð9:240 Þ
n m
c nm
XX an b
r12 ¼ bnm 2 m U46 ðcnm x3 Þ sin an x1 sin bm x2
n m
cnm

and the state of displacement is given by


1 XX an
u1 ¼ bnm 2 U46 ðcnm x3 Þ sin an x1 cos bm x2 ;
2l n m cnm
1 X X b
u2 ¼ bnm 2m U46 ðcnm x3 Þ cos an x1 sin bm x2 ;
2l n m cnm
 ð9:25Þ
0 1 1  2m
u3 ¼ u3 þ b 0 x3
2l 1  m
X X bnm
 U16 ðcnm x3 Þ cos an x1 cos bm x2 ;
n m
cnm

where we have introduced the notations in Sect. A.4.1.1. Because of the symmetry
of the displacements with respect to the Ox1 and Ox2 axes, we have taken

u01 ¼ u02 ¼ 0; x01 ¼ x02 ¼ x03 ¼ 0; ð9:26Þ

the displacement of rigid body u03 being impossible to determinate; hence, the
displacement u3 is obtained only as a displacement relative to another point,
considered as fixed.
We notice that the mean loading b0 leads to a displacement which tends to
infinite for x3 ! 1; this result is due to the fact that we dealt with an infinite solid
(which can be considered as a prism of infinite length acted upon by a simple axial
load). But we can admit that the result thus obtained is valid for points at a finite
distance.
The convergence of the above series depends on the factor bnm . But, even in the
case of concentrated loads, the sum of the series may be approximated with a small
number of terms, because of the exponential factor. The series corresponding to
the displacements are convergent after the factor bnm =cnm ; hence, the sum of these
series may be approximated with a smaller number of terms that in the case of the
stresses.
Taking into account (9.14), we can write the stresses corresponding to the
separation plane in the form.
404 9 Elastic Space. Elastic Half-Space


b0
r11 ðx1 ; x2 ; 0Þ ¼ p3 ðx1 ; x2 Þ  ð1  2mÞ
1m
X X b 2 
m
þ bnm cos an x1 cos bm x2 ;
n m
cnm
 ð9:27Þ
b0
r22 ðx1 ; x2 ; 0Þ ¼ p3 ðx1 ; x2 Þ  ð1  2mÞ
1m
X X an  2 
þ bnm cos an x1 cos bm x2 ;
n m
cnm
X X an b
m
r12 ðx1 ; x2 ; 0Þ ¼ ð1  2mÞ bnm sin an x1 sin bm x2 ; ð9:270 Þ
n m
c2nm

where we took into consideration the difference between the external loading and
the mean one. As well, the deformation of the separation plane is given by
1  m X X bnm
u3 ðx1 ; x2 ; 0Þ ¼ u03  cos an x1 cos bm x2 ð9:28Þ
l n m
cnm

and the displacements in this plane read


1  2m X X an
u1 ðx1 ; x2 ; 0Þ ¼ bnm sin an x1 cos bm x2 ;
2l n m
c2nm
ð9:280 Þ
1  2m X X bm
u2 ðx1 ; x2 ; 0Þ ¼ bnm cos an x1 sin bm x2 :
2l n m
c2nm

These results may be used to calculate the pressure (the normal stress r33 ) and
the soil settlement (the displacement u3 ) in case if isolated formulations, having an
approximately uniform distribution, or in case of a net of foundation beams, for
which the ratio between the spans in a direction is approximately equal to unity. If
the loading is not rigorously periodic, then the approximation must be evaluated in
each case.
Analogically, one may consider also other cases of loading with a normal
loading having other properties of symmetry with respect to the co-ordinate axes.
Cases of periodic tangential loads can be studied in the same way.

9.2.2 Action of a Local Load

We consider now the elastic half-space x3  0 acted upon by local normal or


tangential loads; we use the results which will be given in Sect. 10.1.2 for the
elastic eighth-space.
9.2 Elastic Half-Space 405

9.2.2.1 Normal Load

Let be the case of a normal load


Z 1Z 1
p3 ðx1 ; x2 Þ ¼ a3 ða1 ; a2 Þ cos a1 x1 cos a2 x2 da1 da2 ; ð9:29Þ
0 0

where
Z 1 Z 1
4
a3 ða1 ; a2 Þ ¼ 2 p3 ðn1 ; n2 Þ cos a1 n1 cos a2 n2 dn1 dn2 ; ð9:290 Þ
p 0 0

acting on the plane x3 ¼ 0. One uses double Fourier integrals, even with respect to
the variables x1 ; x2 , which may represent, as well, distributions. The function
a3 ða1 ; a2 Þ, which characterizes the normal load, is even with respect to a1 and a2 .
The considered load is thus symmetric with respect to the Ox1 and Ox2 axes.
Taking
1
A1 ða2 ; a3 Þ ¼ A2 ða3 ; a1 Þ ¼ 0; A3 ða1 ; a2 Þ ¼ a3 ða2 ; a3 Þ; ð9:30Þ
c3
we observe that the system of integral Eq. (10.21) is verified; the first two
equations do not intervene (indeed, they may be considered as verified, because
a1 ða1 ; a2 Þ, a2 ða1 ; a2 Þ are arbitrary functions).
One obtains the state of stress.
Z 1Z 1 " 
a1 2
r11 ¼ a3 ða1 ; a2 Þ U4 ðc3 x3 Þ
0 0 c3
2 #
a2
þm U6 ðc3 x3 Þ cos a1 x1 cos a2 x2 da1 da2 ;
c3
Z 1Z 1 " 
a2 2 ð9:31Þ
r22 ¼ a3 ða1 ; a2 Þ U4 ðc3 x3 Þ
0 0 c3
2 #
a1
þm U6 ðc3 x3 Þ cos a1 x1 cos a2 x2 da1 da2 ;
c3
Z 1Z 1
r33 ¼ a3 ða1 ; a2 ÞU2 ðc3 x3 Þ cos a1 x1 cos a2 x2 da1 da2 ;
0 0
Z 1 Z 1
a2
r23 ¼ a3 ða1 ; a2 Þ
U3 ða3 x3 Þ cos a1 x1 sin a2 x2 da1 da2 ;
c3
Z0 1 Z0 1
a1
r31 ¼ a3 ða1 ; a2 Þ U3 ða3 x3 Þ sin a1 x1 cos a2 x2 da1 da2 ; ð9:310 Þ
0 c3
Z 10 Z 1
a1 a2
r12 ¼ a3 ða1 ; a2 Þ 2 U46 ða3 x3 Þ sin a1 x1 sin a2 x2 da1 da2
0 0 c3
406 9 Elastic Space. Elastic Half-Space

and the state of displacement


Z Z
1 1 1 a1
u1 ¼ a3 ða1 ; a2 Þ 2 U46 ða3 x3 Þ sin a1 x1 cos a2 x2 da1 da2 ;
2l 0 0 c3
Z Z
1 1 1 a2
u2 ¼ a3 ða1 ; a2 Þ 2 U46 ða3 x3 Þ cos a1 x1 sin a2 x2 da1 da2 ; ð9:32Þ
2l 0 0 c3
Z 1Z 1
1 a 3 ða 1 ; a2 Þ
u3 ¼ u03 þ U16 ða3 x3 Þ cos a1 x1 cos a2 x2 da1 da2 ;
2l 0 0 c3

where the displacement of rigid body u03 is non-zero, because of the symmetry
assumed for the load.
Taking into account the load (9.29), one gets the normal stresses

r11 ðx1 ; x2 ; 0Þ ¼ p3 ðx1 ; x2 Þ


Z 1 Z 1 2
a2
 ð1  2mÞ a3 ða1 ; a2 Þ cos a1 x1 cos a2 x2 da1 da2 ;
0 0 c3
ð9:33Þ
r22 ðx1 ; x2 ; 0Þ ¼ p3 ðx1 ; x2 Þ
Z 1 Z 1 2
a1
 ð1  2mÞ a3 ða1 ; a2 Þ cos a1 x1 cos a2 x2 da1 da2
0 0 c3

and the tangential stress


Z 1 Z 1
a1 a2
r12 ðx1 ; x2 ; 0Þ ¼ ð1  2mÞ a3 ða1 ; a2 Þ sin a1 x1 sin a2 x2 da1 da2
0 0 c23
ð9:330 Þ
for the separation plane, the external load being put into evidence.
What concerns the tangential displacements, one obtains
Z Z
1  2m 1 1 a1
u1 ðx1 ; x2 ; 0Þ ¼ a3 ða1 ; a2 Þ sin a1 x1 cos a2 x2 da1 da2 ;
2l 0 0 c2
Z 1Z 1 3 ð9:34Þ
1  2m a2
u2 ðx1 ; x2 ; 0Þ ¼ a3 ða1 ; a2 Þ cos a1 x1 sin a2 x2 da1 da2 ;
2l 0 0 c23

the deformations of the separation plane (normal displacement) being specified by


Z Z
0 1m 1 1 1
u3 ðx1 ; x2 ; 0Þ ¼ u3  a3 ða1 ; a2 Þ cos a1 x1 cos a2 x2 da1 da2 : ð9:340 Þ
l 0 0 c3

The state of stress and displacement is thus completely stated.


Any case of normal load with other properties of symmetry may be analogically
studied.
9.2 Elastic Half-Space 407

9.2.2.2 Tangential Load

If the elastic half-space x3  0 is acted upon by local tangential loads, antisym-


metric with respect to the Ox1 and Ox2 axes, of the form
Z 1Z 1
t31 ðx1 ; x2 Þ ¼ a03 ða1 ; a2 Þ cos a1 x1 sin a2 x2 da1 da2 ;
0 0
Z 1Z 1 ð9:35Þ
t32 ðx1 ; x2 Þ ¼ a003 ða1 ; a2 Þ sin a1 x1 cos a2 x2 da1 da2 ;
0 0

then one may use the results in Sect. 10.1.3, taking

A01 ða2 ; a3 Þ ¼ A001 ða2 ; a3 Þ ¼ 0; A02 ða3 ; a1 Þ ¼ A002 ða3 ; a1 Þ ¼ 0; ð9:36Þ


 
0 1 0 1 00
A3 ða1 ; a2 Þ ¼ c3 a ða1 ; a2 Þ þ a3 ða1 ; a2 Þ ;
a2 3 a1

1 1  
A003 ða1 ; a2 Þ ¼ ð1 þ mÞa22  ma21 a03 ða1 ; a2 Þ ð9:360 Þ
mc3 a2

1 2

2 00
 ð1 þ mÞa1  ma2 a3 ða1 ; a2 Þ :
a1

One observes that the system of integral equations (10.55), (10.550 ) is verified;
indeed, the first two equations of both systems do not intervene, because
a01 ða2 ; a3 Þ, a001 ða2 ; a3 Þ; a02 ða3 ; a1 Þ; a002 ða3 ; a1 Þ are arbitrary functions.
The formulae (10.62)–(10.630 ) allow thus to express the state of stress and
displacement. For instance, the stress normal to the separation plane x3 ¼ 0 is
given by
Z 1Z 1
 0
r33 ¼ x3 a1 a3 ða1 ; a2 Þ
0 0

þa2 a003 ða1 ; a2 Þ ec3 x3 sin a1 x1 sin a2 x2 da1 da2 ; ð9:37Þ

while the displacement normal to this plane reads


Z Z 
1 1 1 a1 a2 1 0
u3 ¼ u03 þ x2 x01  x1 x02 þ a ða1 ; a2 Þ
2l 0 0 c23 a2 3

1
þ a003 ða1 ; a2 Þð1  2m þ c3 x3 Þ ec3 x3 sin a1 x1 sin a2 x2 da1 da2 ; ð9:38Þ
a1
One obtains similar results for other cases of tangential loading.
408 9 Elastic Space. Elastic Half-Space

9.2.3 Applications

We shall consider, in the following, some particular cases with periodic or local
normal loads, which may be studied starting from the previous general results.

9.2.3.1 Uniformly Distributed Isolated Foundations

Let us search the state of stress in the earth, considered to be an elastic half-space
acted upon by a normal periodic load resulting from uniformly distributed isolate
foundations.
We choose square isolate foundations (of side a), so that between their axes be
a distance equal to 2a in both directions; the ratio between the area covered by the
foundation and the total area is 1=4, hence we may consider that the foundations
are isolated. The foundations act with a uniform distributed load of compression p.
Taking the periods L1 ¼ L2 ¼ 2a, we represent the load in the form
p pX 1
p3 ðx1 ; x2 Þ ¼   ð1Þðn1Þ=2 cos an x1
4 p n n
pX 1
 ð1Þðm1Þ=2 cos bm x2
p m m
4p X X 1
þ 2 ð1ÞðnþmÞ=2 cos an x1 cos bm x2 ; n; m ¼ 1; 3; 5; . . .
p n m nm
ð9:39Þ

The normal stress r33 at the depth a under the centre of a foundation, hence for
x1 ¼ x2 ¼ 0; x3 ¼ a is given by (one observes that the two simple series have the
same sum)
p X 1 þ np np
r33 ð0; 0; aÞ ¼   2p ð1Þðn1Þ=2 e
4 n
n
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
4p X X 1 þ p n2 þ m2 ppffiffiffiffiffiffiffiffiffiffi
þ 2 ð1ÞðnþmÞ=2 e n2 þm2
; n; m ¼ 1; 3; 5; . . .
p n m nm
ð9:40Þ
Taking one term in the expansions into series, it results
r33 ð0; 0; aÞ ffi 0:250p  0:114p  0:026p ¼ 0:390p: ð9:400 Þ

We obtain the same result if we apply, e.g., the method of the corner points, on a
longer and more complicate way, using the tables of pressure in the earth at the
corners of the rectangular foundations.
If one calculates the pressure under the corners of the rectangle, one obtains
9.2 Elastic Half-Space 409

 a a 
r33  ;  ; x3 ffi 0:250p ð9:41Þ
2 2
at any depth.
We notice the rapidity of convergence of the series which give the soil set-
tlement, because of a supplementary increasing factor at the denominator.

9.2.3.2 Boussinesq’s Problem

Let us consider the case studied for the first time by J. Boussinesq [1]: the elastic
half-space acted upon by a normal concentrated force.
Let firstly be the case of a normal load, uniformly distributed on a rectangle

p for jx1 j\a1 ; jx2 j\a2 ;
p3 ðx1 ; x2 Þ ¼ ð9:42Þ
0 for jx1 j [ a1 ; jx2 j [ a2 ;

it results
4p
a3 ða1 ; a2 Þ ¼  sin a1 a1 sin a2 a2 : ð9:420 Þ
p2 a1 a2
Making p ! 1 and a1 ; a2 ! 0 with 4pa1 a2 ¼ P, one obtains the results corre-
sponding to a concentrated force; thus, the Fourier coefficient reads
P
a3 ða1 ; a2 Þ ¼  : ð9:4200 Þ
p2
The normal stress r33 becomes
Z Z
P 1 1
r33 ¼  2 ð1 þ c3 x3 Þec3 x3 cos a1 x1 cos a2 x2 da1 da2 ; ð9:43Þ
p 0 0

while the deformation of the separation plane reads


Z Z
ð1  mÞP 1 1 1
u3 ðx1 ; x2 ; 0Þ ¼ u03 þ cos a1 x1 cos a2 x2 da1 da2 : ð9:44Þ
p2 l 0 0 c3

Calculating the above integrals (see Sect. A.4.2.3), one obtains

3P x33
r33 ¼  ð9:430 Þ
2p r 5
and
1mP
u3 ðx1 ; x2 ; 0Þ ¼ u03 þ ; ð9:440 Þ
2pl r3
where
410 9 Elastic Space. Elastic Half-Space

qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r3 ¼ x21 þ x22 : ð9:45Þ

9.2.3.3 Another Case of Loading

Let be now the elastic half-space x3  0 acted upon by a normal load, symmetric
with respect to both co-ordinate axes Ox1 and Ox2 , of the form
p
p3 ðx1 ; x2 Þ ¼  : ð9:46Þ
r33

This case of loading may approximate, in a certain way, a concentration of normal


loads at a point of the separation plane. It corresponds
2p
a3 ða1 ; a2 Þ ¼ c: ð9:460 Þ
p 3
The normal stresses are given by
 x 2 
p 2 x21 x23
r11 ¼  3 2ð1  mÞ  3ð1  2mÞ 15 4 ;
r r r
   
p x1 2 x212 x23
r22 ¼  3 2ð1  mÞ  3ð1  2mÞ 15 4 ; ð9:47Þ
r r r
 x 2 x 4 
p 3 3
r33 ¼  3 1 þ 6 15 ;
r r r

while the tangential stresses read


 x 2 
x2 x3 3
r23 ¼ 3p 1  5 ;
r5 r
 x 2 
x3 x1 3
r31 ¼ 3p 5 1  5 ; ð9:470 Þ
r r
 x 2 
x1 x2 3
r12 ¼ 3p 5 1  2m  5 :
r r
The displacements have the form
 x 2 
p x1 3
u1 ¼  1  2m  3 ;
2l r 3 r
 x 2 
p x2 3
u2 ¼  1  2m  3 ;
2l r 3 r ð9:48Þ
 x 2 
0 p x3 3
u3 ¼ u3  1  2m þ 3 :
2pl r 3 r
:
9.2 Elastic Half-Space 411

For the separation plane one obtains the stresses


" 2 #
p x2
r11 ðx1 ; x2 ; 0Þ ¼  3 2ð1  mÞ  3ð1  2mÞ ;
r3 r3
" 2 # ð9:49Þ
p x1
r22 ðx1 ; x2 ; 0Þ ¼  3 2ð1  mÞ  3ð1  2mÞ ;
r3 r3

x1 x2
r12 ðx1 ; x2 ; 0Þ ¼ 3ð1  2mÞp ð9:490 Þ
r35

and the displacements


1 1 1  2m p
u1 ðx1 ; x2 ; 0Þ ¼ u2 ðx1 ; x2 ; 0Þ ¼ ; u3 ¼ u03 : ð9:50Þ
x1 x2 2l r33
In cylindrical co-ordinates (the problem is axi-symmetric), the state of stress is
given by (see Sect. A.2.1.2)
 
p r 2 z2
rrr ¼  3 2ð1  mÞ  15 4 ;
R R
  z 2 
p
rhh ¼ 3 1  4m  3ð1  2mÞ ; ð9:51Þ
R R
  z 2  z 4 
p
rzz ¼  3 1 þ 6 15 ;
R R R
  z 2 
zr
rhz ¼ rrh ¼ 0; rzr ¼ 3p 5 1  5 ; ð9:510 Þ
R R

while the state of displacement reads


  z 2 
p r
ur ¼ 1  2m  3 ; uh ¼ 0;
2l R3 R
  z 2  ð9:52Þ
0 p z
uz ¼ uz  1  2m þ 3 :
2l R3 R
In spherical co-ordinates, the stresses are given by
p  
rRR ¼ 2 3
1  m  ð5  mÞ cos2 u ;
R
p  
ruu ¼  3 1 þ ð1  2mÞ cos2 u ; ð9:53Þ
R
p  
rhh ¼ 3 1  4m  3ð1  2mÞ cos2 u ;
R
p
ruh ¼ rhR ¼ 0; rRu ¼ ð1  mÞ sin 2u; ð9:530 Þ
R3
412 9 Elastic Space. Elastic Half-Space

while the displacements are of the form


1 p  
uR ¼ u0z cos u þ 1  2m  ð5  4mÞ cos2 u ;
2l R2
ð9:54Þ
1  2m p
uu ¼ u0z sin u þ sin 2u; uh ¼ 0:
2l R2
The formulae (9.510 ) and (9.530 ) show that h is a principal direction, the cor-
responding principal normal stress being rhh (the same in both systems of co-
ordinates). The other two principal directions are contained in the plane ðR; uÞ; for
instance, the angles made by these directions with the direction R are given by
4ð1  mÞ sin 2u
tan 2w ¼ : ð9:55Þ
9 þ ð11  4mÞ cos 2u

9.2.4 Methods of the Theory of Distributions

Let be the elastic half-space x3  0 acted upon by the external loads pj ðx1 ; x2 Þ,
j ¼ 1; 2; 3, on the separation plane x3 ¼ 0. We search the fundamental solution, in
the frame of the theory of distributions, for displacements, as well as for stresses.

9.2.4.1 Fundamental Solution for Displacements

We start from Lamé’s equations (5.10) in the absence of volume forces, i.e.
lDuj þ ðk þ lÞh;j ¼ 0; j ¼ 1; 2; 3; ð9:56Þ

and we put the boundary conditions on the separation plane


lim r3j ¼ l lim ðuj;3 þ u3;j Þ ¼ pj ; j ¼ 1; 2; ð9:57Þ
x3 !þ0 x3 !þ0

lim r33 ¼ lim ðkh þ 2lu3;3 Þ ¼ p3 ; ð9:570 Þ


x3 !þ0 x3 !þ0

as well as the conditions of regularity at infinite


lim rjk ¼ 0; lim uj ¼ 0; j; k ¼ 1; 2; 3: ð9:58Þ
x3 !1 x3 !1

We introduce the matrix (6.123), whose elements satisfy the systems of dif-
ferential equations
lDukj þ ðk þ lÞhj;k ¼ 0; j; k ¼ 1; 2; 3; ð9:59Þ
9.2 Elastic Half-Space 413

where
hj ¼ ukj;k ; j ¼ 1; 2; 3: ð9:590 Þ
The corresponding boundary conditions on the separation plane read
1
lim ðu31;1 þ u11;3 Þ ¼  dðx1 ; x2 Þ;
x3 !þ0 l
lim ðu21;3 þ u31;2 Þ ¼ 0; ð9:60Þ
x3 !þ0

lim ðkh1 þ 2lu31;3 Þ ¼ 0;


x3 !þ0

lim ðu32;1 þ u12;3 Þ ¼ 0;


x3 !þ0
1
lim ðu22;3 þ u32;2 Þ ¼  dðx1 ; x2 Þ; ð9:600 Þ
x3 !þ0 l
lim ðkh2 þ 2lu32;3 Þ ¼ 0;
x3 !þ0

lim ðu33;1 þ u13;3 Þ ¼ 0;


x3 !þ0

lim ðu23;3 þ u33;2 Þ ¼ 0; ð9:6000 Þ


x3 !þ0

lim ðkh3 þ 2lu33;3 Þ ¼ dðx1 ; x2 Þ;


x3 !þ0

while the conditions of regularity at infinite are


lim ujk ¼ 0; j; k ¼ 1; 2; 3: ð9:61Þ
x3 !1

Applying the double Fourier transform with respect to the variables x1 and x2
and considering x3 as a parameter, we may write the equations (3.70) with (3.700 )
for j ¼ 1 in the form
2  
d 2
l  c 3 F ½ u ðx ; x
k1 1 2 3 ; x Þ   ðk þ lÞa k a1 F½u11 ðx1 ; x2 ; x3 Þ
dx23

d
þ a2 F½u21 ðx1 ; x2 ; x3 Þ þ i F½u31 ðx1 ; x2 ; x3 Þ ¼ 0; k ¼ 1; 2; ð9:62Þ
dx3
2 
d 2 d
l  c 3 F½u31 ðx1 ; x2 ; x3 Þ  iðk þ lÞ ½a1 F½u11 ðx1 ; x2 ; x3 Þ
dx23 dx3
d
þ a2 F½u21 ðx1 ; x2 ; x3 Þ þ i F½u31 ðx1 ; x2 ; x3 Þ ¼ 0; ð9:620 Þ
dx3
where F½ ¼ Fx1 ½Fx2 ½ and ak ; k ¼ 1; 2, are complex variables in the space of
transformations; applying the Fourier transforms to the boundary conditions (9.60)
as well, we obtain
414 9 Elastic Space. Elastic Half-Space

 
d
l lim F½u11 ðx1 ; x2 ; x3 Þ  ia1 F½u31 ðx1 ; x2 ; x3 Þ ¼ 1;
x3 !þ0 dx3
 
d
lim F½u21 ðx1 ; x2 ; x3 Þ  ia2 F½u31 ðx1 ; x2 ; x3 Þ ¼ 0;
x3 !þ0 dx3
 ð9:63Þ
d
2l lim F½u31 ðx1 ; x2 ; x3 Þ  ki lim a1 F½u11 ðx1 ; x2 ; x3 Þ
x3 !þ0 dx3 x3 !þ0

d
þa2 F½u21 ðx1 ; x2 ; x3 Þ þ i F½u31 ðx1 ; x2 ; x3 Þ ¼ 0:
dx3
Multiplying the Eq. (9.62) by ak ; k ¼ 1; 2, respectively and adding them, we
have

d2  
l 2
a1 F½u11 ðx1 ; x2 ; x3 Þ þ a2 F½u21 ðx1 ; x2 ; x3 Þ
dx3
 
 ðk þ 2lÞc23 a1 F½u11 ðx1 ; x2 ; x3 Þ þ a2 F½u21 ðx1 ; x2 ; x3 Þ
d
¼ iðk þ lÞc23 F½u31 ðx1 ; x2 ; x3 Þ; ð9:64Þ
dx3
on the other hand, the Eq. (9.620 ) may be written
d
ðk þ lÞ ½a1 F½u11 ðx1 ; x2 ; x3 Þ þ a2 F½u21 ðx1 ; x2 ; x3 Þ
dx3
 
d2 2
¼ i ðk þ 2lÞ 2 F½u31 ðx1 ; x2 ; x3 Þ  lc3 F½u31 ðx1 ; x2 ; x3 Þ : ð9:640 Þ
dx3

Thus, from the Eq. (9.64) and (9.640 ), one obtains the differential equations which
must be satisfied by the Fourier transform F½u31 ðx1 ; x2 ; x3 Þ, namely

d4 2 d
2
F½u 31 ðx 1 ; x 2 ; x3 Þ  2c 3 F½u31 ðx1 ; x2 ; x3 Þ
dx43 dx23
þ c43 F½u31 ðx1 ; x2 ; x3 Þ ¼ 0; ð9:65Þ

whose general solution is of the form


F½u31 ðx1 ; x2 ; x3 Þ ¼ ðA þ Bx3 Þec3 x3 þ ðC þ Dx3 Þec3 x3 ; ð9:650 Þ
the coefficients of which depend on the parameters aj ; j ¼ 1; 2. The conditions of
regularity for x3 ! 1 lead to C ¼ D ¼ 0.
Taking into account that
lim F½u31 ðx1 ; x2 ; x3 Þ ¼ A;
x3 !þ0
9.2 Elastic Half-Space 415

as well as the first two conditions (9.63), the Eq. (9.640 ) gives

2lðk þ lÞc23 A ¼ 2lðk þ 2lÞc3 B  iðk þ lÞa3 ð9:66Þ

for x3 ! þ0. Substituting the Fourier transform F½u31 ðx1 ; x2 ; x3 Þ given by (9.650 )
into the Eq. (9.640 ) and integrating, one obtains

iðk þ lÞ½a1 F½u11 ðx1 ; x2 ; x3 Þ þ a2 F½u21 ðx1 ; x2 ; x3 Þ


¼ f2ðk þ 2lÞB  ðk þ lÞ½c3 A þ ð1 þ c3 x3 ÞBgec3 x3 ; ð9:67Þ

where the integration constant vanishes since


lim ½a1 F½u11 ðx1 ; x2 ; x3 Þ þ a2 F½u21 ðx1 ; x2 ; x3 Þ ¼ 0: ð9:670 Þ
x3 !1

Passing to the limit for x3 ! þ0 in the relation (9.67) and taking into account the
third boundary condition (9.63) we get
lB ¼ ðk þ lÞc3 A: ð9:660 Þ

Thus, the Eqs. (9.66), (9.660 ) give the coefficients


1 a1 i a1
A¼ ; B¼ : ð9:68Þ
2ðk þ lÞ c3 2l c3
From the Eqs. (9.62), we obtain the equation
2 
d 2
 c3 ½a2 F½u11 ðx1 ; x2 ; x3 Þ  a1 F½u21 ðx1 ; x2 ; x3 Þ ¼ 0; ð9:69Þ
dx23

whose general solution is of the form


a2 F½u11 ðx1 ; x2 ; x3 Þ  a1 F½u21 ðx1 ; x2 ; x3 Þ ¼ Lec3 x3 þ Mec3 x3 ; ð9:690 Þ
where L and M are parameters to be determined; from the first two conditions
(9.63) and from the condition of regularity for x3 ! 1, one obtains
1 a2
L¼ ; M ¼ 0: ð9:680 Þ
l c3
Ultimately, from the relations (9.67), (9.690 ) and (9.650 ) and taking into account
the values obtained for the integration coefficients, the Fourier transforms read
416 9 Elastic Space. Elastic Half-Space

 
1 k a21 1 a21 c3 x3
F½u11 ðx1 ; x2 ; x3 Þ ¼ 1  x 3 e ;
lc3 2ðk þ lÞ c23 2 c3
 
1 a1 a 2 k c3 x3
F½u21 ðx1 ; x2 ; x3 Þ ¼  þ x 3 e ; ð9:70Þ
2l c23 ðk þ lÞc3
 
i a1 1 1
F½u31 ðx1 ; x2 ; x3 Þ ¼ þ x3 ec3 x3 :
2 c3 ðk þ lÞc3 l
For the computation of the Fourier transforms we remark that
  qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1
Fx1 ¼ 2K0 a1 x22 þ x23 ; ð9:71Þ
R

where K0 is the modified Bessel function of the second species and zero order;
then we have
 qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 1
Fx K0 a1 x22 þ x23 ¼ ec3 x3 ; Re a1 [ 0; ð9:710 Þ
p 2 c3

whence
 
1 1 1
F ¼ ec3 x3 ; x3 [ 0: ð9:7100 Þ
2p R c3

On the other hand, we may write


   
o 1 1 a1
F ¼ ia1 F ¼ 2pi ec3 x3 : ð9:72Þ
ox1 R R c3
Since
Z 1
1 c3 x3 1
e dx3 ¼ 2 ec3 x3 ;
0 c3 c3

there follows
Z 1   Z 1  
o 1 o 1 a1
F dx3 ¼ F dx3 ¼ 2pi 2 ec3 x3 : ð9:73Þ
x3 ox 1 R x3 ox 1 R c3
In the following, we shall express the inverse Fourier transforms with the aid of
the functions 1=R and lnðR þ x3 Þ; x3 [ 0, between which there exists the differ-
ential relation
o 1
lnðR þ x3 Þ ¼ ; ð9:74Þ
ox3 R
also, we may write
9.2 Elastic Half-Space 417

Z 1
1
dx3 ¼  lnðR þ x3 Þ; ð9:740 Þ
0 R
a relation to which the partial differentiation with respect to x1 and x2 can be
applied.
Thus, we may write the double Fourier transforms
 2 
1 o a2
F 2 lnðR þ x3 Þ ¼ 21 ec3 x3 ;
2p ox1 c3
 2  ð9:75Þ
1 o a1 a2 c3 x3
F lnðR þ x3 Þ ¼ 2 e ;
2p ox1 ox2 c3

whence we deduce also that


Z 1 2 
1 o a21 c3 x3
F 2
lnðR þ x 3 Þdx 3 ¼ e ;
2p x3 ox1 c33
Z 1  ð9:750 Þ
1 o2 a1 a2
F lnðR þ x3 Þdx3 ¼ 3 ec3 x3 :
2p x3 ox1 ox2 c3

Finally, we obtain
 Z 1 2
1 2 k o
u11 ðx1 ; x2 ; x3 Þ ¼  lnðR þ x3 Þdx3
4pl R k þ l x3 ox21
  
o2 1 1 x2
 x3 2 lnðR þ x3 Þ ¼ 1 þ 12
ox1 4pl R R
 
1  2m x21
þ 1 ;
R þ x3 RðR þ x3 Þ
 Z 1 2
1 k o
u21 ðx1 ; x2 ; x3 Þ ¼  lnðR þ x3 Þdx3 ð9:76Þ
4pl k þ l x3 ox1 x2
 " #
o2 1 x1 x2 1 1  2m
þ x3 lnðR þ x3 Þ ¼  ;
ox1 x2 4pl R R2 ðR þ x3 Þ2
 
1 l o o 1
u31 ðx1 ; x2 ; x3 Þ ¼ lnðR þ x3 Þ  x3
4pl k þ l ox1 ox1 R
 
1 x1 x3 1  2m
¼ þ :
4pl R R2 R þ x3

Likewise from the equations (9.59), (9.590 ) and the boundary conditions (9.600 ),
(9.6000 ), respectively, we obtain
418 9 Elastic Space. Elastic Half-Space

" #
1 x1 x2 1 1  2m
u12 ðx1 ; x2 ; x3 Þ ¼  ;
4pl R R2 ðR þ x3 Þ2
   
1 1 x21 1  2m x22
u22 ðx1 ; x2 ; x3 Þ ¼ 1þ 2 þ 1 ; ð9:760 Þ
4pl R R R þ x3 RðR þ x3 Þ

1 x2 x3 1  2m
u32 ðx1 ; x2 ; x3 Þ ¼ þ :
4pl R R2 R þ x3

1 x1 x3 1  2m
u13 ðx1 ; x2 ; x3 Þ ¼  ;
4pl R R2 R þ x3
 ð9:7600 Þ
1 x2 x3 1  2m
u23 ðx1 ; x2 ; x3 Þ ¼  :
4pl R R2 R þ x3
 
1 1 x2
u33 ðx1 ; x2 ; x3 Þ ¼ 2ð1  mÞ þ 32 :
4pl R R

9.2.4.2 Fundamental Solutions for Stresses

The fundamental solutions for stresses are given by the components of the matrix
(6.1320 ), which may be obtained starting from the components of the matrix U; by
means of the relations (6.138), one obtains.
(   )
1 x1 1 x21 1  2m x21 ð3R þ x3 Þ
s111 ðx1 ; x2 ; x3 Þ ¼ 1  2m  3 2  3 2 ;
2p R R2 R ðR þ x3 Þ2 R ðR þ x3 Þ
( 
1 x2 1 x21
s112 ðx1 ; x2 ; x3 Þ ¼ 1  2m  3 : ð9:77Þ
2p R R2 R2
 )
1  2m x21 ð3R þ x3 Þ
 3 2 ;
ðR þ x3 Þ2 R ðR þ x3 Þ
(   )
1 x3 1 x21 1  2m x21 ð2R þ x3 Þ
s113 ðx1 ; x2 ; x3 Þ ¼ 1  2m  3 2  3 2 ;
2p R R2 R ðR þ x3 Þ2 R ðR þ x3 Þ
(   )
1 x1 1 x22 2ð1  2mÞ x22 ð3R þ x3 Þ
s221 ðx1 ; x2 ; x3 Þ ¼ 1  2m  2  3 2 ;
2p R R2 R ðR þ x3 Þ2 R ðR þ x3 Þ
9.2 Elastic Half-Space 419

( 
1 x2 1 x22
s222 ðx1 ; x2 ; x3 Þ ¼ 1  2m  3 2
2p R R2 R
 ) ð9:770 Þ
1  2m x22 ð3R þ x3 Þ
 3 2 ;
ðR þ x3 Þ2 R ðR þ x3 Þ
   
1 x3 1 x22 1  2m x22 ð2R þ x3 Þ
s223 ðx1 ; x2 ; x3 Þ ¼ 1  2m  3 2  1 2 ;
2p R R2 R x3 ðR þ x3 Þ R ðR þ x3 Þ

3 x23 xk
s33k ðx1 ; x2 ; x3 Þ ¼  ; k ¼ 1; 2; 3; ð9:7700 Þ
2p R5
3 x2 x3 xk
s23k ðx1 ; x2 ; x3 Þ ¼  ; k ¼ 1; 2; 3; ð9:78Þ
2p R5
3 x3 x1 xk
s31k ðx1 ; x2 ; x3 Þ ¼  ; k ¼ 1; 2; 3; ð9:780 Þ
2p R5
(   )
1 x2 1 x21 1  2m x21 ð3R þ x3 Þ
s121 ¼ 13 2  1 2 ;
2p R R2 R ðR þ x3 Þ2 R ðR þ x3 Þ
(   )
1 x1 1 x22 1  2m x22 ð3R þ x3 Þ
s122 ¼ 13 2  1 2 ; ð9:7800 Þ
2p R R2 R ðR þ x3 Þ2 R ðR þ x3 Þ
" #
1 x1 x2 x3 2R þ x3
s123 ¼ 3 2  ð1  2mÞ ;
2p R3 R ðR þ x3 Þ2

9.2.4.3 Case of Arbitrary Loads

In the case of a loading expressed by the matrix


2 3
p1 ðx1 ; x2 Þ
p 4 p2 ðx1 ; x2 Þ 5; ð9:79Þ
p3 ðx1 ; x2 Þ

the state of displacement is given by the convolution product


u U
p; ð9:80Þ
where the matrices u and U are given by (6.132). In components, one has
ui ðx1 ; x2 ; x3 Þ ¼ uij ðx1 ; x2 ; x3 Þ
pj ðx1 ; x2 Þ; i ¼ 1; 2; 3: ð9:800 Þ
Analogically, the state of stresses becomes
420 9 Elastic Space. Elastic Half-Space

r ¼ S
p; ð9:81Þ

where the matrices r and S are given by (6.1320 ). In components one has
rij ðx1 ; x2 ; x3 Þ ¼ sijk ðx1 ; x2 ; x3 Þ
pk ðx1 ; x2 Þ; i; j ¼ 1; 2; 3: ð9:810 Þ
Let be now a concentrated force F, of components

pj ðx1 ; x2 Þ ¼ Fj dðx1  x01 ; x2  x02 Þ; j ¼ 1; 2; 3; ð9:82Þ

acting upon the half-space x3  0 on the separation plane, at the point Aðx01 ; x02 ; 0Þ.
The state of displacement is given by

ui ðx1 ; x2 ; x3 Þ ¼ uij ðx1 ; x2 ; x3 Þ


Fj dðx1  x01 ; x2  x02 Þ; i ¼ 1; 2; 3;

hence

ui ðx1 ; x2 ; x3 Þ ¼ Fj uij ðx1  x01 ; x2  x02 ; x3 Þ; i ¼ 1; 2; 3: ð9:820 Þ


Analogically, the state of stress reads

rij ðx1 ; x2 ; x3 Þ ¼ sijk ðx1 ; x2 ; x3 Þ


Fk dðx1  x01 ; x2  x02 Þ; i; j ¼ 1; 2; 3;

hence

rij ðx1 ; x2 ; x3 Þ ¼ Fk sijk ðx1  x01 ; x2  x02 ; x3 Þ; i; j ¼ 1; 2; 3: ð9:8200 Þ


We assume now, for the sake of simplicity, that the concentrated force acts at
the origin (x01 ¼ x02 ¼ 0), so that
ui ðx1 ; x2 ; x3 Þ ¼ Fj uij ðx1 ; x2 ; x3 Þ; i ¼ 1; 2; 3; ð9:83Þ

rij ðx1 ; x2 ; x3 Þ ¼ Fk sijk ðx1 ; x2 ; x3 Þ; i; j ¼ 1; 2; 3: ð9:830 Þ


If F1 ¼ F2 ¼ 0; F3 ¼ P 6¼ 0, the state of displacements is given by

P xi x3 1  2m
ui ðx1 ; x2 ; x3 Þ ¼  ; i ¼ 1; 2; ð9:84Þ
4pl R R2 R þ x3
 
P 1 x23
u3 ðx1 ; x2 ; x3 Þ ¼ 2ð1  mÞ þ 2 ; ð9:840 Þ
4pl R R

for x3 ¼ 0 one obtains the deformation of the separation plane, getting again the
formula (9.440 ) for u03 ¼ 0.
9.2 Elastic Half-Space 421

The state of stress reads


( 
P x3 1 x2i
rii ðx1 ; x2 ; x3 Þ ¼ 1  2m  3 2
2p RR2 R
 )
1  2m x2i ð2R þ x3 Þ
 1 2 ; i ¼ 1; 2;
ðR þ x3 Þ2 R ðR þ x3 Þ

3P x33
r33 ðx1 ; x2 ; x3 Þ ¼  ; ð9:85Þ
2p R5
3P x23 xi
r3i ðx1 ; x2 ; x3 Þ ¼  ; i ¼ 1; 2;
" 2p R5 # ð9:850 Þ
P x3 2R þ x3
r12 ðx1 ; x2 ; x3 Þ ¼  3  ð1  2mÞ :
2pR R2 ðR þ x3 Þ2

These results correspond to the problem of Boussinesq [1], i.e., the problem of a
normal concentrated force.
If F2 ¼ F3 ¼ 0; F1 ¼ P 6¼ 0, the state of displacement is given by
   
P 1 x21 1  2m x21
u1 ðx1 ; x2 ; x3 Þ ¼ 1þ 2 þ 1 ;
4pl R R R þ x3 RðR þ x3 Þ
" #
P x1 x2 1 1  2m
u2 ðx1 ; x2 ; x3 Þ ¼  ; ð9:86Þ
4pl R R2 ðR þ x3 Þ2

P x1 x3 1  2m
u3 ðx1 ; x2 ; x3 Þ ¼ þ ;
4pl R R2 R þ x3

for x3 ¼ 0, one obtains the deformation of the separation plane


1  2m Px1
u3 ðx1 ; x2 ; 0Þ ¼ : ð9:860 Þ
4pl R2
The state of stress reads
(   )
P x1 1 x21 1  2m x2i ð3R þ x3 Þ
rii ðx1 ; x2 ; x3 Þ ¼ 1  2m  3 2  3 2 ;
2p R R2 R ðR þ x3 Þ2 R ðR þ x3 Þ
i ¼ 1; 2;

3P x1 x23
r33 ðx1 ; x2 ; x3 Þ ¼  ; ð9:87Þ
2p R5
422 9 Elastic Space. Elastic Half-Space

3P x3 x1 xi
r3i ðx1 ; x2 ; x3 Þ ¼  ; i ¼ 1; 2;
( 2p R5
  )
P x2 1 x21 1  2m x21 ð3R þ x3 Þ
r12 ðx1 ; x2 ; x3 Þ ¼  13 2  1 2 :
2p R R2 R ðR þ x3 Þ2 R ðR þ x3 Þ
ð9:870 Þ
The only non-zero stress on the separation plane is

mP x2 x21
r12 ðx1 ; x2 ; 0Þ ¼ 13 2 : ð9:8700 Þ
p R3 R
These results correspond to the problem of Cerruti [12], i.e. the problem of a
tangential concentrated force.

9.2.4.4 Concentrated Moments

Let us now consider an elastic half-space x3  0 acted upon at the point Aðx01 ; x02 ; 0Þ
by a directed concentrated moment of magnitude M, specified by the unit vectors u
and F0 ; the state of displacement is obtained in the form
M o
ui ðx1 ; x2 ; x3 Þ ¼  u ðF0 Þ;
0 ou i
i ¼ 1; 2; 3; ð9:88Þ
ju  F j

where o=ou is the directional derivative and ui ðF0 Þ; i ¼ 1; 2; 3, are the dis-
placements corresponding to the action of the concentrated force F0 (of unit
magnitude), in accordance with the formulae (9.820 ).
In particular, if u ¼ i1 and F0 ¼ i3 , where i1 and i3 are the unit vectors of the
Ox1 and Ox3 axes, respectively, we obtain a directed concentrated moment
assumed to be applied at the origin; the moment will cause a positive rotation in
the Ox3 x1 -plane. The equivalent load is given by
o
pi ðx1 ; x2 Þ ¼ 0; i ¼ 1; 2; p3 ðx1 ; x2 Þ ¼ M dðx1 ; x2 Þ ð9:89Þ
ox
and the resulting state of displacement is
9.2 Elastic Half-Space 423

o
u1 ðx1 ; x2 ; x3 Þ ¼ M u13 ðx1 ; x2 ; x3 Þ
ox1
   
M 1 x3 x21 1  2m x21 ð2R þ x3 Þ
¼ 13 2  1 2 ;
4pl R R2 R R þ x3 R ðR þ x3 Þ
o
u2 ðx1 ; x2 ; x3 Þ ¼ M u23 ðx1 ; x2 ; x3 Þ
ox1
" # ð9:90Þ
M x1 x2 x3 2R þ x3
¼ 3 2  ð1  2mÞ ;
4pl R3 R ðR þ x3 Þ2
o
u3 ðx1 ; x2 ; x3 Þ ¼ M u33 ðx1 ; x2 ; x3 Þ
ox1
 
M x1 x23
¼ 2ð1  mÞ þ 3 :
4pl R3 R2
Likewise, let x3  0 be an elastic half-space acted upon, on the separation plane,
at the point Aðx01 ; x02 ; 0Þ, by a rotational concentrated moment of magnitude M; we
assume that the moment causes a positive rotation, may be written in the form
1 o
p1 ðx1 ; x2 Þ ¼ M dðx1  x01 ; x2  x02 Þ;
2 ox2
1 o ð9:91Þ
p2 ðx1 ; x2 Þ ¼  M dðx1  x01 ; x2  x02 Þ;
2 ox1
p3 ðx1 ; x2 Þ ¼ 0;

the resulting state of displacement will be



1 o
u1 ðx1 ; x2 ; x3 Þ ¼ M u11 ðx1  x01 ; x2  x02 ; x3 Þ
2 ox2

o 0 0
 u12 ðx1  x1 ; x2  x2 ; x3 Þ
ox1
M x2
¼  ;
4pl ½ðx1  x01 Þ2 þ ðx2  x02 Þ2 þ x23 3=2

1 o
u2 ðx1 ; x2 ; x3 Þ ¼ M u21 ðx1  x01 ; x2  x02 ; x3 Þ
2 ox2
 ð9:92Þ
o 0 0
 u22 ðx1  x1 ; x2  x2 ; x3 Þ
ox1
M x1
¼ ;
4pl ½ðx1  x1 Þ þ ðx2  x02 Þ2 þ x23 3=2
0 2

1 o
u3 ðx1 ; x2 ; x3 Þ ¼ M u31 ðx1  x01 ; x2  x02 ; x3 Þ
2 ox2

o 0 0
 u32 ðx1  x1 ; x2  x2 ; x3 Þ ¼ 0:
ox1
424 9 Elastic Space. Elastic Half-Space

We shall now consider that the elastic half-space x3  0 is acted upon at the origin
O, on the separation plane, by a concentrated moment of linear dipole type, of
magnitude D, the support of which being in the direction of the Ox1 -axis. The
equivalent load is expressed by
o
p1 ðx1 ; x2 Þ ¼ D dðx1 ; x2 Þ; pi ðx1 ; x2 Þ ¼ 0; i ¼ 2; 3; ð9:93Þ
ox1
hence, the state of displacement reads.
o
ui ðx1 ; x2 ; x3 Þ ¼  Dui1 ðx1 ; x2 ; x3 Þ
ox1
(   )
D xi 1 x21 1  2m x21 ð3R þ x3 Þ
¼ 13 2  3 2 ; i ¼ 1; 2;
4pl R R2 R ðR þ x3 Þ2 R ðR þ x3 Þ
ð9:94Þ
o
u3 ðx1 ; x2 ; x3 Þ ¼  D u31 ðx1 ; x2 ; x3 Þ
ox1
   
D 1 x3 x21 1  2m x21 ð3R þ x3 Þ
¼ 1  3 þ 1  :
4pl R R2 R2 R þ x3 R2 ðR þ x3 Þ

Other cases of loading may be treated in a similar way. By applying the result
presented in Sect. 9.2.3.2, one may easily write, in addition, the state of stress in
each of the considered cases to.

References

A. Books

1. Bussinesq, J.: Applications des potentiels à l’étude de l’équilibre et du mouvement des


solides déformables, Paris (1885)
2. F}oppl, A., F}oppl, L.: Drang und Zwang Eine h}ohere Festigkeitslehre für Ingenieure, I. Verlag
von H. Oldenbourg, München (1920)
3. Haimovici, M.: Teoria Elasticitătßii (Theory of Elasticity). Ed. Did. Ped., Bucuresßti (1969)
4. Kecs, W., Teodorescu, P.P.: Applications of the Theory of Distributions in Mechanics.
Ed. Acad., Bucuresßti, Abacus Press, Tunbridge Wells, Kent (1974)
5. Kecs, W., Teodorescu, P.P.: Vvedenie v teoriyu obobshchonnykh funktsii s prilozheniiami v
tekhnike (Introduction to the Theory of Distributions with Applications in Technics), Izd.
‘‘Mir’’, Moskva (1978)
6. Kirchhoff, G.R.: Gesammelte Abhandlungen, Leipzig (1882)
7. Kr}oner, E.: Kontinuumstheorie der Versetzungen und Eigenspannungen. Springer, Berlin
(1958)
8. Love, A.-E.-H.: A Treatise on the Mathematical Theory of Elasticity. Univ. Press, Cambridge
(1892)
9. Lur’e, A.I.: Prostranstvennye zadachi teorii uprugosti (Spatial Problems of the Theory of
Elasticity). Gostekhizdat, Moskva (1955)
References 425

10. Sneddon, I.N., Berry, D.S.: The Classical Theory of Elasticity, in Handbuch der Physik VI.
Elastizität und Plastizität. Springer, Berlin (1958)
11. Teodorescu, P.P.: Probleme spatßiale în teoria elasticitătßii (Spatial Problems in the Theory of
Elasticity). Ed. Acad., Bucuresßti (1970)

B. Papers

12. Cerruti, V.: Ricerche intorno all’equilibrio dei corpi elastici isotropi. Atti R. Accad. Lincei,
Mem. Cl. Sci. Fis., Mat. e Nat. 13, 81 (1852)
}
13. Hertz, H.: Uber die Berührung fester elasticher Körper. J. Für Reine Angew. Math. 92, 156
(1882)
14. Kecs, W.: Sur les problèmes concernant le demi-espace élastique. Bull. Math. Soc. Sci. Math.
Roum. 6(54), 157 (1962)
15. Kecs, W.: Les solutions généralisées des problèmes concernant l’espace élastique infini. Rev.
Roum. Math. Pures Appl. 11, 27 (1966)
16. Kecs, W.: Les solutions généralisées de l’espace élastique sous l’action de certaines charges
variables. Rev. Roum. Math. Pures Appl. 15, 7 (1970)
17. Lai Pham The and Mandel, J.: Force concentrée agissant au sein d’un semiespace élastique,
homogène, isotrope, limité par un plan dont les points restent fixes. C. Rend. Acad. Sci., Paris
260, 4417 (1965)
18. Lamb, H.: Proc. London Math. Soc. 34, 276 (1902)
19. Love, A.-E.-H.: The stress produced in a semi-infinite solid by pressure of part of the
boundary. Phil. Trans. Roy. Soc. London Ser. A 228, 377 (1929)
20. Lur’e, A.I.: Nekotorye kontaktnye zadachi teorii uprugosti (Some contact problems of the
theory of elasticity). Prikl. Mat. Mekh. 5, 383 (1941)
21. Mindlin, R.D.: Physics 7, 195 (1936)
22. Schleicher, F.: Bauingenieur 7 (1926)
23. Schleicher, F.: Bauingenieur 14, 242 (1933)
24. S
ßandru, N.: O sosredotochennykh vnutrennikh nagruzkakh v uprugom poluprostranstve s
pakreplenymi krayami (On the concentrated loads in the interior of the elastic half-space with
a free separation plane). Bull. Math. Soc. Sci. Math. Phys. Roum. 5(53), 205 (1961)
25. S
ßandru, N.: Asupra actßiunii unei fortße concentrate în spatßiul elastic nemărginit (On the action
of a concentrated force in the elastic infinite space). Com. Acad. Rom. XIII, 1019 (1963)
26. S
ßandru, N.: O deistvii peremennykh sil v neogranichennom uprugom prostranstve (On the
action of a variable force in the elastic space). Bull. Acad. Pol. Sci., Ser. Sci. Techn. 12, 45
(1964)
27. Tanimoto, B.: J. Shinshu Univ., Nagano (1957)
28. Tedone, O.: Saggio di una teoria generale delle equazioni dell’equilibrio elastico per un
corpo isotropo. Ann. Mat. Pura Appl. Ser. III 10, 13 (1904)
29. Teodorescu, P.P.: Asupra unor probleme spatßiale ale teoriei elasticitătßii (On some spatial
problems of the theory of elasticity). St. Cerc. Mec. Apl. VIII. 1101 (1957)
}
30. Teodorescu, P.P.: Uber die Berechnung des elastichen Halbraumes unter örtlicher Belastung.
Bull. Math. Soc. Sci. Math. Phys. Roum. 2(50), 113 (1958)
31. Teodorescu, P P.: Uber} die Berechnung des periodisch belasteten elastischen Halbraumes.
Rev. Méc. Appl. IV, 141 (1959)
32. Teodorescu, P.P.: Fonctions de tension dans le problème tridimensionnel de la théorie de
l’élasticité. Bull. Math. Soc. Sci. Math. Phys. Roum. 3(51), 499 (1959)
}
33. Teodorescu, P.P.: Uber einige räumliche Probleme der Elastizitätstheorie. Apl. Mat., Praha
IV, 225 (1959)
426 9 Elastic Space. Elastic Half-Space

34. Teodorescu, P.P.: Trei probleme tridimensionale ale teoriei elasticitătßii (Three tridimensional
problems of the theory of elasticity). An. Univ. Bucuresßti, Ser. ßst. Nat., Mat.-Mec. XV, 1–17
(1966)
35. Teodorescu, P.P.: Asupra actßiunii sarcinilor concentrate în teoria elasticitătßii (On the action
of concentrated loads in the theory of elasticity). An. Univ. Bucuresßti, Ser. ßst. Nat., Mat.-Mec.
XV, 2–15 (1966)
36. Teodorescu, P.P.: Einige Teilprobleme der dreidimensionalen Elastizitätstheorie. Wiss.
Zeitschrift Techn. Univ. Dresden 16, 77 (1967)
37. Teodorescu, P.P.: Asupra utilizării unor familii de functßii-potentßial reale în teoria elasticitătßii
(On the use of certain families of real potential functions in the theory of elasticity). An.
Univ. Bucuresßti, Ser. ßst. Nat., Mat.-Mec. XVI, 59 (1967)
38. Teodorescu, P.P.: Sur l’application des fonctions potentiel dans la théorie de l’élasticité.
Rend. Sem. Mat. e Fis. di Milano XXXVIII, 231 (1968)
39. Teodorescu, P.P.: Considérations concernant l’introduction des fonctions potentiel de
déplacement dans les problèmes en espace de la théorie de l’élasticité (cas statique). An.
Univ. Bucuresßti, Ser. Mat.-Mec. XVIII, 127 (1969)
40. Teodorescu, P.P.: Consideratßii în legătură cu utilizarea functßiilor potential în teoria
elasticitătßii (Considerations concerning the use of potential functions in the theory of
elasticity). Lucr. ßst. Inst. Ped., Galatßi 5, 27 (1971)
41. Teodorescu, P.P.: Sur l’introduction des fonctions-potentiel en élasticité linéaire. An. Univ.
Bucuresßti, Mat.-Mec., 20, 131 (1971)
42. Terezawa, K.: J. Coll. Sci. Univ. Tokyo
43. Thomson, W.: (Lord Kelvin), Cambridge and Dublin Math. J. (1848)
44. Vodička, V.: Ein durch allgemeine Masseukräfte beauspruchtes unendliches medium.
ZAMM 39, 2 (1959)
45. Weitsman, Y.: Quart. Appl. Math. XXV, 213 (1967)
Chapter 10
Elastic Eighth-Space. Elastic
Quarter-Space

In what follows we deal with the study of the elastic eighth-space subjected to the
action of a normal or tangential load; the results thus obtained will then be par-
ticularized for the case of the elastic quarter-space. We remember that they have
been used in Sect. 9.2 for the elastic half-space. We will take advantage of the
stress functions introduced in [3, 4, 13, 16].

10.1 Elastic Eighth-Space

The problem of the state of strain and stress in the interior of an elastic eighth-
space has been only a few times studied. But the problem is of interest, because
one can thus obtain informations concerning such a local state around a solid angle
with three faces of an elastic solid; if the three plane faces are not orthogonal one
each other, then the problem may be studied similarly, passing to a system of
oblique Cartesian co-ordinate axes. Interesting results are known for the corre-
sponding plane case (the elastic quarter-plane [51], the elastic wedge). A particular
case of loading with an internal concentrated force has been considered by
W. A. Hijab [2].
The elastic eighth-space, acted upon by a periodic load on one of its faces, may
practically constitute the end beam in the case of a continuous wall-beam of
infinite height, having an infinity of equal spans and identically acted upon. The
results corresponding to a local loading may be as well used for the study of the
local effect (corner effect) in the support zone of a wall-beam with a single span.
We will consider the elastic eighth-space xi  0; i ¼ 1; 2; 3, acted upon by
normal loads, using the results given in [6, 8] or by tangential loads, using the
results given in [7]. We will present also a particular case of loading, which leads
to results in finite form, as it has been shown in [10, 14].

P. P. Teodorescu, Treatise on Classical Elasticity, Mathematical and Analytical 427


Techniques with Applications to Engineering, DOI: 10.1007/978-94-007-2616-1_10,
Ó Springer Science+Business Media Dordrecht 2013
428 10 Elastic Eighth-Space. Elastic Quarter-Space

10.1.1 Action of a Periodic Normal Load

Let be the elastic eighth-space xi  0; i ¼ 1; 2; 3, acted upon by a normal load


p3 ðx1 ; x2 Þ; periodic (in two directions) on the face x3 ¼ 0. To fix the ideas, we
represent this load by means of a double Fourier series, even with respect to the
variables x1 and x2 , of the form (9.2.1), (9.2.10 ).

10.1.1.1 General Considerations

The elastic eighth-space xi  0; i ¼ 1; 2; 3, may be thought as a part of the elastic


half-space, acted upon by the periodic load (9.2.1), (9.2.10 ), symmetric with
respect to the Ox1 and Ox2 axes. The tangential stresses vanish on the faces x1 ¼ 0
and x2 ¼ 0, while the normal stresses are given by
m
r11 ð0; x2 ; x3 Þ ¼ b0
1m
X X bnm  
þ 2
a2n ð1  cnm x3 Þ þ 2mb2m ecnm x3 cos bm x2 ;
n m
cnm
m ð10:1Þ
r22 ðx1 ; 0; x3 Þ ¼ b0
1m
X X bnm  
þ 2
b2m ð1  cnm x3 Þ þ 2ma2n ecnm x3 cos an x1 ;
n m
cnm

where we took into account the formulae (9.2.11).


The state of stress thus calculated corresponds to that which appears in the
elastic eighth-space, acted upon by the periodic load (9.2.1), (9.2.10 ) on the face
x3 ¼ 0 and by the normal loads p01 ðx2 ; x3 Þ ¼ r11 ð0; x2 ; x3 Þ and p02 ðx3 ; x1 Þ ¼
r22 ðx1 ; 0; x3 Þ on the faces x1 ¼ 0 and x2 ¼ 0, respectively. By superposing the
state of stress corresponding to the elastic eighth-space acted upon only by the
normal loads p01 ðx2 ; x3 Þ and p02 ðx3 ; x1 Þ on the faces x1 ¼ 0 and x2 ¼ 0, respectively,
one obtains the state of stresses corresponding to the problem enounced above.
To study the elastic eighth-space acted upon by the aperiodic (local) loads
p01 ðx2 ; x3 Þ, p02 ðx3 ; x1 Þ by means of Fourier integrals, one observes that the Fourier
coefficients bnm verify the conditions
10.1 Elastic Eighth-Space 429

Z  
1 X X 
 bnm  2 2
 cnm x3 
 a n ð1  c nm x 3 Þ þ 2mb m e dx3
0 
n m
c2nm 
("    2 Z 1 #
XX an 2 b
 jbnm j þ2m m x3 ecnm x3 dx3
n m
c nm c nm 0
 2 Z 1 )
an
þcnm ecnm x3 dx3 ð10:2Þ
cnm 0
( "   2 #  2 )
XX 1 an 2 bm 1 an
¼ jbnm j þ2m þ 2 ¼ finite;
n m
cnm cnm cnm cnm cnm
Z 1 X X 

 bnm  2 2
 cnm x3 
 b n ð1  c nm x 3 Þ þ 2ma n e dx3 ¼ finite:
0  n m
c2nm 

We mention that these conditions are verified for any case of distributed loading.

10.1.1.2 Discussion

Considering a case of periodic normal load, represented by a Fourier series odd


with respect to the variable x1 , e.g., and using the same procedure, one remarks
that the normal stress vanishes for x1 ¼ 0. Thus, one obtains the tangential stresses
0 0
r12 ð0; x2 ; x3 Þ ¼ t12 ðx2 ; x3 Þ and r13 ð0; x2 ; x3 Þ ¼ t13 ðx2 ; x3 Þ and we are obliged to
study the elastic eighth-space acted upon by a local tangential load on the face
x1 ¼ 0. These considerations remain valid for any other case of periodic tangential
load.
But the load considered above, e.g., on the face x1 ¼ 0, is local along the Ox3 -
axis and periodic along the Ox2 -axis. Thus, we are lead to the study of the elastic
eighth-space acted upon by a normal load periodic on one direction and local on
the other one, e.g. by the load
X Z 1
p3 ð x 1 ; x 2 Þ ¼ cos bm x2 bm ðaÞ cos ax1 da ð10:3Þ
m 0

on the face x3 ¼ 0. One obtains thus the normal stresses


X Z 1
bm ðaÞ 2
r11 ðx1 ; x2 ; x3 Þ ¼ cos bm x2 a ð1  cm x3 Þ
m 0 c2m

þ 2mb2m ecm x3 cos axda;
X Z 1 ð10:4Þ
bm ðaÞ 2
r22 ðx1 ; x2 ; x3 Þ ¼ cos bm x2 bm ð1  cm x3 Þ
m 0 c2m

þ2ma2 ecm x3 cos axda;
430 10 Elastic Eighth-Space. Elastic Quarter-Space

where
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
cm ¼ a2 þ b2m : ð10:40 Þ

One obtains thus normal stresses periodic on one direction and local on the
other one on the faces xi ; i ¼ 1; 2. By an analogous reasoning, we conclude that
one must study the elastic quarter-space xi  0; i ¼ 1; 3, or the elastic quarter-
space xi  0; i ¼ 2; 3, acted upon by local loads on both directions.
In conclusion, the problem of the elastic eighth-space acted upon by an arbi-
trary normal periodic load may be reduced, using the principle of superposition of
effects, to the study of the same domain, acted upon by local loads, and to the
study of an elastic quarter-space, acted upon in an analogous manner. Thus, it is
sufficient to consider only the case of a local load.

10.1.2 Action of a Local Normal Load

Let us study the elastic eighth-space xi  0; i ¼ 1; 2; 3, acted upon by the normal


loads
Z Z
1 1 1
p1 ðx2 ; x3 Þ ¼ a1 ðx2 ; x3 Þ cos a2 x2 cos a3 x3 dx2 dx3 ; . . . ð10:5Þ
4 1 1
on the faces xi ¼ 0; i ¼ 1; 2; 3, respectively. We use double Fourier integrals,
even with respect to the variables xi ; i ¼ 1; 2; 3, which may represent distribu-
tions too; the functions a1 ðx2 ; x3 Þ; a2 ðx3 ; x1 Þ; a3 ðx1 ; x2 Þ are known.

10.1.2.1 Stresses Functions. Boundary Conditions

We put the boundary conditions


x1 ¼ 0 : r11 ¼ p1 ðx2 ; x3 Þ; r12 ¼ r13 ¼ 0; . . . ð10:6Þ
As well, the state of stress must vanish for xi ! 1; i ¼ 1; 2; 3.
Using the representation in Sect. 5.2.2.5, we choose stress functions of the form
10.1 Elastic Eighth-Space 431

Z Z
1 1 1 0
F11 ¼ A ða2 ; a3 Þ
4 1 1 1

þc1 x1 A02 ða2 ; a3 Þ ec1 x1 cos a2 x2 cos a3 x3 da2 da3
Z Z
1 1 1 0
þ B ða3 ; a1 Þ
4 1 1 1

þc2 x2 B02 ða3 ; a1 Þ ec2 x2 cos a3 x3 cos a1 x1 da3 da1
Z Z
1 1 1 0
þ C ða1 ; a2 Þ
4 1 1 1

þc3 x3 C20 ða1 ; a2 Þ ec3 x3 cos a1 x1 cos a2 x2 da1 da2 ; . . .; ð10:7Þ

where we introduced the notations


qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
c1 ¼ a22 þ a23 ; c2 ¼ a23 þ a21 ; c3 ¼ a21 þ a22 ð10:70 Þ

and where the integrals are considered in the sense used in the theory of regu-
larization of the Fourier transforms in the theory of distributions.
The integrable functions A0i ða2 ; a3 Þ; B0i ða3 ; a1 Þ; Ci0 ða1 ; a2 Þ; A00i ða2 ; a3 Þ; . . .,
000
Ci ða1 ; a2 Þ; a1 ; a2 ; a3 2 ð1; 1Þ; i ¼ 1; 2, must be specified by the boundary
conditions (10.6). These functions are even with respect to the variables
a1 ; a2 ; a3 . The conditions at infinite are automatically verified, the method used
corresponding to Fourier transformations.
Using the results given at the above mentioned subsection, we obtain the
normal stresses
Z Z
1 1 1 2
r11 ¼  c A1 þ 2ð2 þ mÞA02
4 1 1 1

 ð2  c1 x1 ÞA  2 ec1 x1 cos a2 x2 cos a3 x3 da2 da3
Z Z
1 1 1  2
þ a B1 2ð2 þ mÞa22 B02
4 1 1 1

 2 c2 x2 ec2 x2 cos a3 x3 cos a1 x1 da3 da1
þ a21 B
Z Z
1 1 1  2
þ a C1  2ð2 þ mÞa23 C20
4 1 1 1

 2 c3 x3 ec3 x3 cos a1 x1 cos a2 x2 da1 da2 ; . . .;
þ a21 C ð10:8Þ
432 10 Elastic Eighth-Space. Elastic Quarter-Space

Z Z
1 1 1 1  2 2 
r23 ¼  a2 a3 A1 þ c41 A02  c21 ð2a23  ma22 ÞA002
4 1 1 a2 a3
 c x
þ ð2a22  ma23 ÞA000 2 2
2 þ a 2 a 3 A2 c 1 x 1 e
1 1
sin a2 x2 sin a3 x3 da2 da3
Z 1Z 1
1 c2  2   
 a B1  ða23  ma21 ÞB02 þ ð1 þ mÞa23 þ ma21 B002
4 1 1 a3 3
  c x
 ð3 þ mÞa23 þ 2a21 B000 2
2 þ a 3 B2 c 2 x 2 e
2 2
sin a3 x3 cos a1 x1 da3 da1
Z 1Z 1
1 
c3 2   
 a2 C1  ða22  ma21 ÞC20  ð3 þ mÞa22 þ 2a21 C200
4 1 1 a2
 
þ ð1 þ mÞa22 þ a21 C2000 þ a22 C  2 c3 x3 ec3 x3 cos a1 x1 sin a2 x2 da1 da2 ; . . .;
ð10:80 Þ

with the notations


 i ¼ A0 þ A00 þ A000 ; i ¼ 1; 2; . . .
A ð10:9Þ
i i i

Putting the boundary conditions for the tangential stresses and taking into
account (10.80 ), one obtains
   
 1 þ ð1 þ mÞa2 þ a2 A0  ð3 þ mÞa2 þ 2a2 A00  ða2  ma2 ÞA000 ¼ 0; . . .;
a22 A 2 3 2 2 3 2 2 3 2
   
 1 þ ð1 þ mÞa2 þ a2 A0  ða2  ma2 ÞA00  ð3 þ mÞa2 þ 2a2 A000 ¼ 0; . . .;
a23 A 3 2 2 3 2 2 2 2 2

wherefrom



 1 ¼  2 þ m a2 þ 2ma2 A00  a2 þ 2ma2 A000 ; . . .;
A 3 2 2 2 3 2
a23  a22


 ð10:10Þ
1
A02 ¼
2 2
2a23 þ ma22 A002  2a22 þ ma23 A000
2 ; . . .;
m a3  a 2

we may write


 2 ¼
2 þ m a2 A00  a2 A000 ; . . .
A ð10:100 Þ
2 2 3 2 2 2
m a3  a2

too. Introducing the notations


1 c1
2 00
C 1 ð a2 ; a3 Þ ¼  2 2
a3 A2  a22 A000
2 ;
m ð 2 þ m Þ a3  a2
1 c2
2 000
C 2 ð a3 ; a1 Þ ¼  2 2
a1 B2  a23 B02 ; ð10:11Þ
m ð 2 þ m Þ a1  a3
1 c3
2 0
C 3 ð a1 ; a2 Þ ¼  2 2
a2 C2  a21 C200 ;
m ð 2 þ m Þ a2  a1
10.1 Elastic Eighth-Space 433

we obtain the state of stress in the form


Z Z
1 1 1
r11 ¼ c C1 ð1 þ c1 x1 Þ ec1 x1 cos a2 x2 cos a3 x3 da2 da3
4 1 1 1
Z Z " 
1 1 1 a1 2
þ c2 C2 ð1  c2 x2 Þ
4 1 1 c2
 2 #
a3
þ2m ec2 x2 cos a3 x3 cos a1 x1 da3 da1
c2
Z Z " 
1 1 1 a1 2
þ c3 C3 ð1  c3 x3 Þ
4 1 1 c3
 2 #
a2
þ2m ec3 x3 cos a1 x1 cos a2 x2 da1 da2 ; . . .; ð10:12Þ
c3
Z Z
1 1 1 a2 a3
r23 ¼  C1 ð1  2m  c1 x1 Þ ec1 x1 sin a2 x2 sin a3 x3 da2 da3
4 1 1 c1
Z Z
1 1 1
þ a3 C2 c2 x2 ec2 x2 sin a3 x3 cos a1 x1 da3 da1
4 1 1
Z Z
1 1 1
þ a2 C3 c3 x3 ec3 x3 cos a1 x1 sin a2 x2 da1 da2 ; . . .; ð10:120 Þ
4 1 1
Due to the factors ec1 x1 ; ec2 x2 ; ec3 x3 , the functions Ci ; i ¼ 1; 2; 3, may be
considered even locally integrable and bounded for xi [ 0; i ¼ 1; 2; 3. The
relations
Z 1 Z 1
2 2
jec1 x1 jdx1 ¼ ¼ finite; . . .; jx1 ec1 x1 jdx1 ¼ 2 ¼ finite; . . .
1 c 1 1 c 1

allow to represent the corresponding functions by means of the Fourier integrals


Z 1
c1 x1 1 1
e ¼ c1 2
cos a1 x1 da1 ; . . .;
p 1 a
Z 1 2 ð10:13Þ
c1 x1 1 a  2a21
x1 e ¼ cos a1 x1 da1 ; . . .;
p 1 a4
with the notation

a2 ¼ a21 þ a22 þ a23 ; ð10:130 Þ


434 10 Elastic Eighth-Space. Elastic Quarter-Space

if results
Z 1 2
2 a1
ð1  c1 x1 Þec1 x1 ¼ c1 cos a1 x1 da1 ; . . .;
p a4
Z1
1 ð10:1300 Þ
2 1
ð1  c1 x1 Þec1 x1 ¼ c31 4
cos a1 x1 da1 ; . . .
p 1 a

Taking into account the relations (10.13), (10.1300 ), the representation (10.12)
leads to
Z Z
1 1 1
r11 ¼ c C1 ð1 þ c1 x1 Þ ec1 x1 cos a2 x2 cos a3 x3 da2 da3
4 1 1 1
Z Z Z 1 2
1 1 1 a2
þ C2 a21 4
cos a2 x2 da2
2p 1 1 1 a
Z 1
1
þma23 2
cos a2 x2 da2 cos a3 x3 cos a1 x1 da3 da1
1 a
Z Z Z 1 2
1 1 1 a1
þ C3 a21 4
cos a3 x3 da3
2p 1 1 1 a
Z 1
1
þma22 2
cos a3 x3 da3 cos a1 x1 cos a2 x2 da1 da2 ; . . . ð10:1400 Þ
1 a

Inverting the order of integration (which is possible, because for


xi [ 0; i ¼ 1; 2; 3, there exists an integrable majorant if Ci ; i ¼ 1; 2; 3, are
only integrable functions), we obtain
Z Z
1 1 1
r11 ¼ ½c C1 ð1 þ c1 x1 Þec1 x1 þ u1  cos a2 x2 cos a3 x3 da2 da3 ; . . .;
4 1 1 1
ð10:15Þ
where following Fourier transforms have been introduced
Z
2 1 a21 2
u1 ða2 ; a3 ; x1 Þ ¼ ða C2 þ a23 C3 Þ cos a1 x1 da1
p 1 a4 2
Z
2m 1 1 2
þ ða C2 þ a22 C3 Þ cos a1 x1 da1 ; . . .; ð10:16Þ
p 1 a2 3
where a2 ; a3 and a3 ; a1 as well a1 ; a2 , respectively, are fixed parameters.
Let be a sequence of integrable functions C1n ða2 ; a3 Þ; . . ., so that
C1n ða2 ; a3 Þ;  K; . . . ð10:17Þ

and

C1n ða2 ; a3 Þ ! C10 ða2 ; a3 Þ; . . . ð10:170 Þ


10.1 Elastic Eighth-Space 435

Thus it is obvious that


 0 
C ða2 ; a3 Þ\K; . . .; ð10:1700 Þ
1

hence, the functions C1 ða2 ; a3 Þ; . . . are locally integrable. If the functions


un1 ða2 ; a3 ; x1 Þ; . . . are the Fourier transforms corresponding to (10.15), we get
 n 
u ða2 ; a3 ; x1 Þ\K; . . .; ð10:18Þ
1

i.e.,
un1 ða2 ; a3 ; x1 Þ ! u01 ða2 ; a3 ; x1 Þ; . . .; ð10:180 Þ

hence, the convergence is uniform with respect to x1 ; x2 ; x3 , respectively.


Let be rn11 ðx1 ; x2 ; x3 Þ; . . . the relations (10.12) and (10.15), respectively, for
n
C1 ða2 ; a3 Þ; . . . Using the formula (10.12), we observe that one has to do with a
uniform convergence of the functions rn11 ðx1 ; x2 ; x3 Þ; . . . (in the normal sense) for
xi [ h [ 0; i ¼ 1; 2; 3. The functions rn11 ð0; x2 ; x3 Þ; . . . become generalized
functions (in the sense of the theory of distributions) for x1 ¼ 0; . . ., so that their
convergence must be considered in a corresponding sense.
Applying the theory of Fourier transforms in distributions, one concludes that
rn11 ðx1 ; x2 ; x3 Þ, considered as a generalized function in x2 and x3 , converges to the
function r011 ðx1 ; x2 ; x3 Þ, in the sense of the theory of distributions with respect to
x1 ; moreover, rn11 ðx1 ; x2 ; x3 Þ is even with respect to x2 and x3 . Analogically,
rn22 ðx1 ; x2 ; x3 Þ and rn33 ðx1 ; x2 ; x3 Þ are distributions in x3 and x1 , the convergence
being uniform with respect to x2 , or in x1 and x2 , the convergence being uniform
with respect to x3 , respectively.
Similarly, one may show that the function rn23 ðx1 ; x2 ; x3 Þ, considered as a dis-
tribution in x3 ; x1 or x1 ; x2 , respectively, converges uniformly with respect to x2 or
x3 , respectively, to the function r023 ðx1 ; x2 ; x3 Þ, in the sense of the theory of dis-
tributions; one can state the same thing for the functions rn31 ðx1 ; x2 ; x3 Þ and
rn12 ðx1 ; x2 ; x3 Þ, considered to be distributions too. Moreover, the corresponding
boundary conditions (10.6) are satisfied.
Passing to the limit for the normal stresses given by (10.15), (10.16), one
obtains the relations
Z 1Z 1( Z
2 1 a21  2
c1 C1 ða2 ; a3 Þ: þ a C2 ða3 ; a1 Þ
1 1 p 1 a4 2
Z
2
 2m 1 1  2
þ a3 C3 ða1 ; a2 Þ da1 þ a C2 ða3 ; a1 Þ
p 1 a2 3
)

þ a22 C3 ða1 ; a2 Þ da1 cos a2 x2 cos a3 x3 da2 da3
Z 1 Z 1
¼ a1 ða2 ; a3 Þ cos a2 x2 cos a3 x3 da2 da3 ; . . . ð10:19Þ
1 1
436 10 Elastic Eighth-Space. Elastic Quarter-Space

Taking into account the isomorphism of the distributions space and that of the
Fourier transforms, one obtains the equations
Z
2 1 a21  2 
c1 C1 ða2 ; a3 Þ þ 4
a2 C2 ða3 ; a1 Þ þ a23 C3 ða1 ; a2 Þ da1
p 1 a
Z
2m 1 1  2 
þ 2
a3 C2 ða3 ; a1 Þ þ a23 C3 ða1 ; a2 Þ da1 ¼ a1 ða2 ; a3 Þ; . . . ð10:190 Þ
p 1 a
If there exists a system of even bounded functions C1 ða2 ; a3 Þ, C2 ða3 ; a1 Þ,
C3 ða1 ; a2 Þ which verifies the system of integral equations (10.190 ), then the state of
stress in the elastic eighth-space is given by the formulae (10.12); as it can be seen,
the boundary conditions are verified in the sense of the uniform convergence of
distributions.
We remark that the system (10.190 ), considered as homogeneous, is indepen-
dent on the loading of the elastic eighth-space, this one appearing only in the right-
hand member of the equations.
In what follows we deal only with distributions, i.e., the case in which
Ci ; i ¼ 1; 2; 3, are even, bounded, locally integrable functions. As in the cor-
responding plane case [5], one can show that the system of equations (10.190 ) has a
unique solution in the class of distributions.

10.1.2.2 System of Integral Equations

Let be the kernel

a22 a23 a21


K ð a1 ; a2 ; a3 Þ ¼ þ m ¼ K ða1 ; a3 ; a2 Þ: ð10:20Þ
a4 a2
Thus, the system of integral equations (10.190 ) may be written in the form
Z
2 1
c1 C1 ða2 ; a3 Þ þ Kða3 ; a1 ; a2 ÞC2 ða3 ; a1 Þda1
p 1
Z 1
2
þ Kða2 ; a3 ; a1 ÞC3 ða1 ; a2 Þda1 ¼ a1 ða2 ; a3 Þ; . . .; ð10:21Þ
p 1
as it can be seen, the system has certain properties of symmetry, which suggest a
simplification of the computation, based on properties of mathematical and
mechanical order.
Let firstly be the case of a loading with an axial symmetry of third order with
respect to the axis x1 ¼ x2 ¼ x3 (equally inclined towards the three co-ordinate
axes); this symmetry is characterized by the relations
a1 ða2 ; a3 Þ ¼ a2 ða3 ; a1 Þ ¼ a3 ða1 ; a2 Þ: ð10:22Þ
Thus, the last two equations (10.21) become
10.1 Elastic Eighth-Space 437

Z
2 1
c1 C2 ða2 ; a3 Þ þ Kða3 ; a1 ; a2 ÞC3 ða3 ; a1 Þda1
p 1
Z 1
2
þ Kða2 ; a3 ; a1 ÞC1 ða1 ; a2 Þda1 ¼ a2 ða2 ; a3 Þ;
p 1
Z ð10:210 Þ
2 1
c1 C3 ða2 ; a3 Þ þ Kða3 ; a1 ; a2 ÞC1 ða3 ; a1 Þda1
p 1
Z
2 1
þ Kða2 ; a3 ; a1 ÞC2 ða1 ; a2 Þda1 ¼ a3 ða2 ; a3 Þ:
p 1
Subtracting the Eqs. (10.21), written in the form (10.210 ), two by two, and taking
into account (10.22), one obtains the system

c1 ½C2 ða2 ; a3 Þ  C3 ða2 ; a3 Þ


Z
2 1
þ Kða3 ; a1 ; a2 Þ½C3 ða3 ; a1 Þ  C1 ða3 ; a1 Þda1
p 1
Z 1
2
þ Kða2 ; a3 ; a1 Þ½C1 ða1 ; a2 Þ  C2 ða1 ; a2 Þda1 ¼ 0; . . . ð10:23Þ
p 1
Taking into account the uniqueness of the system (10.21), it results that the
homogeneous system (10.23) admits only trivial solutions. Hence,
C1 ða2 ; a3 Þ ¼ C2 ða3 ; a1 Þ ¼ C3 ða1 ; a2 Þ; ð10:230 Þ
whence the system (10.21) may be substitute by the equation
Z
2 1
c1 C1 ða2 ; a3 Þ þ Kða3 ; a1 ; a2 ÞC1 ða3 ; a1 Þda1
p 1
Z 1
2
þ Kða2 ; a3 ; a1 ÞC1 ða1 ; a2 Þda1 ¼ a1 ða2 ; a3 Þ: ð10:24Þ
p 1
Let be now a case of loading symmetric with respect to the bisecting plane
x1 ¼ x2 ; this property is characterized by the relations
a1 ða2 ; a3 Þ ¼ a2 ða3 ; a1 Þ; a3 ða1 ; a2 Þ ¼ a3 ða2 ; a1 Þ: ð10:25Þ
By subtracting the Eqs. (10.21) between them and taking into account
(10.25), after a convenient change of variables, one obtains the system
438 10 Elastic Eighth-Space. Elastic Quarter-Space

Z
2 1
c1 ½C1 ða2 ; a3 Þ  C2 ða3 ; a2 Þ þ Kða3 ; a1 ; a2 Þ½C2 ða3 ; a1 Þ
p 1
Z 1
2
 C1 ða1 ; a3 Þda1 þ Kða2 ; a3 ; a1 Þ½C3 ða1 ; a2 Þ  C3 ða2 ; a1 Þda1 ¼ 0;
p 1
Z ð10:26Þ
2 1
c2 ½C3 ða1 ; a2 Þ  C3 ða2 ; a1 Þ þ Kða2 ; a3 ; a1 Þ½C1 ða2 ; a3 Þ
p 1
Z
2 1
 C2 ða3 ; a2 Þda3 þ Kða1 ; a2 ; a3 Þ½C2 ða3 ; a1 Þ  C1 ða1 ; a3 Þda3 ¼ 0;
p 1

which leads to
C1 ða2 ; a3 Þ ¼ C2 ða3 ; a2 Þ; C3 ða1 ; a2 Þ ¼ C3 ða2 ; a1 Þ: ð10:260 Þ
This result represents the mathematical consequence of certain mechanical
properties; one remarks thus that, for an element of the bisecting plane x1 ¼ x2 one
may write
1
r ¼ ðr11 þ r22 Þ  r12 ; ð10:27Þ
2
pffiffiffi
2 1
rkx3 ¼ ðr23  r31 Þ; r? x3 ¼  ðr11  r22 Þ; ð10:270 Þ
2 2
where the normal stress and the tangential stresses parallel and normal to the Ox3 -
axis, respectively, are put in evidence; the formulae (2.2.1) and (2.2.4) have been
used. For these tangential stresses, one obtains
pffiffiffi Z 1
2 a2 a3
rkx3 ¼  ½ð1  2m  c1 x1 Þec1 x1 C1 ða2 ; a3 Þ sin a2 x2
8 1 c1
ð1  2m  c1 x2 Þec1 x2 C2 ða3 ; a2 Þ sin a2 x1  sin a3 x3 da2 da3
pffiffiffi Z 1 Z 1
2
 a3 c1 ½x1 ec1 x1 C1 ða2 ; a3 Þ cos a2 x2  x2 ec1 x2 C2 ða3 ; a2 Þ cos a2 x1 
8 1 1
sin a3 x3 da2 da3
pffiffiffi Z 1 Z 1
2
 a1 c3 x3 ec3 x3 ½C3 ða1 ; a2 Þ sin a1 x1 cos a2 x2
8 1 1
C3 ða2 ; a1 Þ cos a2 x1 sin a1 x2 da1 da1 ; ð10:28Þ
10.1 Elastic Eighth-Space 439

Z Z ("  
1 1 1 a3 2
r?x3 ¼ c1 ð12mc1 x1 Þ
8 1 1 c1
 2 # " 
a2 c1 x1 a3 2
þ2 c1 x 1 e C1 ða2 ;a3 Þcosa2 x2  ð12mþc1 x2 Þ
c1 c1
 2 # )
a2
þ2 c1 x2 ec1 x2 C2 ða3 ;a2 Þcosa2 x1 cosa3 x3 da2 da3
c1
Z Z "   2 #
1 1 1 a1 2 a2
 c3 ð1c3 x3 Þþ2m ec3 x3 ½C3 ða1 ;a2 Þcosa1 x1 cosa2 x2
8 1 1 c3 c3
C3 ða2 ;a1 Þcosa2 x1 cosa1 x2 da1 da2 ð10:280 Þ

These tangential stresses must vanish for a symmetric loading (one makes
x1 ¼ x2 ), which leads necessarily to the relations (10.260 ). This must constitute, at
the same time, a demonstration of the uniqueness of the solution of the system
(10.26).
In particular, if the bisecting plane x2 ¼ x3 is, as well, a symmetry plane for the
loading, i.e., if one has the relations
a2 ða3 ; a1 Þ ¼ a3 ða1 ; a3 Þ; a1 ða2 ; a3 Þ ¼ a1 ða3 ; a2 Þ ð10:29Þ

too, then the third bisecting plane x3 ¼ x1 must also be a plane of symmetry for the
loading; indeed, there result the relations
a3 ða1 ; a2 Þ ¼ a1 ða1 ; a1 Þ; a2 ða3 ; a1 Þ ¼ a2 ða1 ; a3 Þ; ð10:290 Þ

so that the axis x1 ¼ x2 ¼ x3 is a symmetry axis of third order too. We consider


above a more general case: a case of loading with a symmetry axis of third order
for which the bisecting planes are not planes of symmetry.
For an antisymmetric loading with respect to the bisecting plane x1 ¼ x2 one
has the relations
a1 ða2 ; a3 Þ ¼ a2 ða3 ; a2 Þ; a3 ða1 ; a2 Þ ¼ a3 ða2 ; a1 Þ: ð10:30Þ

Taking into account the relations (10.30), the equations allow to write
440 10 Elastic Eighth-Space. Elastic Quarter-Space

Z
2 1
c1 ½C1 ða2 ; a3 Þ þ C2 ða3 ; a2 Þ þ Kða3 ; a1 ; a2 Þ½C1 ða1 ; a3 Þ
p 1
Z 1
2
þ C2 ða3 ; a1 Þda1 þ Kða2 ; a3 ; a1 Þ½C3 ða1 ; a2 Þ þ C3 ða2 ; a1 Þda1 ¼ 0;
p 1
Z
2 1
c3 ½C3 ða1 ; a2 Þ þ C3 ða2 ; a1 Þ þ Kða2 ; a3 ; a1 Þ½C1 ða2 ; a3 Þ
p 1
Z
2 1
þ C2 ða3 ; a2 Þda3 þ Kða1 ; a2 ; a3 Þ½C1 ða1 ; a3 Þ þ C2 ða3 ; a1 Þda3 ¼ 0;
p 1
ð10:31Þ
which leads to
C1 ða2 ; a3 Þ ¼ C2 ða3 ; a2 Þ; C3 ða1 ; a2 Þ ¼ C3 ða2 ; a1 Þ: ð10:310 Þ

The normal stress (10.27) on the bisecting plane x1 ¼ x2 will be of the form
Z Z ("  
1 1 1
a3 2
r¼ c1 ð1 þ 2m þ a1 x1 Þ
8 1 1 c1
 2 # " 
a2 a3 2
þ2 ec1 x1 C1 ða2 ; a3 Þ cos a2 x2 þ ð1 þ 2m þ c1 x2 Þ
c1 c1
 2 # )
a2 c1 x2
þ2 e C2 ða3 ; a2 Þ cos a2 x1 cos a3 x3 da2 da3
c1
Z 1Z 1
1
 a2 c1 ½x1 ec1 x1 C1 ða2 ; a3 Þ sin a2 x2 þ x2 ec1 x2 C2 ða3 ; a2 Þ sin a2 x1  cos a3 x3 da2 da3
8 1 1
Z Z "   2 #
1 1 1 a1 2 a2
þ c ð1  c3 x3 Þþ2m ec3 x3 ½C3 ða1 ; a2 Þ cos a1 x1 cos a2 x2
8 1 1 3 c3 c3
þ C3 ða2 ; a1 Þ cos a2 x1 cos a1 x2 da1 da2
Z Z
1 1 1 a1 a2
þ ð1  2m  c3 x3 Þec3 x3 ½C3 ða1 ; a2 Þ sin a1 x1 sin a2 x2
8 1 1 c3
þ C3 ða2 ; a1 Þ sin a2 x1 sin a1 x2 da1 da2 :
ð10:3100 Þ

Equating this stress to zero for an antisymmetric loading (one makes x1 ¼ x2 ),


there result with necessity the relations (10.310 ). This constitutes, at the same time,
a proof of uniqueness of the solution of the system (10.31).
The uniqueness of the solution of the systems (10.26) and (10.31) being proved,
the uniqueness of the solution of the system (10.21) is proved too.
In particular, if the bisecting plane x1 ¼ x2 is, as well, a plane of antisymmetry
for the loading, i. e. if we have
a2 ða3 ; a1 Þ ¼ a3 ða1 ; a3 Þ; a1 ða2 ; a3 Þ ¼ a1 ða3 ; a2 Þ; ð10:32Þ
then the third bisecting plane x3 ¼ x1 is a plane of antisymmetry for the loading
too; one obtains also the relations
10.1 Elastic Eighth-Space 441

a3 ða1 ; a2 Þ ¼ a1 ða2 ; a1 Þ; a2 ða3 ; a1 Þ ¼ a2 ða1 ; a3 Þ; ð10:320 Þ

the axis x1 ¼ x2 ¼ x3 being thus a symmetry axis of third order. This case is
contained in the general one considered above.
If the functions a1 ða2 ; a3 Þ; a2 ða3 ; a1 Þ; a3 ða1 ; a2 Þ are even with respect to the
variables a2 ; a3 or a3 ; a1 or a1 ; a2 , respectively, the system (10.21) shows that the
functions C1 ða2 ; a3 Þ; C2 ða3 ; a1 Þ; C3 ða1 ; a2 Þ have the same property. Thus, the
system (10.21) may be written in the form
Z
4 1
c1 C1 ða2 ; a3 Þ þ K ða3 ; a1 ; a2 ÞC2 ða3 ; a1 Þda1
p 0
Z 1
4
þ K ða2 ; a3 ; a1 ÞC3 ða1 ; a2 Þda1 ¼ a1 ða2 ; a3 Þ; . . . ð10:33Þ
p 0
In the case of a symmetry with respect to the axis x1 ¼ x2 ¼ x3 , we are led to
the integral equation
Z
4 1
c 1 C 1 ð a2 ; a3 Þ þ K ða3 ; a1 ; a2 ÞC1 ða3 ; a1 Þda1
p 0
Z
4 1
þ K ða2 ; a3 ; a1 ÞC1 ða1 ; a2 Þda1 ¼ a1 ða2 ; a3 Þ: ð10:34Þ
p 0
If, in this last case, we decompose the loading with respect to the properties of
symmetry or of antisymmetry with respect to the bisecting planes x1 ¼ x2 ;
x2 ¼ x3 ; x3 ¼ x1 , one can replace the Eq. (10.34) by the equations
Z
4 1
c 1 C 1 ð a2 ; a3 Þ  K ða3 ; a1 ; a2 ÞC1 ða1 ; a3 Þda1
p 0
Z
4 1
 K ða2 ; a3 ; a1 ÞC1 ða2 ; a1 Þda1 ¼ a1 ða2 ; a3 Þ; ð10:35Þ
p 0
the upper signs corresponding to the symmetry, while the lower ones correspond to
the antisymmetry.
In general, an arbitrary case of loading may be decomposed in two cases of
loading if one considers the properties of symmetry and of antisymmetry with
respect to the plane x1 ¼ x2 ; we have studied above these two cases of loading.
Using the relations (10.260 ) and (10.310 ), one may replace the system (10.33) by
two subsystems, having each one two integral equations
Z
4 1
c 1 C 1 ð a2 ; a 3 Þ  K ða3 ; a1 ; a2 ÞC1 ða1 ; a3 Þda1
p 0
Z 1
4
þ K ða2 ; a3 ; a1 ÞC3 ða1 ; a2 Þda1 ¼ a1 ða2 ; a3 Þ;
p 0
Z ð10:36Þ
4 1
c 3 C 3 ð a1 ; a 2 Þ þ K ða2 ; a3 ; a1 ÞC1 ða2 ; a3 Þda3
p 0
Z
4 1
 K ða1 ; a2 ; a3 ÞC1 ða1 ; a3 Þda3 ¼ a3 ða1 ; a2 Þ:
p 0
442 10 Elastic Eighth-Space. Elastic Quarter-Space

The system with the upper signs corresponds to the symmetric loading, while
that with lower signs corresponds to the antisymmetric loading. Although one must
integrate two systems of integral equations, this procedure presents the advantage
that any subsystem has only two equations, possessing a form of the same degree
of difficulty as the Eqs. (10.33).
In the case of a total symmetry (symmetry with respect to all bisecting planes)
or of a total antisymmetry (with respect to the same planes), the system (10.36) is
reduced to the Eqs. (10.35).
The Eqs. (10.36) allow to express C3 ða1 ; a2 Þ with respect to C1 ða2 ; a3 Þ in the
form
Z
1 4 1
C3 ða1 ; a2 Þ ¼ a3 ð a1 ; a2 Þ  K ða2 ; a3 ; a1 ÞC1 ða2 ; a3 Þda3
c3 p 0
Z 1
4
 K ða1 ; a2 ; a3 ÞC1 ða1 ; a3 Þda3 : ð10:37Þ
p 0

Eliminating C3 ða1 ; a2 Þ between the Eqs. (10.36), one obtains two integral equa-
tions of the form
Z 1 Z 1
C1 ða2 ; a3 Þ  K1 ða2 ; a3 ; a1 ÞC1 ða1 ; a3 Þda1 þ K2 ða2 ; a3 ; bÞC1 ða2 ; bÞdb
Z 1 Z 10 0
1
 Kða2 ; a3 ; a1 ; bÞC1 ða1 ; bÞda1 db ¼ a1 ða2 ; a3 Þ  Aða2 ; a3 Þ;
0 0 c 1
ð10:38Þ
with the notations
 
4 4 a21 a22 ma23
K1 ða2 ; a3 ; a1 Þ ¼ K ð a3 ; a1 ; a2 Þ ¼ þ 2 ; ð10:39Þ
pc1 pc1 a4 a
" Z Z 1 #
16ma22 2 1 a21 da1 2 da1
K2 ða2 ;a3 ;bÞ¼  2 a
þma2
2
p c1 3 0 c a4 c2 þb2 2 0 c3 a4 c2 þb2
3 3 3
" Z Z 1 #
16b2 2 1 a41 da1 a 2
da 1
 2 a
þma22

1
2
p c1 3 0 c3 a4 c23 þb2 0 c3 a4 c2 þb2
3
( "
8 a23 c21 þb2 ðb2 þa22 Þ a23 b2
c þa2
¼ 2
2 2 2 þ
3
c1 4a22 þa23 þ3b2 ln 1
p c1 a3 b a2 a23 b2 a3
qffiffiffiffiffiffiffiffiffiffiffiffiffiffi #
qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
a22 þb2 þa2
2 2 2 2 2
 a2 þb 4a2 þ3a3 þb ln
b
10.1 Elastic Eighth-Space 443

(
ma2 1  c þa2
þ 2 2 2ð1mÞa22 þa23 ln 1
a3 b c1 a3
qffiffiffiffiffiffiffiffiffiffiffiffiffiffi ))
2ð1mÞa22 þb2 a22 þb2 þa2
 qffiffiffiffiffiffiffiffiffiffiffiffiffiffi ln ð10:390 Þ
b
a22 þb2

16
Kða2 ; a3 ; a1 ; bÞ ¼  ða21 a23
pa4 c1 c3 ðc21 þ b2 Þ
 
þ ma22 a2 Þ ma21 ðc21 þ b2 Þ þ a22 b2 ; ð10:3900 Þ
Z 1 2
1 a1
Aða2 ; a3 Þ ¼  a2 a1 ða2 ; a1 Þda1
pc1 3 0 c3 a4
Z 1
2 1
þma2 a ða ; a Þda1 :
4 1 2 1
ð10:39000 Þ
0 c3 a

We reduced thus the problem to the integration of two integral equations having
ðiÞ
each one only one unknown function. The sequence of functions C1 ða2 ; a3 Þ; i ¼
0; 1; 2; . . . with

ð0Þ 1
C 1 ð a2 ; a3 Þ ¼ a 1 ð a 2 ; a 3 Þ  Að a 2 ; a 3 Þ ð10:40Þ
c1
and
Z 1
ðiþ1Þ 1 ðiÞ
C1 ða 2 ; a 3 Þ ¼ a1 ða2 ; a3 Þ  Aða2 ; a3 Þ  K1 ða2 ; a3 ; a1 ÞC1 ða1 ; a3 Þda1
c1 0
Z 1 Z 1Z 1
ðiÞ ðiÞ
 K2 ða2 ; a3 ; bÞC1 ða2 ; bÞdb  K ða2 ; a3 ; a1 ; bÞC1 ða1 ; bÞda1 db
0 0 0

ð10:400 Þ
may approximate the solution of the equations (10.38).

10.1.2.3 State of Strain and Stress

If a1 ða2 ; a3 Þ; a2 ða3 ; a1 Þ; a3 ða1 ; a2 Þ are even functions with respect to the variables
a2 ; a3 or a3 ; a1 or a1 ; a2 , respectively, then one may write the normal loads (10.5)
in the form
Z 1Z 1
p1 ð x 2 ; x 3 Þ ¼ a1 ða2 ; a3 Þ cos a2 x2 cos a3 x3 da2 da3 ; . . . ð10:41Þ
0 0

By means of the U functions introduced in Sect. A.4.1.1, the state of stress reads
444 10 Elastic Eighth-Space. Elastic Quarter-Space

Z 1 Z 1
r11 ¼ c1 C1 ða2 ; a3 ÞU2 ðc1 x1 Þ cos a2 x2 cos a3 x3 da2 da3
0 0
Z 1Z 1 "   2 #
a1 2 a3
þ c2 C2 ða3 ; a1 Þ U 4 ð c 2 x2 Þ þ m U6 ðc2 x2 Þ cos a3 x3 cos a1 x1 da3 da1
0 0 c2 c2
Z 1Z 1 "  2
a1
þ c3 C3 ða1 ; a2 Þ U 4 ð c 3 x3 Þ
0 0 c3
 2 #
a2
þm U4 ðc3 x3 Þ cos a1 x1 cos a2 x2 da1 da2 ; . . .
c3

ð10:42Þ
Z 1Z 1
a 2 a3
r23 ¼  C1 ða2 ; a3 ÞU46 ðc1 x1 Þ sin a2 x2 sin a3 x3 da2 da3
c1
Z0 1 Z0 1
þ a3 C2 ða3 ; a1 ÞU3 ðc2 x2 Þ sin a3 x3 cos a1 x1 da3 da1
0 0
Z 1 Z 1
þ a2 C3 ða1 ; a2 ÞU3 ðc3 x3 Þ cos a1 x1 sin a2 x2 da1 da2 ; . . . ð10:420 Þ
0 0

The state of displacement may be expressed in the form


Z Z
1 1 1
u1 ¼ u01  x2 x03 þ x3 x02  C1 ða2 ; a3 ÞU16 ðc1 x1 Þ cos a2 x2 cos a3 x3 da2 da3
2l 0 0
Z 1Z 1
1 a1
þ C2 ða3 ; a1 ÞU46 ðc2 x2 Þ cos a3 x3 sin a1 x1 da3 da1
2l 0 c
Z 1 Z0 1 2
1 a1
þ C3 ða1 ; a2 ÞU46 ðc3 x3 Þ sin a1 x1 sin a2 x2 da1 da2 ; . . .; ð10:43Þ
2p 0 0 c3

where the motion of rigid body may be specified by certain conditions of fixity. For
instance, if the displacement and the rigid body rotation of the elastic eighth-space
vanish, then the rigid body rotations vanish too, while the rigid body displacements
are given by
Z Z
0 1m 1 1
u1 ¼ C1 ða2 ; a3 Þ cos a2 x2 cos a3 x3 da2 da3 ; . . . ð10:430 Þ
l 0 0

One obtains remarkable results for the separation plans. Thus, the normal
stresses are given by
10.1 Elastic Eighth-Space 445

Z Z " 
1 1
a2 2
r11 ðx1 ; x2 ; 0Þ ¼ p1 ðx1 ; x2 Þ þ c1 C1 ða2 ; a3 Þ ð1  2m þ c1 x1 Þ
0 0 c1
 2 #
a3
þ2 c1 x1 ec1 x1 cos a2 x2 da2 da3
c1
Z 1Z 1 "  2 #
a3
þ c2 C2 ða3 ; a1 Þ 1  2 ð1  2m  c2 x2 Þec2 x3 cos a1 x1 da3 da1
0 0 c1
Z 1Z 1  2
a2
 ð1  2mÞ c3 C3 ða1 ; a2 Þ cos a1 x1 cos a2 x2 da1 da2 ;
0
Z 1Z 1
0 "c3   #
a2 2
r22 ðx1 ; x2 ; 0Þ ¼ p1 ðx1 ; x2 Þ þ c1 C1 ða2 ; a3 Þ 1  2 ð1  2m
0 0 c1
Z 1Z 1 " 
a3 2
c1 x1 Þec1 x1 cos a2 x2 da2 da3 þ c2 C2 ða3 ; a1 Þ ð1  2m þ a2 x2 Þ
0 0 c1
 2 #
a1
þ2 a2 x2 ea2 x2 cos a1 x1 da3 da1
c2
Z 1Z 1  2
a1
 ð1  2mÞ c3 C3 ða1 ; a2 Þ cos a1 x1 cos a2 x2 da1 da2
0 0 c3
ð10:44Þ
in the separation plane x3 ¼ 0, while the tangential stress reads
Z 1Z 1
r23 ðx1 ; x2 ; 0Þ ¼ x1 a2 c1 C1 ða2 ; a3 Þec1 x1 sin a2 x2 da2 da3
0 Z 0 Z
1 1
þ x2 a1 c2 C2 ða3 ; a1 Þec2 x2 sin a1 x1 da3 da1
0 0Z
1Z 1
a1 a2
 ð1  2mÞ C3 ða1 ; a2 Þ sin a1 x1 sin a2 x2 da1 da2 :
0 0 c3
ð10:440 Þ

What concerns the state of strain, the displacements in the plane x3 ¼ 0 are given by
Z Z
1 1 1
0
u1 ðx1 ; x2 ; 0Þ ¼ u1  C1 ða2 ; a3 Þ½2ð1  mÞ þ c1 x1 ec1 x1 cos a2 x2 da2 da3
2l 0 0
Z Z
1 1 1 a1
þ C2 ða3 ; a1 Þð1  2m  c2 x2 Þec2 x2 sin a1 x1 da3 da1
2l 0 c
Z 0 Z2
1  2m 1 1 a1
þ C3 ða1 ; a2 Þ sin a1 x1 cos a2 x2 da1 da2 ;
2l c3
Z 10 Z 10
1 a2
u2 ðx1 ; x2 ; 0Þ ¼ u02 þ C1 ða2 ; a3 Þð1  2m  c1 x1 Þec1 x1 sin a2 x2 da2 da3
2l 0 0 c1
Z Z
1 1 1
 C2 ða3 ; a1 Þ½2ð1  mÞ þ c2 x2 ec2 x2 cos a1 x1 da3 da1
2l 0 0
Z Z
1  2m 1 1 a2
 C3 ða1 ; a2 Þ cos a1 x1 sin a2 x2 da1 da2 ; ð10:45Þ
2l 0 0 c3
446 10 Elastic Eighth-Space. Elastic Quarter-Space

while the displacement normal to this plane reads


Z Z
0 1m 1 1
u3 ðx1 ; x2 ; 0Þ ¼ u3  C3 ða1 ; a2 Þ cos a1 x1 cos a2 x2 da1 da2 : ð10:450 Þ
2l 0 0

The state of strain and stress is thus completely specified.


We also remark that the unknown functions C1 ða1 ; a2 Þ; C2 ða3 ; a1 Þ; C3 ða1 ; a2 Þ
are in direct proportion with the Fourier coefficients of the displacements normal to
the separation planes u1 ð0; x2 ; x3 Þ; u2 ðx1 ; 0; x3 Þ; u3 ðx1 ; x2 ; 0Þ (except the rigid body
displacements u01 ; u02 ; u03 ), which gives them a certain physical significance. On the
other hand, this allows to solve the mixed boundary value problem, in which on
the separation planes are given the normal displacements, the tangential stresses
vanishing.

10.1.3 Action of a Local Tangential Load

We consider now the elastic eighth-space xi  0; i ¼ 1; 2; 3, acted upon by the


tangential stresses
Z Z
1 1 1 0
t12 ðx2 ; x3 Þ ¼ a ða2 ; a3 Þ cos a2 x2 sin a3 x3 da2 da3 ; . . .;
4 1 1 1
Z 1Z 1 ð10:46Þ
1
t13 ðx2 ; x3 Þ ¼ a001 ða2 ; a3 Þ sin a2 x2 cos a3 x3 da2 da3 ; . . .
4 1 1
on the faces xi ¼ 0; i ¼ 1; 2; 3, respectively. One uses double Fourier integrals,
which may represent distributions too, such chosen that the representation be
antisymmetric with respect to all the co-ordinate planes.

10.1.3.1 Stress Functions. Boundary Conditions

We put the boundary conditions


x1 ¼ 0 : r11 ¼ 0; r12 ¼ t12 ðx1 ; x2 Þ; r13 ¼ t13 ðx2 ; x3 Þ; . . .; ð10:47Þ

as well, the state of stress must vanish for xi ! 1; i ¼ 1; 2; 3.


Proceeding as at the previous subsection, we choose stress functions of the form
Z Z
1 1 1 0 
F11 ¼ A1 ða2 ; a3 Þ þ c1 x1 A02 ða2 ; a3 Þ ec1 x1 sin a2 x2 sin a3 x3 da2 da3
4 1 1
Z Z
1 1 1 0 
þ B ða3 ; a1 Þ þ c2 x2 B02 ða3 ; a1 Þ ec2 x2 sin a3 x3 sin a1 x1 da3 da1
4 1 1 1
Z Z
1 1 1 0 
þ C1 ða1 ; a2 Þ þ c3 x3 C20 ða1 ; a2 Þ ec3 x3 sin a1 x1 sin a2 x2 da‘ da2 ; . . .;
4 1 1
ð10:48Þ
10.1 Elastic Eighth-Space 447

with the notations (10.70 ). The stress functions in two variables are taken equal to
zero. The unknown parameters are integrable functions, which must be specified
by boundary conditions; the conditions at infinite are automatically verified,
because we deal with Fourier transforms.
Calculating the normal and tangential stresses, as in the preceding subsection,
and putting the boundary conditions for the normal stresses, we get the relations


 1 þ 2 ð1 þ mÞA0  A00 þ A000 ¼ 0; . . .;
A ð10:49Þ
2 2 2

with the notations (10.9).


Eliminating A 1 ða2 ; a3 Þ; B  1 ða1 ; a2 Þ by means of the relations (10.49)
 1 ða3 ; a1 Þ; C
and introducing the notations

1 c41 
C10 ða2 ; a3 Þ ¼  2
A2 ;
ð 2 þ m Þ a2 a3
1 c42 
C20 ða3 ; a1 Þ ¼  B2 ; ð10:50Þ
ð2 þ mÞ2 a3 a1
1 c43 
C30 ða1 ; a2 Þ ¼  2
C2 ;
ð 2 þ m Þ a1 a2

1 c21
2 00
C100 ða2 ; a3 Þ ¼  a A  a22 A000
2 ;
2 þ m a2 a3 3 2
1 c22
2 000
C200 ða3 ; a1 Þ ¼  a1 B2  a23 B02 ; ð10:500 Þ
2 þ m a3 a1
1 c23
2 0
C300 ða1 ; a2 Þ ¼  a2 C2  a21 C200 ;
2 þ m a1 a2
the components of the stress tensor read
Z Z
1 1 1 a2 a3 0
r11 ¼ C ða2 ; a3 Þc1 x1 ec1 x1 sin a2 x2 sin a3 x3 da2 da3
4 1 1 c21 1
Z Z ( 
1 1 1 a3 a1 a1 2
þ ½2ð1 þ mÞ  c2 x2 C20 ða3 ; a1 Þ
4 1 1 c22 c22
)
 2C200 ða3 ; a1 Þ ec2 x2 sin a3 x3 sin a1 x1 da3 da1

Z Z ( 
1 1 1
a1 a2 a1 2
 2
½2ð1 þ mÞ  c3 x3 C30 ða1 ; a2 Þ
4 1 1 c3 c3
)
þ 2C3000 ða1 ; a2 Þ ec3 x3 sin a1 x1 sin a2 x2 da1 da2 ; . . .; ð10:51Þ
448 10 Elastic Eighth-Space. Elastic Quarter-Space

Z Z ( (  2 )
1 1
1 a2 a3
r23 ¼ C10 ða2 ; a3 Þ 1 ½ 2ð 1 þ m Þ  c 1 x 1 
4 1 1 c21

a22  a23 00
þ C1 ða2 ; a3 Þ ec1 x1 cos a2 x2 cos a3 x3 da2 da3
c21
Z Z ("   #
1 1 1 a1 a3 2 a23  a21 0
þ ð1  c2 x2 Þ þ m C2 ða3 ; a1 Þ
4 1 1 c2 c2 c22
)
þ C200 ða3 ; a1 Þ ec2 x2 sin a3 x3 sin a1 x1 da3 da1

Z Z ("   #
1 1 1
a1 a2 2 a21  a22 0
þ ð1  c3 x3 Þ  m C3 ða1 ; a2 Þ
4 1 1 c1 c3 c23
)
 C300 ða1 ; a2 Þ ec3 x3 sin a1 x1 cos a2 x2 da1 da2 ; . . . ð10:510 Þ

Taking into account (10.13), one can use the representations (10.1300 ), by means
of even functions or the representations
Z
1 1 a1
ec1 x1 ¼ sin a1 x1 da1 ; . . .;
p 1 a2
Z 1 ð10:13000 Þ
c1 x1 2 a1
x1 e ¼ c1 4
sin a1 x1 da1 ; . . .;
p 1 a

by means of odd functions of the variable.


If we put the boundary conditions for the tangential stresses and take into
account (10.13) and (10.13000 ), we invert the order of integration and make con-
siderations analogous to those at the previous subsection, then the parameters
Ci0 ; Ci00 ; i ¼ 1; 2; 3, are given by the system of equations

ð1 þ mÞa22  ma23 0 1
C1 ða2 ; a3 Þ þ C100 ða2 ; a3 Þ
c31 c1
Z 2 2
1 1 1 a1 a2
2 2
0
þ 2 4  m a3  a1 C2 ða3 ; a1 Þda1
p 1 c22 a2 a
Z 1
1 1 00
 C ða3 ; a1 Þda1
p 1 a2 2
Z 1  
1 1 a21 a22 a23
þ m  2 4 m þ 2 C30 ða1 ; a2 Þda1
p 1 a2 c3 a
Z 1 2 2
1 a1  a2 00 1
þ 2 2
C3 ða1 ; a2 Þda1 ¼ a01 ða2 ; a3 Þ; . . . ð10:52Þ
p 1 c3 a a3
10.1 Elastic Eighth-Space 449

ð1 þ mÞa23  ma22 0 1
3
C1 ða2 ; a3 Þ  C100 ða2 ; a3 Þ
c1 c1
Z 1   Z
1 1 2 2
a1 a3 a22 0 1 1 a23  a21 00
þ m  2 m þ C ða
2 3 1 ; a Þda 1 þ C2 ða3 ; a1 Þda1
p 1 a2 c42 a2 p 1 c22 a2
Z 2 2
1 1 1 a1 a3
2 2
0
þ 2 2 þ m a1  a2 C3 ða1 ; a2 Þda1
p 1 a2 c23 a
Z 1
1 1 00 1
þ C ða1 ; a2 Þda1 ¼ a001 ða2 ; a3 Þ; . . . ð10:520 Þ
p 1 a2 3 a2
If there exists a system of bounded symmetric functions C10 ða2 ; a3 Þ;
C100 ða2 ; a3 Þ; . . .;
C300 ða1 ; a2 Þ, which verifies the system of integral equations (10.52),
0
(10.52 ), then the state of stress in the elastic eighth-space is given by (10.51),
(10.510 ); the boundary conditions are verified in the sense of the uniform con-
vergence of the distributions.
We remark that the system (10.52), (10.520 ), considered as a homogeneous one,
is independent on the loading of the elastic eighth-space with tangential loads,
because these ones intervene only in the right-hand member.
In what follows we suppose to be in the class of generalized functions, i.e. in
the case of which C10 ða2 ; a3 Þ; C20 ða2 ; a3 Þ; . . .; C300 ða1 ; a2 Þ are locally integrable
bounded even functions. As in the previous case, the system (10.52), (10.520 ) has a
unique solution in the class of distributions.
If the functions a01 ða2 ; a3 Þ; a001 ða2 ; a3 Þ; . . .; a003 ða1 ; a2 Þ are even with respect to
the variable a2 and odd with respect to the variable a3 or with respect to a3 ; a1 or to
a1 ; a2 , respectively, then the system (10.52), (10.520 ) shows that the functions
C10 ða2 ; a3 Þ; C100 ða2 ; a3 Þ; . . .; C300 ða1 ; a2 Þ are even with respect to the variables a2 ; a3
or a3 ; a1 or a1 ; a2 , respectively.
Summing and subtracting the corresponding equations of the system (10.52),
(10.520 ) and taking into account the above remarks, one obtains the equivalent
system
Z   Z
1 0 4 1 a41 a22 0 4 1 a21 00
C1 ða2 ; a3 Þ þ m þ C ða 3 ; a 1 Þda 1  C ða3 ; a1 Þda1
c1 p 0 c42 a2 a2 2 p 0 c22 a2 2
Z   Z
4 1 a41 a23 0 4 1 a21 00
þ m þ C ða 1 ; a 2 Þda 1 þ C ða1 ; a2 Þda1
p 0 c43 a2 a2 3 p 0 c23 a2 3
1 1
¼ a01 ða2 ; a3 Þ þ a001 ða2 ; a3 Þ; ð10:53Þ
a3 a2
450 10 Elastic Eighth-Space. Elastic Quarter-Space

a22  a23 0 2
ð1 þ 2mÞ 2
C1 ða2 ; a3 Þ þ C100 ða2 ; a3 Þ
c1 k
Z 1 2 2
4

1 a1 a 2 2 2
4 0
þ a 1 þ 2a 3 ma 3 C2 ða3 ; a1 Þda1
p 0 c42 a2 a2
Z Z
4 1 a23 00 4 1 1 a21 a23
2
 C ða3 ; a1 Þda1  a1 þ 2a2 ma2 C30 ða1 ; a2 Þda1
2 4
p 0 c22 a2 2 p 0 c43 a2 a2
Z
4 1 a22 00
 C ða1 ; a2 Þda1
p 0 c23 a2 3
1 1
¼ a01 ða2 ; a3 Þ  a001 ða2 ; a3 Þ; . . . ð10:530 Þ
a3 a2

10.1.3.2 System of Integral Equations

Let be the kernels


 
a41 a22
K1 ða1 ; a2 ; a3 Þ ¼ m þ ;
c42 a2 a2
ð10:54Þ
a41 a21 a22
2 2
4 a21
K2 ða1 ; a2 ; a3 Þ ¼ 4 2 a 1 þ 2a 3  ma 3 ; K 3 ða 1 ; a 2 ; a 3 Þ ¼ ;
c2 a a2 c22 a2

the system of integral equations (10.53), (10.530 ) becomes


Z Z
1 0 4 1 4 1
C1 ða2 ;a3 Þ þ K1 ða1 ;a2 ;a3 ÞC20 ða3 ;a1 Þda1  K3 ða1 ;a2 ;a3 ÞC200 ða3 ;a1 Þda1
c1 p 0 p 0
Z Z
4 1 0 4 1
þ K1 ða1 ;a3 ;a2 ÞC3 ða1 ;a2 Þda1 þ K3 ða1 ;a3 ;a2 ÞC300 ða1 ;a2 Þda1
p 0 p 0
1 1
¼ a01 ða2 ;a3 Þ þ a001 ða2 ;a3 Þ;...; ð10:55Þ
a3 a2
Z
a22  a23 0 2 00 4 1
ð1 þ 2mÞ C 1 ð a2 ; a3 Þ þ C 1 ð a2 ; a3 Þ þ K2 ða1 ; a2 ; a3 ÞC20 ða3 ; a1 Þda1
c31 c1 p 0
Z Z
4 1 00 4 1
 K2 ða3 ; a2 ; a1 ÞC2 ða3 ; a1 Þda1  K2 ða1 ; a3 ; a2 ÞC30 ða1 ; a2 Þda1
p 0 p 0
Z
4 1 1 1
 K3 ða2 ; a3 ; a1 ÞC300 ða1 ; a2 Þda1 ¼ a01 ða2 ; a3 Þ  a001 ða2 ; a3 Þ; . . .
p 0 a3 a2
ð10:550 Þ
We remark that this system has some properties of symmetry, which suggest
various simplifications of computation, based on some mathematical and
mechanical properties.
10.1 Elastic Eighth-Space 451

Let be, e.g., the case of a loading with an axial symmetry of third order with
respect to the x1 ¼ x2 ¼ x3 axis; this symmetry is characterized by the relations

a01 ða2 ; a3 Þ ¼ a02 ða2 ; a3 Þ ¼ a03 ða2 ; a3 Þ;


ð10:56Þ
a001 ða2 ; a3 Þ ¼ a002 ða2 ; a3 Þ ¼ a003 ða2 ; a3 Þ:

By considerations analogous to those made in case of a normal load, one obtains


the relations

C10 ða2 ; a3 Þ ¼ C20 ða2 ; a3 Þ ¼ C30 ða2 ; a3 Þ;


ð10:560 Þ
C100 ða2 ; a3 Þ ¼ C200 ða2 ; a3 Þ ¼ C300 ða2 ; a3 Þ:
If we decompose the loading after the properties of symmetry or antisymmetry
with respect to the bisecting planes x1 ¼ x2 ; x2 ¼ x3 ; x3 ¼ x1 , we can replace the
system (10.55), (10.550 ) by the equations
Z Z
1 0 4 1 0 4 1
C ða2 ;a3 Þ  K1 ða1 ;a2 ;a3 ÞC1 ða1 ;a3 Þda1  K3 ða1 ;a2 ;a3 ÞC100 ða1 ;a3 Þda1
c1 1 p 0 p 0
Z Z
4 1 4 1
 K1 ða1 ;a3 ;a2 ÞC10 ða2 ;a1 Þda1  K3 ða1 ;a3 ;a2 ÞC100 ða2 ;a1 Þda1
p 0 p 0
1 1
¼ a01 ða2 ;a3 Þ  a01 ða3 ;a2 Þ; ð10:57Þ
a3 a2
Z
a2  a2 2 4 1
ð1 þ 2mÞ 2 3 3 C10 ða2 ; a3 Þ þ C100 ða2 ; a3 Þ  K2 ða1 ; a2 ; a3 ÞC10 ða1 ; a3 Þda1
c1 c1 p 0
Z Z
4 1 4 1
 K3 ða3 ; a2 ; a1 ÞC100 ða1 ; a3 Þda1  K2 ða1 ; a3 ; a2 ÞC10 ða2 ; a1 Þda1
p 0 p 0
Z
4 1 1
 K3 ða2 ; a3 ; a1 ÞC100 ða2 ; a1 Þda1 ¼ a01 ða2 ; a3 Þ  a01 ða3 ; a2 Þ: ð10:570 Þ
p 0 a3
We consider now properties of symmetry or antisymmetry with respect to the
bisecting plane x1 ¼ x2 , which are characterized by

a01 ða2 ; a3 Þ ¼ a002 ða3 ; a2 Þ; a001 ða2 ; a3 Þ ¼ a02 ða3 ; a2 Þ;


ð10:58Þ
a03 ða1 ; a2 Þ ¼ a003 ða2 ; a1 Þ;

we get thus the relations

C10 ða2 ; a3 Þ ¼ C20 ða3 ; a2 Þ; C100 ða2 ; a3 Þ ¼ C200 ða3 ; a2 Þ;


ð10:580 Þ
C30 ða1 ; a2 Þ ¼ C30 ða2 ; a1 Þ; C300 ða1 ; a2 Þ ¼ C300 ða2 ; a1 Þ;

the upper signs corresponding to the symmetric case, while the lower ones cor-
respond to the antisymmetric case. The remarks of mechanical order which can be
made, analogous to those in the case of a normal loading, for an element of the
452 10 Elastic Eighth-Space. Elastic Quarter-Space

bisecting plane x1 ¼ x2 lead to the uniqueness of the solution of the system


(10.55), (10.550 ). The respective system may be replaced, in this case, by two
subsystems, having each one four integral equations
Z Z
1 0 4 1 0 4 1
C ða2 ;a3 Þ  K1 ða1 ;a2 ;a3 ÞC1 ða1 ;a3 Þda1  K3 ða1 ;a2 ;a3 ÞC100 ða1 ;a3 Þda1
c1 1 p 0 p 0
Z Z
4 1 4 1
þ K1 ða1 ;a3 ;a2 ÞC30 ða1 ;a2 Þda1 þ K3 ða1 ;a3 ;a2 ÞC300 ða1 ;a2 Þda1
p 0 p 0
1 1
¼ a01 ða2 ;a3 Þ þ a001 ða2 ;a3 Þ; ð10:59Þ
a3 a
Z 12 Z 1
1 0 4 4
C ða1 ;a2 Þ þ K1 ða3 ;a1 ;a2 ÞC10 ða2 ;a3 Þda3  K3 ða3 ;a1 ;a2 ÞC100 ða2 ;a3 Þda3
c3 3 p 0 p 0
Z Z
4 1 4 1
 K1 ða3 ;a2 ;a1 ÞC10 ða1 ;a3 Þda3  K3 ða3 ;a2 ;a1 ÞC100 ða1 ;a3 Þda3
p 0 p 0
1 1
¼ a03 ða1 ;a2 Þ  a003 ða1 ;a2 Þ;
a2 a1
Z
a2  a2 2 4 1
ð1 þ 2mÞ 2 3 3 C10 ða2 ; a3 Þ þ C100 ða2 ; a3 Þ  K2 a1 ; a2 ; a3 C10 ða1 ; a3 Þda1
c1 c1 p 0
Z Z
4 1 4 1
 K3 ða3 ; a2 ; a1 ÞC100 ða1 ; a3 Þda1  K2 ða1 ; a3 ; a2 ÞC10 ða1 ; a2 Þda1
p 0 p 0
Z
4 1 1 1
 K3 ða2 ; a1 ; a3 ÞC100 ða1 ; a2 Þda1 ¼ a01 ða2 ; a3 Þ  a001 ða2 ; a3 Þ; ð10:590 Þ
p 0 a3 a2
Z
a2  a2 2 4 1
ð1 þ 2mÞ 1 3 2 C30 ða1 ; a2 Þ þ C300 ða1 ; a2 Þ þ K2 ða3 ; a1 ; a2 ÞC10 ða2 ; a3 Þda3
c3 c3 p 0
Z Z
4 1 4 1
 K3 ða2 ; a1 ; a3 ÞC100 ða2 ; a3 Þda3  K2 ða3 ; a2 ; a1 ÞC10 ða1 ; a3 Þda3
p 0 p 0
Z
4 1 1 1
 K3 ða1 ; a2 ; a3 ÞC100 ða1 ; a3 Þda3 ¼ a03 ða1 ; a2 Þ  a003 ða1 ; a2 Þ
p 0 a2 a1
The system with upper signs corresponds to the symmetric loading, while that
with lower signs corresponds to the antisymmetric one. Although we have thus to
integrate two systems of integral equations, this procedure has the advantage that
each system has only four equations of a form which is not more complicated that
the form of the Eqs. (10.55), (10.550 ).
The Eqs. (10.59), (10.590 ) allow to express C30 ða1 ; a2 Þ and C300 ða1 ; a2 Þ with
regard to C10 ða2 ; a3 Þ; C100 ða2 ; a3 Þ, reducing thus the general problem to the inte-
gration of only two systems of two equations each one. These equations may be
integrated by successive approximations, as in the case of a normal loading.
10.1 Elastic Eighth-Space 453

10.1.3.3 State of Strain and Stress

Using the notations in Sect. A.4.1.1, we may express the state of stress in the form
Z 1Z 1
a2 a3 0
r11 ¼ C ða2 ; a3 ÞU3 ðc1 x1 Þ sin a2 x2 sin a3 x3 da2 da3
0 0 c21 1
Z 1Z 1 " 
a1 a3 a1 2 0
 C2 ða3 ; a1 ÞU56 ðc2 x2 Þ
0 0 c22 c2
#
þ C200 ða3 ; a1 ÞU6 ða2 x2 Þ sin a3 x3 sin a1 x1 da3 da1

Z Z " 
1 1
a1 a2 a1 2 0
 C3 ða1 ; a2 ÞU56 ðc3 x3 Þ
0 0 c23 c3
#
 C300 ða1 ; a2 ÞU6 ða3 x3 Þ sin a1 x1 sin a2 x2 da1 da2 ; ð10:60Þ

Z Z ( 
1 1
a2 a3 2 0 1
r23 ¼ C3 ða2 ; a3 ÞU56 ðc1 x1 Þ þ mC10 ða2 ; a3 Þ
0 0 c21 2
)
a2  a2
þ 2 2 3 C100 ða2 ; a3 Þ U6 ðc1 x1 Þ cos a2 x2 cos a3 x3 da2 da3
c1
Z 1 Z 1 (  2
a1 a3 0 1 a23  a21 0
þ C2 ða3 ; a1 ÞU4 ðc2 x2 Þ þ m C2 ða3 ; a1 Þ
0 0 c2 c2 2 c22
)
þ C200 ða3 ; a1 Þ U6 ðc2 x2 Þ cos a3 x3 sin a1 x1 da3 da1

Z Z ( 
1 1
a1 a2 2 0 1 a21  a22 0
þ C3 ða1 ; a2 ÞU4 ðc3 x3 Þ  m C3 ða1 ; a2 Þ
0 0 c3 c3 2 c23
)
þ C300 ða1 ; a2 Þ U6 ðc3 x3 Þ sin a1 x1 cos a2 x2 da1 da2 ; . . .; ð10:600 Þ

the state of displacement being given by


454 10 Elastic Eighth-Space. Elastic Quarter-Space

Z Z
1 1 1 a2 a3 0
u1 ¼u01 x2 x03 þx3 x02 þ C ða2 ;a3 ÞU26 ðc1 x1 Þsina2 x2 sina3 x3 da2 da3
2l 0 0 c31 1
Z Z (  2
1 1 1 a3 a1 
þ C20 ða3 ;a1 ÞU56 ðc2 x2 Þþ mC20 ða3 ;a1 Þ
2l 0 0 c22 c2
)
00

þC2 ða3 ;a1 Þ U6 ðc2 x2 Þ sina3 x3 cosa1 x1 da3 da1

Z ( 
1Z 1
1 a2 a1 2 0 
þ C3 ða1 ;a2 ÞU56 ðc3 x3 Þ þ mC30 ða1 ;a2 Þ
2l 0 0 c23 c3
)

C300 ða1 ;a2 Þ U6 ðc3 x3 Þ cosa1 x1 sina2 x2 da1 da2 ;...; ð10:61Þ

where the motion of rigid body must be specified by certain conditions of fixity.
For instance, if the displacement and the rigid body rotation of the elastic eighth-
space vanish, then the rigid body displacements vanish too, while the rigid body
local rotations are given by
Z Z
1 1 1 a22  a23  0 
x01 ¼  2
mC1 ða2 ; a3 Þ  C100 ða2 ; a3 Þ da2 da3 ; . . . ð10:610 Þ
2l 0 0 c 1

One obtains
Z Z "  2
1 1
a 1 a2 a1
r11 ðx1 ; x2 ; 0Þ ¼ 2 ð1 þ m Þ C30 ða1 ; a2 Þ
0 0 c23 c3
#
þ C300 ða1 ; a2 Þ sin a1 x1 sin a2 x2 da1 da2 ;

Z Z "  2 ð10:62Þ
1 1
a 1 a2 a2
r22 ðx1 ; x2 ; 0Þ ¼ 2 2
ð1 þ m Þ C30 ða1 ; a2 Þ
0 0 c3 c3
#
 C300 ða1 ; a2 Þ sin a1 x1 sin a2 x2 da1 da2 ;

for the normal stresses in the plane x3 ¼ 0 (one of the separation planes), while the
corresponding tangential stress reads
Z Z ("   #
1 1 1 a1 a2 2 0
r12 ðx1 ; x2 ; 0Þ ¼ m  2ð 1 þ m Þ C1 ða1 ; a2 Þ
m 0 0 c23
)
a21  a22 00
þ C3 ða1 ; a2 Þ cos a1 x1 cos a2 x2 da1 da2 : ð10:620 Þ
c23
10.1 Elastic Eighth-Space 455

What concerns the state of strain, the displacements in the plane x3 ¼ 0 are
given by
u1 ðx1 ; x2 ; 0Þ ¼  x2 x03
Z Z
1 1 1 a2 ma22  a21 0 00
þ C3 ða 1 ; a2 Þ  2C 3 ð a1 ; a 2 Þ cos a1 x1 sin a2 x2 da1 da2 ;
l 0 0 c23 c23
u2 ðx1 ; x2 ; 0Þ ¼  x1 x03
Z Z
1 1 1 a1 ma21  a22 0 00
þ C3 ða 1 ; a2 Þ þ 2C 3 ð a1 ; a 2 Þ sin a1 x1 cos a2 x2 da1 da2 ;
l 0 0 c23 c23
ð10:63Þ
while the displacement normal to this plane may be written in the form
Z Z
1  2m 1 1 a1 a2 0
u3 ðx1 ; x2 ; 0Þ ¼  x1 x02 þ x2 x01 ¼ C ða1 ; a2 Þ sin a1 x1 sin a2 x2 da1 da2
2l 0 0 c23 3
Z Z ( "  2  2 #
1 1 1 a2 a2 a3
þ 2m  ð2  c1 x1 Þ C10 ða2 ; a3 Þ
2l 1 1 c21 c1 c1
) Z Z ("  
00 c1 x1 1 1 1 a3 a1 2
þ 2C1 ða2 ; a3 Þ e sin a2 x2 da2 da3 þ 2m
2l 1 1 c22 c2
 2 # )
a3
 ð2  c2 x2 Þ C20 ða3 ; a2 Þ  2C200 ða3 ; a1 Þ ec2 x2 sin a1 x1 da3 da1 :
c2
ð10:630 Þ

Thus, the state of strain and stress is completely specified.


We notice that one may express the unknown functions C10 ða2 ; a3 Þ;
C1 ða2 ; a3 Þ; ::; C300 ða1 ; a2 Þ by means of the displacements tangent to the separation
00

planes (neglecting the rigid body rotation x03 ). One may thus solve the mixed
boundary value problem in which on the separation planes are given the tangential
displacements, the normal stresses vanishing.

10.1.4 Particular Cases. Application

We consider two particular cases which are not contained in the above general
study; we deal than with an interesting application.

10.1.4.1 Particular Cases

Let thus be the elastic eighth-space acted upon on x3 ¼ 0 by a normal uniformly


distributed load of compression p; the state of stress is given by
456 10 Elastic Eighth-Space. Elastic Quarter-Space

r11 ¼ r22 ¼ 0; r33 ¼ p; r23 ¼ r31 ¼ r12 ¼ 0; ð10:64Þ

while the state of displacement reads


mp mp p
u1 ¼  x 1 ; u2 ¼  x 2 ; u3 ¼ x3 : ð10:640 Þ
2ð1 þ mÞl 2ð1 þ mÞl 2ð1 þ mÞl
Analogically, one can consider the case of the elastic eighth-space acted upon
on the faces x3 ¼ 0 and x1 ¼ 0, e.g., by a tangential load of intensity t, parallel to
the Ox1 -axis. The state of stress reads

r11 ¼ r22 ¼ r33 ¼ 0; r23 ¼ 0; r31 ¼ r32 ¼ t; ð10:65Þ

while the state of displacements is given by

t t
u1 ¼ x 3 ; u3 ¼ x1 : ð10:650 Þ
2l 2l

10.1.4.2 Application

We consider now the elastic eighth-space acted upon by a normal load symmetric
with respect to the axis x1 ¼ x2 ¼ x3 and with respect to the bisecting planes
x2 ¼ x3 ; x1 ¼ x2 ; x2 ¼ x3
p
p1 ð x 2 ; x 3 Þ ¼  ; . . .; ð10:66Þ
r13

with the notations


qffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r1 ¼ x22 þ x23 ; r2 ¼ x23 þ x21 ; r3 ¼ x21 þ x22 : ð10:660 Þ

This case of loading may approximate, in a certain form, a concentrated load


acting at the vertex of the elastic eighth-space and directed along the line
x1 ¼ x2 ¼ x3 , equally inclined on the three faces of the solid angle.
We have
2 2 2
a1 ða2 ; a3 Þ ¼ pc1 ; a2 ða3 ; a1 Þ ¼ pc2 ; a3 ða1 ; a2 Þ ¼ pc3 ; ð10:67Þ
p p p
taking into account the results given in Sect. A.4.2, one may easily integrate the
system (10.21) or the Eq. (10.34), obtaining
p
C1 ða2 ; a3 Þ ¼ C2 ða3 ; a1 Þ ¼ C3 ða1 ; a2 Þ ¼ : ð10:68Þ
ð1 þ mÞp
It is one of the cases in which the system of integral equations (10.21) can be
exactly integrated.
10.1 Elastic Eighth-Space 457

The state of stress is given by


x 2
p 1
r11 ¼ 3 13 ; . . .; ð10:69Þ
R R
x2 x3
r23 ¼ 3p ; . . .; ð10:690 Þ
R5
where
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
R¼ x21 þ x22 þ x23 ; ð10:70Þ

corresponding to an incompressible state of strain (the sum of the normal stresses


vanishes).
If we put also the conditions of fixity, it results the state of displacement
p xi
ui ¼  ; i = 1,2,3: ð10:6900 Þ
2l R3
In cylindrical co-ordinates, the state of stress is given by
r 2  z 2
p 3 p p
rr3 r3 ¼  3 1  3 ; rhh ¼  3 ; rzz ¼  3 1  3 ; ð10:71Þ
R R R R R
zr3
rhz ¼ rr3 h ¼ 0; rzr3 ¼ 3p ; ð10:710 Þ
R5
while the state of displacement reads
p r3 p z
ur3 ¼  3
; uh ¼ 0; uz ¼  : ð10:7100 Þ
2l R 2l R3
In spherical co-ordinates, one obtains
p p
rR ¼ 2 3
; ru ¼ rh ¼  3 ; ð10:72Þ
R R
ruh ¼ rhR ¼ rRu ¼ 0 ð10:720 Þ

and the state of displacement becomes


p 1
uR ¼  ; uu ¼ uh ¼ 0: ð10:7200 Þ
2l R2
Let be now the normal stress r33 ; it is positive in the interior of the cone C, of
equation

r32 ¼ 2x23 ; ð10:73Þ

and negative in the exterior of this cone; the stresses r33 vanish on the conical
surface. One can make analogous remarks for the stresses r11 and r22 . In partic-
ular, all three normal stresses vanish on the axis x1 ¼ x2 ¼ x3 :
458 10 Elastic Eighth-Space. Elastic Quarter-Space

The maximum normal stresses (positive) are obtained on the Ox3 -axis; for
x3 ¼ x03 one has
rffiffiffi
max 12 3 p p
r33 ¼
ffi 0:372
0 3 ð10:74Þ
25 5 x0 3 x
3 3

on the cylinder 3r32 ¼ 2ðx03 Þ2 .


These stresses are minimum (negative) in the
exterior of the cone C. For x3 ¼ x03 one has
2 p p
rmin
33 ¼  pffiffiffi ffi 0:036 0 3 ð10:740 Þ
25 5 ðx03 Þ3 ðx3 Þ

on the cylinder r3 ¼ 2x03 . Along the Ox3 -axis, the normal stress has a hyperbolical
variation
p
r33 ð0; 0; x3 Þ ¼ 2 : ð10:7400 Þ
x33

If, for x3 ¼ x03 , we draw a diagram r33 vs r3 , then one obtains:

1. for r3 ¼ 0 a maximum
p
rmax
33 ¼ 2 ; ð10:75Þ
ðx03 Þ3

2. for r3 ¼ 0:5x03 a point of inflection


rffiffiffi
4 2 p p
r33 ¼ ffi 1:090 0 3 ; ð10:750 Þ
3 3 ðx03 Þ3 ðx3 Þ
pffiffiffi 0
3. for r3 ¼ 2x3 ffi 1:41x03 the stress r33 vanishes;
4. for r3 ¼ 2x03 a minimum
2 p p
rmin
33
¼ pffiffiffi
0 3
ffi 0 3; ð10:7500 Þ
25 5 ðx3 Þ ðx3 Þ
pffiffiffiffiffiffiffi 0
5. for r3 ¼ 6:5x3 ffi 2:55x03 a point of inflection
2 p p
r33 ¼  pffiffiffiffiffiffiffi
0 3
ffi 0:029 0 3 ; ð10:75000 Þ
25 7:5 ðx3 Þ ðx3 Þ
6. for x3 ! 1 the stress r33 tends asymptotically to zero.
The normal stress rr3 r3 has an analogous variation. One observers that it is
negative in the interior of the cone C0 , of equation

2r32 ¼ x23 ; ð10:730 Þ


10.1 Elastic Eighth-Space 459

and positive in the exterior of it; obviously, it vanishes on it. We remark that the
cone C0 is interior to the cone C. Along the Ox3 -axis ðr3 ¼ 0Þ, the normal stress
has a hyperbolical variation
p
rr3 r3 ¼  : ð10:76Þ
R3
For x3 ¼ x03 one obtains a minimum (negative)
16 pffiffiffi p p
rmin ¼ 5 0 3 ffi 0:286 0 3 ; ð10:760 Þ
r r
3 3 125 ðx3 Þ ðx3 Þ

on the cylinder 2r3 ¼ x03 . The maximum normal stresses are obtained in the
exterior of the cone C 0 ; for x3 ¼ x03 , one obtains
rffiffiffi
max 8 2 p p
rr3 r3 ¼ ffi 0:202 0 3 ð10:7600 Þ
25 5 ðx03 Þ3 ðx3 Þ

on the cylinder 2r32 ¼ 3ðx03 Þ2 .


Drawing a diagram rr3 r3 vs r3 , we remark that, for x3 ¼ x03 , we have:

1. for r3 ¼ 0 a minimum
p
rmin
r3 r3 ¼  ; ð10:77Þ
ðx03 Þ3
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffi
2. for r3 ¼ ð1=2Þ 6  30x03 ffi 0:36x03 a point of inflection
pffiffiffiffiffi
8 30  4 p p
rr3 r3 ¼ pffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffi 0 3 ffi 0:545 0 3 ; ð10:770 Þ
5 ð13  2 30Þ 10  30 ðx3 Þ ðx3 Þ
pffiffiffi
3. for r3 ¼ 2=2x03 ffi 0:71x03 the normal stress vanishes;
pffiffiffiffiffiffiffiffi 0
4. for r3 ¼ 3=2x3 ffi 1:22x03 a maximum
rffiffiffi
max 8 2 p p
rr3 r3 ¼ ffi 0:202 0 3 ; ð10:7700 Þ
25 5 ðx03 Þ3 ðx3 Þ
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffi 0
5. r3 ¼ ð1=2Þ 6 þ 30x3 ffi 1:69x03 a point of inflection
pffiffiffiffiffi
8 4 þ 30 p p
rmin
r3 r3 ¼ pffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffi ffi 0:161 0 3 ; ð10:77000 Þ
5 ð13 þ 2 30Þ 10 þ 30 ðx03 Þ3 ðx3 Þ

6. for r3 ! 1 the stress rr3 r3 tends asymptotically to zero.


The tangential stresses are always positive in the interior of the elastic eighth-
space, eventually vanishing on the separation planes and being equal to zero on the
axis x1 ¼ x2 ¼ x3 .
460 10 Elastic Eighth-Space. Elastic Quarter-Space

Concerning the tangential stress r12 which acts in the plane x3 ¼ x03 we remark
that one obtains a maximum at the intersection of the common tangent of the cones
r1 ¼ 2x1 and r2 ¼ 2x2 with this plane, hence at the point x1 ¼ x2
pffiffiffi
¼ ð 3=3Þx03 ffi 0:58x03 ; it results
rffiffiffi
max 9 3 p p
r12 ¼ 3
ffi 0:279 0 3 : ð10:78Þ
25 5 ðx3 Þ
0
ðx3 Þ
Varying x3 , one obtains a maximum
x1 x2
rmax
12 ¼ 3p : ð10:780 Þ
r15

for x3 ¼ 0; e.g. for x1 ¼ x01 ; x2 ¼ x01 =2 it results


48 pffiffiffi p p
rmax
12 ¼ 5 0 3 ffi 0:859 0 3 : ð10:7800 Þ
125 ðx1 Þ ðx1 Þ

Analogically, one can consider the stress r31 in the plane x3 ¼ x03 ; one observes
that at the point x1 ¼ x03 =2; x2 ¼ 0 one obtains
48 pffiffiffi p p
rmax
31 ¼ 5 0 3 ffi 0:859 0 3 : ð10:79Þ
125 ðx3 Þ ðx3 Þ
A variation along the variable x3 , we have
48 pffiffiffi p p
rmax
31 ¼ 5 0 3 ffi 0:859 0 3 ð10:790 Þ
125 ðx1 Þ ðx1 Þ

for x1 ¼ x01 ; x2 ¼ 0; x3 ¼ x01 =2


rffiffiffi
9 3 p p
rmax
31 ¼ 3
ffi 0:279 0 3 ð10:7900 Þ
25 5 ðx2 Þ
0
ðx2 Þ
pffiffiffi
for x2 ¼ x02 ; x1 ¼ x3 ¼ ð 3=3Þx02 ffi 0:58x02
Drawing a diagram along the r3 -axis for r3r3 (in the plane x3 ¼ x03 ), one has:

1. for r3 ¼ 0 the tangential stress vanishes;


2. for r3 ¼ 0:5x03 a maximum (positive)
48 pffiffiffi p p
rmax
3r3 ¼ 5 0 3 ffi 0:859 0 3 ; ð10:80Þ
125 ðx3 Þ ðx3 Þ

pffiffiffi 0
3. . for r3 ¼ 3=2 x3 ffi 0:78x03 a point of inflection
10.1 Elastic Eighth-Space 461

rffiffiffi
48 3 p p
r3r3 ¼ ffi 0:758 0 3 ; ð10:800 Þ
49 5 ðx03 Þ3 ðx3 Þ
4. for r3 ! 1 the tangential stress tends asymptotically to zero.
The relations (10.72), (10.720 ) of the stresses in spherical co-ordinates allow to
specify the principal directions of the stress tensor. The principal direction which
leads to a positive maximum for rR is given by the radius R; any other direction
normal to the radius R is a principal direction (leading to a minimum ruu ¼ rhh ),
the quadric of Cauchy being a two sheet hyperboloid of rotation. The relations
(10.7800 ) of the displacement show that the displacement vector is directed along
the vector radius R.
We also remark that the vertex of the elastic eighth-space is a singular point. It
is thus useful to eliminate this point, by cutting up an eighth-sphere of radius R0 ,
having the centre at this point; this sphere will be acted upon only by the normal
stresses rRR ¼ rR0 . Projecting these stresses on the Ox3 -axis and summing for the
eighth-sphere, one obtains the resultant
Z
p p
rR0 cos udA0 ¼ ; ð10:81Þ
A0 2 R0

in the negative sense of Ox3 , taking into account that the elementary area is given
by dA0 ¼ R20 sin ududh. As well, the total load on x3 ¼ 0 is given by
Z 1 Z p=2
p p p
RdRdh ¼ ; ð10:810 Þ
R0 0 R3 2 R0

being directed in the positive sense of Ox3 . A relation of global equilibrium is thus
obtained.
Projecting all the stresses which act on the eighth-sphere on the straight-line L
of equation x1 ¼ x2 ¼ x3 and summing, one obtains the resultant
Z pffiffiffi
p 3 p
rR0 cosðR; LÞd A0 ¼ ; ð10:82Þ
A0 2 R0

where one took into consideration


1
cosðR; LÞ ¼ pffiffiffi ½sin uðsin h þ cos hÞ þ cos u: ð10:83Þ
3
Analogous considerations can be made in cylindrical co-ordinates. One can thus
eliminate the vertex and one of the edges, e.g. x1 ¼ x2 ¼ 0, by means of the
quarter cylinder r3 ¼ r30 ; on this surface will act the normal stresses r0r3 r3 and the
tangential stresses r03r3 . Projecting these stresses on the Ox1 -axis and summing,
one obtains the resultant
462 10 Elastic Eighth-Space. Elastic Quarter-Space

Z 1 Z p=2
p
r0r3 r3 r30 sin hdx3 dh ¼ ; ð10:84Þ
0 0 r30

in the negative sense of this axis; as well, the total load on x1 ¼ 0 is given by
Z 1Z 1
p p
dx dx ¼ 0 ;
3 2 3
ð10:840 Þ
0 r 3 r3
0 r3

in the positive sense of the Ox1 -axis. A relation of global equilibrium is thus put in
evidence. The resultant of the load along the bisectrix of the solid angle x1 ¼
x2 ¼ 0 is
pffiffiffi
p 2 p
2 0 ffi 1:41 0 : ð10:85Þ
r3 2 r3

10.2 Elastic Quarter-Space

Throwing to infinite one of the faces of the elastic eighth-space, i.e. the face
x3 ¼ 0, one obtains the elastic quarter-space x1  0; x2  0. We study the action of
normal or tangential loads, particularizing the results obtained in the previous
section.
As well, one can adapt to this domain the results obtained in Sect. 10.1.1.1
concerning the action of a periodic load. We conclude that the problem of the
elastic quarter-space acted upon by an arbitrary load, periodic on two directions,
may be reduced, by means of the principle of superposition of effects, to the study
of the same domain acted upon by a load, local on the semiinfinite direction and
periodic on the infinite one, and to the study of an elastic half-space, analogically
acted upon. This case of loading of an elastic quarter-space cannot be reduced to a
simpler one. We also mention the case of the elastic quarter-space acted upon by a
load, local on the infinite direction and periodic on the infinite one, which may be
reduced to the study of an elastic half-space, analogically loaded. In the following,
we will consider only the case of local loads, using the results given by us for
normal loads [8, 10], for tangential loads [11. 14] and for a certain particular case
of loading [9, 13].

10.2.1 Action of a Local Normal Load

Let be the elastic quarter-space x1  0; x2  0 acted upon by the normal loads


10.2 Elastic Quarter-Space 463

Z 1 Z 1
p1 ðx2 ; x3 Þ ¼ a1 ða2 ; a3 Þ cos a2 x2 cos a3 x3 dx2 dx3 ;
Z0 1 Z0 1 ð10:86Þ
p2 ðx3 ; x1 Þ ¼ a2 ða3 ; a1 Þ cos a3 x3 cos a1 x1 dx3 dx1
0 0

on the faces x1 ¼ 0 and x2 ¼ 0, respectively. One uses double Fourier integrals,


even with respect to x1 ; x2 ; x3 , which may represent generalized functions too. As
well, a1 ða2 ; a3 Þ; a2 ða3 ; a1 Þ are even functions with respect to a2 ; a3 or a3 ; a1 ,
respectively. A symmetry of loading with respect to Ox1 and Ox2 axes is assumed.
The boundary conditions are put in the form

x1 ¼ 0 : r11 ¼ p1 ðx2 ; x3 Þ; r12 ¼ r13 ¼ 0;


ð10:87Þ
x2 ¼ 0 : r22 ¼ p2 ðx3 ; x1 Þ; r23 ¼ r21 ¼ 0;

the state of stress must vanish for xi ! 1; i ¼ 1; 2; 3.


To solve the problem we use the potential functions (10.7), (10.70 ) with
Ci ða1 ; a2 Þ ¼ Ci00 ða1 ; a2 Þ ¼ Ci000 ða1 ; a2 Þ ¼ 0; i ¼ 1; 2. Taking into account
0

(10.11), we have C3 ða1 ; a2 Þ ¼ 0 too. We may thus obtain all the results for the
elastic quarter-space, starting from the results in Sect. 10.1.2.

10.2.1.1 State of Strain and Stress

The state of stress is thus given by


Z 1 Z 1
r11 ¼ c1 C1 ða2 ; a3 ÞU2 ðc1 x1 Þ cos a2 x2 cos a3 x3 da2 da3
0 0
Z 1Z 1 "   2 #
a1 2 a3
þ c2 C2 ða3 ; a1 Þ U4 ðc2 x2 Þþm U6 ðc2 x2 Þ cos a3 x3 cos a1 x1 da3 da1 ;
0 0 c2 c2
Z 1Z 1 "   2 #
a2 2 a3
r22 ¼ c1 C1 ða2 ; a3 Þ U4 ðc1 x1 Þþm U6 ða1 x1 Þ cos a2 x2 cos a3 x3 da2 da3
0 0 c1 c1
Z 1Z 1
þ c2 C2 ða3 ; a1 ÞU2 ðc2 x2 Þ cos a3 x3 cos a1 x1 da3 da1 ; ð10:88Þ
0 0
Z 1Z 1 "  2  2 #
a3 a2
r33 ¼ c1 C1 ða2 ; a3 Þ U4 ðc1 x1 Þþm U6 ðc1 x1 Þ cos a2 x2 cos a3 x3 da2 da3
0 0 c 1 c1
Z 1Z 1 "  2  2 #
a3 a1
þ c2 C2 ða3 ; a1 Þ U4 ða2 x2 Þþm U6 ða2 x2 Þ cos a3 x3 cos a1 x1 da1 da3 ;
0 0 c2 c2
464 10 Elastic Eighth-Space. Elastic Quarter-Space

Z 1 Z 1
a2 a3
r23 ¼  C1 ða2 ; a3 ÞU46 ðc1 x1 Þ sin a2 x2 sin a3 x3 da2 da3
0 0 c1
Z 1 Z 1
þ a3 C2 ða3 ; a1 ÞU3 ðc2 x2 Þ sin a3 x3 cos a1 x1 da3 da1 ;
0 0
Z 1Z 1
r31 ¼ a3 C1 ða2 ; a3 ÞU3 ðc1 x1 Þ cos a2 x2 sin a3 x3 da2 da3
00
Z 1Z 1 ð10:880 Þ
a3 a1
 C2 ða3 ; a1 ÞU46 ðc2 x2 Þ sin a3 x3 sin a1 x1 da3 da1 ;
0 0 c2
Z 1Z 1
r12 ¼ a2 C1 ða2 ; a3 ÞU3 ðc1 x1 Þ sin a2 x2 cos a3 x3 da2 da3
0 0
Z 1Z 1
þ a1 C2 ða3 ; a1 ÞU3 ðc2 x2 Þ cos a3 x3 sin a1 x1 da3 da1 ;
0 0

while the state of displacement is given by


Z Z
1 1 1
u1 ¼ u01  x2 x01 þ x3 x02  C1 ða2 ; a3 ÞU16 ðc1 x1 Þ cos a2 x2 cos a3 x3 da2 da3
2l 0 0
Z Z
1 1 1 a1
þ C2 ða3 ; a1 ÞU46 ðc2 x2 Þ cos a3 x3 sin a1 x1 da3 da1 ;
2l 0 0 c2
Z Z
1 1 1 a2
u2 ¼ u02  x3 x01 þ x1 x03 þ C1 ða2 ; a3 ÞU46 ðc1 x1 Þ sin a2 x2 cos a3 x3 da2 da3
2l 0 0 c1
Z Z
1 1 1
 C2 ða3 ; a1 ÞU16 ðc2 x2 Þ cos a3 x3 sin a1 x1 da3 da1 ;
2l 0 0
Z Z
1 1 1 a3
u3 ¼ u03  x1 x02 þ x2 x01 þ C1 ða2 ; a3 ÞU46 ðc1 x1 Þ cos a2 x2 sin a3 x3 da2 da3
2l 0 0 c1
Z Z
1 1 1 a3
þ C2 ða3 ; a1 ÞU46 ðc2 x2 Þ sin a3 x3 cos a1 x1 da3 da1 :
2l 0 0 c2
ð10:89Þ
The motion of rigid body is specified as in Sect. 10.1.2.3 with u03 ¼ 0.
What concerns the stresses and the displacements corresponding to the sepa-
ration planes, one obtains remarkable results. Thus, the normal stresses in the
plane x1 ¼ 0 are given by
10.2 Elastic Quarter-Space 465

Z 1 Z 1  2
a3
r22 ð0; x2 ; x3 Þ ¼ p1 ðx2 ; x3 Þ  ð1  2mÞ c1 C1 ða2 ; a3 Þ cos a2 x2 cos a3 x3 da2 da3
0 c1
Z Z "0 
1 1
a3 2
þ c2 C2 ða3 ; a1 Þ ð1  2m þ c2 x2 Þ
0 0 c2
 2 #
a1
þ2 c2 x2 ec2 x2 cos a3 x3 da3 da1 ; ð10:90Þ
c2
Z 1Z 1  2
a2
r33 ð0; x2 ; x3 Þ ¼ p1 ðx2 ; x3 Þ  ð1  2mÞ c1 C1 ða2 ; a3 Þ cos a2 x2 cos a3 x3 da2 da3
0 0 c1
Z 1Z 1 "   #
a3 2
þ c2 C2 ða3 ; a1 Þ 1  2 ð1  2m  c2 x2 Þec2 x2 cos a3 x3 da3 da1
0 0 c2

and the tangential stress reads


Z 1Z 1
a2 a3
r23 ð0; x2 ; x3 Þ ¼  ð1  2mÞ C1 ða2 ; a3 Þ sin a2 x2 sin a3 x3 da2 da3
0 0 c1
Z 1Z 1
þ a3 C2 ða3 ; a1 Þc2 x2 ec2 x2 sin a3 x3 da3 da1 : ð10:900 Þ
0 0

The displacements in this plane may be written in the form


Z 1 Z 1
1  2m a2
u2 ð0; x2 ; x3 Þ ¼ u02 þ C1 ða2 ; a3 Þ sin a2 x2 sin a3 x3 da2 da3
2l 0 0 c1
Z 1Z 1
1
 C2 ða3 ; a1 Þ½2ð1  mÞ þ c2 x2 ec2 x2 cos a3 x3 da3 da1 ; ð10:91Þ
2l 0 0
Z Z
1  2m 1 1 a3
u3 ð0; x2 ; x3 Þ ¼ C1 ða2 ; a3 Þ cos a2 x2 sin a3 x3 da2 da3
2l c1
Z 10 Z 10
1 a3
þ C2 ða3 ; a1 Þ(1  2m  c2 x2 )ec2 x2 sin a3 x3 da3 da1 ;
2l 0 0 c2
ðÞ
while the displacement normal to it is given by
Z Z
0 1m 1 1
u1 ð0; x2 ; x3 Þ ¼ u1 þ C1 ða2 ; a3 Þ cos a2 x2 cos a3 x3 da2 da3 : ð10:910 Þ
l 0 0

We also remark that the unknown functions C1 ða2 ; a3 Þ; C2 ða3 ; a1 Þ are in direct
proportion to the Fourier coefficients of the displacements normal to the separation
planes u1 ð0; x2 ; x3 Þ; u2 ðx1 ; 0; x3 Þ, respectively (neglecting the displacements of
rigid body u01 ; u02 ), which confers them a certain physical significance. On the other
hand, this allows to solve the mixed fundamental problem in which, on the sep-
aration planes are given the normal stresses, the tangential stresses vanishing.
466 10 Elastic Eighth-Space. Elastic Quarter-Space

10.2.1.2 System of Integral Equations

By means of the kernel (10.20), the system of integral equations which is verified
by the unknown functions Ci ; i ¼ 1; 2, may be written in the form
Z
4 1
c1 C1 ða2 ; a3 Þ þ Kða3 ; a1 ; a2 ÞC2 ða3 ; a1 Þda1 ¼ a1 ða2 ; a3 Þ;
p 0
Z 1 ð10:92Þ
4
c2 C2 ða3 ; a1 Þ þ Kða3 ; a1 ; a2 ÞC1 ða2 ; a3 Þda2 ¼ a2 ða3 ; a1 Þ:
p 0
We observe that the third integral equation (10.33) does not intervene; as a matter
of fact, it may be considered as verified, because the function a3 ða1 ; a2 Þ may be
considered as arbitrary.
One remarks that the system (10.92) has some properties of symmetry, leading
to a simplification of the computation. In general, an arbitrary loading can be
decomposed in two cases of loading, corresponding to the properties of symmetry
or antisymmetry with respect to the plane x1 ¼ x2 , which are characterized by the
relations
a1 ða2 ; a3 Þ ¼ a2 ða3 ; a2 Þ; ð10:93Þ
one obtains thus
C1 ða2 ; a3 Þ ¼ C2 ða3 ; a2 Þ; ð10:930 Þ

the upper signs corresponding to the symmetric case, while the lower ones to the
antisymmetric one. The system (10.92) may be thus replaced by two integral
equations
Z
4 1
c1 C1 ða2 ; a3 Þ  Kða3 ; a1 ; a2 ÞC1 ða1 ; a3 Þda1 ¼ a1 ða2 ; a3 Þ; ð10:94Þ
p 0
each one of a single unknown.
The Eqs. (10.92) allow to express C2 ða3 ; a1 Þ with regard to C1 ða2 ; a3 Þ in the
form
Z
1 4 1
C2 ða3 ; a1 Þ ¼ a2 ða3 ; a1 Þ  Kða3 ; a1 ; a2 ÞC1 ða2 ; a3 Þda2 : ð10:95Þ
c2 p 0
Eliminating C2 ða3 ; a1 Þ between the integral equations (10.92), we get an inte-
gral equation of the form
Z 1
1
C1 ða2 ; a3 Þ þ K 0 ða2 ; a3 ; bÞC1 ðb; a3 Þdb ¼ a1 ða2 ; a3 Þ þ A0 ða2 ; a3 Þ; ð10:96Þ
0 c 1

where, taking into account (10.390 ), one can write

K 0 ða2 ; a3 ; bÞ ¼ K2 ða3 ; a2 ; bÞ; ð10:97Þ


10.2 Elastic Quarter-Space 467

Z 1 2 Z 1
4 a1 a2 ða3 ; a1 Þ a2 ða3 ; a1 Þ
A0 ða2 ; a3 Þ ¼  a22 4
da 1 þ ma 2
3 2
da 0
1 : ð10:97 Þ
pc1 0 c 2 a 0 c 2 a

Thus, we reduce the problem to the integration of a single integral equation with
ðiÞ
only one unknown function. The sequence of functions C1 ða2 ; a3 Þ; i ¼ 0; 1; 2; . . .,
with
1 ð0Þ
a1 ða2 ; a3 Þ þ A0 ða2 ; a3 Þ;
C1 ¼ ð10:98Þ
c1
Z 1
ðiþ1Þ 1 0 ðiÞ
C1 ða2 ; a3 Þ ¼ a1 ða2 ; a3 Þ þ A ða2 ; a3 Þ  K 0 ða2 ; a3 ; bÞC1 ðb; a3 Þdb;
c1 0
ð10:980 Þ

approximates the solution of the Eq. (10.96).


A case of loading antisymmetric with respect to the axes Ox1 and Ox2 , of the
form
Z 1Z 1
p1 ðx2 ; x3 Þ ¼ a1 ða2 ; a3 Þ cos a2 x2 sin a3 x3 da2 da3 ;
Z0 1 Z0 1 ð10:99Þ
p2 ðx3 ; x1 Þ ¼ a2 ða3 ; a1 Þ sin a3 x3 cos a1 x1 da3 da1
0 0

may be similarly studied.

10.2.2 Action of a Local Tangential Load

Let be the elastic quarter-space x1  0; x2  0, acted upon by the tangential loads


Z 1Z 1
t12 ðx2 ; x3 Þ ¼ a01 ða2 ; a3 Þ cos a2 x2 sin a3 x3 da2 da3 ;
Z0 1 Z0 1 ð10:100Þ
t13 ðx2 ; x3 Þ ¼ a001 ða2 ; a3 Þ sin a2 x2 cos a3 x3 da2 da3 ;
0 0
Z 1Z 1
t23 ðx3 ; x1 Þ ¼ a02 ða3 ; a1 Þ cos a3 x3 sin a1 x1 da3 da1 ;
Z0 1 Z0 1 ð10:1000 Þ
t21 ðx3 ; x1 Þ ¼ a002 ða3 ; a1 Þ sin a3 x3 cos a1 x1 da3 da1
0 0

on the faces x1 ¼ 0; x2 ¼ 0, respectively; an antisymmetry of loading with respect


to the Ox1 and Ox2 axes, respectively, has been considered. The double Fourier
integrals may represent distributions too.
468 10 Elastic Eighth-Space. Elastic Quarter-Space

We put the boundary conditions in the form

x1 ¼0 : r11 ¼ 0; r12 ¼ t12 ðx2 ; x3 Þ; r13 ¼ t13 ðx2 ; x3 Þ;


ð10:101Þ
x2 ¼0 : r22 ¼ 0; r23 ¼ t23 ðx3 ; x1 Þ; r21 ¼ t21 ðx3 ; x1 Þ;

the stresses vanishing for xi ! 1; i ¼ 1; 2; 3.


To solve the problem, we use the stress functions (10.48), with Ci0 ða1 ; a2 Þ;
Ci ða1 ; a2 Þ; Ci000 ða1 ; a2 Þ ¼ 0; i ¼ 1; 2:
00

10.2.2.1 State of Strain and Stress

Using the results obtained in Sect. 10.1.3, we get the state of stress
Z 1Z 1
a2 a3 0
r11 ¼ C ða2 ; a3 ÞU3 ðc1 x1 Þ sin a2 x2 sin a3 x3 da2 da3
0 0 c21 1
Z 1Z 1 " 
a3 a1 a1 2 0
 C2 ða3 ; a1 ÞU56 ðc2 x2 Þ
0 0 c22 c2
#
þ C200 ða3 ; a1 ÞU6 ðc2 x2 Þ sin a3 x3 sin a1 x1 da3 da1 ;

Z Z " 
1 1
a2 a3 a2 2 0
r22 ¼  C1 ða2 ; a3 ÞU56 ðc1 x1 Þ
0 0 c21 c1
#
 C100 ða2 ; a3 ÞU6 ðc1 x1 Þ sin a2 x2 sin a3 x3 da2 da3
Z 1 Z 1 ð10:102Þ
a3 a1 0
þ C ða3 ; a1 ÞU3 ðc2 x2 Þ sin a3 x3 sin a1 x1 da3 da1 ;
0 0 c22 2
Z 1Z 1 " 
a2 a3 a3 2 0
r33 ¼  C1 ða2 ; a3 ÞU56 ðc1 x1 Þ
0 0 c21 c1
#
þ C100 ða2 ; a3 ÞU6 ðc1 x1 Þ sin a2 x2 sin a3 x3 da2 da3

Z Z " 
1 1
a3 a1 a3 2 0
 C2 ða3 ; a1 ÞU56 ðc2 x2 Þ
0 0 c22 c2
#
 C200 ða3 ; a1 ÞU6 ðc2 x2 Þ sin a3 x3 sin a1 x1 da3 da1
10.2 Elastic Quarter-Space 469

Z Z (  "
1 1
a2 a3 2 0 1
r23 ¼ C1 ða2 ; a3 ÞU56 ðc1 x1 Þ þ mC10 ða2 ; a3 Þ
0 0 c21 2
# )
a22  a23 00
þ C2 ða2 ; a3 Þ U6 ðc1 x1 Þ cos a2 x2 cos a3 x3 da2 da3
c21
Z 1 Z 1 (  2 "
a1 a3 1 a2  a2
þ C20 ða3 ; a1 ÞU4 ðc2 x2 Þ þ m 3 2 1 C20 ða3 ; a1 Þ
0 0 c2 c2 2 c2
# )
þ C200 ða3 ; a1 Þ U6 ðc2 x2 Þ cos a3 x3 sin a1 x1 da3 da1 ;

Z Z (  "
1 1
a2 a3 2 0 1 a22  a23 0
r31 ¼ C1 ða2 ; a3 ÞU4 ðc1 x1 Þ  m C1 ða2 ; a3 Þ
0 0 c1 c1 2 c21
# )
þ C100 ða2 ; a3 Þ U6 ðc1 x1 Þ sin a2 x2 cos a3 x3 da2 da3 ð10:1020 Þ

Z Z (  "
1 1
a3 a1 2 0 1
þ C2 ða3 ; a1 ÞU56 ðc2 x2 Þ þ mC20 ða3 ; a1 Þ
0 0 c22 2
# )
a23
 a21 00
þ C2 ða3 ; a1 Þ U6 ðc2 x2 Þ cos a3 x3 cos a1 x1 da3 da1 ;
c22
Z 1 Z 1 ( 2 "
a3 a2 0 1 a22  a23 0
r12 ¼ C1 ða2 ; a3 ÞU4 ðc1 x1 Þ þ m C1 ða2 ; a3 Þ
0 0 c1 c1 2 c21
# )
þ C100 ða2 ; a3 Þ U6 ðc1 x1 Þ cos a2 x2 sin a3 x3 da2 da3

Z Z (  "
1 1
a3 a1 2 0 1 a23  a21 0
þ C2 ða3 ; a1 ÞU4 ðc2 x2 Þ  m C2 ða3 ; a1 Þ
0 0 c2 c2 2 c22
# )
þ C200 ða3 ; a1 Þ U6 ðc2 x2 Þ sin a3 x3 cos a1 x1 da3 da1

and the state of displacement


470 10 Elastic Eighth-Space. Elastic Quarter-Space

Z 1 Z 1
1 a2 a3 0
u1 ¼ x3 x02 þ C ða2 ; a3 ÞU26 ðc1 x1 Þ sin a2 x2 sin a3 x3 da2 da3
0 2l
0 c31 1
Z Z (  2
1 1 1 a3 a1
þ 2
C20 ða3 ; a1 ÞU56 ðc2 x2 Þ þ ½mC20 ða3 ; a1 Þ
2l 0 0 c 2 c2
)
þ C200 ða3 ; a1 ÞU6 ðc2 x2 Þ sin a3 x3 cos a1 x1 da3 da1 ;

Z ( 
Z
1 1
a2 2 0
1
a3
u2 ¼  x3 x01 þ C1 ða2 ; a3 ÞU56 ðc1 x1 Þ
2l 0 0 c1 c21
)
 0 00

þ mC1 ða2 ; a3 Þ  C1 ða2 ; a3 Þ U6 ðc1 x1 Þ cos a2 x2 sin a3 x3 da2 da3
Z 1 Z 1
a3 a1 0
þ C ða3 ; a1 ÞU26 ðc2 x2 Þ sin a3 x3 sin a1 x1 da3 da1 ; ð10:103Þ
0 0 c32 2
Z Z ( 
0 0 1 1 1 a2 a3 2 0
u3 ¼ x 2 x 1  x 1 x 2 þ 2
C1 ða2 ; a3 ÞU56 ðc1 x1 Þ
2l 0 0 c1 c1
)
 0 00

þ mC1 ða2 ; a3 Þ þ C1 ða2 ; a3 Þ U6 ðc1 x1 Þ sin a2 x2 cos a3 x3 da2 da3

Z Z ( 
1 1 1
a3 2 0
a1
þ C2 ða3 ; a1 ÞU56 ðc2 x2 Þ
2l 0 0 c2c22
)
 0 00

þ mC2 ða3 ; a1 Þ  C2 ða3 ; a1 Þ U6 ðc2 x2 Þ cos a3 x3 sin a1 x1 da3 da1 ;

where the motion of rigid body was specified (only x01 ; x02 6¼ 0).
The normal stresses corresponding to the plane x1 ¼ 0 are
Z 1Z 1 "  2
a2 a3 a2
r11 ð0; x2 ; x3 Þ ¼ 2 2
ð1 þ mÞ C10 ða2 ; a3 Þ
0 0 c1 c1
#
 C100 ða2 ; a3 Þ sin a2 x2 sin a3 x3 da2 da3 ;

Z Z "  2 ð10:104Þ
1 1
a2 a3 a3
r22 ð0; x2 ; x3 Þ ¼ 2 2
ð1 þ mÞ C10 ða2 ; a3 Þ
0 0 c1 c1
#
 C100 ða2 ; a3 Þ sin a2 x2 sin a3 x3 da2 da3 ;
10.2 Elastic Quarter-Space 471

while the tangential stress reads


Z Z ("  #
1 1
a2 a3 2 0
r23 ð0; x2 ; x3 Þ ¼ 1  2ð1 þ mÞ C1 ða2 ; a3 Þ
0 0 c21

a22  a23 00
m C1 ða2 ; a3 Þ cos a2 x2 cos a3 x3 da2 da3 ; ð10:1040 Þ
c21
The displacements in the plane x1 ¼ 0 become
Z Z
0 1 1 1 a3 ma23  a22 0
u2 ð0; x2 ; x3 Þ ¼  x3 x1 þ C1 ða2 ; a3 Þ
2l 0 0 c21 c21

 2C100 ða2 ; a3 Þ cos a2 x2 sin a3 x3 da2 da3 ;
Z Z ð10:105Þ
1 1 1 a2 ma22  a23 0
u3 ð0; x2 ; x3 Þ ¼ x2 x01 þ C1 ða2 ; a3 Þ
2l 0 0 c21 c21

þ 2C100 ða2 ; a3 Þ sin a2 x2 cos a3 x3 da2 da3 ;

while the displacement normal to this plane reads


Z Z
1  2m 1 1 a2 a3 0
u1 ð0; x2 ; x3 Þ ¼ x3 x02 þ C ða2 ; a3 Þ sin a2 x2 sin a3 x3 da2 da3
2l 0 0 c21 1
Z Z ("     #
1 1 1 a3 a3 2 a1 2 
þ 2
2m  ð2  c2 x2 Þ C20 ða3 ; a1 Þ
2l 0 0 c2 c2 c2
)

þ 2C200 ða3 ; a1 Þ ec2 x2 sin a3 x3 da3 da1 : ð10:1050 Þ

We notice also here that the unknown functions C10 ða2 ; a3 Þ; C100 ða2 ; a3 Þ;
C20 ða3 ; a1 Þ;
C200 ða3 ; a1 Þ are in direct proportion to the displacements tangential to
the separation planes (neglecting the rigid rotations x01 ; x02 ). This allows the
solving of the mixed fundamental problem in which on the separation planes are
given the tangential displacements, the normal stresses vanishing.

10.2.2.2 System of Integral Equations

By means of the kernels (10.54), one may write the system of integral equations
for the unknown functions C10 ða2 ; a3 Þ; C100 ða2 ; a3 Þ; C20 ða3 ; a1 Þ; C200 ða3 ; a1 Þ in the
form
472 10 Elastic Eighth-Space. Elastic Quarter-Space

Z Z
1 0 4 1 4 1
C1 ða2 ; a3 Þ þ K1 ða1 ; a2 ; a3 ÞC20 ða3 ; a2 Þda1  K3 ða1 ; a2 ; a3 ÞC200 ða3 ; a1 Þda1
c1 p 0 p 0
1 1
¼ a01 ða2 ; a3 Þ þ a001 ða2 ; a3 Þ; ð10:106Þ
a3 a
Z 12 Z 1
1 0 4 4
C ða3 ; a1 Þ þ K1 ða2 ; a1 ; a3 ÞC10 ða2 ; a3 Þda2 þ K3 ða2 ; a1 ; a3 ÞC100 ða2 ; a3 Þda2
c2 2 p 0 p 0
1 1
¼ a02 ða3 ; a1 Þ þ a002 ða3 ; a1 Þ;
a1 a3
Z
a2  a 2 2 4 1
ð1 þ 2mÞ 2 2 3 C10 ða2 ; a3 Þ þ C100 ða2 ; a3 Þ þ K2 ða1 ; a2 ; a3 ÞC20 ða3 ; a1 Þda1
c1 c1 p 0
Z
4 1 1 1
 K3 ða3 ; a2 ; a1 ÞC200 ða3 ; a1 Þda1 ¼ a01 ða2 ; a3 Þ  a001 ða2 ; a3 Þ; ð10:1060 Þ
p 0 a3 a2
Z
a23  a21 0 2 00 4 1
ð1 þ 2mÞ C2 ða3 ; a1 Þ þ C1 ða2 ; a3 Þ  K2 ða2 ; a1 ; a3 ÞC10 ða2 ; a3 Þda2
c22 c2 p 0
Z
4 1 1 1
 K3 ða3 ; a1 ; a2 ÞC100 ða2 ; a3 Þda2 ¼ a02 ða3 ; a1 Þ  a002 ða3 ; a1 Þ:
p 0 a1 a3

This system of four integral equations has a unique solution in the class of
generalized functions.
An arbitrary case of loading may be decomposed in two cases of loading after
the properties of symmetry or of antisymmetry with respect to the plane x1 ¼ x2 ,
which are characterized by the relations
a01 ða2 ; a3 Þ ¼ a002 ða3 ; a2 Þ; a001 ða2 ; a3 Þ ¼ a02 ða3 ; a2 Þ: ð10:107Þ

It results
C10 ða2 ; a3 Þ ¼ C200 ða3 ; a2 Þ; C100 ða2 ; a3 Þ ¼ C200 ða3 ; a2 Þ; ð10:1070 Þ

the upper signs corresponding to the symmetric case, while the lower ones the
antisymmetric one. In this case, the system of equations (10.106), (10.1060 ) may be
replaced by two subsystems
Z Z
1 0 4 1 0 4 1
C ða2 ;a3 Þ  K1 ða1 ;a2 ;a3 ÞC1 ða1 ;a3 Þda1  K3 ða1 ;a2 ;a3 ÞC100 ða1 ;a3 Þda1
c1 1 p 0 p 0
1 1
¼ a01 ða2 ;a3 Þ þ a001 ða2 ;a3 Þ; ð10:108Þ
a3 a2
Z
a2  a2 2 4 1
ð1 þ 2mÞ 2 3 3 C10 ða2 ;a3 Þ þ C100 ða2 ;a3 Þ  K2 ða1 ;a2 ;a3 ÞC10 ða1 ;a3 Þda1
c1 c1 p 0
Z
4 1 1 1
 K3 ða3 ;a2 ;a1 ÞC100 ða1 ;a3 Þda1 ¼ a01 ða2 ;a3 Þ  a001 ða2 ;a3 Þ;
p 0 a3 a2
having each one only two unknown functions, which, obviously, is an advantage
for the computation; the upper signs correspond to the symmetric case and the
10.2 Elastic Quarter-Space 473

lower ones to the antisymmetric one. Eliminating, e. g., C20 ða3 ; a1 Þ and
C200 ða3 ; a1 Þbetween the Eqs. (10.106), (10.1060 ), one obtains a system of two
integral equations having C10 ða2 ; a3 Þ and C100 ða2 ; a3 Þ as unknown functions; this
system may be integrated by successive approximations.
The case of a loading symmetric with respect to the axes Ox1 and Ox2 , of the
form
Z 1Z 1
t12 ðx2 ; x3 Þ ¼ a01 ða2 ; a3 Þ cos a2 x2 cos a3 x3 da2 da3 ;
0 0
Z 1Z 1 ð10:109Þ
00
t13 ðx2 ; x3 Þ ¼ a1 ða2 ; a3 Þ sin a2 x2 sin a3 x3 da2 da3 ;
0 0
Z 1 Z 1
t21 ðx3 ; x1 Þ ¼ a02 ða3 ; a1 Þ sin a3 x3 sin a1 x1 da3 da1 ;
0 0
Z 1 Z 1 ð10:1090 Þ
t23 ðx3 ; x1 Þ ¼ a002 ða3 ; a1 Þ cos a3 x3 cos a1 x1 da3 da1 ;
0 0

may be, similarly, studied.

10.2.3 Particular Cases. Applications

Let us deal with two particular cases which are not contained in the above general
theory. As well we shell consider two applications.

10.2.3.1 Particular Cases

Let be the elastic quarter-space acted upon on the face x1 ¼ 0 by a uniform


distributed normal load of compression, of intensity p. The state of stress is given
by
r11 ¼ p; r22 ¼ r33 ¼ 0; ð10:110Þ

r23 ¼ r31 ¼ r12 ¼ 0; ð10:1100 Þ

while the state of displacement reads


p p p
u1 ¼ x1 ; u2 ¼ m x2 ; u3 ¼ m x3 : ð10:111Þ
E E E
Analogically, we consider the elastic quarter-space acted upon by a uniform
distributed tangential load on the face x1 ¼ 0; we assume that the load is parallel to
the Ox3 -axis. Thus, the state of stress becomes
474 10 Elastic Eighth-Space. Elastic Quarter-Space

r11 ¼ r22 ¼ r33 ¼ 0; ð10:112Þ

r23 ¼ r12 ¼ 0; r31 ¼ t; ð10:1120 Þ


while the state of displacement becomes
t t
u1 ¼ x3 ; u2 ¼ 0; u3 ¼ x1 : ð10:113Þ
2l 2l

10.2.3.2 Applications

We will deal with two cases of normal loading, where we put in evidence prop-
erties of symmetry or antisymmetry with respect to the bisecting plane x1 ¼ x2 ,
obtaining results in a finite form.
The symmetric load
p 
p1 ðx2 ; x3 Þ ¼  ð3  2mÞx22 þ 4mx23 ;
r15
p  ð10:114Þ
p2 ðx3 ; x1 Þ ¼  5 ð3  2mÞx21 þ 4mx23
r2

leads to
2p  2 
a1 ða2 ; a3 Þ ¼ 2a2 þ ð1 þ 2mÞa23 ;
pc1
ð10:1140 Þ
2p  2 
a2 ða3 ; a1 Þ ¼ 2a1 þ ð1 þ 2mÞa23 ;
pc2
the Eq. (10.94) with the upper sign leads to
2
C1 ða2 ; a3 Þ ¼ C2 ða3 ; a1 Þ ¼ p: ð10:115Þ
p
One obtains the state of stress
 x 2 x 2 x 2 
p 2m 3 1 3
r11 ¼ 3 3 1   ð1  2mÞ  35 ;
R 3 R R R
 x 2 x 2 x 2 
p 2m 3 2 3
r22 ¼ 3 3 1   ð1  2mÞ  35 ; ð10:116Þ
R 3 R R R
x 2 x 4
p 3 3
r33 ¼  3 1 þ 2m  6ð2 þ mÞ þ15 ;
R R R
10.2 Elastic Quarter-Space 475

x 2
x2 x3 3
r23 ¼ 3p 3 þ 2m  5 ;
R5 R
x 2
x3 x1 3
r31 ¼ 3p 5 3 þ 2m  5 ; ð10:1160 Þ
R R
x 2
x1 x2 3
r12 ¼ 3p 5 3  5
R R

and the state of displacement


x 2
3p x1 3
u1 ¼  1  ;
2l R3 R
x 2
3p x2 3
u2 ¼  1  ; ð10:117Þ
2l R3 R
x 2
p x3 3
u3 ¼  1 þ 4m  3 :
2l R3 R
In cylindrical co-ordinates, one obtains the normal stresses
 z 2  z 4
p  m
rr3 r3 ¼ 3 3 2 1 þ  ð7 þ 2mÞ þ5 ;
R 3 R R
 z 2
p 2m
rhh ¼ 3 3 1   ð1  2mÞ ; ð10:118Þ
R 3 R
 z 2  z 4
p
rzz ¼  3 1 þ 2m  6ð2 þ mÞ þ15
R R R

and the tangential stresses


 z 2
zr3
rhz ¼ rr3 h ¼ 0; rzr3 ¼ 3p 3 þ 2m  5 ; ð10:1180 Þ
R5 R

the displacements are given by


 z 2  z 2
3p r3 p z
ur 3 ¼  1 ; uh ¼ 0; uz ¼  1 þ 4m  3 : ð10:119Þ
2l R3 R 2l R3 R
As well, in spherical co-ordinates one obtains the state of stress
p  
rRR ¼ 2 3
3 þ m  ð5  mÞ cos2 u ;
R
p  
ruu ¼  3 1 þ 2m  ð1  2mÞ cos2 u ; ð10:120Þ
R
p  
rhh ¼  3 3  2m  3ð1  2mÞ cos2 u ;
R
p
ruh ¼ rhR ¼ 0; rRu ¼ ð1 þ mÞ sin 2u ð10:1200 Þ
R3
476 10 Elastic Eighth-Space. Elastic Quarter-Space

and the state of displacement


1 p  
uR ¼  1 þ 4m  ð5  4mÞ cos2 u ;
4l R2
ð10:121Þ
1  2m p
uu ¼  sin 2u; uh ¼ 0:
2l R2
The state of strain is compressible, because

r11 þ r22 þ r33 ¼ rr3 r3 þ rhh þ rzz ¼ rRR þ ruu þ rhh


 z 2
p p
¼ 2ð1 þ mÞ 3 1  3 ¼ 2ð1 þ mÞ 3 ð1  3 cos2 uÞ:
R R R
ð10:122Þ
0 0
The formulae (10.118 ) and 10.120 ) show that, at an arbitrary point, the
direction h is a principal direction, the corresponding principal normal stress being
rhh ; the other principal stresses are contained in the plane ðR; uÞ. For instance, the
angles formed by these principal directions with the direction R are given by
2rRu 2ð1 þ mÞ sin 2u
tan 2w ¼ ¼ : ð10:123Þ
rRR  ruu 7 þ 4m  ð11  4mÞ cos2 u
In the case of antisymmetric loads of the form
p 2 p
p1 ðx2 ; x3 Þ ¼ ðx  2x23 Þ; p2 ðx3 ; x1 Þ ¼  5 ðx21  2x23 Þ; ð10:124Þ
r15 2 r2

to which correspond the Fourier coefficients


2p 2 2p 2
a1 ða2 ; a3 Þ ¼ a ; a2 ða3 ; a1 Þ ¼  a ; ð10:1240 Þ
pc1 3 pc2 3
the system (10.94) with the lower sign leads to
2p
C1 ða2 ; a3 Þ ¼ C2 ða3 ; a1 Þ ¼ : ð10:125Þ
ð1  2mÞp
One obtains the state of stress
x 2  
p 3 3  x1  2 x21  x22
r11 ¼ 3 1  3  25 ;
R R 1  2m R R2
x 2  
p 3 3 x2 2 x21  x22
r22 ¼  3 1  3  2þ5 ; ð10:126Þ
R R 1  2m R R2

x2  x2 5 x3 2
r33 ¼ 3p 1 5 2 1  ;
R 1  2m R
10.2 Elastic Quarter-Space 477

 
3p x2 x3 x21  x22
r23 ¼ 2m þ 5 ;
1  2m R5 R2
 
3p x3 x1 x21  x22 ð10:1260 Þ
r31 ¼ 2m  5 ;
1  2m R5 R2
15p x1 x2
2
r12 ¼ 7
x1  x22
1  2m R
and the state of displacement
 
p x1 3 x21  x22
u1 ¼  2þ ;
2l R3 1  2m R2
 
p x2 3 x21  x22
u2 ¼ 2  ; ð10:127Þ
2l R3 1  2m R2
3p x3
2
u3 ¼  x  x22 ;
2ð1  2mÞl R5 1
In cylindrical co-ordinates one obtains the normal stresses
  z 2  z 2 
p 3 r3 2
rr3 r3 ¼ 3 1  3 þ 35 cos 2h;
R R 1  2m R R
 z 2
p
rhh ¼  3 1  3 cos 2h; ð10:128Þ
R R

r32 5  z 4
rzz ¼ 3p 5 1  cos 2h
R 1  2m R

and the tangential stresses


6m zr3
rhz ¼ p 5 sin 2h;
1  2m R
3p zr3 r 2
3
rzr3 ¼ 2m  5 cos 2h; ð10:1280 Þ
1  2m R5 R

p 3m r3 2
rr3 h ¼ 2 3 1 þ sin 2h;
R 1  2m R

the displacements are given by



p r3 3 r3 2
ur3 ¼  2þ cos 2h;
2l R3 1  2m R
ð10:129Þ
p r3 3p zr32
uh ¼ sin 2h; u z ¼  cos 2h:
2l R3 2l R5
478 10 Elastic Eighth-Space. Elastic Quarter-Space

In spherical co-ordinates, one obtains the state of stress


2ð5  mÞ p
rRR ¼ sin u cos 2h;
1  2m R3 ð10:130Þ
p p
ruu ¼  3 ð3  cos2 uÞ cos 2h; rhh ¼  3 ð1  3 cos2 uÞ cos 2h;
R R
p
ruh ¼ 2 cos u sin 2h;
R3
ð10:1300 Þ
2ð1 þ mÞ p 1þm p
rhR ¼ sin u sin 2h; rRu ¼  sin 2u cos 2h
1  2m R3 1  2m R3
and the state of displacement
 
p 1 6m 2
uR ¼  5 þ sin u sin2 u cos 2h;
4l R2 1  2m
  ð10:131Þ
p 1 3m p 1
uu ¼  1 þ sin 2u cos 2h; uh ¼ sin u sin 2u:
2l R2 1  2m 2l R2
The state of strain is compressible, because

r11 þ r22 þ r33 ¼ rr3 r3 þ rhh þ rzz ¼ rRR þ ruu þ rhh


1 þ m x21  x22 1 þ m r32 1þm p
¼6 p 5
¼6 p cos 2h ¼ 6 sin2 u cos 2h
1  2m R 1  2m R5 1  2m R3
ð10:132Þ
As in Sect. 10.1.4.2, one may eliminate the singularity which appears at the
origin of the co-ordinate axes by means of a quarter-sphere, acted upon by a
normal load rRR and by tangential loads rRu and rRh .

References

A. Books

1. Teodorescu, P.P.: Probleme spatßiale în teoria elasticitătßii (Spatial Problems in the Theory of
Elasticity). Ed. Acad. Rom., Bucuresßti, (1970)

B. Papers

2. Hijab, W.A.: Concentrated force in eighth-space with mixed boundary conditions. J. für reine
angew. Math. 225, 85 (1967)
3. Teodorescu, P.P.: Fonctions de tension dans le problème tridimensionnel de la théorie de
l’élasticité. Bull. Math. Soc. Sci. Math. Phys. Roum. 3(51), 499 (1959)
References 479

}
4. Teodorescu, P.P.: Uber einige räumliche Probleme der Elastizitätstheorie. Aplik. Mat. Praha.
4, 225 (1959)
5. Teodorescu, P.P.: Sur le problème du quart de plan élastique. Aplik. Mat. Praha 6, 359 (1961)
6. Teodorescu, P.P.: Sur le problème du demi–quart d’espace èlastique. Bull. Acad. Pol. Sci.
Sér. Sci. Techn. 12, 599 (1964)
7. Teodorescu, P.P.: Sur le problème du demi–quart d’espace élastique soumis à l’action d’une
charge tangentielle. Rev. Roum. Sci. Techn. Sér.. Méc. Appl. 10, 81 (1965)
8. Teodorescu, P.P.: Sur l’effet de coin en élasticité tridimensionnelle. Arch. Mech. Stos. 17,
249 (1962)
9. Teodorescu, P.P.: Sur le calcul du quart d’espace élastique. Rev. Roum. Sci. Techn., sér.
Méc. Appl. 11, 1141 (1966)
10. Teodorescu, P.P.: Trei probleme tridimensionale ale teoriei elasticitătßii (Three tridimensional
problems of the theory of elasticity). An. Univ. Bucuresßti, ser. ßst. nat., mat.–mec. 15, 17
(1966)
11. Teodorescu, P.P.: Sur le problème du quart d’espace élastique. I. Charges normales. Bull.
Acad. Pol. Sci., sér. Sci. Techn. 14, 487 (1966)
12. Teodorescu, P.P.: Sur le problème du quart d’espace élastique. II. Charges trangentielles.
Bull. Acad. Pol. Sci., sér. Sci. Techn. 14, 497 (1966)
13. Teodorescu, P.P.: Asupra utilizării unor familii de functßii potentßial reale în teoria elasticitătßii
(On the use of certain families of real potential functions in the theory of elasticity). An.
Univ. Bucuresßti, ser. ßst. nat., mat.–mec. 16, 59 (1967)
14. Teodorescu, P.P.: Einige Teilprobleme der dreidimensionalen Elastizitätstheorie. Wiss.
Zeitschsift Techn. Univ. Dresden 16, 77 (1967)
15. Teodorescu, P.P.: Sur le problème du demi–quart d’espace élastique soumis à l’action d’une
charge tangentielle. In: Problems of Fluid Flow Machines, Panst. wgdaw. nauk., Warszawa,
pp. 883 (1968)
16. Teodorescu, P.P.: Sur l’application des fonctions potentiel dans la théorie de l’élasticité.
Rend. Sem. Mat. Fis., Milano. 38, 231 (1968)
Chapter 11
Elastic Parallelepiped. Elastic Strip.
Elastic Layer. Thick Plate

In what follows we make a study of the elastic parallelepiped subjected to the


action of a normal load; the result thus obtained will be then particularized for the
elastic strip and for the elastic layer. One makes then various considerations
concerning the finite thick plate. To solve the above mentioned problems, we shall
use representations with the stress functions introduced in Sect. 5.3.2.5.

11.1 Elastic Parallelepiped

The first fundamental problem for an elastic parallelepiped has been formulated, in
1852, by G. Lamé [4], who considered it to be very important both from the
theoretical and the practical point of view, emphasizing—at the same time—the
difficulties which appear in its solving. In his book, G. Lamé stimulates the young
researchers to deal with this problem, for which he only sketches some results. At
the same time, twice (in 1846 and 1858), the Academy of Sciences in Paris
proposed this problem, endowing it with a prize; it became a classical problem of
the theory of elasticity.
This problem has not been sufficiently studied. There have been various con-
siderations: theoretical, as well experimental; from the last point of view, the
researchers dealt with the problem of compression of the concrete cubes. We
mention thus the classical treatise of A. Föppl and L. Föppl [2].
The Russian school of elasticity tackled this problem by various approximate
methods of computation. We mention thus the papers of M. M. Filonenko-Borodich
[14–18], based on a variational method of Castigliano and on a choice of a system
of functions which have been previous considered by the same author [1]; one used
triple Fourier series, obtaining results in the first or in the second approximation for
some cases of loading. A. I. Meshkov [28, 29] takes again this methods of com-
putation for the oblique parallelepiped, considering loads which lead to torsion too.

P. P. Teodorescu, Treatise on Classical Elasticity, Mathematical and Analytical 481


Techniques with Applications to Engineering, DOI: 10.1007/978-94-007-2616-1_11,
Ó Springer Science+Business Media Dordrecht 2013
482 11 Elastic Parallelepiped. Elastic Strip. Elastic Layer. Thick Plate

The same method of computation has been applied by E. S. Kononenko [24, 25] to
study the state of stress of concrete cubes subjected to compression. V. P. Netrebko
[32, 33] deals with the torsion of the elastic parallelepiped acted upon by tangential
loads. Ē. N. Baı̆da [9–11], using a method of computation analogous to that indi-
cated by Boussinesq and Galerkin, makes some considerations concerning the
isotropic and the anisotropic elastic parallelepiped. M. Mishonov [31] deals with
the elastic parallelepiped acted upon by arbitrary volume forces, using triple
Fourier series; by a certain procedure, one also obtains results for superficial loads
as a limit case.
In what follows, we will use the results given by us [38] and presented at the
1960 IUTAM Congress.

11.1.1 Stress Functions. Boundary Conditions

Let be the elastic parallelepiped jxi j  ai , of dimensions 2ai ; i ¼ 1; 2; 3, where the


axes of symmetry are the co-ordinate axes. We assume that the faces xi ¼ ai are
acted upon by the normal loads p1 ðx2 ; x3 Þ; p2 ðx3 ; x1 Þ; p3 ðx1 ; x2 Þ, respectively,
symmetric with respect to the three planes of co-ordinates.
Any other case of loading with normal or tangential loads, having various
properties of symmetry or antisymmetry, may be similarly studied. The results for
an arbitrary case of loading may be obtained using the principle of superposition of
effects.

11.1.1.1 Stress Functions

To determine the state of strain and stress in the elastic parallelepiped thus acted
upon, we will introduce stress functions of the form (where the simple series which
intervene, as well as the free terms are put in evidence)
X 
F11 ¼ L01 x21 þ L02 x22 þ L03 x23 þ A01m0 cosh bm x1 þ bm x1 A02m0 sinh bm x1 cos bm x2
m
X 
þ A010n cosh cn x1 þ cn x1 A020n sinh cn x1 cos cn x3
n
XX 
þ A01mn cosh kmn x1 þ kmn x1 A02mn sinh kmn x1 cos bm x2 cos cn x3
m n
X 
þ B01n0 cosh cn x2 þ cn x2 B02n0 sinh cn x2 cos cn x3
n
X  
þ B010l cosh al x2 þ al x2 B020l sinh al x2 cos al x1
l
XX 
þ B01nl cosh lnl x2 þ lnl x2 B02nl sinh lnl x2 cos cn x3 cos al x1
n l
X 
0 0
þ C1l0 cosh al x3 þ al x3 C2l0 sinh al x3 cos al x1
l
11.1 Elastic Parallelepiped 483

X 
0 0
þ C10m cosh bm x3 þ bm x3 C20m sinh bm x3 cos bm x2
m X
X 
0 0
þ C1lm cosh mlm x3 þ mlm x3 C2lm sinh mlm x3 cos al x1 cos bm x2 ; . . .;
l lm

where ð11:1Þ

lp mp np
al ¼ ; bm ¼ ; cn ¼ ; l; m; n ¼ 1; 2; 3; . . .; ð11:2Þ
a1 a2 a3
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
kmn ¼ b2m þ c2n ; lnl ¼ c2n þ a2l ;
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð11:20 Þ
mlm ¼ a2l þ b2m ; l; m; n ¼ 1; 2; 3; . . .;

choosing the periods Li ¼ 2ai ; i ¼ 1; 2; 3; on the three directions.


The functions Fij ; i ¼
6 j; i; j ¼ 1; 2; 3, are equated to zero.

11.1.1.2 Boundary Conditions

The boundary conditions are put in the form


x1 ¼ a1 : r11 ¼ p1 ðx2 ; x3 Þ; r12 ¼ r13 ¼ 0; . . .; ð11:3Þ
the external loads being expressed in the form
X X
p1 ðx2 ; x3 Þ ¼a000 þ a0m0 cos bm x2 þ a00n cos cn x3
m n
XX
þ a0mn cos bm x2 cos cn x3 ; . . .; ð11:4Þ
m n

a0mn ; a00nl ; a000


lm being loading coefficients.
Starting from (11.1), we obtain the normal stresses
X
r11 ¼ 2½ð1 þ mÞL01  ðL001 þ L000 0 0
1 Þ þ ð2 þ mÞðL2 þ L3 Þ  b2m ffA1m0
m
  
2 ð1 þ mÞA02m0  ðA002m0 þ A000
2m0 Þ cosh bm x1 þ bm x1 A2m0 sinh bm x1 cos bm x2
X   
 c2n A10n  2 ð1 þ mÞA020n  ðA0020n þ A00020n Þ cosh cn x1
n
 XX  
þcn x1 A20n sinh cn x1 cos cn x3  k2mn A1mn  2 ð1 þ mÞA02mn
m n
 
ðA002mn þ A000
2mn Þ cosh kmn x1 þ kmn x1 A2mn sinh kmn x1 cos bm x2 cos cn x3
X X  
þ 2ð2 þ mÞ c2n B02n0 cosh cn x2 cos cn x3 þ k2l B10l þ 2ð2 þ mÞB020l cosh al x2
n l
 X X  
þal x2 sinh al x2 cos al x1 þ a2l B1nl þ 2ð2 þ mÞl2nl B02nl cosh lnl x2
n l
484 11 Elastic Parallelepiped. Elastic Strip. Elastic Layer. Thick Plate

 X
þlnl x2 a2l B2nl sinh lnl x2 cos cn x3 cos al x1 þ a2l f½C1l0
0
 l
þ2ð2 þ mÞC2l0 cosh al x3 þ al x3 C2l0 sinh al x3 cos al x1
X
0
þ2ð2 þ mÞ b2m C20m cosh bm x2
X X m 
0
þ a2l C1lm þ 2ð2 þ mÞm2lm C2lm cosh mlm x3
l m

þmlm x3 a2l C2lm sinh mlm x3 cos al x1 cos bm x2 ;. .. ð11:5Þ
and the tangential stresses
X  
r23 ¼ x3 b3m mðA02m0 þ A002m0 Þ A000
2m0 coshbm x1 sinbm x2
m
X   X X 1 
þ x2 c2n mðA020n þ A000 00
20n Þ 2A20n coshcn x1 sincn x3  b2m c2n A1mn
m m n
bm cn

mk4mn A02mn þ k2mn ð2c2n  mb2m ÞA002mn þ k2mn ð2b2m  mc2n ÞA000
2mn coshkmn x1
 X 
þkmn x1 b2m c2n A2mn sinhkmn x1 sinbm x2 sincn x3 þ c2n B1n0 þ B02n0
n
 
 ð1 þ mÞB002n0 þð3 þ mÞB000 2n0 sinhcn x2 þ cn x2 B2n0 coshcn x2 sincn x3
X   X X l 
 x3 a2l mðB020l þ B0020l Þ B000
20l sinhal x2 cosal x1 þ
nl
c2n B1nl
l n l
cn
  
þðc2n  ma2l ÞB02nl  ½ð1þ mÞc2n þ ma2l B002nl þ ð3 þ mÞc2n þ 2a2l B000
2nl sinhlnl x2
 X 
0 000
þlnl x2 c2n B2nl coshlnl x2 sincn x3 cosal x1  x2 a2l mðCl0 þ Cl0 Þ
l
 X 
00 0 00
2Cl0 sinhal x3 cosal x1 þ b2m C10m þ C20m þ ð3þ mÞC20m
m
000
 
ð1þ mÞC20m sinhbm x3 þ bm x3 C20m coshbm x3 sinbm x2
X X mlm 
0 00
þ b2m C1lm þ ðb2m  ma2l ÞC2lm þ ½ð3 þ mÞb2m þ 2a2l C2lm
m n
b m
000

½ð1þ mÞb2m þ ma2l C2lm sinhmlm x3
2

þmlm x3 bm C2lm coshmlm x3 cosal x1 sinbm x2 ;...; ð11:50 Þ

with the notations


Aim0 ¼ A0im0 þ A00im0 þ A000
im0 ; i ¼ 1; 2; m ¼ 1; 2; 3; . . .;
Ai0n ¼ A0i0n þ A00i0n þ A000
i0n ; i ¼ 1; 2; n ¼ 1; 2; 3; . . .; ð11:6Þ
Aimn ¼ A0imn þ A00imn þ A000
imn ; i ¼ 1; 2; m; n ¼ 1; 2; 3; . . .;

Concerning the free terms, the boundary conditions for the normal stresses lead to

2½ð1 þ mÞL01  ðL001 þ L000 0 0 0


1 Þ þ ð2 þ mÞðL2 þ L3 Þ ¼ a00 ; . . .; ð11:7Þ
11.1 Elastic Parallelepiped 485

because these coefficients do not intervene in the expressions of the tangential


stresses, one may chose them in the form

L001 ¼ L000 000 0 0 00


1 ¼ L2 ¼ L2 ¼ L3 ¼ L3 ¼ 0; ð11:8Þ

a000 a0000 a000


L01 ¼ ; L002 ¼ ; L000 ¼ 00
: ð11:80 Þ
2ð1 þ mÞ 2ð1 þ mÞ 3 2ð1 þ mÞ
Putting the boundary conditions for the tangential stresses, one obtains

mðA02m0 þ 2A002m0 Þ ¼ 0; ð11:9Þ

A1m0  ð1 þ mÞA02m0 þ ð3 þ mÞA002m0 þ A000


2m0 þ bm a1 A2m0 coth bm a1 ¼ 0; . . .;
mðA020n þ A000 00
20n Þ  2A20n ¼ 0; ð11:90 Þ
A10n  ð1 þ mÞA020n þ A0020n þ ð3 þ mÞA000
20n þ cn a1 A20n coth cn a1 ¼ 0; . . .;
 
b2m A1mn sinh kmn a1  ð1 þ mÞb2m þ mc2n sinh kmn a1
  
b2m kmn a1 cosh kmn a1 A02mn þ ð3 þ mÞb2m þ 2c2n sinh kmn a1
 
þb2m kmn a1 cosh kmn a1 A002mn þ ðb2m  mc2n Þ sinh kmn a1

þb2m kmn a1 cosh kmn a1 A000
2mn ¼ 0;
2
 2 2

cn A1mn sinh kmn a1  ð1 þ mÞcn þ mbm sinh kmn a1
 
c2n kmn a1 cosh kmn a1 A02mn þ ðc2n  mb2m Þ sinh kmn a1
  
þc2n kmn a1 cosh kmn a1 A002mn þ ð3 þ mÞc2n þ 2b2m sinh kmn a1

þc2n kmn a1 cosh kmn a1 A000
2mn ¼ 0; . . .
ð11:900 Þ
It results that
1
A1m0 ¼ ð2 þ mÞ½A02m0  3A002m0  bm a1 ðA02m0 þ A002m0 Þ coth bm a1 ;
2 ð11:10Þ
m
A2m0 ¼  ðA02m0 þ A002m0 Þ; . . .;
000
2
1
A10n ¼ ð2 þ mÞ½A020n  3A000 0 000
20n  cn a1 ðA20n þ A20n Þ coth cn a1 ;
2 ð11:100 Þ
m
A0020n ¼ ðA020n þ A000
20n Þ; . . .;
2
ð2 þ mÞ
A1mn ¼ ½ð1  kmn a1 coth kmn a1 Þðc2n A002mn  b2m A000
2mn Þ
mðc2n  b2m Þ
þ 2mðb2m A002mn  c2n A000
2mn Þ; ð11:1000 Þ
1
A02mn ¼ ½ð2c2n þ mb2m ÞA002mn  ð2b2m þ mc2n ÞA000
2mn ; . . .;
mðc2n  b2m Þ
486 11 Elastic Parallelepiped. Elastic Strip. Elastic Layer. Thick Plate

as well
1
A2m0 ¼ ð2 þ mÞðA02m0 þ A002m0 Þ;
2
1
A20n ¼ ð2 þ mÞðA020n þ A000
20n Þ; ð11:11Þ
2
1 c2 A00  b2m A000 2mn Þ
A2mn ¼ ð2 þ mÞ n 2mn ;...
m c2n  b2m
We introduce the notations

k2mn c2n A002mn  b2m A000


A0mn ¼ 2mn
; m; n ¼ 1; 2; 3; . . .; ð11:12Þ
mð2 þ mÞ c2n  b2m

with the particular cases

b2m c2n
A0m0 ¼ A000 0
2m0 ; A0n ¼ A000 ; m; n ¼ 1; 2; 3; . . . ð11:120 Þ
mð2 þ mÞ mð2 þ mÞ 20n
Thus, the state of stress may be written in the form
X
r11 ¼ a000 þ A0m0 ½ð1 þ bm a1 coth bm a1 Þ cosh bm x1  bm x1 sinh bm x1  cos bm x2
m
X
þ A00n ½ð1 þ cn a1 coth cn a1 Þ cosh cn x1  cn x1 sinh cn x1  cos cn x3
n
XX
þ A0mn ½ð1 þ kmn a1 coth kmn a1 Þ cosh kmn x1
m n
X
 kmn x1 sinh kmn x1  cos bm x2 cos cn x3 þ 2l B0n0 cosh cn x2 cos cn x3
n
X
þ B00l ½ð1  al a2 coth al a2 Þ cosh al x2 þ al x2 sinh al x2  cos al x1
l
(
2
XX al
þ B0nl ½ð1  lnl a2 coth lnl a2 Þ cosh lnl x2 þ lnl x2 sinh lnl x2 
n l
lnl

2 )
c X
0
þ2m n cosh lnl x2 cos cn x3 cos al x1 þ Cl0 ½ð1  al a3 coth al a3 Þ cosh al x3
lnl l
X
0
þ al x3 sinh al x3  cos al x1 þ 2l C0m cosh bm x3 cos bm x2
m
(
2
XX al
0
þ Clm ½ð1  mlm a3 cothmlm a3 Þ cosh mlm x3 þ mlm x3 sinh mlm x3 
l m
mlm

2 )
b
þ2m m cosh mlm x3 cos al x1 cos bm x2 ; . . .;
mlm
ð11:13Þ
11.1 Elastic Parallelepiped 487

XX bm cn
r23 ¼  A0mn
½ð1  kmn a1 coth kmn a1 Þ cosh kmn x1 þ kmn x1 sinh kmn x1
m n k2mn
X
 2m cosh kmn x1  sin bm x2 sin cn x3  B0n0 ðcn a2 coth cn a2 sinh cn x2
n
XXcn
 cn x2 cosh cn x2 Þ sin cn x3  B0nl
ðl a2 coth lnl a2 sinh lnl x2
n l
lnl nl
X
0
 lnl x2 cosh lnl x2 Þ sin cn x3 cos al x1  C0m ðbm a3 coth bm a3 sinh bm x3
m
XX bm
0
 bm a3 cosh bm x3 Þ sin bm x2  Clm ðmlm a3 coth mlm a3 sinh mlm x3
l m
mlm
 mlm x3 cosh mlm x3 Þ cos al x1 sin bm x2 ; . . .
ð11:130 Þ
Putting the boundary conditions for the normal stresses too, one obtains fol-
lowing relations which must be verified by the three double sequences of
parameters.
X
X

0 bm a1 0 c n a1
Am0 cosh bm a1 þ cos bm x2 þ A0n ðcosh cn a1 þ cos cn x3
m
sinh bm a1 n
sinh cn a1
XX

kmn a1
þ A0mn cosh kmn a1 þ cos bm x2 cos cn x3
m n
sinh kmn a1
X X
þ 2m B0n0 cosh cn x2 cos cn x3 þ B0nl cos lp½ð1  al a2 coth al a2 Þcosh al x2
n l
(
2
XX al
þ al x2 sinh al x2  þ 2l B0nl cos lp ½ð1  lnl a2 coth lnl a2 Þcosh lnl x2
n l
lnl

2 )
cn
þlnl x2 sinh lnl x2  þ 2l cosh lnl x2 cos cn x3
lnl
X
0
þ Cl0 cos lp½ð1  al a3 coth al a3 Þ cosh al x3 þ al x3 sinh al x3 
l
X
0
þ 2m C0m coshbm x3 cos bm x2
m
(
2
XX al
0
þ Clm cos lp ½ð1  mlm a3 coth mlm a3 Þ cosh mlm x3 þ mlm x3 sinh mlm x3 
l m
mlm

2 )
b
þ2m m cosh mlm x3 cos bm x2 ¼ p1 ðx2 ; x3 Þ;.. .
mlm
ð11:14Þ
488 11 Elastic Parallelepiped. Elastic Strip. Elastic Layer. Thick Plate

By means of the Fourier series expansions


1 2kmn X cos lp cos al x1
cosh kmn x1 ¼ sinh kmn a1 þ sinh kmn a1 2
;
kmn a1 a1 2
l a l þ bm þ c n
2

1 2b X cos lp cos al x1
cosh bm x1 ¼ sinh bm a1 þ m sinh bm a1 ; ð11:15Þ
bm a 1 a1 l a2l þ b2m
1 2c X cos lp cos al x1
cosh cn x1 ¼ sinh cn a1 þ n sinh cn a1 ; . . .;
c n a1 a1 l
a2l þ c2n

1
kmn x1 sinh kmn x1 ¼ ðkmn a1 cosh kmn a1  sinh kmn a1 Þ
kmn a1
" #
kmn X cos lpcos al x1 a2l  ðb2m þ c2n Þ
þ2 sinh kmn a1 2
kmn a1 coth kmn a1 þ 2 ;
a1 2 2
l al þ b m þ c n al þ b2m þ c2n
1
bm x1 sinh bm x1 ¼ ðb a1 cosh bm a1  sinh bm a1 Þ
b m a1 m
!
bm X cos lp cos al x1 a2l  b2m
þ 2 sinh bm a1 bm a1 coth bm a1 þ 2 ;
a1 l a2l þ b2m al þ b2m
1
cn x1 sinh cn x1 ¼ ðc a1 cosh cn a1  sinh cn a1 Þ
c n a1 n
c X cos lp cos al x1 a2l  c2n

þ 2 n sinh cn a1 cn a 1 coth c n a 1 þ ;...


a1 l
a2l þ c2n a2l þ c2n
ð11:150 Þ
and taking into account that the double and triple series which intervene thus in the
relations (11.14) are absolute convergent with respect to each index (we assume
that the sequences of coefficients k2mn Amn ; b2m Am0 ; c2n A0n ; . . .; with the notations
sinh kmn a1
Amn ¼ A0mn cos mp cos np ;
kmn a1
sinh bm a1
Am0 ¼ A0m0 cos mp ; ð11:16Þ
bm a1
sinh cn a1
A0n ¼ A00n cos np ; . . .;
c n a1
are bounded) and the conditions in which the expansions into Fourier aeries
represent the same functions, one obtains the relations

kmn a1 cos mp 0 c
A0mn cosh kmn a1 þ þ 4m Bn0 sinh cn a2 2 n
sinhkmn a1 a2 bm þ c2n
"

4 cos mp X 0 al 2 b2m
þ Bnl lnl cos lp sinh lnl a2  2 2
a2 l
lnl al þ b2m þ c2n

2 #
c 1 4m cos mp 0 b
þm n þ C0m sinh bm a3 2 m
lnl a2l þ b2m þ c2n a3 bm þ c2n
11.1 Elastic Parallelepiped 489

"
2
4 cos np X 0 al c2n
þ Clm mlm cos lp sinh mlm a3  2 2
a3 l
mlm al þ b2m þ c2n

2 #
bm 1
þm ¼ a0mn ;...; ð11:17Þ
mlm a2l þ b2m þ c2n

4b2m cos mp X 0
bm a1 1
A0m0 cosh bm a1 þ þ B0l al cos lp sinh al a2  2
sinh bm a1 a2 l al þ b2m
2

X
2
0 sinh bm a3 0 b sinh mlm a3
þ 2mC0m þ 2m Clm cos lp m ¼ a0m0 ; . . .;
bm a3 l
m lm m lm a 3

c n a1 sinh cn a2
A00n cosh cn a1 þ þ 2mB0n0 ð11:170 Þ
sinh cn a1 c n a2
X
2
c sinh lnl a2
þ 2m B0nl cos lp n
l
lnl lnl a2
2
4c cos np X 1
0 0
þ n Cl0 al cos lp sinh al x3  2 ¼ a0n ; . . .
a3 l
2
al þ c2n

11.1.2 Infinite System of Linear Equations. State of Strain


and Stress

Using the above results, one may obtain a solution of the problem, reducing it to
the solving of an infinite system of linear algebraic equations with constant
coefficients.

11.1.2.1 Infinite System of Linear Equations

Taking into account the notations (11.16), one may write the relations (11.17),
(11.170 ) in the form

X
kmn a1 a2i b2m þ mc2n ða2i þ b2m þ c2n Þ
Amn coth kmn a1 þ k a
mn 1 þ 4 B ni  2 2
sinh2 kmn a1 i a i þ b2 þ c 2 m n
X a2i c2n þ mb2m ða2i þ b2m þ c2n Þ
þ4 Cim  2 2 ¼ ð1Þmþn a0mn ; m; n ¼ 1; 2; 3; . . .;
i ai þ b2m þ c2n
ð11:18Þ
490 11 Elastic Parallelepiped. Elastic Strip. Elastic Layer. Thick Plate


X
bm a1 a2 b2
Am0 coth bm a1 þ 2
bm a 1 þ 4 Bni  i m 2
sinh bm a1 i a2i þ b2 m
X b2
þ 2m Cim 2 m 2 ¼ ð1Þm a0m0 ; m ¼ 1; 2; 3; . . .;
i a i þ bm

ð11:180 Þ
c n a1 X c2
A0n coth cn a1 þ 2
cn a1 þ 2m Bni 2 n 2
sinh cn a1 i
ai þ c n
X a2 c 2
þ4 Ci0  i n 2 ¼ ð1Þn a00n ; n ¼ 1; 2; 3; . . .;
i a2i þ c2n

where in the above sums one takes i = 0 two.


Let us introduce the coefficients

4b2m c2n 4ma2l


fmnl ¼  2 þ ; ð11:19Þ
a2l þ b2m þ c2n a2l þ b2m þ c2n

with m; n ¼ 0; 1; 2; 3; . . .; l ¼ 0; 1; 2; 3; . . . and

2la2l
fmnl ¼ ; ð11:190 Þ
a2l þ b2m þ c2n

with m ¼ 0; n ¼ 0; 1; 2; 3; . . . or n ¼ 0; m ¼ 0; 1; 2; 3; . . . and l ¼ 0; 1; 2; 3; . . .; by
circular permutations one obtains the coefficients gnlm and hlmn .
One observes that, introducing the above coefficients and taking into account
that
km0 ¼ bm ; k0n ¼ cn ; . . .; ð11:20Þ
one may write each group of relations as only one relation; the corresponding
indices may vanish.
Introducing the notations given in Sect. A.4.1.2, we remark that the sequences
of parameters Amn ; Bmn ; Cmn ; are given by a system of linear algebraic equations
with a triple infinity of unknowns, of the form
X X
vðkmn a1 ÞAmn þ himn Bni þ gnim Cim ¼ ð1Þmþn a0mn ;
i i
X X
hljn Ajn þ vðlnl a2 ÞBnl þ fjnl Clj ¼ ð1Þnþl a00nl ; ð11:21Þ
j j
X X
gklm Amk þ fmkl Bkl þ vðmlm a3 ÞClm ¼ ð1Þlþm a000
lm ;
k k

with l; m; n; i; j; k ¼ 0; 1; 2; 3; . . .; the parameters A00 ; B00 ; C00 do not intervene in


computation.
The study of the static problem of the elastic parallelepiped is thus reduced to
the study of an infinite system of linear algebraic equations, as in the dynamic
11.1 Elastic Parallelepiped 491

case, considered by S. Kaliski [22, 23]. The system may be studied, e.g., using the
Banach–Schauder theorem of fixed point, as in [22].
As a matter of fact, the results thus obtained generalize thus given by us [6] for
the elastic rectangle. The functions and notations which have been introduced
correspond to those used in the mentioned plane case; these functions have been
plotted into diagrams, which leads to a simplification of the computation.
From a practical point of view, one can proceed otherwise too. Eliminating the
sequence of coefficients Amn , one obtains a linear system with a double infinity of
unknowns
X XX
anli Bni þ bnlij Cij ¼ a00nl ;
i i j
XX X ð11:22Þ
bkilm Bkl þ cilm Cim ¼ a000
im ;
k l l

where
X hijn hljn
anli ¼  ; anli ¼ anil ; i 6¼ l;
j
vðkjn a1 Þ
ð11:23Þ
X h2ljn
anll ¼ vðlnl a2 Þ  ;
j
vðkjn a1 Þ

gnij hljn
bnlij ¼  ; i 6¼ l;
vðkjn a1 Þ
ð11:230 Þ
gnij hljn
bnllj ¼ fjnl  ;
vðkjn a1 Þ
X gklm gkim
cilm ¼  ; cilm ¼ clim ; i 6¼ l;
k
vðkmk a1 Þ
X ð11:2300 Þ
g2klm
cllm ¼ vðmlm a3 Þ  ;
k
vðkmk a1 Þ
X hljn
a00nl ¼ ð1Þnþl a00nl 
 ð1Þjþn a0jn
;
j
vðkjn a1 Þ
X ð11:23000 Þ
gklm
a000
 lm ¼ ð1Þlþm a000
lm  ð1Þmþk a0mk :
k
vðk mk a1 Þ

Obviously, such a system of linear equations must be solved by various procedures


of successive approximations.
492 11 Elastic Parallelepiped. Elastic Strip. Elastic Layer. Thick Plate

11.1.2.2 State of Strain and Stress

Taking into account the relations (11.13), (11.130 ), (11.20), the notations (11.16) and
the functions introduced in Sect. A.4.1.2, one can write the state of stress in the form
XX
r11 ¼ a000 þ ð1Þmþn Amn W2 ðkmn ; x1 ; a1 Þ cos bm x2 cos cn x3
m n "

XX nþl al 2
þ ð1Þ Bnl W4 ðlnl ; x2 ; a2 Þ
n
lnl

2
l #
c
þ m n W5 ðlnl ; x2 ; a2 Þ cos cn x3 cos al x1
lnl
"

XX lþm al 2
þ ð1Þ Clm W4 ðmlm ; x3 ; a3 Þ
m
mlm
l

2 #
bm
þm W5 ðmlm ; x3 ; a3 Þ cos al x1 cos bm x2 ; . . .; ð11:24Þ
mlm

XX bm cn
r23 ¼  ð1Þmþn Amn W45 ðkmn ; x1 ; a1 Þ sin bm x2 sin cn x3
m n k2mn
XX cn
þ ð1Þnþl Bnl W3 ðlnl ; x2 ; a2 Þ sin cn x3 cos al x1
n l
lnl
XX bm
þ ð1Þlþm Clm W3 ðmlm ; x3 ; a3 Þ cos al x1 cos bm x2 ; . . .; ð11:240 Þ
l m
mlm

the sign sum being extended for the case in which one of the indices vanishes (the
two indices cannot vanish simultaneously).
For the state of displacement one obtains, similarly,
1  0 
u1 ¼ a00  mða0000 þ a000
00 Þ x1
2ð1 þ mÞl
XX Amn
þ ð1Þmþn W16 ðkmn ; x1 ; a1 Þ cos bm x2 cos cn x3
m n
kmn
XX Bnl al
þ ð1Þnþl W45 ðlnl ; x2 ; a2 Þ cos cn x3 sin al x1
n l
lnl lnl
XX Clm al
þ ð1Þlþm W45 ðmlm ; x3 ; a3 Þ sin al x1 cos bm x2 ; . . .; ð11:25Þ
l m
mlm mlm

where the displacements and the rotations of rigid body vanish, because of sym-
metry reasons with respect to the co-ordinate planes.
In use of the action of volume forces which has analogous properties of sym-
metry, one may use a representation by means of triple Fourier series
11.1 Elastic Parallelepiped 493

XXX
F1 ¼ a0lmn sinal x1 sin bm x2 sin cn x3 ; . . . ð11:26Þ
l m n

For the components of the stress tensor, one obtains easily a system of particular
integrals of the same form, which verifies the equilibrium and the Beltrami–Michell
equations. Putting the boundary conditions, the triple Fourier series become double
Fourier series and one has to solve a problem analogous to that considered above.

11.2 Elastic Strip. Elastic Layer

We deal, in what follows, with the study of other two infinite bodies: the elastic
strip and the elastic layer, using analogous methods of computations.

11.2.1 Elastic Strip

Throwing to infinite two opposite faces of the elastic parallelepiped, one obtains
the elastic strip jxi j  ai ; i ¼ 1; 2; which we consider to be acted upon by the
periodic loads (11.5).

11.2.1.1 System of Linear Equations

To solve the problem, one may use the representation (11.1)–(11.20 ), taking
L03 ¼ L003 ¼ L000 0 0 000
3 ¼ 0; C1lm ¼ C2lm ¼ . . . ¼ C2lm ¼ 0. The sequences of parameters
to be determined are obtained by putting the boundary conditions
x1 ¼ a1 : r11 ¼ p1 ðx2 ; x3 Þ; r12 ¼ r13 ¼ 0;
ð11:27Þ
x2 ¼ a2 : r22 ¼ p2 ðx3 ; x1 Þ; r21 ¼ r23 ¼ 0:

The study of the problem leads thus to the study of a system of linear algebraic
equations with a double infinity of unknowns of the form
X
vðkmn a1 ÞAmn þ himn Bni ¼ ð1Þmþn a0mn ;
i
X ð11:28Þ
hljn Ajn þ vðlnl a2 ÞBnl ¼ ð1Þnþl a00nl ;
j

with l; m; n; i; j ¼ 0; 1; 2; 3; . . .; the parameters A00 ; B00 do not intervene in


computation.
One can eliminate a sequence of unknowns (e.g., Amn ), obtaining for the
sequence Bnl the system of linear equations
494 11 Elastic Parallelepiped. Elastic Strip. Elastic Layer. Thick Plate

X
a00nl ;
anli Bni ¼  ð11:29Þ
i

with the notations (11.23), (11.2300 ).

11.2.1.2 State of Strain and Stress

Using the results in Sect. 11.1.2.2, we obtain the state of stress


XX
r11 ¼ a000 þ ð1Þmþn Amn W2 ðkmn ; x1 ; a2 Þ cos bm x2 cos cn x3
m n "

XX nþl al 2
þ ð1Þ Bnl W4 ðlnl ; x2 ; a2 Þ
n
lnl

2
l #
c
þ m n W5 ðlnl ; x2 ; a2 Þ cos cn x3 cos al x1 ;
lnl
"

XX mþn bm 2
00
r22 ¼ a00 þ ð1Þ Amn W4 ðkmn ; x1 ; a1 Þ
m n
kmn

2 #
cn
þm W5 ðkmn ; x1 ; a1 Þ cos bm x2 cos cn x3
kmn
XX ð11:30Þ
þ ð1Þnþl Bnl W2 ðlnl ; x2 ; a2 Þ cos cn x3 cos al x1 ;
n l "

XX mþn cn 2
0 00
r33 ¼ mða00 þ a00 Þ þ ð1Þ Amn W4 ðkmn ; x1 ; a1 Þ
m n
kmn

2 #
bm
þm W5 ðkmn ; x1 ; a1 Þ cos bm x2 cos cn x3
kmn
"

XX nþl cn 2
þ ð1Þ Bnl W4 ðlnl ; x2 ; a2 Þ
n
lnl

2
l #
al
þm W5 ðlnl ; x2 ; a2 Þ cos cn x3 cos al x1 ;
lnl
11.2 Elastic Strip. Elastic Layer 495

XX bm cn
r23 ¼  ð1Þmþn Amn W45 ðkmn ; x1 ; a1 Þ sin bm x2 sin cn x3
m n k2mn
XX cn
þ ð1Þnþl Bnl W3 ðlnl ; x2 ; a2 Þ sin cn x3 cos al x1
n l
l nl
XX c
r31 ¼ ð1Þmþn Amn n W5 ðkmn ; x1 ; a1 Þ cos bm x2 sin cn x3
m n
kmn
XX c al ð11:300 Þ
 ð1Þnþl Bnl n 2 W45 ðlnl ; x2 ; a2 Þ sin cn x3 sin al x1 ;
n l
lnl
XX b
r12 ¼ ð1Þmþn Amn m W3 ðkmn ; x1 ; a1 Þ sin bm x2 cos cn x3
m n
kmn
XX al
þ ð1Þnþl Bnl W3 ðlnl ; x2 ; a2 Þ sin cn x3 sin al x1
n l
lnl

and the state of displacement


1  
u1 ¼ ð1  mÞa000  ma0000 x1
2l
1 XX Amn
þ ð1Þmþn W16 ðkmn ; x1 ; a1 Þ cos bm x2 cos cn x3
2l m n kmn
1 XX Bnl al
þ ð1Þnþl W45 ðlnl ; x2 ; a2 Þ cos cn x3 sin al x1
2l n l lnl lnl
1  
u2 ¼ ð1  mÞa0000  ma000 x2
2l
1 XX Amn bm ð11:31Þ
þ ð1Þmþn W45 ðkmn ; x1 ; a1 Þ sin bm x2 cos cn x3
2l m n kmn kmn
1 XX Bnl
þ ð1Þnþl W16 ðlnl ; x2 ; a2 Þ cos cn x3 cos al x1 ;
2l n l lnl
XX Amn cn
u3 ¼ ð1Þmþn W45 ðkmn ; x1 ; a1 Þ cos bm x2 sin cn x3
m n
kmn kmn
XX Bnl cn
þ ð1Þnþl W45 ðlnl ; x2 ; a2 Þ sin cn x3 cos al x1 ;
n l
lnl lnl

we took into account that the displacement must be finite in the direction of the
Ox3 -axis (this fact has an influence on the normal stresses too).
496 11 Elastic Parallelepiped. Elastic Strip. Elastic Layer. Thick Plate

11.2.2 Elastic Layer

Throwing to infinite two opposite faces of an elastic strip, one obtains the elastic
layer jx3 j  a3 ; we shall consider the action of a periodic load, as well as that of a
local load.

11.2.2.1 Elastic Layer Acted Upon by Periodic Loads

We shall deal with the elastic layer subjected to the action of a periodic normal
load, symmetric with respect to the Ox1 an Ox2 co-ordinate axes; any other case of
normal or tangential load may be studied similarly. In what follows, we shall use
the results obtained by us in [38], results which generalize those known for the
wall-beams acted similarly [6].
We will consider both a case of a symmetric load with respect to the Ox1 x2 -
plane and the case of an antisymmetric load with respect to the same plane.
Let thus be the elastic layer jx3 j  a3 ; acted upon by a normal load of the form
(the simple Fourier series are contained in the double Fourier series).
XX
p3 ðx1 ; x2 Þ ¼ a000
00 þ a000
lm cos al x1 cos bm x2 : ð11:32Þ
l m

We introduce stress functions of the form


XX
F11 ¼ L01 x21 þ 0
ðC1lm cosh mlm x3
l m
0
þ mlm x3 C2lm sinh mlm x3 Þ cos al x1 cos bm x2 ; . . .; ð11:33Þ

which are obtained from (11.1), taking L02 ¼ L03 ¼ . . . ¼ L000 0


3 ¼ 0 and A1mn ¼
0 000
A2mn ¼ . . . ¼ B2nl ¼ 0; to determine the unknown coefficients, we put the
boundary conditions
x3 ¼ a3 : r33 ¼ p3 ðx1 ; x2 Þ; r31 ¼ r32 ¼ 0: ð11:34Þ
We may introduce the results given in Sect. 11.1, taking
a000
Amn ¼ Bnl ¼ 0; Clm ¼ ð1Þlþm lm
; ð11:35Þ
vðmlm a3 Þ

The third Eq. (11.21) is thus verified; what concerns the first two equations, they
are replaced by conditions of global equilibrium on the planes x1 ¼ a1 and
x2 ¼ a2 .
It results the state of stress
11.2 Elastic Strip. Elastic Layer 497

"

m X X a000 al 2
000 lm
r11 ¼ a00 þ W4 ðmlm ; x3 ; a3 Þ
1m l m
vðmlm a3 Þ mlm

2 #
bm
þm W5 ðmlm ; x3 ; a3 Þ cos al x1 cos bm x2 ;
mlm
"

m X X a000 bm 2
000 lm ð11:36Þ
r22 ¼ a00 þ W4 ðmlm ; x3 ; a3 Þ
1m l m
vðmlm a3 Þ mlm

2 #
al
þm W5 ðmlm ; x3 ; a3 Þ cos al x1 cos bm x2 ;
mlm
X X a000
r33 ¼ a000
00 þ
lm
W2 ðmlm ; x3 ; a3 Þ cos al x1 cos bm x2 ;
l m
vðm lm a3 Þ

XX a000 bm
lm
r23 ¼ W3 ðmlm ; x3 ; a3 Þ cos al x1 sin bm x2
l m
vðmlm a3 Þ mlm
XX a000 al
lm
r31 ¼ W3 ðmlm ; x3 ; a3 Þ sin al x1 cos bm x2 ; ð11:360 Þ
l m
vðmlm a3 Þ mlm
X X a000 al b
lm m
r12 ¼ W45 ðmlm ; x3 ; a3 Þ sin al x1 sin bm x2
l m
vðmlm a3 Þ m2lm

and the state of displacement


1 X X a000lm al
u1 ¼ W45 ðmlm ; x3 ; a3 Þ sin al x1 cos bm x2 ;
2l l m vðmlm a3 Þ m2lm
1 X X a000lm bm
u2 ¼ W45 ðmlm ; x3 ; a3 Þ cos al x1 sin bm x2 ;
2l l m vðmlm a3 Þ m2lm
ð11:37Þ
1  2m 000
u3 ¼ a x3
2ð1  mÞl 00
1 X X a000 lm 1
þ W16 ðmlm ; x3 ; a3 Þ cos al x1 sin bm x2 ;
2l l m vðmlm a3 Þ mlm

we took into account that the displacements in the direction of the Ox1 and Ox2
axes must be finite and that they have some properties of symmetry with respect to
these axes.
The convergence of these series depend on the factor a000lm ; even in the case of
concentrated loads, the sum of the series may be approximated with a small
number of terms, due to the hyperbolic lines at the denominator. If we approach
the separation planes x3 ¼ a3 ; then the expansion in series becomes slowly
convergent. Taking into account 11.32, we may write the stresses corresponding to
these planes in the form
498 11 Elastic Parallelepiped. Elastic Strip. Elastic Layer. Thick Plate

1  2m 000
r11 ðx1 ; x2 ; a3 Þ ¼ p3 ðx1 ; x2 Þ  a
1  m 00
X X b
2
 ð1  2mÞ m
a000
lm cos al x1 cos bm x2
l m
m lm
X X a2 þ mb2 4mlm a3
 l m
a000 cos al x1 cos bm x2 ;
l m m2lm 2mlm a3 þ sinh 2mlm a3 lm
1  2m 000
r22 ðx1 ; x2 ; a3 Þ ¼ p3 ðx1 ; x2 Þ  a
1  m 00
X X al
2
 ð1  2mÞ a000
lm cos al x1 cos bm x2
l m
m lm
X X b2 þ ma2 4mlm a3
 m
2
l
a000
lm cos al x1 cos bm x2 ;
l m m lm 2m lm a 3 þ sinh 2m lm a 3

ð11:38Þ
X X al b
r12 ðx1 ; x2 ; a3 Þ ¼  ð1  2mÞ m
a000
lm sin al x1 sin bm x2
l m m2lm
XX 4mlm a3 al bm 000
þ ð1  mÞ a sin al x1 sin bm x2 ;
l m
2mlm a3 þ sinh 2mlm a3 m2lm lm
ð11:380 Þ

where intervenes the difference between the external load and the mean one. Thus,
the computation is easy, the above expansions being rapid convergent.
The series which correspond to the displacements are convergent after the
factor a000
lm =mlm ; hence they may be calculated with a smaller number of terms than
the stresses. As well, the deformation of the separation planes is given by
1  2m 000
u3 ðx1 ; x2 ; a3 Þ ¼  a
2ð1  mÞl 00
1 XX 2 sinh2 mlm a3 a000
lm
 ð1  mÞ sin al x1 sin bm x2 :
l l m
sinh 2m lm a 3 þ 2m lm a 3 m lm

ð11:39Þ
In case of a loading symmetric with respect to Ox1 and antisymmetric with
respect to Ox2
XX
p3 ðx1 ; x2 Þ ¼ a000
lm sin al x1 cos bm x2 ; ð11:40Þ
l m

we start from stress functions of the form


11.2 Elastic Strip. Elastic Layer 499

XX
0 0
F11 ¼ ðC1lm cosh mlm x3 þ mlm x3 C2lm sinh mlm x3 Þ sin al x1 cos bm x2 ; . . .
l m
ð11:400 Þ
For a load antisymmetric with respect to Ox1 and symmetric with respect to Ox2
XX
p3 ðx1 ; x2 Þ ¼ a000
lm cos al x1 sin bm x2 ; ð11:41Þ
l m

we use stress functions of the form


XX
0 0
F11 ¼ ðC1lm cosh mlm x3 þ mlm x3 C2lm sinh mlm x3 Þ cos al x1 sin bm x2 ; . . .
l m
ð11:410 Þ
As well, a load antisymmetric with respect to both axes of co-ordinates
XX
p3 ðx1 ; x2 Þ ¼ a000
lm sin al x1 sin bm x2 ð11:42Þ
l m

leads to the stress functions


XX
0 0
F11 ¼ ðC1lm cosh mlm x3 þ mlm x3 C2lm sinh mlm x3 Þ sin al x1 sin bm x2 ; . . .
l m
ð11:420 Þ
The results which one obtains are similar to those given above.
Let be the same elastic layer jx3 j  a3 ; acted upon by a normal load of the form
(11.32), antisymmetric with respect to the Ox1 x2 -plane; the free term is taken equal
to zero, by reasons of global equilibrium of the elastic layer. As in the previous
case, using the notations introduced in Sect. A.4.1.2, one may express the state of
stress in the form
"

X X a000 al 2 0
lm
r11 ¼ w4 ðmlm ; x3 ; a3 Þ
l m
v0 ðmlm a3 Þ mlm

2 #
bm 0
þm w5 ðmlm ; x3 ; a3 Þ cos al x1 cos bm x2 ;
mlm
"

X X a000 bm 2 0
r22 ¼ lm
w4 ðmlm ; x3 ; a3 Þ ð11:43Þ
v 0 ðm a Þ m
l m lm 3 lm

2 #
al 0
þm w5 ðmlm ; x3 ; a3 Þ cos al x1 cos bm x2 ;
mlm
X X a000
r33 ¼ lm
w0 ðm ; x ; a Þ cos al x1 cos bm x2 ;
0 ðm a Þ 2 lm 3 3
l m
v lm 3
500 11 Elastic Parallelepiped. Elastic Strip. Elastic Layer. Thick Plate

XX a000 bm 0
lm
r23 ¼ W3 ðmlm ; x3 ; a3 Þ cos al x1 sin bm x2 ;
l m
v0 ðm lm a3 Þ mlm
XX a000 al 0
lm
r31 ¼ W3 ðmlm ; x3 ; a3 Þ sin al x1 cos bm x2 ; ð11:430 Þ
l m
v0 ðm a Þ
lm 3 lm m
XX a000 al bm 0
lm
r12 ¼  W ðmlm ; x3 ; a3 Þ sin al x1 sin bm x2 ;
l m
v0 ðmlm a3 Þ m2lm 45

the state of displacement being given by


1 X X a000 lm al 0
u1 ¼ W ðmlm ; x3 ; a3 Þ sin al x1 cos bm x2 ;
2l l m v0 ðmlm a3 Þ m2lm 45
1 X X a000 lm bm 0
u2 ¼ W ðmlm ; x3 ; a3 Þ cos al x1 sin bm x2 ; ð11:44Þ
2l l m v ðmlm a3 Þ m2lm 45
0

1 X X a000 1 0
u3 ¼ u03 þ lm
W ðmlm ; x3 ; a3 Þ cos al x1 sin bm x2 ;
2l l m v ðmlm a3 Þ mlm 16
0

where u03 must be specified by certain conditions of support.


The stresses corresponding to the separation planes are given by
X X b
2
r11 ðx1 ; x2 ; a3 Þ ¼  p3 ðx1 ; x2 Þ  ð1  2mÞ m
a000
lm cos al x1 cos bm x2
l m
m lm
X X a2 þ mb2 4mlm a3
 l m
a000 cos al x1 cos bm x2 ;
l m m2lm sinh 2mlm a3  2mlm a3 lm

X X al
2
r22 ðx1 ; x2 ; a3 Þ ¼  p3 ðx1 ; x2 Þ  ð1  2mÞ a000
lm cos al x1 cos bm x2
l m
mlm
X X b2 þ ma2 4mlm a3
 m
2
l
a000
lm cos al x1 cos bm x2 ;
l m m lm sinh 2m lm a 3  2m lm a 3

ð11:45Þ
X X al b
r23 ðx1 ; x2 ; a3 Þ ¼  ð1  2mÞ m
a000
lm sin al x1 sin bm x2
l m m2lm
XX 4mlm a3 al bm 000
 ð1  mÞ a sin al x1 sin bm x2 :
l m
sinh 2mlm a3  2mlm a3 m2lm lm
ð11:450 Þ
As well, the deformation of the separation planes is given by
11.2 Elastic Strip. Elastic Layer 501

1 XX 2 cosh2 mlm a3 a000


u3 ðx1 ; x2 ; a3 Þ ¼ u03 þ ð1  mÞ lm
cos al x1 cos bm x2 :
l l m
sinh 2m lm a 3  2m lm a 3 m lm

ð11:46Þ

11.2.2.2 Elastic Layer Acted Upon by a Local Load

In the case of an elastic layer subjected to the action of a normal load one obtains
results analogous to thus presented in the preceding subsection. We will use the
results obtained by us [37].
Let thus be the elastic layer jx3 j  a3 ; acted upon by a normal load of the form
Z 1Z 1
p3 ðx1 ; x2 Þ ¼ a3 ða1 ; a2 Þ cos a1 x1 cos a2 x2 da1 da2 ; ð11:47Þ
0 0

symmetric with respect to the co-ordinate axes Ox1 and Ox2 . We introduce the
stress functions
Z 1Z 1
F11 ¼ ½C 0 ða1 ; a2 Þ cosh m3 x3
0 0
þm3 x3 C 00 ða1 ; a2 Þ sinh m3 x3  cos a1 x1 cos a2 x2 da1 da2 ; . . .; ð11:48Þ

the variable parameters which are introduced must verify boundary conditions of
the form (11.34). We have denoted
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
m3 ¼ a21 þ a22 : ð11:49Þ

One obtains thus the state of stress


Z 1Z 1 "

a3 ða1 ; a2 Þ a1 2
r11 ¼ W4 ðm3 ; x3 ; a3 Þ
0 0 vðm3 a3 Þ m3

2 #
a2
þm W5 ðm3 ; x3 ; a3 Þ cos a1 x1 cos a2 x2 da1 da2 ;
m3
Z 1Z 1 "

a3 ða1 ; a2 Þ a2 2 ð11:50Þ
r22 ¼ W4 ðm3 ; x3 ; a3 Þ
0 0 vðm3 a3 Þ m3

2 #
a1
þm W5 ðm3 ; x3 ; a3 Þ cos a1 x1 cos a2 x2 da1 da2 ;
m3
Z 1Z 1
a3 ða1 ; a2 Þ
r33 ¼ W2 ðm3 ; x3 ; a3 Þ cos a1 x1 cos a2 x2 da1 da2 ;
0 0 vðm3 a3 Þ
502 11 Elastic Parallelepiped. Elastic Strip. Elastic Layer. Thick Plate

Z 1 Z 1
a3 ða1 ; a2 Þ a2
r23 ¼ W3 ðm3 ; x3 ; a3 Þ cos a1 x1 sin a2 x2 da1 da2 ;
0 0 vðm3 a3 Þ m3
Z 1Z 1
a3 ða1 ; a2 Þ a1
r31 ¼ W3 ðm3 ; x3 ; a3 Þ sin a1 x1 cos a2 x2 da1 da2 ; ð11:500 Þ
0 0 vðm3 a3 Þ m3
Z 1Z 1
a3 ða1 ; a2 Þ a1 a2
r12 ¼ W45 ðm3 ; x3 ; a3 Þ sin a1 x1 sin a2 x2 da1 da2
0 0 vðm3 a3 Þ m23

and the state of displacement


Z Z
1 1 1 a3 ða1 ; a2 Þ a1
u1 ¼ W45 ðm3 ; x3 ; a3 Þ sin a1 x1 cos a2 x2 da1 da2 ;
2l 0 0 vðm3 a3 Þ m23
Z Z
1 1 1 a3 ða1 ; a2 Þ a2
u2 ¼ W45 ðm3 ; x3 ; a3 Þ cos a1 x1 sin a2 x2 da1 da2 ; ð11:51Þ
2l 0 0 vðm3 a3 Þ m23
Z Z
1 1 1 a3 ða1 ; a2 Þ 1
u3 ¼ W16 ðm3 ; x3 ; a3 Þ cos a1 x1 sin a2 x2 da1 da2 ;
2l 0 0 vðm3 a3 Þ m3

the displacements and rotations of rigid body are taken equal to zero because of the
symmetry of the load.
For the separation planes, there correspond the stresses
r11 ðx1 ; x2 ; a3 Þ ¼ pðx1 ; x2 Þ
Z 1 Z 1
2
a2
 ð1  2mÞ a3 ða1 ; a2 Þ cos a1 x1 cos a2 x2 da1 da2
0 0 m3
Z 1Z 1 2
a1 þ ma22 4m3 a3
 a3 ða1 ; a2 Þ cos a1 x1 cos a2 x2 da1 da2 ;
0 0 m23 sinh 2m3 a3 þ 2m3 a3
r22 ðx1 ; x2 ; a3 Þ ¼ pðx1 ; x2 Þ
Z 1 Z 1
2
a1
 ð1  2mÞ a3 ða1 ; a2 Þ cos a1 x1 cos a2 x2 da1 da2
0 0 m3

Z 1 Z 1
a22 þ ma21 4m3 a3
 a3 ða1 ; a2 Þ cos al x1 cos a2 x2 da1 da2 ; ð11:52Þ
0 0 m23 sinh 2m3 a3 þ 2m3 a3
Z 1Z 1
a1 a2
r12 ðx1 ; x2 ; a3 Þ ¼  ð1  2mÞ a3 ða1 ; a2 Þ sin a1 x1 sin a2 x2 da1 da2
0 0 m23
Z 1Z 1
4m3 a3 a1 a2
þ ð1  mÞ 2
a3 ða1 ; a2 Þ sin a1 x1 sin a2 x2 da1 da2 ;
0 0 sinh 2m3 a3 þ 2m3 a3 m3

ð11:520 Þ
while the deformations of the separation planes are given by
11.2 Elastic Strip. Elastic Layer 503

Z1 Z1
1 2 sinh2 m3 a3 a3 ða1 ; a2 Þ
u3 ðx1 ; x2 ; a3 Þ ¼  ð1  mÞ cos a1 x1 cos a2 x2 da1 da2 :
l sinh 2m3 a3 þ 2m3 a3 m3
0 0

ð11:53Þ
For a load symmetric with respect to the Ox1 -axis and antisymmetric with
respect to the Ox2 -axis
Z 1Z 1
p3 ðx1 ; x2 Þ ¼ a3 ða1 ; a2 Þ sin a1 x1 cos a2 x2 da1 da2 ð11:54Þ
0 0

we use stress functions of the form


Z 1Z 1
F11 ¼ ½C 0 ða1 ; a2 Þ cosh m3 x3
0 0
þm3 x3 C00 ða1 ; a2 Þ sinh m3 a3  sin a1 x1 cos a2 x2 da1 da2 ; . . . ð11:540 Þ

If the load is antisymmetric with respect to the Ox1 -axis and symmetric with
respect to the Ox2 -axis
Z 1Z 1
p3 ðx1 ; x2 Þ ¼ a3 ða1 ; a2 Þ cos a1 x1 sin a2 x2 da1 da2 ; ð11:55Þ
0 0

then the stress function is of the form


Z 1Z 1
 0
F11 ¼ C1 ða1 ; a2 Þ cosh m3 x3
0 0

þm3 x3 C20 ða1 ; a2 Þ sinh m3 a3 cos a1 x1 sin a2 x2 da1 da2 ; . . . ð11:550 Þ

In the case of a load antisymmetric with respect to both axes of co-ordinates


Z 1Z 1
p3 ðx1 ; x2 Þ ¼ a3 ða1 ; a2 Þ sin a1 x1 sin a2 x2 da1 da2 ð11:56Þ
0 0

we introduce stress functions of the form


Z 1Z 1
 0
F11 ¼ C1 ða1 ; a2 Þ cosh m3 x3
0 0

þm3 x3 C20 ða1 ; a2 Þ sinh m3 x3 sin a1 x1 sin a2 x2 da1 da2 ; . . . ð11:560 Þ

One obtains thus results similar to those above.


504 11 Elastic Parallelepiped. Elastic Strip. Elastic Layer. Thick Plate

We can make analogous considerations for the elastic layer jx3 j  a3 acted upon
by the normal load (11.47), antisymmetric with respect to the Ox1 x2 -plane.
We obtain thus the state of stress
Z 1Z 1 "

a3 ða1 ; a2 Þ a1 2 0
r11 ¼ W4 ðm3 ; x3 ; a3 Þ
0 0 v0 ðm3 a3 Þ m3

2 #
a2 0
þm W5 ðm3 ; x3 ; a3 Þ cos a1 x1 cos a2 x2 da1 da2 ;
m3
Z 1Z 1 "

a3 ða1 ; a2 Þ a2 2 0
r22 ¼ W4 ðm3 ; x3 ; a3 Þ ð11:57Þ
0 0 v0 ðm3 a3 Þ m3

2 #
a1 0
þm W5 ðm3 ; x3 ; a3 Þ cos a1 x1 cos a2 x2 da1 da2 ;
m3
Z 1Z 1
a3 ða1 ; a2 Þ 0
r33 ¼ W ðm3 ; x3 ; a3 Þ cos a1 x1 cos a2 x2 da1 da2 ;
0 0 v0 ðm3 a3 Þ 2
Z 1Z 1
a3 ða1 ; a2 Þ a2 0
r23 ¼ W ðm3 ; x3 ; a3 Þ cos a1 x1 sin a2 x2 da1 da2 ;
0 0 v0 ðm3 a3 Þ m3 3
Z 1Z 1
a3 ða1 ; a2 Þ a1 0
r31 ¼ W ðm3 ; x3 ; a3 Þ sin a1 x1 cos a2 x2 da1 da2 ; ð11:570 Þ
0 0 v0 ðm3 a3 Þ m3 3
Z 1Z 1
a3 ða1 ; a2 Þ a1 a2 0
r12 ¼  W ðm3 ; x3 ; a3 Þ sin a1 x1 sin a2 x2 da1 da2
0 0 v0 ðm3 a3 Þ m23 45

and the state of displacement


Z Z
1 1 1 a3 ða1 ; a2 Þ a1 0
u1 ¼ W ðm3 ; x3 ; a3 Þ sin a1 x1 cos a2 x2 da1 da2 ;
2l 0 0 v0 ðm3 a3 Þ m23 45
Z Z
1 1 1 a3 ða1 ; a2 Þ a2 0
u2 ¼ W ðm3 ; x3 ; a3 Þ cos a1 x1 sin a2 x2 da1 da2 ; ð11:58Þ
2l 0 0 v0 ðm3 a3 Þ m23 45
Z Z
1 1 1 a3 ða1 ; a2 Þ 1 0
u3 ¼ W ðm3 ; x3 ; a3 Þ cos a1 x1 cos a2 x2 da1 da2 :
2l 0 0 v0 ðm3 a3 Þ m3 16
11.2 Elastic Strip. Elastic Layer 505

For the separation planes correspond the stresses


r11 ðx1 ; x2 ; a3 Þ ¼  p3 ðx1 ; x2 Þ
Z 1 Z 1
2
a2
 ð1  2mÞ a3 ða1 ; a2 Þ cos a1 x1 cos a2 x2 da1 da2
0 0 m3
Z 1Z 1 2
a1 þ ma22 4m3 a3
 a3 ða1 ; a2 Þ cos a1 x1 cos a2 x2 da1 da2 ;
0 0 m23 sinh 2m3 a3  2m3 a3
r22 ðx1 ; x2 ; a3 Þ ¼  p3 ðx1 ; x2 Þ
Z 1 Z 1
2
a1
 ð1  2mÞ a3 ða1 ; a2 Þ cos a1 x1 cos a2 x2 da1 da2 ð11:59Þ
0 0 m3
Z 1Z 1 2
a2 þ ma21 4m3 a3
 2
a3 ða1 ; a2 Þ cos al x1 cos a2 x2 da1 da2 ;
0 0 m3 sinh 2m 3 a3  2m3 a3

Z 1Z 1
a1 a2
r12 ðx1 ; x2 ; a3 Þ ¼  ð1  2mÞ a3 ða1 ; a2 Þ sin a1 x1 sin a2 x2 da1 da2
0 0 m23
Z 1Z 1
4m3 a3 a1 a2
 ð1  mÞ 2
a3 ða1 ; a2 Þ sin a1 x1 sin a2 x2 da1 da2
0 0 sinh 2m a
3 3  2m 3 3 m3
a
ð11:590 Þ

and the deformation of those planes is given by


Z 1 Z 1
1 2 cosh2 m3 a3 a3 ða1 ; a2 Þ
u3 ðx1 ; x2 ; a3 Þ ¼ ð1  mÞ cos a1 x1 cos a2 x2 da1 da2 :
l 0 0 sinh 2m3 a3  2m3 a3 m3
ð11:60Þ

11.3 Thick Plate

The first approximation which can be made concerning the thick plate is that of
considering it as infinite, i.e. as an elastic layer. The first researches of the elastic
layer subjected to the action of local loads are due to Dougall [13] and L. Orlando
[34], who use the method of reflection. We mention an interesting paper of
J. N. Sneddon [35] too. We mention also a presentation of those problems made by
A. I. Lur’e [5] in his treatise, which contains the studies [26, 27] made by himself
in this direction too.
In the case of finite plates one uses usual the hypothesis of linear element
(normal to the middle plane, which remains linear, normal to the deformed middle
plane, without linear deformations) of G. R. Kirchhoff [3]; the solving of the
problem, on the basis of this supplementary hypothesis (of the nature of the
hypotheses of strength of materials), is reduced to the integration of the Lagrange-
Sophie Germain partial differential equation, with certain boundary conditions,
corresponding to the supports of the plate.
506 11 Elastic Parallelepiped. Elastic Strip. Elastic Layer. Thick Plate

Fig. 11.1 Rectangular thick


x3
plate: sectional efforts
x2
M11
N12
M12 N11
T11

a3
M22 T22

N21 O x1
M21

a3
N 22

The thick plate for which this hypothesis is no more verified has been less
studied; but such plates can intervene in engineering. We mention thus the study of
B. G. Galerkin [19], concerning the simply supported circular thick plate, pub-
lished in 1932, and the study, published in 1935 by the same author [20], con-
cerning the thick rectangular plate. One gives both qualitative and quantitative
results, with some displacement functions; but these functions have a form more
complicated as necessary. A direct solution, in displacements too, of the problem
has given, in 1933, S. Woinowski-Krieger [41], searching solutions in form of
trigonometric series expansions, which verify Lamé’s equations. Other studies in
this direction have been made by E. Almansi [8]. In 1955, L. Bonneau [12] dealt
again with the problem, especially with the integration of the equations of equi-
librium and of those of continuity of Beltrami for the thick plate, making various
considerations of practical order. As well, M. Misßicu [30] presented some results
with a general character, as an application to the representation of the solution in
displacements of the spatial problem of the theory of elasticity by means of some
monogenic functions.
M. Haimovici [21] used Galerkin’s method, starting from functions of the form
X
n
Ak ðx1 ; x2 ÞPk ðx3 Þ; ð11:61Þ
k¼1

where Pk ðx3 Þ is a complete system of function on the thickness of the plate. By


means of Lamé’s equations, one obtains for Ak a system of 3n differential equa-
tions of elliptic type. The potential energy converges to the potential energy of the
exact three-dimensional theory for n ! 1. We [40] indicated a computation
method in the frame of the bidimensional problems (which are reduced to the
determination of functions of only two variables) of the theory of elasticity. As
well, we have shown [39] how one may use biharmonic polynomials (elementary
methods of computations) to study a rectangular thick plate; in what follows, we
give some results in this direction.
11.3 Thick Plate 507

11.3.1 General Considerations

Let be a rectangular thick plate of dimensions 2a1 and 2a2 and of thickness 2a3 ; we
denote the ratios a1 ¼ a1 =a3 ; a2 ¼ a2 =a3 and a ¼ a1 =a2 ¼ a1 =a2 : Our study is, in
general, valid in the frame of Saint-Venant’s principle, hence we shall assume that
a1 ; a2  2.

11.3.1.1 Sectional Efforts

In the computation, one uses, usually, the sectional efforts along the thickness of
the plate (Fig. 11.1).
We introduce thus the bending moments
Z a3
M11 ðx1 ; x2 Þ ¼  r11 ðx1 ; x2 ; x3 Þx3 dx3 ;
a
Z a33 ð11:62Þ
M22 ðx1 ; x2 Þ ¼  r22 ðx1 ; x2 ; x3 Þx3 dx3
a3

and the moment of torsion


Z a3
M12 ðx1 ; x2 Þ ¼ M21 ðx1 ; x2 Þ ¼  r12 ðx1 ; x2 ; x3 Þx3 dx3 ; ð11:620 Þ
a3

the shearing forces are given by


Z a3
T11 ðx1 ; x2 Þ ¼  r13 ðx1 ; x2 ; x3 Þdx3 ;
a3
Z a3 ð11:63Þ
T22 ðx1 ; x2 Þ ¼  r23 ðx1 ; x2 ; x3 Þdx3 :
a3

In the middle plane appear the axial forces


Z a3
N11 ðx1 ; x2 Þ ¼ r11 ðx1 ; x2 ; x3 Þdx3 ;
a
Z a33 ð11:630 Þ
N22 ðx1 ; x2 Þ ¼ r22 ðx1 ; x2 ; x3 Þdx3
a3

and the sliding forces


Z a3
N12 ðx1 ; x2 Þ ¼ N21 ðx1 ; x2 Þ ¼ r12 ðx1 ; x2 ; x3 Þdx3 ; ð11:6300 Þ
a3

On a normal section, parallel to the co-ordinate axes, we introduce also the


generalized shearing forces
508 11 Elastic Parallelepiped. Elastic Strip. Elastic Layer. Thick Plate

0
T11 ðx1 ; x2 Þ ¼ T11 ðx1 ; x2 Þ þ M12;2 ðx1 ; x2 Þ;
0 ð11:64Þ
T22 ðx1 ; x2 Þ ¼ T22 ðx1 ; x2 Þ þ M21;1 ðx1 ; x2 Þ;

where the comma indicates the differentiation with respect to the variable at the
right of it.
The concentrated reactions at the vertices of the plate are given by
V ¼ 2M12 ða1 ; a2 Þ: ð11:65Þ

11.3.1.2 Stress Functions. Boundary Conditions

We consider a thick plate acted upon by a uniform distributed normal load of


intensity p. We decompose this case of loading in a symmetric one with respect to
the Ox1 x2 -plane (a simple compression), for which we have
p
r11 ¼ r22 ¼ 0; r33 ¼  ; ð11:66Þ
2
r23 ¼ r31 ¼ r12 ¼ 0 ð11:660 Þ

and in an antisymmetric one with respect to the same plane. The last case remains
to be studied.
We choose the stress functions

F11 ¼ A1 x21 x3 þ A2 x22 x3 þ A3 x33 þ A4 x41 x3 þ A5 x42 x3 þ A6 x21 x22 x3


1 1
þ A7 x21 x33 þ A8 x22 x33  A1 þ A5 þ A6 þ A7 þ A8 x53 ; . . .; ð11:67Þ
5 3

using the biharmonic polynomials considered in Sect. 5.2.1.3. One must thus
determine the constants Ai ; Bi ; Ci ; i ¼ 1; 2; . . .; 8:
The functions F23 and F31 contain x1 and x2 , respectively, at the zeroth power,
so that they cannot be odd with respect to these variables. But r23 and r31 are odd
functions with respect to x2 and to x1 , respectively, because of the conditions of
symmetry. Hence, we must take
F23 ¼ F31 ¼ 0: ð11:68Þ
The harmonic function F12 must be odd with respect to both variables; we may
take
F12 ¼ E1 x1 x2 þ E2 ðx31 x2  x1 x32 Þ: ð11:680 Þ
The normal stresses are given by
11.3 Thick Plate 509

r11 ¼ 2½ð2 þ mÞðA1 þ A2 þ A3 Þ  D1 x3 þ 2½ð2 þ mÞð6A4 þ A6 þ 3A7 Þ  6D4 x21 x3


þ 2½ð2 þ mÞð6A5 þ A6 þ 3A8 Þ  D6 x22 x3


A6
 2 ð2 þ mÞ 2 þ A4 þ A5 þ A7 þ A8 þ D7 x33 ;
3
r22 ¼ 2½ð2 þ mÞðB1 þ B2 þ 3B3 Þ  D2 x3
þ 2½ð2 þ mÞð6B4 þ B6 þ 3B7 Þ  D6 x21 x3
þ 2½ð2 þ mÞð6B5 þ B6 þ 3B8 Þ  6D5 x22 x3


B6
 2 ð2 þ mÞ 2 þ B4 þ B5 þ B7 þ B8 þ D8 x33 ; ð11:69Þ
3
r33 ¼ 2f½ð2 þ mÞðC1 þ C2 þ 3C3 Þ  3D3 x3
þ 2½ð2 þ mÞð6C4 þ C6 þ 3C7 Þ  3D7 gx21 x3
þ 2½ð2 þ mÞð6C5 þ C6 þ 3C8 Þ  3D8 x22 x3


C6
 2 ð2 þ mÞ 2 þ C4 þ C5 þ C7 þ C8
3


D6
 2 D4 þ D5 þ þ D7 þ D8 x33
3

and the tangential stresses are expressed in the form


r23 ¼  fð2 þ mÞ½C1 þ C2 þ 3C3  ðL þ E1 Þ  ðD1  D2 þ 3D3 Þgx2
 fð2 þ mÞ½6ðA4 þ C4 Þ þ A6 þ C6 þ 3ðA7 þ C7 Þ  ðM þ 3E2 Þ
 6ð2D4 þ D7 Þgx21 x2 þ fð2 þ mÞ½2C6 þ 6ðA4  B5 þ C4 þ C5 Þ
þ A6  B6 þ 3ðA7  B8 þ C7 þ C8 Þ  2½6D4 þ D6 þ 3ðD7 þ D8 Þgx23 x2

1 1
 ð2 þ mÞ 2ðC5  B5 Þ þ ðC6  B6 Þ þ C8  B8 þ ðM þ 3E2 Þ þ 4D5 x32 ;
3 3
r31 ¼  ½ð2 þ mÞðC1 þ C2 þ 3C3 þ L þ E1 Þ  ðD2  D1 þ 3D3 Þx1
ð11:690 Þ
 fð2 þ mÞ½6ðB5 þ C5 Þ þ B6 þ C6 þ 3ðB8 þ C8 Þ  ðM þ 3E2 Þ
 6ð2D5 þ D8 Þgx1 x22 þ fð2 þ mÞ½2C6 þ 6ðB5  A4 þ C4 þ C5 Þ
þ B6  A6 þ 3ðB8  A7 þ C7 þ C8 Þ  2½6D5 þ D6 þ 3ðD7 þ D8 Þgx23 x1

1 1
 ð2 þ mÞ 2ðC4  A4 Þ þ ðC6  A6 Þ þ C7  A7 þ ðM þ 3E2 Þ þ 4D4 x21
3 3
r12 ¼  2fð2 þ mÞ½2C6 þ 6ðA4 þ B5 þ C4 þ C5 Þ þ A6 þ B6 
þ 3ðA7 þ B8 þ C7 þ C8 Þ  2½D6 þ 6ðD4 þ D5 Þ þ 3ðD7 þ D8 Þgx1 x2 x3 ;

where L and M are other constants which must be determined and where we have
used the notation
Di ¼ Ai þ Bi þ Ci ; i ¼ 1; 2; . . .; 8: ð11:70Þ
510 11 Elastic Parallelepiped. Elastic Strip. Elastic Layer. Thick Plate

x3 3 p 3 p 3ν
p α ξ α ξ α αξ ξ
8 1 1 8 2 2 4 1 2 1 2
σS
σ11E

0.448a3
11

2+ν p

a3
20 5

O x1, x2

0.775a3

a3
p
2

σ11 2+ν p σ33 τ13 τ23 τ12


20 (x1 >0) (x 2 >0) ( x1, x2 >0)

Fig. 11.2 Rectangular thick plate: stresses

To specify the parameters which have been introduced, we put the boundary
conditions

x1 ¼ a1 : N11 ¼ 0; N12 ¼ 0; M11 ¼ 0;


ð11:71Þ
x2 ¼ a2 : N22 ¼ 0; N21 ¼ 0; M22 ¼ 0;
p
x3 ¼ a3 : r33 ¼  ; r31 ¼ r32 ¼ 0; ð11:710 Þ
2
there are less equations than constants to be determined, but the latter ones may be
grouped so that the state of stress be uniquely determined.

11.3.2 State of Strain and Stress

By the aid of the coefficients determined as it has been shown above, the state of
strain and stress in the considered thick plate may be specified.

11.3.2.1 State of Stress

The state of stress will be expressed in the form

rij ¼ rSij þ rEij ; i; j ¼ 1; 2; 3; ð11:72Þ

where we have denoted by S the results obtained by the methods of the strength of
materials and by E the corrections given by the theory of elasticity.
We introduce the reduced co-ordinates
xi
ni ¼ ; i ¼ 1; 2; 3: ð11:73Þ
ai
11.3 Thick Plate 511

The state of stress is given by (Fig. 11.2)


3
rSii ¼  ð1 þ mÞa2i pð1  n2i Þn3 ; i ¼ 1; 2; rS33 ¼ 0 ð11:74Þ
8
and by

1 3 p
rE11 ¼ rE22 ¼ ð2 þ mÞp  n23 n3 ; rE33 ¼  ð1 þ n3 Þ2 ð2  n3 Þ ð11:740 Þ
8 5 4

for the normal stresses, where we took into account (11.66) too; for the tangential
stresses, one may write
3 3
rS23 ¼ a2 pð1  n23 Þn2 ; rS31 ¼ a1 pð1  n23 Þn1 ;
8 8 ð11:75Þ
S 3m
r12 ¼  a1 a2 pn1 n2 n3
4
as well as

rE23 ¼ rE31 ¼ rE12 ¼ 0: ð11:750 Þ


What concerns the normal stresses r11 and r22 , the theory of elasticity brings
the same correction for all the cross sections. In the limit cases a1 ¼ 2 or a2 ¼ 2,
one may write
 
rS11 ð0; n2 ; 1Þa1 ¼2 ¼ rS22 ðn1 ; 0; 1Þa2 ¼2 ¼ ð1:50. . .2:25Þp; ð11:76Þ

as well

rE11 ðn1 ; n2 ; 1Þ ¼ rE22 ðn1 ; n2 ; 1Þ ¼ ð1:100. . .0:125Þp: ð11:760 Þ

Hence, the maximal difference is less than 7 %, whatever be m, so that the dia-
grams of the stresses r11 and r22 remain very near to the linear ones.
The normal stress r33 is equal to zero in the frame of the strength of materials,
but in the frame of the theory of elasticity has a variation after a parabola of third
order, being equal to p=2 in the middle plane.
The tangential stresses have variatiations analogous to those obtained by the
methods of the strength of materials.

11.3.2.2 State of Efforts

The efforts on the cross section are given by


N11 ðx1 ; x2 Þ ¼ N22 ðx1 ; x2 Þ ¼ N12 ðx1 ; x2 Þ ¼ N21 ðx1 ; x2 Þ ¼ 0; ð11:77Þ
1
Mii ðx1 ; x2 Þ ¼ ð1 þ mÞpa2i ð1  n2i Þ; i ¼ 1; 2; ð11:78Þ
4
512 11 Elastic Parallelepiped. Elastic Strip. Elastic Layer. Thick Plate

1
M12 ðx1 ; x2 Þ ¼ M21 ðx1 ; x2 Þ ¼ mpa1 a2 n1 n2 ; ð11:780 Þ
2
1
Tii ðx1 ; x2 Þ ¼  pai ni ; i ¼ 1; 2; ð11:79Þ
2
1
Tii0 ðx1 ; x2 Þ ¼  ð1  mÞpai ni ; i ¼ 1; 2: ð11:790 Þ
2
The reactions of the four sides read

0 1 0 1
T11 ða1 ; x2 Þ ¼  ð1  mÞa1 ; T22 ðx1 ; a2 Þ ¼  ð1  mÞa2 ð11:80Þ
2 2
and the reactions concentrated at the vertices become
V ¼ mpa1 a2 : ð11:800 Þ
The global equilibrium of the plate may by easily verified.

11.3.2.3 State of Displacement

Taking into account the conditions of symmetry in the case of the motion of rigid
body, it results the state of displacement in the form
u1 1
E ¼  ð1 þ mÞp½a21 ð3  n21 Þ  3ma22 ð1  n22 Þ þ ð2  mÞn23 n3 n1
a1 8
3 m
þ ð2 þ 9m  m2 Þpn3 n1 þ pn1 ;
40 2 ð11:81Þ
u2 1
E ¼  ð1 þ mÞp½a2 ð3  n2 Þ  3ma21 ð1  n21 Þ þ ð2  mÞn23 n3 n2
2 2
a2 8
3 m
þ ð2 þ 9m  m2 Þpn3 n2 þ pn2 ;
40 2


u3  u03 1 6
E ¼ ð1 þ mÞp 3m½a21 ð1  n21 Þ þ a22 ð1  n22 Þ  ð1 þ mÞ  n23 n23
a3 16 5

p 3 3 1
 1 þ n3 n3 þ mð1 þ mÞa21 a22 pð1  n21 Þð1  n22 Þ þ ð1 þ mÞp½a41 ð6  n21 Þn21
2 5 16 32
3
þ a22 ð6  n22 Þn22  þ ð8 þ m þ m2 Þpða21 n21 þ a22 n22 Þ: ð11:810 Þ
80

The displacement u03 is determined by conditions of support. One observer that


one cannot put conditions of simple support, so that one must imagine that the
reactions (11.80), (11.800 ) have another provenance.
The deformed middle surface is given by
11.3 Thick Plate 513

u3 ðx1 ; x2 ; 0Þ  u03 3
E ¼ mð1 þ mÞa21 a22 pð1  n21 Þð1  n22 Þ
a3 16
1
þ ð1 þ mÞp½a41 ð6  n21 Þn21 þ a42 ð6  n22 Þn22 
32
3
þ ð8 þ m þ m2 Þpða21 n21 þ a22 n22 Þ: ð11:82Þ
80
The conditions of simple support at the vertices
n1 ¼ 1; n2 ¼ 1; n3 ¼ 0 : u3 ¼ 0 ð11:83Þ

lead to
pa3
u03 ¼ ½25ð1 þ mÞða41 þ a42 Þ þ 6ð8 þ m þ m2 Þpða21 þ a22 Þ; ð11:830 Þ
160E
the displacement of rigid body being thus completely determined.

References

A. Books

1. Filonenko-Borodich, M.M.: Teoriya uprugosti (Theory of Elasticity). 4th edn. Fizmatgiz,


Moskva (1959)
2. Föppl, A., Föppl, L.: Drang und Zwang. Eine Höhere Festigkeitslehre für Ingenieure. I.Druck
und Verlag von R. Oldenbourg, München (1920)
3. Kirchhoff, G.R.: Gesammelte Abhandlungen. Leipzig (1882)
4. Lamé, G.: Leçons sur la théorie mathématique de l’élasticité des corps solides. Paris (1852)
5. Lur’e, A.I.: Prostranstvennye zadachi teorii uprugosti (Spatial Problems of the Theory of
Elasticity). Fizmatgiz, Moskva (1955)
6. Teodorescu, P.P.: Probleme plane în teoria elasticitătßii (Plane Problems in the Theory of
Elasticity). I. Ed. Acad., Bucuresßti (1961)
7. Teodorescu, P.P.: Probleme spatßiale în teoria elasticitătßii (Spatial Problems in the Theory of
Elasticity). I. Ed. Acad., Bucuresßti (1970)

B. Papers

8. Almansi, E.: Sulle deformazioni delle piastre elastiche. Atti R. Accad. dei Lincei, Rend., Cl.
Sci. fis., mat. nat., ser. 6, 17, 12 (1933)
9. Baı̆da, Ē.N.: Obshchee reshenie uravneniı̆ ravnoveciya i padacha o parallelipipede dlya
anizotropnogo i izotropnogo tela (General solution of the equilibrium equations and the
problem of the parallelepiped for anisotropic and isotropic bodies). Sb. Dokl. Leningr. inzh.-
stroit. inst. 611 (1958)
514 11 Elastic Parallelepiped. Elastic Strip. Elastic Layer. Thick Plate

10. Baı̆da, Ē.N.: Zadacha o uprugo-deformirevannem sostoyanii ortotropnogo i ipotropnogo


parallelepipedov (Problem of the state of elastic deformation of the orthotropic and isotropic
parallelepiped). Izd. vycih. uchebn. paved. stroit. i arkh. 6 (1959)
11. Baı̆da, Ē.N.: Obshchie resheniya teorii urpugosti i padachi o parallelepipede i tsilindre
(General solution of the theory of elasticity and of the problem of the parallelepiped and of
the cylinder). Gosstroı̆pdat (1961)
12. Bonneau, L.: L’équation de Lagrange et les plaques rectangulaires. Ann. des Ponts et
Chaussées 125, 403 (1955)
13. Dougall, J.: An analytical theory of the equilibrium of an isotropic elastic plate. Trans. Roy.
Soc. Edinborough, 41, 8 (1904)
14. Filonenko-Borodich, M.M.: Ob odnoı̆ sisteme funktsii i ee prilozheniyakh v teorii uprugosti
(On a system of functions and on their applications in the theory of elasticity). Prikl. mat. i
mekh. 10, 193 (1946)
15. Filonenko-Borodich, M.M.: Zadacha o ravnovesii uprugogo parallelipipeda pri zadannykh
nagruzkakh na ego granyakh (Problem of the equilibrium of the elastic parallelepiped for
given lods on its boundary). Prikl. mat. i mekh. 15, 137 (1951)
16. Filonenko-Borodich, M.M.: Dve padachi o ravnovesii uprugogo parallelepipeda (Two
equilibrium problems of the elastic parallelepiped). Prikl. mat. i mekh. 15, 563 (1951)
17. Filonenko-Borodich, M.M.: Nekotorye obobshcheniya padachi Lame dlya uprugogo
parallelepipeda (Some general Lamé problems for the elastic parallelepiped). Prikl. mat. i
mekh. 17, 465 (1953)
18. Filonenko-Borodich, M.M.: O zadache Lame dlya parallelepipeda v obschem sluchae
poverkhnostnykh nagrupok (On the Lamé problem of the parallelepiped in the general case of
surface loading). Prikl. mat. i mekh. 21, 550 (1957)
19. Galerkin, B.G.: Sur l’équilibre d’une plaque circulaire épaisse et d’une plaque en forme de
secteur circulaire. C. Rend. Acad. Sci. Paris 194, 1440 (1932)
20. Galerkin, B.G.: Napryazhennoe sostoyanie pri ipgibe priamougol’noı̆ plity i teorii plit tonkikh
(State of stress of a rectangular plate subjected to bending and the theory of thin plates). Tr.
Leningr. ni-pa soor. 22 (1935)
21. Haimovici, M.: On the bending of elastic plates. Bull. Pol. Acad. Sci. Sér. Sci. Techn. XIV,
605 (1966)
22. Kaliski, S.: The dynamical problem of the rectangular parallelepiped. Arch. Mech. Stos. 10,
329 (1958)
23. Kaliski, S.: Non-steady forced vibration of a rectangular parallelepiped. Arch. Mech. Stos.
10, 727 (1958)
24. Kononenko, E.S.: Zadacha o szhatii parallelepipeda mezhdu zhestkimi plitami bez
skol’zheniya (Problem on the compression of a parallelepiped between rigid plates without
sliding). Sb. Issled. po teorii sooruzheniı̆ 6 (1954)
25. Kononenko, E.S.: Raspredelenie napryazhenii v uprugom prizmaticheskom obraztse pri
ispytanii na statike s nalichiem sil treniya na tortsakh (Stresses in a prismatic domain
subjected to forces of friction). Sb. Issled. po teorii soor. 437 (1957)
26. Lur’e, A.I.: K zadache o ravnovesii plastiny peremennoı̆ tolshchiny (On the problem of the
equilibrium of a plate of variable thickness). Tr. Leningr. industr. inst. 57 (1936)
27. Lur’e, A.I.: K teorii tolstykh plit (On the theory of thick plates). Prikl. mat. i mekh. 6, 151
(1942)
28. Meshkov, A.I.: Ravnovesie uprugogo parallelepipeda (Equilibrium of the elastic
parallelepiped). Vestnik Mosk. Univ., ser. mat.-mekh, astron. fiz. him. 12, 35 (1957)
29. Meshkov, A.I.: Obshchee reshenie padachi o kruchenii kosougol’nogo parallelepipeda
(General solution of the problem of the torsion of the oblique parallelepiped). Vestnik Mosk.
Univ., ser. mat.-mekh, astron. fiz. him. 14, 43 (1959)
30. Misßicu, M.: On the solving of the spatial problem of the theory of elasticity. Applications to
the theory of plates. Rev. Méc. Appl. II, 171 (1957)
References 515

31. Mishonov, M.: Obshch metod za reshenie na prostranstvenata zadacha v teoriya na


elastichnostta za parallelepipeda (General method of solving the spatial problem of the theory
of elasticity for the parallelepiped). B’’lgarska Akad. na nauk., Izv. na tekhn. inst. Sofiya, 9–
10, 15 (1960)
32. Netrebko, V.P.: Kruchenie uprugogo parallelepipeda (Torsion of the elastic parallelepiped).
Vest. Mosk. Univ. ser. fiz.-mat. i estestv. nauk, 15 (1954)
33. Netrebko, V.P.: Stesnennoe kruchenie uprugogo parallelepipeda (State of torsion of the
elastic parallelepiped). Vest. Mosk. Univ. ser. fiz.-mat. i estestv. nauk, 11 (1956)
34. Orlando, L.: Sulla deformazione di un solido isotropo limitato da due piani paralleli per
tensioni superficiali date. Rend. del circ. mat. di Palermo, 66 (1905)
35. Sneddon, J.N.: The elastic stresses produced in a thick plate by the applications of pressure to
the free surfaces. Proc. Cambridge Phil. Soc. 42, 260 (1946)
}
36. Teodorescu, P.P.: Uber die Berechnung dicker ebener Platten unter örtlicher Belastung. Bull.
Math. Soc. Sci. Math. Phys. de Roum. 2, 463 (1958)
37. Teodorescu, P.P.: On the strength calculation of thick plane plates. Rev. Méc. Appl. IV, 323
(1959)
38. Teodorescu, P.P.: Sur le problème du parallelepipede élastique. Arch. Mech. Stos. XII, 705
(1960)
39. Teodorescu, P.P.: Uber } die Anwendung der elementaren Rechenmethoden auf die
dreidimensionalen Probleme der Elastizitätstheorie. Rev. Méc. Appl. V, 649 (1960)
40. Teodorescu, P.P.: Consideratßii în legătură cu calculul plăcilor plane groase (Considerations
concerning the computation of thick plates). Bul. ßst. Inst. Polit. Cluj, XI, 371 (1968)
41. Woinowski-Krieger, S.: Der Spannungszustand in dicken elastischen Platten. Ing. Arch. 4,
203 (1933)
Chapter 12
Dynamical Problems of Elastic Bodies

Hereafter, we shall study only a few problems of elastodynamics, i.e.: axisym-


metrical problems, considerations on progressive waves, free, forced and charac-
teristic vibrations, as well as waves propagations in an infinite medium.

12.1 Axisymmetrical Problems

In what follows we deal with the axially symmetrical problem, in cylindrical


co-ordinates, by making use of the results given in Sects. 2.2.4.2, 3.2.5.2 and
2.3.2.4 for the state of strain and stress, respectively, and the constitutive law in
this system of co-ordinates.
Owing to the axial symmetry, all the results supplied by the above-mentioned
subsections must be independent of the variable h. Thus, we obtain
uh ¼ 0; chz ¼ crh ¼ 0; rhz ¼ rrh ¼ 0; ð12:1Þ

which will bring certain simplifications in the computation.


Formulations of this problem in displacements were given by A.-E.-H. Love [9]
in the static case and by W. Noll [24] in the case of small stationary motions;
M. Predeleanu [28] resumed the problem, in 1958, and gave a representation of the
Somigliana-Iacovache type and a representation generalizing the Love represen-
tation in the general dynamic case.

12.1.1 Formulation in Displacements of the Limit Problem

We shall state the axially symmetrical problem of elastodynamics in displace-


ments and specify the corresponding differential equations, as well as the limit
conditions that are imposed.

P. P. Teodorescu, Treatise on Classical Elasticity, Mathematical and Analytical 517


Techniques with Applications to Engineering, DOI: 10.1007/978-94-007-2616-1_12,
Ó Springer Science+Business Media Dordrecht 2013
518 12 Dynamical Problems of Elastic Bodies

12.1.1.1 The Differential Equations of the Problem

Taking into account (12.1), the results given in the above-mentioned subsections
lead to the following relations between strains and displacements
our ur ouz
err ¼ ; ehh ¼ ; ezz ¼ ; ð12:2Þ
or r oz
ouz our
czr ¼ þ ; ð12:20 Þ
or oz
to the cubical dilatation
our ur ouz 1 o ouz
h¼ þ þ ¼ ðrur Þ þ ; ð12:3Þ
or r oz r or oz
to the equations of motion
orrr orzr 1
þ þ ðrrr  rhh Þ þ Fr ¼ q€ur ;
or oz r
ð12:4Þ
orrz orzz 1
þ þ rzr þ Fz ¼ q€uz
or oz r
and to Hooke’s law (4.1.110), to which is added
1
czr ¼ rzr : ð12:5Þ
G
By eliminating the strains and the stress, we find Lamé’s equations
1 oh 1 oh
h02 uR þ ¼ 0; h02 uz þ ¼ 0; ð12:6Þ
1  2m or 1  2m oz
in the absence of the volume forces; here, beside the classical d’Alembert’s
operators, we introduced the operators

1 o2
h0k ¼ M0  ; k ¼ 1; 2; ð12:7Þ
c2k ot2

with
1
M0 ¼ M  ; ð12:8Þ
r2
where Laplace’s operator is given by
 
1o o o2 o2 1 o o2
M¼ r þ 2¼ 2þ þ 2 ð12:80 Þ
r or or oz or r or oz
12.1 Axisymmetrical Problems 519

and where ck  k ¼ 1; 2, are the propagation velocities given by (5.3.4), (5.3.40 ).


Between the newly-introduced differential operators, we can write the relations

2ð1  mÞh01 ¼ M0 þ ð1  2mÞh02 : ð12:70 Þ

12.1.1.2 Limit Conditions

For t  t0 we shall lay the boundary conditions in stresses in the form

pnr ¼ rrr cosðn; rÞ þ rzr cosðn; zÞ;


ð12:9Þ
pnz ¼ rrz cosðn; rÞ þ rzz cosðn; zÞ

or in displacement in the form


n n
ur ¼ ur ; uz ¼ uz ð12:90 Þ

n being the unit vector of the external normal to the area element considered.
At the time t ¼ t0 , we can consider initial conditions for the whole body, in
displacements and displacements velocities, in the form

ur ¼ u0r ; uz ¼ u0z ; ð12:10Þ

u_ r ¼ u_ 0r ; u_ z ¼ u_ 0z ð12:100 Þ

or, in stresses and stress rates,

rrr ¼ r0rr ; rhh ¼ r0hh ; rzz ¼ r0zz ; rzr ¼ r0zr ; ð12:11Þ

r_ rr ¼ r_ 0rr ; r_ hh ¼ r_ 0hh ; r_ zz ¼ r_ 0zz ; r_ zr ¼ r_ 0zr : ð12:110 Þ

We can thus formulate the three fundamental problems of elastodynamics for the
problem with axial symmetry with two space variables.

12.1.2 Solutions by Potential Functions

Hereafter, we shall give a representation of the Somigliana-Iacovache type, due to


M. Predeleanu [27], as well as the Predeleanu representation that generalizes the
one given by Love in the static case.
520 12 Dynamical Problems of Elastic Bodies

12.1.2.1 Representation of the Somigliana-Iacovache Type

If the displacements are expressed in the form


1 oU 1 oU
u r ¼ F1  ; uz ¼ F2  ; ð12:12Þ
2 or 2 oz
we shall remark that the Eqs. (12.6) lead to
   
o 1m 1 oF1 F1 oF2 1 €
h02 F1 ¼ DU  þ þ  2U ;
or 1  2m 1  2m or r oz 2c2
   
o 1m 1 oF1 F1 oF2 1 €
h 2 F2 ¼ DU  þ þ  2U ;
oz 1  2m 1  2m or r oz 2c2

these conditions are fulfilled if the potential functions U ¼ Uðr; z; tÞ and Fi


¼ Fi ðr; z; tÞ; i ¼ 1; 2, verify the equations
 
1 oF1 F1 oF2
h1 U ¼ þ þ ; ð12:13Þ
1  m or r oz

h02 F1 ¼ 0; h2 F2 ¼ 0: ð12:130 Þ

The functions F1 ; F2 , of class C2, will be supplied by the Eq. (12.130 ). while the
functions U, of class C3, will result from the Eq. (12.13); thus, the formulae
(12.12) give a representation by displacement potentials of the solution of the
Lamé equations in case of the axially symmetrical problem (A Schaefer type
representation).
If the functions F1 ; F2 are written in the form

F1 ¼ h01 Cr ; F2 ¼ h1 Cz ; ð12:14Þ

then the Eq. (12.130 ) lead to the conditions

h01 h02 Cr ¼ 0; h1 h2 Cz ; ð12:15Þ

the Eq. (12.13) allows, in this case, to choose the function U in the form
 
1 oCr Cr oCz
U¼ þ þ : ð12:140 Þ
1  m or r oz

We obtain thus the representation of the Somigliana-Iacovache type


 
1 o oCr Cr oCz
ur ¼ h01 Cr  þ þ ;
2ð1  mÞ or or r oz
  ð12:150 Þ
1 o oCr Cr oCz
uz ¼ h1 Cz  þ þ ;
2ð1  mÞ oz or r oz
12.1 Axisymmetrical Problems 521

with the help of two potential functions Cr ¼ Cr ðr; z; tÞ; Cz ¼ Cz ðr; z; tÞ that verify
two double waves equations; these functions must be of class C4.
The same results can be obtained by particularization, starting from the general
three-dimensional case.

12.1.2.2 Predeleanu’s Representation

We shall now show that the state of displacement can be expressed with the help of
a single potential function; to this end, we shall introduce the function K, given by

o2 K
¼ Cr : ð12:16Þ
oroz
We see that

o2 o2
M0 ¼ M;
oroz oroz
the first Eq. (12.5) leads to

o2
h1 h2 K ¼ 0;
oroz
from which we deduce
h1 h2 K ¼ g1 þ g2 ;

where g1 ¼ g1 ðr; tÞ; g2 ¼ g2 ðz; tÞ are arbitrary functions. The general integral of
this equation will assume the form
K ¼ K  þ h1 þ h2 ;
where h1 ¼ h1 ðr; tÞ; h2 ¼ h2 ðz; tÞ are particular integrals corresponding to the
functions g1 and g2, respectively, while K ¼ K ðr; z; tÞ is the general integral of
the homogeneous equation
h1 h2 K ¼ 0: ð12:160 Þ
In this case, the notation (12.16) leads to

o2 K
Cr ¼ ; ð12:17Þ
oroz
by introducing (12.17) in the representation (12.15), we observe that the potential
functions K and Cz may be grouped as follows
 
l o2 K ð1  2mÞl
v¼ Cz  2  h2 K  : ð12:18Þ
1m oz 1m
522 12 Dynamical Problems of Elastic Bodies

Thus, with the help of the function v ¼ vðr; z; tÞ, we can write Predeleanu’s
representation in the form

1 o2 v
ur ¼  ;
2l oroz
ð12:19Þ
1m 1 o2 v
uz ¼ h1 v  ;
l 2l oz2
the state of stress is given by
 
o o2 v
rrr ¼ mh2 v  2 ;
oz or
  ð12:20Þ
o 1 ov
rhh ¼ mh2 v  ;
oz r or
 
o o2 v
rzz ¼ ð1  mÞh2 v þ Mv  2 ;
oz oz
 2 
o ov
rrz ¼ ð1  mÞh1 v  2 : ð12:200 Þ
oz oz

The potential function v must belong to the class C4 and verify the equation
h1 h2 v ¼ 0: ð12:21Þ
We mention that this representation is complete.
We remark moreover that we can apply Boggio’s theorem which allows to
break down the waves into longitudinal and transverse ones.

12.2 Progressive Waves. Free and Characteristic


Vibrations

With respect to a given fixed frame of reference, we shall consider the three
equations of motion (5.1.6), the six relations between strain and displacements
(5.1.2) and the constitutive law (4.1.56). This system of 15 equations is associated
with boundary conditions and initial conditions.
Hereafter, we shall deal with waves propagation and elastic vibrations.

12.2.1 Introduction

The tensor Hijkl has the properties (4.1.57), (4.1.570 ) of symmetry, hence the body
under study is a hyperelastic one.
12.2 Progressive Waves. Free and Characteristic Vibrations 523

The tensor H is called strictly (strongly) elliptic if 8x; y 2 E3 ; x; y 6¼ 0.


Hijkl xi yj xk yl [ 0; ð12:22Þ

the tensor H is called positive definite if, for any symmetric tensor e 6¼ 0,
Hijkl eij ekl [ 0: ð12:220 Þ
If H is positive definite, it is also strictly elliptic; the converse is not true.
Indeed, let H be positive definite; then 8x; y 2 E3 ; x; y 6¼ 0, the dydic product
x  y admits the decomposition xi xj ¼ xði yjÞ þ x½i yj . It follows that
Hijkl xi yj xk yl ¼ Hijkl xði yjÞ xðk ylÞ [ 0; thus, the property is proved.
For a homogeneous (H is constant), isotropic elastic body, Hooke’s tensor H is
symmetric, positive definite, hence strictly elliptic too, with the components
Hijkl ¼ kdij dkl þ lðdik djl þ dil djk Þ; ð12:23Þ

where k; l [ 0 are Lamé’s constants.

12.2.1.1 Acoustic Tensor

Let m 2 E3 be a unit vector and let A(m) be the tensor of components Aik ðmÞ,
given by the contraction of H with m, according to the rule
1
Aik ðmÞ ¼ Hijkl mj ml ; ð12:24Þ
q
A(m) is called the acoustic tensor associated with the given elastic body, which
represents a second order tensor field with the properties:
(i) A(m) is symmetric for any m if and only if H is symmetric; indeed, we have

qAik ðmÞ ¼ Hijkl mj ml ¼ Hklij ml mj ¼ qAki ðmÞ:

(ii) A(m) is positive definite for any m if and only if H is strictly elliptic; indeed,
8y 2 E3 ; y 6¼ 0, we have

qAik ðmÞyi yk ¼ Hijkl yi mj yk ml [ 0:

(iii) If H is positive definite, then A(m) is positive definite. The property is


obvious from (ii) and from was had been said in Sect. 12.2.1 about H.
524 12 Dynamical Problems of Elastic Bodies

12.2.1.2 Acoustic Tensor for an Elastic Isotropic Body

Using the relations (12.23) and (12.24), we obtain succesively

qAik ðmÞ ¼ kdij dkl mj ml þ lðdik djl þ dil djk Þmj ml ¼ kmi mk
þ lðdik mj mj þ mi mk Þ ¼ ðk þ 2lÞmi mk þ lðdik  mi mk Þ:

Let us note c21 ¼ ðk þ 2lÞ=q; c22 ¼ l=q, introducing thus the propagation
velocities (5.3.4), (5.3.40 ). Thus, it follows that, for an isotropic elastic body, the
acoustic tensor has the form

AðmÞ ¼ c21 m  m þ c22 ðd  m  mÞ: ð12:25Þ

Let us show that c21 and c22 represent the eigenvalues of the tensor A(m) for any
m and that, for a given fixed m, the corresponding eigenvectors are m and m? ,
respectively; we have noted by m? a vector normal to m. Hence, if m1 and m2 are
two vectors normal to m and if m1  m2 ¼ 0, then the characteristic couples of
A(m) are ðc21 ; mÞ; ðc22 ; m1 Þ; ðc22 ; m2 Þ.
It follows easily that the characteristic equation of A(m), i.e.
det½Aij  kdij  ¼ 0, can be put in the form
 
 m2 þ d m1 m2 m1 m3 
 1
 m2 m1 m2 þ d m2 m3  ¼ 0; ð12:26Þ
 2 
 m3 m1 m3 m2 m2 þ d 
3

where d ¼ ðc22  kÞ=ðc21  c22 Þ; we conclude that d2 ðd þ 1Þ ¼ 0, from which d1 ¼


1; d2 ¼ d3 ¼ 0, i.e. k1 ¼ c21 ; k2 ¼ k3 ¼ c22 . The eigenvectors are given by the
equation ½AðmÞ  kdu ¼ 0. For k ¼ c21 , we get from (12.25) that ðc21  c22 Þ½m  m
du ¼ 0, i.e. u ¼ ðm  uÞm, from which we conclude that u and m are collinear.
Since u and m are unit vectors, it further follows that u ¼ m. For k ¼ c22 , we
have the relation ðc21  c22 Þ½m  mu ¼ 0, i.e. ðm  uÞm ¼ 0, which implies that
m  u ¼ 0; the asserted property is thus proved.

12.2.2 Plane Progressive Waves

The progressive wave represents a special type of wave, which illlustrates the
properties of a given elastic medium concerning the propagation of waves in
general.
12.2 Progressive Waves. Free and Characteristic Vibrations 525

12.2.2.1 General Considerations

Thus, let us consider an infinite elastic medium, characterized by a constant Hooke


tensor, which has a constant density q.
By progressive wave we understand a motion uðx; tÞ : E3 ð1; 1Þ ! E3 , of
the form
uðx; tÞ ¼ aUðx  m  ctÞ; ð12:27Þ

where a and m are unit vectors called direction of motion and direction of
propagation, respectively, and U is a real function of class C2, with second
derivatives not identically zero, i.e. the mapping R
s ! UðsÞ 2 R has the
property

d2 UðsÞ
U 00 ðsÞ ¼ 6¼ 0; ð12:28Þ
ds2
c is a real constant, called velocity of propagation.
If a and m are linearly dependent, which is equivalent to say (the two being unit
vectors) that a ¼ m, the wave is called longitudinal; if a is normal to m, then
the wave is called transverse.
One can see that u and a are collinear at any moment and at any point. Since
u represents the displacement vector, it follows that a, permanently collinear with
u, represents the direction of motion. Let us consider a 2 R; a ¼ const: Let us also
discuss the plane P fx 2 E3 jx  m  ct ¼ ag. From (12.27) it follows that u is
constant on P, having the value uðx; tÞ ¼ aUðaÞ. Hence, all points of this plane, at
every moment t, have the same displacement of direction a and magnitude UðaÞ.
The normal vector to the plane P is m; as a function of t, P moves in the direction
m with the velocity c. These considerations justify the terms: direction of motion,
direction of propagation, velocity of propagation, longitudinal wave and transverse
wave.

12.2.2.2 Elastic Progressive Waves

An elastic progressive wave is a progressive wave, i.e. a motion of the form


(12.27), which satisfies the system (4.1.56), (5.1.2) and (5.1.6), for vanishing
volume forces.
From (12.27) we conclude that ui;j ¼ ai U;j ¼ ai U 0 mj ; u_ i ¼ ai U 0 c; €ui ¼ c2 ai U 00 ;
hence

Gradu ¼ ru ¼ U 0 a  m; u
€ ¼ c2 U 00 a: ð12:29Þ

From (4.1.56) and (5.1.2), we obtain r ¼ He ¼ H½ru þ ðruÞ> =2 ¼ ½Hruþ


HðruÞ> =2 ¼ Hru, since H is symmetric with respect to the last two indices;
hence
526 12 Dynamical Problems of Elastic Bodies

r ¼ U 0 Hða  mÞ: ð12:30Þ


Since H, a and m are constant, from (12.30) and (12.24) we obtain
rij;j ¼ U;j0 Hijkl ak ml ¼ U 00 Hijkl ak mj ml ¼ qU 00 Aik ðmÞak , i.e.

Divr ¼ U 00 Hða  mÞm ¼ qU 00 AðmÞa: ð12:31Þ

12.2.2.3 The Fresnel-Hadamard Propagation Condition

Taking into account (12.28), we find from (5.6) that

AðmÞa ¼ c2 a; ð12:32Þ
a relation called the Fresnel-Hadamard propagation condition (for elastic pro-
gressive waves).
This condition expresses the following physical fact: in order that an elastic
progressive wave, with direction of propagation m, should be propagated, it is
necessary and sufficient that the direction of motion a be an eigenvector for the
acoustic tensor A(m) and the square of the velocity of propagation be the corre-
sponding eigenvalue.
If, H is symmetric, then A(m) is symmetric too (see Sect. 12.2.1.1, point (i));
hence, there exist at least three eigenvectors for any m. But, in general, the
eigenvalues that give the squares of the corresponding velocities of propagation
are not always positive, hence it is possible that the velocities of propagation be
not real.
If in addition, H is strictly elliptic or positive definite, then A is positive definite
and the velocities of propagation are real.
In conclusion, using also the general results of Sect. 1.2.1.1, if H is symmetric
and strictly elliptic, then for any direction of propagation m there exist three
mutually perpendicular directions of propagation and, correspondingly, three real
direction velocities of propagation.
If a is a possible direction of motion corresponding to the direction of propa-
gation m, the velocity of propagation is given by the relation

c2 ¼ a  ðAðmÞaÞ; ð12:33Þ

which is obtained from (12.32) by means of scalar multiplication by the unit


vector a.
The relation (12.33) shows that, for the same direction of propagation, c as well
as -c are possible velocities of propagation.
We infer from the Sect. 12.2.1.2 that, for an isotropic elastic body, only two
types of progressive waves are possible, namely one with a velocity of propagation
given by c21 ¼ ðk þ 2lÞ=q, the wave being longitudinal, and another one having
the velocity of propagation given by c22 ¼ l=q, the wave being transverse; hence,
12.2 Progressive Waves. Free and Characteristic Vibrations 527

in an isotropic elastic body only longitudinal and transverse waves can be


propagated.
The question arises whether in an anisotropic elastic body (with arbitrary H),
longitudinal and transverse waves can be propagated in some direction, because, in
general, no connection of this kind between a and m is manifest in (12.32). The
answer is affirmative and it is known as the Fedorov-Stippes theorem, given in
1964; it asserts that, if H is symmetric and strictly elliptic, then there exist lon-
gitudinal and transverse elastic progressive waves in some direction. Moreover,
along at least three distinct directions, longitudinal waves are propagated, as
showed Kolodner in 1966.

12.2.3 Free and Characteristic Waves

Since H is symmetric with respect to the last two indices, one may write
r ¼ Hru: ð12:34Þ
This relation allows us to reduce the system of 15 equations of elastodynamics to
Lamé’s system of three scalar equations
DivðHruÞ þ F ¼ q€
u; ð12:35Þ

with the unknowns ui ; i ¼ 1; 2; 3


Let D be the set

D ¼ fuju 2 C2 ðXÞ; u; ru 2 C1 ðXÞg; ð12:36Þ

let us denote by L the operator (mapping) L : D ! L2 ðXÞ (L2 ðXÞ represents the
space of square integrable functions on X) given by the relation
Lu ¼ DivðHruÞ; ð12:37Þ
L is called Lamé’s operator.
Thus, Lamé’s equations can be written as
Lu þ q€
u ¼ F; ð12:350 Þ
it is easy to conclude that, if the operator L is symmetric and negative definite, then
Lamé’s equations are hyperbolic in the dynamic case and elliptic in the static case.

12.2.3.1 Free Vibrations of an Elastic Body

The problem of free vibrations of an elastic body is a study of the motions of the
form
528 12 Dynamical Problems of Elastic Bodies

~ðx; tÞ ¼ uðxÞ sinðxt þ uÞ;


u ð12:38Þ

where u represents the amplitude of the vibration, x the circular frequency


(pulsation) and u the plane shift.
For the case when the volume forces vanish (F = 0), we deduce that u(x) sat-
isfies the equation DivðHruÞ ¼ qx2 u, i.e.

Lu ¼ qx2 u: ð12:39Þ
Let us discuss the analogy between the relation (12.39) and the Fresnel-Had-
amard propagation condition.
Let S1, S2 represent a partition of the boundary oX of the body, i.e.
S1 \ S2 ¼ [, S1 [ S2 ¼ oX. The problem of free vibrations can be formulated as
the problem of finding possible combinations of the circular frequency x and of
the amplitude u. such that (12.39) be satisfied for bodies X fixed on the part S1 of
the boundary and free otherwise.
In a mathematical formulation, this means the determination of the so-called
characteristic solutions of the problem of free vibrations, which we shall now
explain.
By characteristic solution we mean an ordered couple ðk; uÞ; k 2 R; u 2 D,
such that
Lu ¼ ku; jjujj ¼ 1; ð12:40Þ

ujS1 ¼ 0; ð12:400 Þ
n
rðuÞjS2 ¼ 0; ð12:4000 Þ

k is called the characteristic value or eigenvalue and u is the corresponding


characteristic displacement (amplitude).
For a given non-negative k, the frequency of free vibrations is given by
pffiffiffiffiffiffiffiffi
x ¼ k=q. Hence, the problem of free vibrations leads to a problem of eigen-
vectors and eigenvalues for Lamé’s operator L.
Further, we say that the boundary is free if S2 ¼ oX and it is fixed if S1 ¼ oX.
Let (k, u) be a characteristic solution. Then:
(i) k ¼ ½u; u and k [ 0, where

Z Z
½u; v ¼ Hijkl ui;j vk;l dX ¼ Hijkl eij ðuÞekl ðvÞdX
ZX ZX ð12:41Þ
¼ rij ðvÞeij ðuÞdX ¼ rkl ðuÞekl ðvÞdX
X X

represents the energetic scalar product and the norm (juj2 ¼ ½u; u), respectively.
Supposing that H is symmetric, positive definite and of class C0 ðXÞ,  we have
12.2 Progressive Waves. Free and Characteristic Vibrations 529

R
denoted by hu; vi ¼ Xu  vdX; jjujj2 hu; ui, the scalar product and the norm in
n
L2 ðXÞ, respectively. If rðuÞ represents the stress vector on the surface element of
external normal n, then
n
rðuÞ ¼ rn ¼ ðHruÞn; ð12:42Þ
using the above notations, one may prove (using the Gauss-Ostrogradski formula)
Betti’s relation, which states that
Z
n
hLu; vi ¼ ½u; v ¼ rðuÞ  vdS; 8u; v 2 D: ð12:43Þ
oX

(ii) k ¼ 0 if and only if u is a rigid displacement of the body; moreover, there exist
characteristic solutions (k, u) with k ¼ 0 if and only if the body is free, in
which case any rigid displacement u with jjujj ¼ 1 is a characteristic dis-
placement corresponding to the eigenvalue k ¼ 0.

12.2.3.2 Characteristic Vibrations

Let ðk1 ; u1 Þ and ðk2 ; u2 Þ be characteristic couples, with k1 6¼ k2 . Then hu1 ; u2 i


¼ ½u1 ; u2  ¼ hLu1 ; u2 i ¼ hu1 ; Lu2 i ¼ 0. Indeed, by assumption, both characteristic
couples satisfy the relations (12.40), (12.4000 ). Hence, Lu1 ¼ k1 u1 ; Lu2 ¼ k2 u2
and rðu1 Þ  u2 ¼ 0; rðu2 Þ  u1 ¼ 0 on oX. Thus, we obtain
hLu1 ; u2 i ¼ k1 hu1 ; u2 i; hLu2 ; u1 i ¼ k2 hu2 ; u1 i: ð12:44Þ
From Betti’s relation, it follows easily that hLu1 ; u2 i; ¼ hu1 ; Lu2 i ¼ ½u1 ; u2 ,
hence k1 hu1 ; u2 i ¼ k2 hu2 ; u1 i, from which hu1 ; u2 i ¼ 0 since k1 6¼ k2 . The
remaining results may be obtained from (12.44).
If the eigenvalues are equal, the above orthogonality result is no longer
obtained. Howeover, it is easy to show that, if u1 ; u2 ; . . .; un are characteristic
amplitudes, corresponding to the same k, then any vector u, which is a linear
combination of the vectors u1 ; u2 ; . . .; un and for which jjujj ¼ 1, is again a
characteristic amplitude, corresponding to the same characteristic value k. Indeed,
from the relations Lui ¼ kui ; i ¼ 1; 2; . . .; n, it follows also that v ¼ ai ui satisfies
the relation (12.40). Owing to the linearity of the operator L, the vector u ¼ v=jjvjj
satisfies, at its turn, the same relation. The conditions (12.400 ), (12.4000 ) are
obviously satisfied too.
530 12 Dynamical Problems of Elastic Bodies

12.2.3.3 Minimum Principles for Characteristic Values

Let K be the family of vector fields


 jjvjj ¼ 1; vj ¼ 0g:
K ¼ fvjv 2 C1 ðXÞ; ð12:45Þ
S1

With this notation, one may state a minimum principle in following form.
Let us consider u1 2 D \ K such that
ju1 j jvj; 8v 2 K; ð12:46Þ

and k1 ¼ ju1 j2 . Then ðk1 ; u1 Þ is a characteristic couple and k1 is the smallest


eigenvalue.
Moreover, we can state a general minimum principle, namely:
Let us consider u1 ; u2 ; . . .; un ; . . . 2 D \ K and K1 ¼ K; K2 ¼ K \ fu1 g? ,
Kr ¼ K \ fu1 ; u2 ; . . .; ur1 g? . Let us suppose that, for any r  1,
ur 2 Kr ; jur j jvj; 8v 2 K: ð12:47Þ

Let us consider, further kr ¼ jur j2 . Then, every couple ðkr ; ur Þ is a characteristic


couple and the sequence of characteristic values increases indefinitely
(0 k1 k2    kn . . .).
One can also show that this general minimum principle generates all charac-
teristic values.
The ordered set fk1 ; k2 ; . . .; kn ; . . .g is also called the spectrum corresponding to
the body X, to the elastic constants H and to the partition S1, S2 of the boundary
oX; the sequence u1 ; u2 ; . . .; un ; . . . of the characteristic amplitudes is called the
sequence of fundamental (normal) modes of vibration.
Another result is the following: if another body occupies the same domain X,
but the partition S01 ; S02 of the boundary is different and satisfies S01 S and/or the
constants H0 of the new body satisfy the inequality Hijkl 0
eij ekl Hijkl eij ekl , then the
corresponding spectra satisfy the inequality k0n kn ; n 2 N . This result points out
the relation between the vibration frequencies of two bodies having the same form
but differing by their materials and/or by the manner in which their boundaries are
fixed.
We end the above considerations by mentioning that the sequence of funda-
mental modes of vibration constitutes a complete sequence in the class of con-
tinuous functions on X. This shows that, for any other problem (with continuous
solutions) on the domain X, we may look for solutions f in the form of the series
P
1
f¼ cn un , cn being the Fourier coefficients corresponding to the sequence
n¼1
u1 ; u2 ; . . .; un ; . . ., that is cn ¼ hf; un i.
The result justifies the procedure of seeking solutions for different problems in
the form of a series of the fundamental modes of vibration.
12.3 Forced and Free Vibrations 531

12.3 Forced and Free Vibrations

The limit problems of elastodynamics (the study of the action of external loads,
variable in time, the study of free and forced vibrations etc.) have made the object
of many researches, even in the last two centuries.
Thus, a problem qualitatively different from that of the elastic space, considered
in Sect. 6.2.4.3, is the problem of the elastic half-space; in this field, Lord Ray-
leigh [10, 29] discovered in 1887 the surface waves, called nowadays by his name,
L. Knopoff [20] gave numerical results concerning the propagation velocity of
Rayleigh’s waves. In the case of a stratified body, formed by an elastic layer,
superposed on an elastic half-space with different mechanical properties, we find
the waves of A. E. H. Love [9].
Another important related problem is that of a point or a linear load, variable in
time and acting upon the separation plane of the elastic half-space; the first study
in this direction was made by H. Lamb [21], whose name was given to this
problem. Interesting results were given, in a preliminary form, in this direction, by
F. Santer [30]; these results have been numerically estimated by K. B. Broberg
[12]. The case of a load, varying in time and applied within the elastic half-space,
was considered by E. Pinney [27], C. C. Pekeris [25] and by C. C. Pekeris and
H. Lifson [26], who later gave numerical results. Other researches into the elastic
half-space were made by G. Eason [16, 17], L. M. Flitman [18], M. Hayes and
R. S. Rivlin [19], H. A. Lang [22] and C. C. Chao, H. H. Bleich and J. Sackmann
[14]; some of the results, especially those about the case of a state of plane strain,
are due to J. W. C. Sherwood [31]. C. C. Chao [13] used the technique of inversion
of the integral transforms, given by L. Cagniard [3], to obtain results in case of a
concentrated tangential force, suddenly applied to the separating plane of the
elastic half-space and maintained constant in time. P. P. Teodorescu [11] showed
how to obtain waves of Rayleigh type, corresponding to a three-dimensional state
of strain and stress (in the classical case, these waves correspond to a plane state of
strain). He indicated moreover how to study systematically, by means of the same
family of stress functions, the elastic parallelepiped and all the infinite domains
derived from it by throwing to infinite one or several of its faces, from the
standpoint of both free and forced vibrations [35, 36].
In connection with the waves propagation, the monographs of L. M. Ber-
ekhovskikh [2], R. M. Davies [4], W. M. Ewing, W. S. Jardetzky and F. Press [5],
H. Kolsky [8] and W. M. Ewing and F. Press [6]. As well, we mention
the monographs of W. Kecs and P. P. Teodorescu [7], I. Beju et al. [1] and
P. P. Teodorescu [11]. We mention moreover the synthesis studies of R. M. Davies
[15] and J. Miklowitz [23], as well as the syntheses study presented by Teodorescu
[34] at a Conference on non-linear vibrations in Berlin.
Hereafter, we shall deal with the problem of free and forced vibrations in the
case of the elastic half-space and give some indications concerning other infinite
spatial domains. We use the results given in [34], as well as the methods of the
theory of distributions as in [7].
532 12 Dynamical Problems of Elastic Bodies

The influence of the volume forces on the results given hereafter can be
introduced with the help of some particular integrals. Thus, can be considered
volume forces such as Fi ðx1 ; x2 ; x3 ; tÞ ¼ F  i ðx1 ; x2 ; x3 Þ sin xt; i ¼ 1; 2; 3; if for
 i ðx1 ; x2 ; x3 Þ we use representations made with the help of Fourier series and
F
integrals, then we can easily compute the corresponding particular integrals in the
same form; we remark that the boundary conditions are of the same form as those
used in the absence of the volume forces. Thus, the problem where non-vanishing
volume forces appear is reduced to the similar problem without volume forces, but
with conventional boundary conditions.
In the steady-state problems it is not necessary to consider initial conditions, the
only limit conditions to be verified being the boundary conditions.

12.3.1 Forced Vibrations

Hereafter, we shall give some results developed for the elastic half-space; we shall
then supply a few computing hints concerning other space domains.

12.3.1.1 Elastic Half-Space

Let be the elastic half-space x3  0, on the separation plane of which acts the
normal periodic load
pðx1 ; x2 ; tÞ ¼ pðx1 ; x2 Þ sin xt; ð12:48Þ
for simplifying the computation, we took a only one term from the expansion with
respect to time. The results thus obtained can help, by superposition of effects, in
the case of a Fourier series or of a Fourier integral with respect to time.
We shall admit that the load is periodic with respect to the variables x1 and x2
and antisymmetric with respect to both axes of co-ordinates Ox1 and Ox2
XX
pðx1 ; x2 Þ ¼ anm sin an x1 sin bm x2 ; ð12:49Þ
n m

with
np mp
an ¼ ;b ¼ ; n; m ¼ 1; 2; 3; . . .; ð12:490 Þ
a1 m a2
where Li ¼ 2ai ; i ¼ 1; 2, are the periods.
The loading cases with other symmetry or antisymmetry properties with respect
to the co-ordinate axes, as well as the cases of tangential loads, can be similarly
studied.
We can choose the stress functions, odd with respect to x1 and x2 , in the form
12.3 Forced and Free Vibrations 533

XX
Fi ðx1 ; x2 ; x3 ; tÞ ¼ sin xt ðA0nm ecnm x3
n m

þ B0nm e dnm x3
Þ sin an x1 sin bm x2 ; i ¼ 1; 2; 3; ð12:50Þ

with the relations


1  2m q 2 q
a2n þ b2m ¼ c2nm þ x ¼ d2nm þ x2 ; ð12:51Þ
2ð1  mÞ l l
we mention that
1 q 2
k2 ¼ c2nm  d2nm ¼ x ð12:510 Þ
2ð1  mÞ l
and that

a2n þ b2m ¼ 2ð1  mÞc2nm  ð1  2mÞd2nm : ð12:5100 Þ


The sequences of coefficients A0nm ; B0nm ; . . .; B000
nm will be determined by the
boundary conditions
x3 ¼ 0 : r33 ¼ pðx1 ; x2 ; tÞ; r31 ¼ r32 ¼ 0: ð12:52Þ
We remark moreover that, owing to the manner of choosing the stress func-
tions, the components of the stress tensor tend to zero for x3 ! 0. The functions
Fij ; i 6¼ j; i; j ¼ 1; 2; 3 of two variables shall be considered equal to zero.
We obtain thus the state of stress
XX 
r11 ¼ sin xt ða2n þ mk2 ÞAnm ecnm x3 þ ða2n  2k2 ÞB0nm
n m

d x
þða2n þ mk ÞðB00nm þB000
2
nm Þ e
nm 3
sin an x1 sin bm x2 ;
XX 
r22 ¼ sin xt ðb2m þ mk2 ÞAnm ecnm x3 þ ðb2m  2k2 ÞB00nm
n m

d x ð12:53Þ
þðb2m þ mk2 ÞðB000 0
nm þBnm Þ e
nm 3
sin an x1 sin bm x2 ;
X X 

r33 ¼  sin xt d2nm þ ð1  mÞk2 Anm ecnm x3 þ ðd2nm þ 2k2 ÞB000
nm
n m


þðd2nm  mk2 ÞðB0nm þB00nm Þ ednm x3 sin an x1 sin bm x2 ;
534 12 Dynamical Problems of Elastic Bodies

XX 
1
r23 ¼ sin xt cnm dnm b2m Anm ecnm x3 þ d2nm b2m
n m
d nm b m
  
1 2 2 1
þ mk ðdnm  b2m Þ B0nm þ d2nm b2m  k2 ð2d2nm þ mb2m Þ B00nm
2 2
 
1
þ d2nm b2m þ k2 ðmd2nm þ 2b2m Þ B000 nm e dnm x3
sin an x1 sin bm x2 ;
2
XX 1 
r31 ¼ sin xt cnm dnm a2n Anm ecnm x3 þ d2nm a2n
n m
dnm an
  
1 2 2 1 2 2
 k ð2dnm  man Þ Bnm þ dnm an þ mk ðdnm  an Þ B00nm
2 0 2 2 2 ð12:530 Þ
2 2
 
2 2 1 2 2 2 000 dnm x3
þ dnm an þ k ðmdnm þ 2an Þ Bnm e sin an x1 cos bm x2 ;
2
XX 1 
r12 ¼  sin xt a2n b2m Anm ecnm x3 þ a2n b2m
n m
an bm
  
1 1
þ k2 ðma2n  2b2m Þ B0nm þ a2n b2m  k2 ð2a2n  mb2m Þ B00nm
2 2
 
1
þ a2n b2m þ mk2 ða2n þ b2m Þ B000 nm e dnm x3
cos an x1 cos bm x2
2

and the state of displacement


1 XX 1
u1 ¼ x0 x2  sin xt a2n Anm ecnm x3
2l n m
an

þ½ða2n  2k2 ÞB0nm þ ða2n þ mk2 ÞðB00nm þB000
nm Þe
dnm x3
cos an x1 sin bm x2 ;
1 XX 1
u2 ¼ x0 x1  sin xt b2 Anm ecnm x3
2l n m
bm m ð12:54Þ

þ½ðb2m  2k2 ÞB00nm þ ðb2m þ mk2 ÞðB000 0
nm þBnm Þe
dnm x3
sin an x1 cos bm x2 ;
1 XX 1
u3 ¼ sin xt fc dnm Anm ecnm x3 þ ½ðd2nm þ 2k2 ÞB000
2l n m
dnm nm nm

þðd2nm  mk2 ÞðB0nm þB00nm Þednm x3 sin an x1 sin bm x2

where we used the notation

Anm ¼ A0nm þ A00nm þ A000


nm : ð12:55Þ
The boundary conditions (12.52) lead to the following system of linear alge-
braic equations
12.3 Forced and Free Vibrations 535

½ð1  mÞc2nm þ md2nm Anm  ½mc2nm  ð1 þ mÞd2nm ðB0nm þ B00nm Þ


þ ð2c2nm  d2nm ÞB000
nm ¼ anm ;
2cnm dnm a2n Anm þ ½2d2nm a2n  ðc2nm  d2nm Þð2d2nm þ ma2n ÞB0nm
þ ½2d2nm a2n þ mðc2nm  d2nm Þðd2nm  a2n ÞB00nm
ð12:56Þ
þ ½2d2nm a2n þ ðc2nm  d2nm Þðmd2nm þ 2a2n ÞB000
nm ¼ 0;
2cnm dnm b2m Anm þ ½2d2nm b2m þ mðc2nm  d2nm Þðd2nm  b2m ÞB00nm
þ ½2d2nm b2m  ðc2nm  d2nm Þð2d2nm þ mb2m ÞB00nm
þ ½2d2nm b2m þ ðc2nm  d2nm Þðmd2nm þ 2b2m ÞB000
nm ¼ 0;

by subtracting the last two equations from the first one, we get

ða2n þ b2m Þ½ð1  mÞc2nm  2cnm dnm þ md2nm Anm  d2nm ½a2n þ b2m
 ð2  mÞðc2nm  d2nm ÞðB0nm þ B00nm Þ  d2nm ½a2n þ b2m
þ 2mðc2nm  d2nm Þ ¼ ða2n þ b2m Þanm : ð12:57Þ

Taking into account the relation (12.5100 ), we can write the first Eqs. (12.56) and
the Eq. (12.57) in the form

½ð1  mÞc2nm þ md2nm Anm þ ½ð1 þ mÞd2nm  mc2nm Bnm


þ ð2 þ mÞðc2nm  d2nm ÞB000
nm ¼ anm ;
 ða2n þ b2m Þ½ð1  mÞc2nm  2cnm dnm þ md2nm Anm ð12:58Þ
þ d2nm ½ð1þ mÞd2nm  mc2nm Bnm
þ ð2 þ mÞd2nm ðc2nm  d2nm ÞB000
nm ¼ ða2n þ b2m Þanm ;

with

Bnm ¼ B0nm þ B00nm þ B000


nm : ð12:550 Þ
These equations lead to
1 2  anm
Anm ¼  an þ b2m þ d2nm ; ð12:59Þ
2 Dnm
where we have introduced the denominator
1 2 2
Dnm ¼ an þ b2m þ d2nm cnm dnm ða2n þ b2m Þ: ð12:60Þ
4
We obtain then

cnm dnm ða2n þ b2m Þ anm ð1 þ mÞd2nm  mc2nm


B000
nm ¼  Bnm ; ð12:61Þ
ð2 þ mÞðc2nm þ d2nm Þ Dnm ð2 þ mÞðc2nm  d2nm Þ
536 12 Dynamical Problems of Elastic Bodies

likewise, we obtain

cnm dnm a2n anm mðc2nm  d2nm Þ þ a2n


B0nm ¼  þ Bnm ;
ð2 þ mÞðc2nm  dnm Þ Dnm ð2 þ mÞðc2nm  d2nm Þ
2
ð12:610 Þ
cnm dnm b2m anm mðc2nm  d2nm Þ þ b2m
B00nm ¼ þ Bnm :
ð2 þ mÞðc2nm  d2nm Þ Dnm ð2 þ mÞðc2nm  d2nm Þ
By substituting in (12.53), (12.530 ) and (12.54), we find the state of stress (the
parameter Bnm is eliminated from the computation)
X X anm 


r11 ¼  sin xt ð1  mÞc2nm þ md2nm a2n þ mðc2nm  d2nm Þ ecnm x3


n m
Dnm

cnm dnm a2n ednm x3 sin an x1 sin bm x2 ;
X X anm 


r22 ¼  sin xt ð1  mÞc2nm þ md2nm b2m þ mðc2nm  d2nm Þ ecnm x3


n m
Dnm ð12:62Þ

cnm dnm b2m ednm x3 sin an x1 sin bm x2 ;
X X anm 

r33 ¼ sin xt ð1  mÞc2nm þ md2nm ecnm x3


n m
D nm

cnm dnm ða2n þ b2m Þednm x3 sin an x1 sin bm x2 ;
X X anm 
r23 ¼  sin xt cnm bm ð1  mÞc2nm
n m
Dnm

 c x 
þmd2nm e nm 3  ednm x3 sin an x1 sin bm x2 ;
X X anm 
r31 ¼  sin xt cnm an ð1  mÞc2nm
n m
D nm ð12:620 Þ

 c x 
2
þmdnm e nm 3
 ednm x3 sin an x1 cos bm x2 ;
X X anm 

r12 ¼ sin xt an bm ð1  mÞc2nm þ md2nm ecnm x3


n m
Dnm
dnm x3

cnm dnm e cos an x1 cos bm x2

and the state of displacement


1 X X anm 

u1 ¼  x03 x2 þ sin xt an ð1  mÞc2nm þ md2nm ecnm x3


2l n m
Dnm
dnm x3

cnm dnm e cos an x1 sin bm x2 ;
1 X X anm 

u2 ¼x03 x1 þ sin xt bm ð1  mÞc2nm þ md2nm ecnm x3


2l n m
Dnm ð12:63Þ
dnm x3

cnm dnm e sin an x1 cos bm x2 ;
1 X X anm 

u3 ¼  sin xt cnm ð1  mÞc2nm þ md2nm ecnm x3


2l n m
D nm
2 2 dnm x3

ðan þ bm Þe sin an x1 sin bm x2 ;
12.3 Forced and Free Vibrations 537

where, for symmetry’s sake, there remains only to be determined the rotation of
rigid body x03 .
If the displacements vanish at t ¼ 0 or if we impose the condition that the
displacements should be periodic functions, we shall have x03 ¼ 0.
The distorted form of the separation plane is supplied by

qx2 X X anm
u3 ðx1 ; x2 ; 0; tÞ ¼ sin xt c sinan x1 sin bm x2 : ð12:64Þ
4l2 n m
Dnm nm

The above series are uniformly convergent, while the series (12.64) is rapidly
convergent.
In case of a local load, we can make use of Fourier integrals with regard to the
variables x1 and x2 .

12.3.1.2 Other Space Domains

Similarly, other space domains, bounded at a finite distance by planes parallel to


the co-ordinate planes, can be considered.
Thus, in case of the elastic layer jx3 j a3 on which acts, on the faces
x3 ¼ a3 , a normal periodic load, symmetric with respect to the plane x3 ¼ 0, in
the form (12.48)–(12.490 ), we introduce the stress functions
XX
Fi ðx1 ; x2 ; x3 ; tÞ ¼ sin xt ðA0nm coshcnm x3
n m
0
þ Bnm coshdnm x3 Þ sin an x1 sin bm x2 ; i ¼ 1; 2; 3; ð12:65Þ

where the parameters cnm ; dnm are related by (12.51). The functions Fij ; i 6¼ j;
i; j ¼ 1; 2; 3, are taken equal to zero.
The sequences of coefficients A0nm ; B0nm ; . . .; B000
nm will be defined by the boundary
conditions
x3 ¼ a3 : r33 ¼ pðx1 ; x2 ; tÞ; r31 ¼ r32 ¼ 0; ð12:66Þ
while the state of stress and the state of displacement will be expressed by means
of uniformly convergent series.
In case of the elastic quarter-space x1  0; x3  0, we shall admit that one
single local normal load in the form
Z 1Z 1
pðx1 ; x2 Þ ¼ aða1 ; a2 Þ cos a1 x1 cos a2 x2 da1 da2 ; ð12:67Þ
0 0

with
Z 1 Z 1
4
aða1 ; a2 Þ ¼ 2 pðn1 ; n2 Þ cos a1 n1 cos a2 n2 dn1 dn2 ; ð12:670 Þ
p 0 0
538 12 Dynamical Problems of Elastic Bodies

is acting only on the plane x3 ¼ 0. We introduce the stress functions


Z 1 Z 1  
F1 ðx1 ; x2 ; x3 ; tÞ ¼ sin xt A01 em1 x3 þ A02 em2 x3 cos a1 x1 cos a2 x2 da1 da2
0 0
Z 1Z 1  
þ sin xt B01 ek1 x3 þ B02 ek2 x3 cos a2 x2 cos a3 x3 da2 da3 ; . . .;
0 0
ð12:68Þ
where the newly introduced parameters verify the relations
1 2 1
a22 þ a23 ¼ k21 þ x ¼ k22 þ 2 x2 ;
c21 c2
ð12:69Þ
1 1
a21 þ a22 ¼ m21 þ 2 x2 ¼ m22 þ 2 x2 :
c1 c2
The functions A01 ða1 ; a2 Þ; A02 ða1 ; a2 Þ; . . .; B000
2 ða2 ; a3 Þ will be determined by the
boundary conditions

x1 ¼ 0 : r11 ¼ 0; r12 ¼ r13 ¼ 0;


ð12:70Þ
x3 ¼ 0 : r33 ¼ pðx1 ; x2 ; tÞ; r31 ¼ r32 ¼ 0:

The problem is reduced, from a mathematical standpoint, to the integration of a


system of two integral equations of the second kind of Fredholm type.
In case of the elastic eights-space xi  0; i ¼ 1; 2; 3, undergoing a load such as
(12.48)–(12.490 ) on the plane x3 ¼ 0, we use stress functions of the form
Z 1Z 1
 0 k x 
F1 ðx1 ; x2 ; x3 ; tÞ ¼ sin xt A1 e 1 1 þ A02 ek2 x1 cos a2 x2 cos a3 x3 da2 da3
0
Z 10 Z 1
 0 , x 
þ sin xt B1 e 1 2 þ B02 e,2 x2 cos a3 x3 cos a1 x1 da3 da1
Z0 1 Z0 1
 0 m x 
þ sin xt C1 e 1 3 þ C20 em2 x3 cos a1 x1 cos a2 x2 da1 da2 ; . . .;
0 0
ð12:71Þ
with the relations (12.69) as well as with
1 2 1
a23 þ a21 ¼ ,21 þ x ¼ ,22 þ 2 x2 : ð12:690 Þ
c21 c2
The functions A01 ða2 ; a3 Þ; A02 ða2 ; a3 Þ; . . .; C2000 ða1 ; a2 Þ will be defined by the
boundary conditions

x1 ¼ 0 : r11 ¼ 0; r12 ¼ r13 ¼ 0;


x2 ¼ 0 : r22 ¼ 0; r21 ¼ r23 ¼ 0; ð12:72Þ
x3 ¼ 0 : r33 ¼ pðx1 ; x2 ; tÞ; r31 ¼ r32 ¼ 0
12.3 Forced and Free Vibrations 539

and the problem is reduced to the integration of a system of three integral equa-
tions of the second kind of Fredholm type.
In case of an elastic half-layer and of an elastic quarter-layer, we shall use
stress functions built up of Fourier series, along a direction normal to the parallel
separation planes and of Fourier integrals in the other directions.
In the case of an elastic strip jx2 j a2 ; jx3 j a3 undergoing normal loads such
as (12.48), with
XX
pðx1 ; x2 Þ ¼ alm sin al x1 sin bm x2 ð12:73Þ
l m

on the separation planes x3 ¼ a3 , we shall use stress functions of the form
XX
F1 ðx1 ; x2 ; x3 ; tÞ ¼ sin xt B01nl sinh l1nl x2
n l

þB02nl sinh l2nl x2 sin cn x3 sin al x3
XX 
0 0
þ sin xt C1lm sinh m1lm x3 þ C2lm sinh m2lm x3 sin al x1 sin bm x2 ; . . .;
l m

ð12:74Þ

with the relations


1 2 1
c2n þ a2l ¼ l21nl þx ¼ l22nl þ 2 x2 ;
c21 c2
ð12:75Þ
1 1
a2l þ b2m ¼ m21lm þ 2 x2 ¼ m22lm þ 2 x2 :
c1 c2
The sequences of coefficients B01nl ; B02nl ; . . .; C2lm
000
are determined by the
boundary conditions

x2 ¼ a2 : r22 ¼ 0; r21 ¼ r23 ¼ 0;


ð12:76Þ
x3 ¼ a3 : r33 ¼ pðx1 ; x2 ; tÞ; r31 ¼ r32 ¼ 0

and, from a mathematical standpoint, the problem is reduced to solving an infinite


system of linear algebraic equations, with a double infinity of unknowns.
In case if an elastic half-strip, the stress functions can be built similarly with the
help of the Fourier series and integrals.
Let us now consider the elastic parallelepiped jxi j ai ; i ¼ 1; 2; 3, undergoing a
load in the form (12.48)–(12.490 ) on the faces jx3 j ¼ a3 . We shall introduce
stress functions in the form
540 12 Dynamical Problems of Elastic Bodies

XX
Fi ðx1 ; x2 ; x3 ; tÞ ¼ sin xt ðA01mn sinhk1mn x1
m n
þ A02mn sinhk2mn x1 Þ sin bm x2 sin cn x3
XX
þ sin xt ðB01nl sinhl1nl x2 þ B02nl sinhl2nl x2 Þ sin cn x3 sin al x1
n
l
XX
0 0
þ sin xt ðC1lm sinhm1lm x3 þ C2lm sinhm2lm x3 Þ sin al x1 sin bm x2 ; . . .;
l m

ð12:77Þ
with the relations (12.75), as well as with the relations
1 2 1
b2m þ c2n ¼ k21mn þ x ¼ k22mn þ 2 x2 : ð12:750 Þ
c21 c2
The sequences of coefficients A01mn ; A02mn ; . . .; C2lm
000
will be determined by the
boundary conditions

x1 ¼ a1 : r11 ¼ 0; r12 ¼ r13 ¼ 0;


x2 ¼ a2 : r22 ¼ 0; r21 ¼ r23 ¼ 0; ð12:78Þ
x3 ¼ a3 : r33 ¼ pðx1 ; x2 ; tÞ; r31 ¼ r32 ¼ 0:

The problem is reduced to the solving of an infinite system of linear algebraic


equations with a triple infinite of unknowns; the regularity of the system may be
proved.

12.3.2 Free Vibrations

Hereafter, we shall give results and computational hints about the free vibrations
of the domains considered in the previous section, by emphasizing chiefly the
Rayleigh waves,

12.3.2.1 Elastic Half-Space

Let be the elastic half-space x3  0, undergoing no load on the separation plane


x3 ¼ 0. As in Sect. 12.3.1.1, we choose stress functions of the form
XX
Fi ðx1 ; x2 ; x3 ; tÞ ¼ ðCnm sinxnm t þ Dnm cosxnm tÞðA0nm ecnm t
n m

þ B0nm ednm t Þ sin al x1 sin bm x2 ; . . .; ð12:79Þ

where the following relations take plane


12.3 Forced and Free Vibrations 541

1  2m q 2 q
a2n þ b2m ¼ c2nm þ x ¼ d2nm þ x2nm : ð12:80Þ
2ð1  mÞ l nm l
The components of the stress tensor tend to zero when x3 ! 1 and the
sequence of coefficients A0nm ; B0nm ; . . .; B000
nm are defined by the boundary conditions

x3 ¼ 0 : r33 ¼ 0; r31 ¼ r32 ¼ 0: ð12:81Þ


We find thus the conditions

½ð1  mÞc2nm þ md2nm Anm  ½mc2nm  ð1 þ mÞd2nm ðB0nm þ B00nm Þ


þ ð2c2nm  d2nm Þ ¼ 0;
2a2n cnm dnm Anm þ ½2a2n d2nm  ðc2nm  d2nm Þð2d2nm þ ma2n ÞB0nm
þ ½2a2n d2nm þ mðc2nm  d2nm Þðd2nm  a2n ÞB00nm
ð12:82Þ
þ ½2a2n d2nm þ ðc2nm  d2nm Þðmd2nm þ 2a2n ÞB000
nm ¼ 0;
2b2m cnm dnm Anm þ ½2b2m c2nm þ mðc2nm  d2nm Þðd2nm  b2m ÞB0nm
þ ½2b2m d2nm  ðc2nm  d2nm Þð2d2nm þ mb2m ÞB00nm
þ ½2b2m d2nm þ ðc2nm  d2nm Þðmd2nm þ 2b2m ÞB000
nm ¼ 0;

so as to obtain non-banal solutions, the following condition must be fulfilled


 2
4Dnm ¼ d2nm þ b2m þ a2n 4cnm dnm ða2n þ b2m Þ ¼ 0: ð12:83Þ
Using the propagation waves velocity, corresponding to Rayleigh’s wave,
xnm
c ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð12:84Þ
a2n þ b2m

and the non-dimensional ratio


c 2
2
v¼ ; ð12:85Þ
c
we find the relations
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiqffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1  2m
cnm ¼ 1 a2n þ b2m ;
2ð1  mÞv
sffiffiffiffiffiffiffiffiffiffiffi ð12:86Þ
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1
dnm ¼ 1 a2n þ b2m :
v

By so doing, the condition (12.83) leads to the equation


542 12 Dynamical Problems of Elastic Bodies

sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffiffiffiffiffi 1  2m
4v v  1 v  ¼ ð2v  1Þ2 ; ð12:87Þ
2ð1  mÞ

which, by squaring, takes the form


1
v3  ð2  mÞv2 þ ð1  mÞv  ð1  mÞ ¼ 0: ð12:870 Þ
8
We obtain thus the same equation for the sequence of eigenvalues as in the case of
a state of plane strain.
Let be the function
1
f ðvÞ ¼ v3  ð2  mÞv2 þ ð1  mÞv  ð1  mÞ ð12:88Þ
8
and its derivative
df ðvÞ
¼ 3v2  2ð2  mÞv þ 1  m; ð12:880 Þ
dv
if we assume that this derivative vanishes, we shall have
1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
v¼ 2  m  m2  m þ 1 1; ð12:89Þ
3
for 0 m 0:5. The form (12.81) of the equation shows that only the roots larger
then 1 are valid; for v  1, we have df ðvÞ=dv [ 0, so that f ðvÞ is a monotonously
growing function. We remark that f ð1Þ ¼ ð1  mÞ=8 \ 0, while f ð1Þ ¼ 1 [ 0,
which means that a single convenient root (greater than 1) of the Eq. (12.81) exists.
Thus, we obtain a single sequence of eigenvalues.
For m ¼ 1=4 (average value, adopted by Rayleigh for Poisson’s ratio), the Eq.
(12.870 ) becomes
3
4v3  7v2 þ 3v  ¼ 0; ð12:90Þ
2
and we obtain the value (one of the roots is 1=4)
1 pffiffiffi
v ¼ ð3 þ 3Þ ffi 1:1830: ð12:900 Þ
4
The relation (12.79) leads to the propagation velocity of Rayleigh waves
c2
c¼ ¼ 0:9194c2 ; ð12:9000 Þ
1:0877
as well, one obtains
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
cnm ¼ 0:8475 a2n þ b2m ; dnm ¼ 0:3933 a2n þ b2m : ð12:90000 Þ
12.3 Forced and Free Vibrations 543

For the limit value m ¼ 0, we obtain the equation


1
v3  2v2 þ v  ¼ 0; ð12:91Þ
8
hence (one of the roots is 1=2)
1 pffiffiffi
v ¼ ð3 þ 5Þ ¼ 1:3090; ð12:910 Þ
4
we get the propagation velocity
c2
c¼ ¼ 0:8740c2 ð12:9100 Þ
1:1441
and
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
cnm ¼ 0:7861 a2n þ b2m ; dnm ¼ 0:4859 a2n þ b2m : ð12:91000 Þ

If we remark that
anm 2Anm
¼ 2 ð12:92Þ
Dnm dnm þ a2n þ b2m

and if we introduce the notation

cnm ðtÞ ¼ Anm ða2n þ b2m ÞðCnm sin xnm t þ Dnm sin xnm tÞ; ð12:93Þ

we may express the state of stress in the form


(    
XX a2n 1 dnm x3
cnm x3
r11 ¼  unm ðtÞ e  1 e
n m a2n þ b2m 2v

m
þ ecnm x3 sin an x1 sin bm x2 ;
2ð1  mÞv
(    
XX b2m 1 dnm x3
r22 ¼  unm ðtÞ e cnm x3
 1  e ð12:94Þ
n m a2n þ b2m 2v

m
þ ecnm x3 sin an x1 sin bm x2 ;
2ð1  mÞv
 
1 XX  
r33 ¼ 1  u ðtÞ ecnm x3  ednm x3 sin an x1 sin bm x2 ;
2v n m nm
544 12 Dynamical Problems of Elastic Bodies

XX cnm bm  cnm x3 
r23 ¼  unm ðtÞ e  ednm x3 sin an x1 sin bm x2 ;
a 2 þ b2
n m n m
XX c an  
r31 ¼ unm ðtÞ nm 2 ecnm x3  ednm x3 cos an x1 sin bm x2 ; ð12:940 Þ
2
an þ bm
n m
XX    
an bm cnm x3 1 dnm x3
r12 ¼ unm ðtÞ e  1  e cos an x1 sin bm x2 ;
n m a2n þ b2m 2v

the state of displacement is supplied by



1 XX an
u1 ¼  x03 x2
þ u ðtÞ ecnm x3
2l n m nm a2n þ b2m
  
1 dnm x3
 1 e cos an x1 sin bm x2 ;
2v

1 XX bm
u2 ¼  x03 x2 þ unm ðtÞ ecnm x3
2l n m a2n þ b2m
   ð12:95Þ
1 dnm x3
 1 e sin an x1 sin bm x2 ;
2v

1 XX cnm
u3 ¼  u ðtÞ ecnm x3
2l n m nm a2n þ b2m

2v dnm x3
 e sin an x1 sin bm x2 ;
2v  1

where the rotation of rigid body x03 must be determined by an additional condition
of fixed point of the elastic half-space.
On the separation plane, appear the normal stresses

1 XX m
r11 ðx1 ; x2 ; 0; tÞ ¼  u ðtÞ
2v n m nm 1m
!
a2n
þ sin an x1 sin bm x2 ;
a2n þ b2m
 ð12:96Þ
1 XX m
r22 ðx1 ; x2 ; 0; tÞ ¼  u ðtÞ
2v n m nm 1m
!
b2m
þ sin an x1 sin bm x2
a2n þ b2m

and the tangential stress


1 XX an bm
r12 ðx1 ; x2 ; 0; tÞ ¼ u ðtÞ cos an x1 cos bm x2 : ð12:960 Þ
2v n m nm a2n þ b2m
12.3 Forced and Free Vibrations 545

We can now compute the displacements


1 XX an
u1 ¼ x03 x2 þ u ðtÞ cos an x1 sin bm x2 ;
4vl n m nm a2n þ b2m
ð12:97Þ
1 XX bm
u2 ¼ x03 x2 þ u ðtÞ sin an x1 cos bm x2 ;
4vl n m nm a2n þ b2m

the distorted form of the separation plane is supplied by


1 XX cnm
u3 ¼ u ðtÞ sin an x1 sin bm x2 : ð12:970 Þ
2ð2v  1Þl n m nm a2n þ b2m

If the stress functions (12.77) are expressed in cos an x1 and cos bm x2 we shall
obtain similar results and the same sequence of eigenvalues.
We come thus to a state of strain and stress corresponding to the Rayleigh
waves and generalizing the classical results suppplied in the case of a state of plane
strain.

12.3.2.2 Other Space Domains

In the case of: the elastic layer, the elastic half-layer or the elastic quarter-layer,
the elastic quarter-space or the elastic eights-space, the elastic strip or the elastic
half-strip, as in the case of the elastic parallelepiped, the study of the free
vibrations can be made, as in the previous subsection, by using the results obtained
in the case of forced vibrations.

References

A. Books

1. Beju, I., Soós, E., Teodorescu, P.P.: Euclidiean Tensor Calculus with Applications. Ed.
Tehnică, Bucuresßti, Abacus Press, Tunbridge Wells, Kent (1983)
2. Brekhovskikh, L.M.: Waves Layered Media. Appl. Math. Mech. 6, Academic Press, New
York (1960)
3. Cagniard, L.: Reflexion et refraction des ondes seismiques progressives. Gauthier-Villars,
Paris (1935)
4. Davies, R.M.: Stress Waves in Solids. Surv. Mech. Cambridge University Press, Cambridge
(1956)
5. Ewing, W.M., Jardetzky, W.S., Press, F.: Elastic Waves in Layered Media. McGraw-Hill
Book Co., Inc., New York (1957)
6. Ewing, W.M., Press, F.: Surface Waves and Guided Waves. Enc. Phys. 47, 75. Springer,
Berlin (1956)
7. Kecs, W., Teodorescu, P.P.: Applications of the Theory of Distributions in Mechanics. Ed.
Acad., Bucuresßti, Abacus Press, Tunbridge Wells, Kent (1974)
546 12 Dynamical Problems of Elastic Bodies

8. Kolsky, H.: Stress Waves in Solids. Oxford University Press, London (1953)
9. Love, A.-E.-H.: Treatise on the Mathematical Theory of Elasticity, 4th edn. Cambridge
University Press, London (1934)
10. Rayleigh, L., Strutt, J.W.: The Theory of Sound, 2nd edn. Dover Publishing, New York
(1945)
11. Teodorescu, P.P.: Dynamics of Linear Elastic Bodies. Ed. Acad., Bucuresßti, Abacus Press,
Tunbridge Wells, Kent (1975)

B. Papers

12. Broberg, K.B.: Shock waves in elastic and elastic-plastic media. Kungl. Fort. Befast,
Stockholm, 12 (1956)
13. Chao, C.C.: Dynamic response of an elastic half-space in tangential surface loadings. J. Appl.
Mech. 27, 559 (1960)
14. Chao, C.C., Bleich, H.H., Sackman, J.: Surface waves in an elastic half-space. ASME-Trans.
J. Appl. Mech. 300 (1961)
15. Davies, R.M.: Stress waves in solids. Appl. Mech. Rev. 6, 1 (1953)
16. Eason, G.: On the torsional impulsive loading of an elastic half-space. Quart. J. Mech. Appl.
Math 17, 279 (1964)
17. Eason, G.: On the torsional impulsive body force within an elastic half-space. Math 2, 75
(1964)
18. Flitman, L.M.: Dynamic problem of the die on an elastic half-space. J. Appl. Math. Mech 23,
997 (1959)
19. Hayes, M., Rivlin, R.S.: Surface waves in deformed elastic materials. Arch. Rat. Mech. Anal
8, 358 (1961)
20. Knopoff, L.: On rayleigh wave velocities. Bull. Seism. Soc. Amer 42, 307 (1952)
21. Lamb, H.: On the propagation of tremors over the surface of an elastic solid. Phil. Trans. Roy.
Soc. London Ser. A, 203, 1 (1904)
22. Lang, H.A.: Surface displacements in an elastic half-space. ZAMM 41, 141 (1961)
23. Miklowitz, J.: Recent developments in elastic wave propagation. Appl. Mech. Rev. 13, 12
(1960)
24. Noll, W.: Verschiebungsfunktionen für elastische Schwingungprobleme. ZAMM 37, 81
(1957)
25. Pekeris, C.C.: The seismic surface pulse. Proc. Nat. Acad. Sci 41, 469 (1955)
26. Pekeris, C.C., Lifson, H.: Motion on the surface of a uniform elastic medium. Bull. Seism.
Soc. Am 29, 1233 (1957)
27. Pinney, E.: Surface motion due to a point in a semi-infinite elastic medium. Bull. Seism. Soc.
Am 44, 571 (1954)
}
28. Predeleanu, M.: Uber die Verschiebungsfunktionen für das achsensymmetrische problem der
Elastodynamik. ZAMM 38, 402 (1958)
29. Rayleigh, L., Strutt, J.W.: On waves propagated along the plane surface of an elastic solid.
Proc. London Math. Soc. 17, 4 (1887)
30. Santer, F.: Der elastische Halbraun bei einer mechanischen Beeinflussung seiner Oberfläche.
ZAMM 30, 203 (1950)
31. Sherwood, J.W.C.: Elastic wave propagaton in a semi-infinite medium. Proc. Phys. Soc.
London 71, 207 (1958)
32. Teodorescu, P.P.: Schwingungen der elastischen Kontinua. Abh. der deutschen Akad. Wiss.,
Berlin, Kl. Math. Phys. Techn. 29 (1965)
}
33. Teodorescu, P.P.: Uber das dreidimensionale problem der elastokinetik. ZAMM 45, 513
(1965)
34. Teodorescu, P.P.: Sur quelques problèmes dynamiques de la théorie de l’élasticité. Rev.
Roum. Math. Pures Appl. 11, 773 (1966)
Chapter 13
Particular Cases of States of Strain
and Stress

The formulations of the fundamental problems by means of the potential functions


is one of the most used methods in the theory of elasticity. In the case of the first
fundamental problem (conditions in displacements on the boundary) one intro-
duces displacement functions, while in the case of the second problem (conditions
in stresses on the boundary) one introduces stress functions. The case of the mixed
problem may be studied by anyone of the two representations.
Obviously, in the case of a dynamic problem, intervenes the temporal variable
too, so that initial conditions must also be put.
Hereafter, we shall consider homogeneous, isotropic and linearly elastic bodies
in the case of infinitesimal deformations and in the absence of volume forces; the
potential functions depend, in general, on three variables (xi ; i ¼ 1; 2; 3; in case of
orthogonal Cartesian coordinates). The computation difficulties led to the study of
more simple problems, in two variables only. Thus, one considered plane problems
(plane state of stress or plane state of strain), as well as antiplane ones. In both
mentioned problems, one assumes, from the very beginning, that some of the
components of the stress or strain tensors vanish. Thus, in what follows, we shall
deal with problems in which one or two of the occurring components (e.g. normal
or tangential stress or linear or angular strain) are considered to be equal to zero.
The plane and the antiplane problems will be taken into consideration too.
A problem where the third variable (let be the variable x3 ) appears only by its
powers and where the potential functions depend only on two variables will be
called a two-dimensional problem. Such results have been given by J. H. Michell
[6], by A.-E.-H. Love [2] for a finite cylinder and by E. Almansi [4] for a thick
plate. Many results in this direction as well as various generalizations have been
given by G. Supino [8–13]. Similar problems have been considered by
Gr. C. Moisil [7], A. Davidescu-Moisil [5] and A. Clebsch [1].
We also dealt with these problems [3, 14–26], which we will use in what
follows.

P. P. Teodorescu, Treatise on Classical Elasticity, Mathematical and Analytical 547


Techniques with Applications to Engineering, DOI: 10.1007/978-94-007-2616-1_13,
Ó Springer Science+Business Media Dordrecht 2013
548 13 Particular Cases of States of Strain and Stress

13.1 Conditions for Stresses

Hereafter we shall put conditions for normal or for tangential stresses, obtaining
general stress functions for the states of stress thus put in evidence.

13.1.1 Case of a Zero Normal Stress

First of all, we deal with the general problem of elastostatics which results by
putting the condition that one normal stress vanishes, namely
r33 ¼ 0: ð13:1Þ

13.1.1.1 General Case

Taking into account the condition (13.1), in the absence of volume forces, the third
equation of equilibrium (3.62) leads to
r31 ¼ K2 ; r23 ¼ K;1 ; ð13:2Þ

the function K ¼ Kðx1 ; x2 ; x3 Þ being arbitrary. This function must be of class C1 :


In that follows, we will suppose that the functions which will be introduced are
continuous and differentiable as it is necessary; we will specify the class to which
belong the functions which are involved in the final results, taking into account the
necessities of computation.
The first two equations of equilibrium become now
r11;1 þ ðr12 þ K;3 Þ;2 ¼ 0; r22;2 þ ðr12  K;3 Þ;1 ¼ 0; ð13:3Þ

wherefrom
r11 ¼ A;2 ; r22 ¼ B;1 ;
ð13:20 Þ
r12 þ K;3 ¼ A;1 ; r12  K;3 ¼ B;2 ;

where A ¼ Aðx1 ; x2 ; x3 Þ; B ¼ Bðx1 ; x2 ; x3 Þ; the tangential stress becomes


1
r12 ¼  ðA;1 þB;2 Þ ð13:30 Þ
2
and one has
1
K;3 ¼  ðA;1  B;2 Þ: ð13:300 Þ
2
We denote, without losing the generality
13.1 Conditions for Stresses 549

 ;3 ;
A¼A  ;3 ;
B¼B ð13:4Þ
 ¼ Aðx
with A  1 ; x2 ; x3 Þ;  ¼ Bðx
B  1 ; x2 ; x3 Þ; so that the relation (3.30 ) becomes

1  
K ¼  ðA ;1  B;2 Þ þ C; ð13:3000 Þ
2
where C ¼ Cðx1 ; x2 Þ is an arbitrary function in two variables only.
Taking into account (13.30 ) and (13.4), the relations (13.2), (13.20 ) and (13.3)
allow us to represent the state of stress in the form
 ;23 ;
r11 ¼ A  ;31 ;
r22 ¼ B ð13:5Þ

1   1  
r23 ¼ ðA ;11  B;12 Þ  C;1 ; r31 ¼  ðA ;22  B;22 Þ þ C;2 ; ð13:50 Þ
2 2
1  
r;12 ¼  ðA ;31 þB;23 Þ: ð13:500 Þ
2
Taking now into account the condition (13.1), the third equation of Beltrami (5.43)
leads to
H;33 ¼ 0; ð13:6Þ

in the absence of the volume forces; it results that

H ¼ r11 þ r22 ¼ x3 H0 þ H0 ; ð13:7Þ


observing the harmonicity of the function H (the Eq. (5.41)), it follows that

DH0 ¼ 0; DH0 ¼ 0: ð13:70 Þ


The first two equations (5.430 ) become
1 1
Dr23 þ H0;2 ; Dr31 þ H0;1 ¼ 0; ð13:8Þ
1þm 1þm
now, (13.2) leads to
1 1
DK;1 ¼ H0;2 ; DK;2 ¼  H0;1 : ð13:80 Þ
1þm 1þm
Hence,
1
DK ¼ Du þ 
hðx3 Þ; ð13:800 Þ
1þm
so that the Eqs. (13.80 ) are of the Cauchy-Riemann form
Du;1 ¼ H0;2 ; Du;2 ¼ H0;1 ; ð13:8000 Þ
550 13 Particular Cases of States of Strain and Stress

the functions u ¼ uðx1 ; x2 Þ being biharmonique


DDu ¼ 0 ð13:9Þ
 ¼ hðx
and the function h  3 Þ being arbitrary.
Taking into account (13.5), the relation (13.7) leads to

A  ;1 ¼ 1 x23 H0 þ x3 H0 þ w;
 ;2 þB ð13:10Þ
2
where w ¼ wðx1 ; x2 Þ is an arbitrary function. Similarly, starting from (13.5),
(13.50 ) and introducing in the first two equations (5.43) and in the third Eq. (5.430 ),
one may write
   
DA ;23 þ 1 x3 H0;11 þ H0;11 ¼ 0; DB  ;13 þ 1 x3 H0;22 þ H0;22 ¼ 0;
1þm 1þm
ð13:11Þ
1    
 ;23 ¼ 1
DA;13 þDB x3 H0;12 þ H0;12 : ð13:110 Þ
2 1þm
Taking into account (13.70 ), the Eqs. (13.11) become
   
DA ;23 ¼ 1 x3 H0;22 þ H0;22 ; DB  ;13 ¼ 1 x3 H0;11 þ H0;11 ;
1þm 1þm
integrating, it results
 
1 1 2

DA x3 H0;2 þ x3 H0;2 þ f1 þ f2;2 ;
1þm 2
  ð13:12Þ
¼ 1
DB
1 2
x3 H0;1 þ x3 H0;1 þ g1 þ g2;1 :
1þm 2
To specify the function f1 ¼ f1 ðx1 ; x3 Þ and g1 ¼ g1 ðx2 ; x3 Þ; we remark that the
Eq. (13.110 ) must be verified too; one has
f1;13 þ g1;23 ¼ 0:

Introducing the arbitrary functions of only one variable h ¼ hðx3 Þ; hi ¼ hi ðx3 Þ;


i ¼ 1; 2; h3 ¼ h3 ðx1 Þ; h4 ¼ h4 ðx2 Þ; one obtains

f1 ¼ x1 
h þ h1 þ h3 ; g1 ¼ x 2 
h þ h2 þ h4 : ð13:13Þ

Applying Laplace’s operator to the relation (13.3000 ) and taking into account
(13.800 ), (13.12) and (13.13), it results that
1 1 1
DK ¼ 
h  ðf2  g2 Þ;12  ðh3;1  h4;2 Þ þ DC ¼ Du þ h:
2 2 1þm
13.1 Conditions for Stresses 551

Therefore, h¼ h ¼ h; where h ¼ hðx3 Þ; an arbitrary constant may be contained in


the other functions. Moreover, the functions h3 and h4 may be contained in the
functions f2 ¼ f2 ðx1 ; x2 Þ and g2 ¼ g2 ðx1 ; x2 Þ; respectively, which are specified by
the relation
2
ðf2  g2 Þ;12 ¼  Du þ 2DC: ð13:14Þ
1þm
Without losing the generality, one may denote

u¼u  0;12 ; C ¼ C
 ;12 ; H0 ¼ H  ;1122 ; ð13:140 Þ

the function C  ¼ Cðx


 1 ; x2 Þ being arbitrary; the functions u
¼u 0 ¼
 ðx1 ; x2 Þ and H

H0 ðx1 ; x2 Þ verify the equations
 ;12 ¼ 0:
u;12 ¼ 0; DH
DD

Taking into account the theorem of Boggio (see A.1.2.7), the functions u and H0
differ from a biharmonic and a harmonic function, respectively, only by a sum of
two functions of only one variable. Introducing in (13.14), one observes that the
latter functions may be neglected; one may thus suppose that the functions u
 and

H0 verify the equations

u0 ¼ 0;
DD 0 ¼ 0
DH ð13:15Þ
and that the Cauchy-Riemann equations
 0;2 ; D
u;1 ¼ H
D  0;1
u;2 ¼ H ð13:150 Þ

hold. Taking into account (13.140 ), the relation (13.14) leads to


2  ;12 ;
f 2  g2 ¼  u þ 2DC
D
1þm
where a sum of two functions of only one variable has been included in the
 (without losing the generality); it results that
arbitrary function C
1  ;12 ;
f2  x;12 ¼ ðg2  x;12 Þ ¼  u þ DC
D ð13:16Þ
1þm
where x ¼ xðx1 ; x2 Þ is an arbitrary function.
Introducing (13.13) and (13.16) in (13.12), one obtains
  
¼ 1 1 2  0 þ DC þx 1
DA x3 H0 þ x3 H  u;2  x1 h þ h1 ;
D
1þm 2 ;122 1 þ m
  
¼ 1 1 2  0  DC þx 1
DB x3 H0 þ x3 H þ u;1 þ x2 h þ h2 ;
D
1þm 2 ;112 1 þ m
ð13:17Þ
552 13 Particular Cases of States of Strain and Stress

where the notation

H0 ¼ H
0
;12 ð13:18Þ
0 ¼ H
has been used, the function H  0 ðx1 ; x2 Þ being harmonic (similar consider-
ations to those above)
 0 ¼ 0:
DH ð13:19Þ
Denoting further
 ;
w¼w ð13:180 Þ
;12

 ¼ wðx
where w  1 ; x2 Þ is an arbitrary function, the relation (13.10) becomes
 
 ;1 ¼ 1 x23 H
 ;2 þ B
A  0 þ x3 H 
0 þ w : ð13:20Þ
2 ;12

Let be
 ¼ U;1 ;
A  ¼ W;2 :
B ð13:200 Þ
The formulae (13.5)–(13.500 ) allow to write the state of stress in the form
r11 ¼ U;123 ; r22 ¼ W;123 ; ð13:21Þ

1 
 ;1122  W;22 ; 1 
 ;1122  W;22 ; ð13:210 Þ
r23 ¼ U;11  2C ;1
r31 ¼  U;11  2C
2 2
1 
r12 ¼  U;11 þ W;22 ;3 ; ð13:2100 Þ
2
where (13.140 ) has been taken into account too.
Introducing (13.200 ) in (13.17) and in (13.20) and integrating, one obtains
  
1 1 2 0  1  1
DU ¼ x3 H0 þ x3 H þ DC þ x þ H0 þ a1  x21 h þ x1 h1 ;
1þm 2 ;22 1 þ m 2
ð13:22Þ
  
1 1 2 1  1
DW ¼ x H0 þ x3 H0  þx
 DC þ H0 þ a2 þ x22 h þ x2 h2 ;
1þm 2 3 ;11 1þm 2

1  0 
U þ W ¼ x23 H 0 þ x 3 H þ w þ b 1 þ b2 ; ð13:220 Þ
2
where one took into account (13.150 ) and where one has introduced the functions
of two variables a1 ¼ a1 ðx2 ; x3 Þ; a2 ¼ a2 ðx1 ; x3 Þ; b1 ¼ b1 ðx2 ; x3 Þ; b2 ¼ b2 ðx1 ; x3 Þ:
Applying the operator D to the relation (13.220 ) and taking into account (13.22),
(13.15) and (13.19), it results
13.1 Conditions for Stresses 553

2  1  þ a1 þ a2
DðU þ WÞ ¼ H0 þ Dx  ðo11  o22 ÞDC
1þm 2
1  þ Dðb þ b Þ;
 0 þ Dw
 ðx21  x22 Þh þ x1 h1 þ x2 h2 ¼ H 1 2
2
which leads to
1
a1 ¼ Db1  x22 h  x2 h2 þ c2 þ c3 ;
2 ð13:23Þ
1
a2 ¼ Db2 þ x21 h  x1 h1 þ c1  c3 ;
2

 1  1m 
D x  w  ðo11  o22 ÞC ¼  H 0  c1  c2 ; ð13:230 Þ
2 1þm

where c1 ¼ c1 ðx1 Þ; c2 ¼ c2 ðx2 Þ; c3 ¼ c3 ðx3 Þ are functions of a single variable; one


observer that the functions may be neglected, because they may be included in the

functions b1 ; b2 and C:
The relation (13.230 ) leads to

x¼x
 þw   1  mH
 þ 1 ðo11  o22 ÞC  ;
0
2 1þm
 ¼ xðx
where x  1 ; x2 Þ is a harmonic function
 ¼0
Dx ð13:24Þ
 ¼H
and H  ðx ; x Þ is a particular integral of the equation
0 0 1 2

 ¼H
DH  0: ð13:240 Þ
0

The relation (13.220 ) allows to express the functions U and W in the form
 
1 1 2  þb þb ;
0 þ w
U¼vþ x3 H0 þ x3 H 1 2
2 2
  ð13:25Þ
1 1 2  þb þb ;
0 þ w
W ¼ v þ x3 H0 þ x3 H 1 2
2 2

where the function v ¼ vðx1 ; x2 ; x3 Þ may be determined by one of the equations


(13.22); one obtains thus

1 1 2  0  1m   ;1122
Dv ¼ x H0 þ x3 H  ð1  mÞH0 þ H0 þ x
 ;22 þ 2C
1þm 2 3 ;22 2ð1 þ mÞ
1   Þ þ 1 Dðb  b Þ  1 ðx2 þ x2 Þh þ x1 h1  x2 h2 þ c :
 ðw w 1 2 3
2 ;11 ;22
2 2 1 2

ð13:250 Þ
554 13 Particular Cases of States of Strain and Stress

Introducing (13.25) in (13.29)–(13.2900 ) one obtains the state of stress in the


form
1 0 þ H  0 Þ ; r22 ¼ v;123 þ 1 ðx3 H 0 þ H  0Þ ;
r11 ¼ v;123 þ ðx3 H ;12 ;12 ð13:26Þ
2 2

 
1 1 1   0  1
r23 ¼ Dv  v;33 þ ðo11  o22 Þ x23 H 0 þ x 3 H þ w þ b2;11
2 2 2 2

 2C  ;1122 ¼  1 v;133 þ 1
ðo11  o22 Þ ð1  mÞH 
0
;1 2 4ð1 þ mÞ
 
1  0 1  1
þm x23 H 0 þ x3 H  x  ;11  x ;22 þ b2;33 ;1  ðx1 h  h1 Þ;
2 ;1 4 2

  ð13:260 Þ
1 1 1 2 0 þ w   b 1
r31 ¼  Dv  v;33 þ ðo11  o22 Þ x3 H0 þ x3 H
2 2 2 2 1;22

 2C  ;1122 ¼ 1 v;233  1
ðo11  o22 Þ ð1  mÞH 
0
;2 2 4ð1 þ mÞ
 
1  0 1  1
þm x23 H 0 þ x3 H þ x  ;11  x ;22  b1;33 ;2 þ ðx2 h  h2 Þ;
2 ;2 4 2

1 1 
r12 ¼  v;11  v;22 þ b2;11 þ b1;22 ð13:2600 Þ
2 2 ;3

where (13.250 ), (13.15), (13.19), (13.24), (13.240 ) have been taken into consider-
ation too.
One remarks that the function v appears only by its derivative with respect to
x3 . Thus, in the right member of the Eq. (13.250 ) one may neglect the functions
which depend only on the variables x1 and x2 ; indeed, these functions lead to
particular integrals which may depend only on the variables x1 and x2 ; which
disappear by differentiation.
Taking into account (13.15), (13.19), the Eq. (13.250 ) allows to write
 
1 1   0 1
Dv ¼  ðo11  o22 Þ x23 H 0 þ x 3 H þ Dðb1  b2 Þ
2ð1 þ mÞ 2 2
1 2 
 x1 þ x22 h þ x1 h1  x2 h2 þ c3 ;
2
wherefrom
 
1 1   0 1
v ¼
v ðx1 o1  x2 o2 Þ x23 H 0 þ x 3 H þ ðb1  b2 Þ
4ð1 þ mÞ 2 2
Z Z Z Z Z Z
1 2 
 x1 þ x22 dx3 hdx3 þ 2 dx3 dx3 dx3 hdx3
2
Z Z Z Z Z Z
þ x1 dx 3 h 1 dx3  x2 dx 3 h 2 dx3 þ dx3 c3 dx3 ;
13.1 Conditions for Stresses 555

v¼
where  vðx1 ; x2 ; x3 Þ is a harmonic function

v ¼ 0;
D ð13:27Þ
the relation (A.102) has been taken into account too.
It results that
Z
1  
v;3 ¼ X  ð x 1 o1  x 2 o2 Þ ð x 3 H  0 Þ þ 1 ðb1  b2 Þ  1 x2 þ x2
0 þH hdx3
;3 1 2
4ð1 þ mÞ 2 2
Z Z Z Z Z Z
þ 2 dx3 dx3 hdx3 þ x1 h1 dx3  x2 h2 dx3 þ c3 dx3 ;

ð13:28Þ
with the notation
X¼
v;3 :

The conditions (13.27) allows to state that also the function X ¼ Xðx1 ; x2 ; x3 Þ is
harmonic
DX ¼ 0: ð13:29Þ
Introducing (13.28) in (13.26), one obtains

1  0 Þ þ 1 ðx3 H
0 þ H 0 þ H
 0Þ ;
r11 ¼ X;12  ðx1 o112  x2 o122 Þðx3 H ;12
4ð1 þ mÞ 2
ð13:30Þ
1 0 þ H 0 Þ þ 1 ðx3 H
0 þ H
 0Þ :
r22 ¼ X;12 þ ðx1 o112  x2 o122 Þðx3 H ;12
4ð1 þ mÞ 2

As well, introducing (13.28) in (13.2600 ), one has


1 1  0 Þ:
0 þ H
r12 ¼  ðo11  o22 ÞX þ ðx1 o1  x2 o2 Þðo11  o22 Þðx3 H
2 8ð1 þ mÞ
ð13:300 Þ
We introduce now (13.28) in (13.260 ); it results
556 13 Particular Cases of States of Strain and Stress

  
1 1 1 2  0 
r23 ¼  X;13 þ ðo11  o22 Þ m x3 H0 þ x3 H þ ð1  mÞH0
2 4ð1 þ mÞ 2 ;1
1 1
 0;1  ðo11  o22 Þx
þ ð x 1 o1  x 2 o2 Þ H  ;1 ;
8ð1 þ mÞ 4
  
1 1 1 2  0 
r31 ¼ X;23  ðo11  o22 Þ m x3 H0 þ x3 H þ ð1  mÞH0
2 4ð1 þ mÞ 2 ;2
1 1
 0;2 þ ðo11  o22 Þx
 ð x 1 o1  x 2 o2 Þ H  ;2 :
8ð1 þ mÞ 4
ð13:3000 Þ
Taking into account (13.240 ), one may write
 Z Z  
 1   1  0;1 þ 2H 0
ðo11  o22 ÞH0 ¼ ðo11  o22 Þ x1 H0 dx1 þ x2 H0 dx2 ¼ x1 H
4 4
Z  Z 
    1  0;
þ x2 H0;11 dx2  x2 H0;2 þ 2H0 þ x1 H0;22 dx1 þ ð x 1 o1  x 2 o2 Þ H
2

where a particular integral of the Eq. (13.240 ) has been used.


As well, the Cauchy-Riemann Eqs. (13.150 ) and the property expressed by the
formula (A.102), lead to
1  0 ¼  1 Dðx1 o2 þ x2 o1 ÞD 1
Dðx1 o1  x2 o2 ÞH u ¼  ðo12 þ o21 ÞD
u ¼ D
u;12 ;
4 4 2
hence,
1  0;
 ;12 ¼ u0  ðx1 o1  x2 o2 ÞH
u
4
where u0 ¼ u0 ðx1 ; x2 Þ is a harmonic function

Du0 ¼ 0:
Thus, the tangential stresses (13.3000 ) become
 
1 m 1 2  0
r23 ¼ X;13 þ ðo11  o22 Þ x3 H0 þ x3 H
2 4ð1 þ mÞ 2 ;1
2m   1
   u0 ;1  ðo11  o22 Þx
u  ;1 ;
2ð1 þ mÞ ;12 4
 
1 m 1 2  0
r31 ¼  X;23  ðo11  o22 Þ x3 H0 þ x3 H
2 4ð1 þ mÞ 2 ;2
2m   1
þ   u0 ;2 þ ðo11  o22 Þx
u  ;2 :
2ð1 þ mÞ ;12 4
13.1 Conditions for Stresses 557

The harmonic function x  is arbitrary, hence the function ðo11  o22 Þx


 has the
same property; it can be included in the function x0 ; which is arbitrary too.
Moreover, the harmonic function x0 may be included in the biharmonic function
u¼u  ;12 ; the harmonic part of which is also arbitrary (indeed, the relations
(13.150 ) involve only Du). Thus, the functions u0 and x  may be taken equal to
zero, without losing the generality of the representation; indeed, they may corre-
spond to the harmonic part of the function u:
Taking into account the notations (13.140 ) and (13.18) and the Eqs. (13.15) and
(13.19), one obtains the state of stress in the final form
1 1
r11 ¼ X;12  ðx1 o1  x2 o2 ÞH þ H;
4ð1 þ mÞ 2
ð13:31Þ
1 1
r22 ¼ X;12 þ ðx1 o1  x2 o2 ÞH þ H;
4ð1 þ mÞ 2
(   )
1 m 1 2 2
r23 ¼  X;13 þ x Du þ 1  u x3 H;2 ;
2 2ð1 þ mÞ 2 3 m ;1
(   ) ð13:310 Þ
1 m 1 2 2
r31 ¼ X;23  x Du þ 1  u x3 H;1 ;
2 2ð1 þ mÞ 2 3 m ;2

1 1
r12 ¼  ðo11  o22 ÞX  ðx1 o2  x2 o1 ÞH: ð13:3100 Þ
2 4ð1 þ mÞ
The function X ¼ Xðx1 ; x2 ; x3 Þ is harmonic (Eq. (13.29)). The sum of the
normal stresses is a linear with respect to x3 function, given by (13.7); the func-
tions H0 ¼ H0 ðx1 ; x2 Þ and H0 ¼ H0 ðx1 ; x2 Þ are harmonic (Eq. (13.70 )). The bi-
harmonic function u ¼ uðx1 ; x2 Þ (Eq. (13.9)) is connected by its Laplacian to the
function H0 (Cauchy-Riemann Eqs. (13.800 )).
One also observes that the function X must be of class C4 ; as well as the
function u; it is sufficient that the functions H0 and H0 be of class C3 :
Moreover, one remarks that the given representation depends only on three
arbitrary functions, i.e.: the biharmonic function X (in three variables), the har-
monic function H0 and the biharmonic function u (the two latter functions in two
variables). Because one must determine a function in three variables too, the
problem is, in general, a quasi-bidimensional problem.
Starting from the equations of the problem, one arrived step by step to the final
representation; hence, this representation is complete (any state of stress may be
represented in this form).

13.1.1.2 Case of a State of Incompressible Deformation

As it has been shown in Sect. 4.1.3.7, if m ¼ 1=2 one has to do with an elastic
incompressible body; as well, if H ¼ 0 (without having m ¼ 1=2), one has a state
558 13 Particular Cases of States of Strain and Stress

of incompressible (isochore) deformation, which does not depend on the


mechanical properties of the body, but on a particular state of loading.
Taking into account (13.7), it follows that, in the case of an isochore state of
deformation, one has

H0 ¼ H0 ¼ 0: ð13:32Þ

The Cauchy-Riemann equations (13.150 ) lead to


Du ¼ C; C ¼ const;
it results that
2ð1 þ mÞ C
u¼ f þ ðx21 þ x22 Þ;
2m 4
where f ¼ f ðx1 ; x2 Þ is a harmonic function
Df ¼ 0:
Introducing in (13.31)–(13.3100 ), one observes that a part of the state of stress is
given by
r11 ¼ r22 ¼ r12 ¼ 0; ð13:33Þ
2m 2m
r23 ¼  Cx1 ; r31 ¼ Cx2 ; ð13:330 Þ
4ð1 þ mÞ 4ð1 þ mÞ
but this state of stress may be obtained by means of the stress function

2m   2
X¼ C x21 þ x22 x3  x33 ; ð13:3300 Þ
4ð1 þ mÞ 3

the last term of which appears because the function must be harmonic.
One can put C ¼ 0 and the state of stress takes the form
r11 ¼ r22 ¼ X;12 ; ð13:34Þ
 
1 1
r23 ¼  X;13  f;1 ¼  X;3 þ f ;
2 2 ;1
  ð13:340 Þ
1 1
r31 ¼ X;23 þ f;2 ¼ X;3 þ f ;
2 2 ;2

1
r12 ¼  ðo11  o22 ÞX: ð13:3400 Þ
2
Hence, the stress function X ¼ Xðx1 ; x2 ; x3 Þ corresponds to an incompressible state
of deformation. The function f must be of class C3 ; in the compressible case, it is
included in the function u:
13.1 Conditions for Stresses 559

13.1.1.3 State of Displacement

Introducing (13.31) in Hooke’s law (4.3), one obtains


1 1
Ee11 ¼ ð1 þ mÞX;12  ðx1 o1  x2 o2 ÞH þ ð1  mÞH;
4 2 ð13:35Þ
1 1
Ee22 ¼ ð1 þ mÞX;12 þ ðx1 o1  x2 o2 ÞH þ ð1  mÞH:
4 2
Ee33 ¼ mH: ð13:350 Þ

Taking into account (2.43) and introducing the notation (13.140 ) and (13.18), with
 ¼ x3 H
H  0;
0 þ H ð13:36Þ

one may integrate the Eqs. (13.35), (13.350 ); it results

2l½u1  ðx2 x03 þ x3 x02 þ u01 Þ

¼ X;2 
1  þ 1m H
ðx1 o1  x2 o2 ÞH  ;2 ;
4ð1 þ mÞ 2ð1 þ mÞ
ð13:37Þ
2l½u2  ðx3 x01 þ x1 x03 þ u02 Þ

¼ X;1 þ
1  þ 1m H
ðx1 o1  x2 o2 ÞH  ;1 ;
4ð1 þ mÞ 2ð1 þ mÞ
 
0 0 0 1 2
2l½u3  ðx1 x2 þ x2 x1 þ u3 Þ ¼ m x3 H0  x3 H þ g; ð13:370 Þ
2

where the function g ¼ gðx1 ; x2 Þ remains to be determined. The arbitrary functions


which appear in (13.37) have not be mentioned, because they do not intervene in
the computation.
With the aid of the Eqs. (13.15), (13.19) and of the notation (13.36), it results
 ¼ 0;
DH ð13:38Þ

then, one has


 ¼ 2ðx1 o2 þ x2 o1 ÞH:
ðo11  o22 Þðx1 o1  x2 o2 ÞH
The third relation (4.30 ) with (2.430 ) is thus easily verified. The first two rela-
tions (4.30 ) with (2.430 ) lead to

1 1  1 
 ðx1 o1  x2 o2 ÞH0 þ ð1  mÞH0  ð2  mÞu þg;1 ¼ 0;
ð1 þ mÞ 4 2 ;2

1 1 1  0 þ ð2  mÞu þg;2 ¼ 0:
 0 þ ð1  mÞH
ð x 1 o1  x 2 o2 Þ H
ð1 þ mÞ 4 2 ;1
560 13 Particular Cases of States of Strain and Stress

Taking into account (13.140 ), (13.150 ) with u0 ¼ 0; it results


   
1m 1 1m 1
g;1 ¼  ;22  D
u u ; g;2 ¼ Duu ;11 ;
1þm 2 ;1 1þm 2 ;2

wherefrom the particular integral


1m
g¼ ðo11  o22 Þ
u:
2ð1 þ mÞ
The displacement u3 is given by
 
2l u3  x1 x02 þ x2 x01 þ u03
 
1 2 1m
¼ m x3 H0  x3 H  ðo11  o22 Þ
u ð13:3700 Þ
2 2ð1 þ mÞ
One observes that the functions H  0 and u
 0; H  are easily obtained from the
0
relations (13.14 ), (13.18), with the supplementary conditions that the two first
functions must be harmonic and the function u  must be biharmonic.
In the case of a state of incompressible deformation there result the
displacements
 
2l u1  x2 x03 þ x3 x02 þ u01 ¼ X;2 ;
 
2l u2  x3 x01 þ x1 x03 þ u02 ¼ X;1 ; ð13:38Þ
 
2l u3  x1 x02 þ x2 x01 þ u03 ¼ g;

where the function g ¼ gðx1 ; x2 Þ is harmonic


Dg ¼ 0 ð13:39Þ
and, together with the function f ¼ f ðx1 ; x2 Þ; verifies the Cauch-Riemann
equations
g;1 ¼ 2f;2 ; g;2 ¼ 2f;1 ; ð13:390 Þ

Hooke’s law is thus identically verified.

13.1.1.4 Other Conditions for a Normal Stress

Analogically, one may study the case in which


r33;3 ¼ 0; ð13:40Þ

as well as the case in which


r33;33 ¼ 0: ð13:400 Þ
13.1 Conditions for Stresses 561

These two problems may be put in connection with the study of plates of mean
thickness, acted upon by loads normal to the middle plane in the case of a state of
stress symmetric or antisimmetric with regard to the middle plane, respectively.
In the first case, the sum of the normal stresses is given by
1
H ¼  x23 DH0 þ x3 H0 þ H0 ð13:41Þ
2
and the normal stress r33 takes the form
1
r33 ¼ H0 þ #; ð13:410 Þ
1þm
where the function H0 ¼ H0 ðx1 ; x2 Þ is biharmonic

DDH0 ¼ 0 ð13:42Þ

and the functions H0 ¼ H0 ðx1 ; x2 Þ; # ¼ #ðx1 ; x2 Þ are harmonic


DH0 ¼ 0; D# ¼ 0: ð13:420 Þ
In the second case, the function H ¼ Hðx1 ; x2 ; x3 Þ is expressed in the form
1 1
H ¼  x33 DH0  x23 DH0 þ x3 H0 þ H0 ð13:43Þ
6 2
and the normal stress r33 takes the form
1
r33 ¼ ðx3 H0 þ H0 Þ þ x3 #0 þ #0 ; ð13:430 Þ
1þm
where the functions H0 ¼ H0 ðx1 ; x2 Þ; H0 ¼ H0 ðx1 ; x2 Þ are biharmonic

DDH0 ¼ 0; DDH0 ¼ 0 ð13:44Þ

and the functions #0 ¼ #0 ðx1 ; x2 Þ; #0 ¼ #0 ðx1 ; x2 Þ are harmonic

D#0 ¼ 0; D#0 ¼ 0: ð13:440 Þ


What concerns the state of stress, one obtains formulae analogue to the for-
mulae (13.31)–(13.3100 ). For instance, in the case (13.40), the formulae analogue to
the formulae (13.31) are

1 1 0 1
r11 ¼ X;12  ðx1 o1  x2 o2 ÞH þ H þ ðH  #Þ;
2ð1 þ mÞ 2 2
 ð13:45Þ
1 1 0 1
r22 ¼ X;12 þ ðx1 o1  x2 o2 ÞH  H þ ðH  #Þ:
2ð1 þ mÞ 2 2
562 13 Particular Cases of States of Strain and Stress

13.1.2 Particular Cases

Starting from the results obtained above for the case of a zero normal stress, one
may put supplementary conditions to the other components of the stress tensor,
obtaining results which can be interesting in various problems.

13.1.2.1 Case of Two Zero Normal Stresses

Let be the supplementary condition


r22 ¼ 0; ð13:46Þ
which leads to
r11 ¼ H: ð13:460 Þ
The representation (13.31) allows to write
1 1
X;12 ¼ H þ ðx1 o1  x2 o2 ÞH: ð13:47Þ
2 4ð1 þ mÞ
Applying Laplace’s operator and taking into account (13.29) and the formula
(A.102), it results
ðo11  o22 ÞH ¼ 0; ð13:48Þ

but this function is harmonic, so that


H;11 ¼ H;22 ¼ 0: ð13:480 Þ

Taking into account (13.7), it results


H ¼ ax1 x2 þ bx1 þ cx2 þ d; ð13:49Þ
where

a ¼ a0 x 3 þ a0 ; b ¼ b0 x 3 þ b0 ; c ¼ c 0 x 3 þ c 0 ; d ¼ d0 x 3 þ d 0 ; ð13:490 Þ
a0 ; a0 ; b0 ; . . .; d0 being arbitrary constants.
The relation (13.47) becomes
1 1
X;12 ¼ ðax1 x2 þ bx1 þ cx2 þ dÞ þ ðbx1  cx2 Þ;
2 4ð1 þ mÞ
13.1 Conditions for Stresses 563

hence
 
1 1 2 2
X¼ ax1 x2 þ bx21 x2 þ cx1 x22 þ 2dx1 x2
4 2
1
þ ðbx21 x2  cx1 x22 Þ þ u1 þ u2 ; ð13:50Þ
8ð1 þ mÞ
where u1 ¼ u1 ðx1 ; x3 Þ; u2 ¼ u2 ðx2 ; x3 Þ.
The condition (13.29) leads to
1  2  1
a x1 þ x22 þ 2ðbx2 þ cx1 Þ þ ðbx2  cx1 Þ þ Du1 þ Du2 ¼ 0;
4 4ð1 þ mÞ

hence
1  2  1
a x1 þ x22 þ 2ðbx2 þ cx1 Þ þ ðbx2  cx1 Þ ¼ d1  d2 ;
4 4ð1 þ mÞ
Du1 ¼ d3;33  d1 ; Du2 ¼ d2  d3;33 :

with d1 ¼ d1 ðx1 Þ;
d2 ¼ d2 ðx2 Þ; d3 ¼ d3 ðx3 Þ; it results
1 2  1
d1 ¼ ax1 þ 2cx1  cx1 þ e0 ; d2
4 4ð1 þ mÞ
1  1
¼  ax22 þ 2bx2  bx2 þ e0 ;
4 4ð1 þ mÞ
e0 being an arbitrary constant.
One may also write
 
1 1 4 1 1
u¼u 1  3
ax þ cx1 þ cx3  e0 x2 þ d3 ;
12 4 1 24ð1 þ mÞ 1 2 1
 
1 1 4 1 1
u¼u 2  3
ax þ bx2  bx3 þ e0 x2  d3 ;
12 4 2 24ð1 þ mÞ 2 2 2
1 ¼ u
where u  1 ðx1 ; x3 Þ; u
2 ¼ u
 2 ðx2 ; x3 Þ are harmonic functions in two variables
u1 ¼ D
D u2 ¼ 0:
Introducing in (13.50), one obtains
1  4     
X¼ a x1  6x21 x22 þ x42 þ 4b x32  3x21 x2 þ 4c x31  3x1 x22  24dx1 x2
48
  1  3   
þ24e0 x21  x22  b x2  3x21 x2  c x31  3x1 x22 þ u1 þ u 2:
24ð1 þ mÞ
Taking into account (13.49), (13.490 ), the Eqs. (13.8000 ) become
Du;1 ¼ a0 x1 þ c0 ; Du;2 ¼ a0 x2  b0 ;
564 13 Particular Cases of States of Strain and Stress

it results
1  
Du ¼ a0 x21  x22 þ c0 x1  b0 x2 þ e0
2
and
1  4  1  1  
u ¼ u0 þ a0 x1  x42 þ c0 x31  b0 x32 þ e0 x21 þ x22 ;
24 6 4
where u0 ¼ u0 ðx1 ; x2 Þ is a harmonic function
Du0 ¼ 0
and e0 is an arbitrary constant.
We introduce the functions in two variables F23 ¼ F23 ðx2 ; x3 Þ; F31 ¼
F31 ðx3 ; x1 Þ by means of the Cauchy-Riemann equations
   
 1;1 ;3 ¼ 2F31;1 ; u
u  1;1 ;1 ¼ 2F31;1 ;
   
 2;2 ;3 ¼ 2F23;2 ; u
u  2;2 ;2 ¼ 2F23;3 :

Let be also the function F12 ¼ F12 ðx1 ; x2 Þ; specified by


2m
F12 ¼  u:
2ð1 þ mÞ 0
Obviously, the functions F23 ; F31 ; F12 are harmonic
DF23 ¼ 0; DF31 ¼ 0; DF12 ¼ 0
Introducing the results thus obtained in (13.31)–(13.3100 ), it results the state of
stress

r11 ¼ ax1 x2 þ bx1 þ cx2 þ d ¼ a0 x1 x2 x3 þ c0 x2 x3


þ b0 x3 x1 þ a0 x1 x2 þ b0 x1 þ c0 x2 þ d0 x3 þ d0 ; ð13:51Þ
13.1 Conditions for Stresses 565

m 2
r23 ¼ ðF12  F31 Þ;1  x3 ða0 x1 þ c0 Þ þ 2x3 ða0 x1 þ c0 Þ
4ð1 þ mÞ
 
2m 1 3 2 1  
 a0 x1 þ c0 x1 þ e0 x1  2b0 x1 x2 þ c0 x21  x22
4ð1 þ mÞ 3 16ð1 þ mÞ
1  3   
þ a0 x1  3x1 x2  6b0 x1 x2 þ 3e0 x21  x22  6d0 x2 ;
2
24
m 2
r31 ¼ ðF23  F12 Þ;2  x3 ða0 x2 þ b0 Þ þ 2x3 ða0 x2 þ b0 Þ
4ð1 þ mÞ
 
2m 1 3 2 1  2 
 a0 x 2 þ b0 x 2  e 0 x 2 þ b0 x1  x22  2c0 x1 x2
4ð1 þ mÞ 3 16ð1 þ mÞ
1  3   
 a0 x2  3x21 x2  3b0 x21  x22  6e0 x1 x2  6d0 x1 ;
24
1  2 
r12 ¼ ðF31  F23 Þ;3  a x1 þ x22 þ 2ðbx2 þ cx1 Þ
4ð1 þ mÞ
1  
þ a x21  x22  2ðbx2  cx1 Þ þ 4e0 :
4

We consider now the following harmonic stress functions


1   1  
F23 ¼  a0 x42  6x22 x23 þ x43  b0 x32  3x2 x23
48ð1 þ mÞ 6ð1 þ mÞ
1 2m   1
þ b0 x2 x3 þ e0 x22  x23  e0 x3 ;
2ð1 þ mÞ 8ð1 þ mÞ 2
m  4  m  
F31 ¼ 2 2
a0 x1  6x1 x3 þ x3  4
a0 x33  3x21 x3
18ð1 þ mÞ 12ð1 þ mÞ
m  3  m 2m   1
 c0 x1  3x1 x3 þ2
c0 x1 x3 þ e0 x21  x23 þ e0 x3 ;
12ð1 þ mÞ 2ð1 þ mÞ 8ð1 þ mÞ 2
1  4 2 2 4
  3 2
  3 2

F12 ¼ a0 x1  6x1 x3 þ x2 þ 4b0 x2  3x1 x2 þ 4c0 x1  3x1 x2
96
1  3   
24d0 x1 x2  þ b0 x2  3x21 x2  c0 x31  3x1 x22 ;
48ð1 þ mÞ

introducing some terms of the tangential stresses in these functions, one may—
without losing the generality—express these stresses in the form
1
r23 ¼ ðF12  F31 Þ;1  ða0 x1 þ 3c0 Þx21 ; ð13:510 Þ
6ð1 þ mÞ
566 13 Particular Cases of States of Strain and Stress


m
r31 ¼ ðF23  F12 Þ;2  x2 ða0 x2 þ b0 Þ þ 2x3 ða0 x2 þ b0 Þ:
4ð1 þ mÞ 3

1 1 1  
 x22 ða0 x2 þ 3b0 Þ  a0 x2 x22 þ 3x23 þ x3 ðb0 x3 þ b0 Þ ;
3 2ð1 þ mÞ 6
 ð13:5100 Þ
m
r12 ¼ ðF31  F23 Þ;3  x2 ða0 x3 þ a0 Þ þ 2x2 ðb0 x3 þ b0 Þ:
4ð1 þ mÞ 2

1 2 0 1 1  2 2
 0 0
 x3 ða0 x3 þ 3a Þ  a0 x3 x3 þ 3x2 þ x2 ða x2 þ b Þ ;
3 2ð1 þ mÞ 6

where the functions F23 ; F31 ; F12 are of class C3 :


Using the formulae (2.43), (2.430 ) and (4.300 ), one obtains the state of
displacement

E½u1  ðx2 x03 þ x3 x02 þ u01 Þ ¼ 2ð1 þ mÞF  23 þ 1 a0 x2 x3 ½3x21  ðx22 þ x23 Þ
6
1 0 1
þ a x2 ð3x21  2x22 Þ þ b0 x3 ð3x21  2x23 Þ
6 6
1 0 2 2 2
þ b ½x1  ðx2 þ x3 Þ
2
þ c0 x1 x2 x3 þ c0 x1 x2 þ d0 x3 x1 þ d0 x1 ;
ð13:52Þ

E½u2  ðx3 x01 þ x1 x03 þ u02 Þ ¼ 2ð1 þ mÞF  31  1 a0 x1 x3 ½x21 þ mð3x22  x23 Þ
6
1 0
 a x1 ½x1 þ 3mðx22  x23 Þ  mb0 x1 x2 x3
2
6
1
 mb0 x1 x2  c0 x3 ½3x21 þ mð3x22  x23 Þ
6
1 0 2
 c ½x1 þ mðx22  x23 Þ  md0 x2 x3  md 0 x2 ;
2
0 0 0
E½u3  ðx1 x2 þ x2 x1 þ u3 Þ ¼ 2ð1 þ mÞF  12  1 a0 x1 x2 ½x21 þ mðx22  3x23 Þ
6
0 1
 ma x1 x2 x3  b0 x1 ½x21  3mðx22  x23 Þ
6
0 1
 mb x1 x3  c0 x2 ½3x21  mðx22  3x23 Þ
6
1
 mc0 x2 x3  ½x21  mðx22  x23 Þ  md0 x3 ;
2
ð13:520 Þ
 23 ¼ F
where the functions in two variables F  23 ðx2 ; x3 Þ; F
 31 ¼ F
 31 ðx3 ; x1 Þ; F
 12 ¼

F12 ðx1 ; x2 Þ are harmonic
13.1 Conditions for Stresses 567

 23 ¼ 0; DF
DF  31 ¼ 0; DF
 12 ¼ 0 ð13:53Þ
and verify Cauchy-Riemann equations
 23;3 ; F23;3 ¼ F
F23;2 ¼ F  23;2 ; ð13:54Þ
 31;1 ; F31;1 ¼ F
F31;3 ¼ F  31;3 ; ð13:540 Þ
 12;2 ; F12;2 ¼ F
F12;1 ¼ F  12;1 : ð13:5400 Þ

13.1.2.2 Antiplane State of Stress

If, in the preceding case, we put the supplementary condition


r23 ¼ 0; ð13:55Þ

then one obtains an antiplane state of stress. This condition leads to


1
ðF12  F31 Þ;1 ¼ ða0 x1 þ 3c0 Þx21 :
6ð1 þ mÞ
The first member of this relation is a harmonic function, but the second one has
not this property, hence
a0 ¼ c0 ¼ 0:
Taking into account that the functions F12 ¼ F12 ðx1 ; x2 Þ and F31 ¼ F31 ðx3 ; x1 Þ
are harmonic, it follows that
1
F31 ¼ k1 ðx21  x23 Þ þ k2 x1 þ k3 x3 þ k4 ;
2
1
F12 ¼ k1 ðx21  x22 Þ þ k2 x1 þ k5 x2 þ k6 ;
2
where k1 ; k2 ; . . .; k6 are arbitrary constants.
Observing that one can use a stress function of the form
F23 ¼ k5 x2  k3 x3
and introducing the terms which depend on the constants k3 ; k5 in the function F23 ;
it results the state of stress

r11 ¼ b0 x3 x1 þ a0 x1 x2 þ b0 x1 þ c0 x2 þ d0 x3 þ d0 ; ð13:56Þ
568 13 Particular Cases of States of Strain and Stress

1 nm  
r31 ¼ F23;2 þ b0 x22  x23  2x3 ða0 x2 þ b0 Þ
2ð1 þ mÞ 2
o
 x 3 ð b0 x 3 þ b0 Þ þ k 1 x 2 ;
nm   ð13:560 Þ
1
r12 ¼  F23;3  a0 x22  x23 þ 2x2 ðb0 x3 þ b0 Þ
2ð1 þ mÞ 2
o
þ x 2 ð a0 x 2 þ b0 Þ  k 1 x 3 :

The state of displacement is given by

E½u1  ðx2 x03 þ x3 x02 þ u01 Þ ¼ 2ð1 þ mÞF  23 þ 1 a0 x2 ð3x21  2x22 Þ


6
1 1
þ b0 x3 ð3x1  2x23 Þ þ b0 ½x21  ðx22 þ x23 Þ
2
6 2
þ c0 x1 x2 þ d0 x3 x1 þ d0 x1 ; ð13:57Þ

1
E½u2  ðx3 x01 þ x1 x03 þ u02 Þ ¼  a0 x1 ½x21 þ 3mðx22  x23 Þ þ mb0 x1 x2 x3
6
1
 mb0 x1 x2  c0 ½x21 þ mðx22  x23 Þ
2
 md0 x2 x3  md 0 x2  2ð1 þ mÞk1 x3 x1 ;
1
E½u3  ðx1 x02 þ x2 x01 þ u03 Þ ¼  ma0 x1 x2 x3  b0 x1 ½x21  3mðx22  x23 Þ ð13:570 Þ
6
1
 mb x3 x1  mc0 x2 x3  d0 ½x21  mðx22  x23 Þ
0
2
 md0 x3 þ 2ð1 þ mÞk1 x1 x2 :

The function F23 must be of class C3 :

13.1.2.3 The Case r11 6¼ 0; r23 6¼ 0

Let be the conditions


r22 ¼ r33 ¼ r31 ¼ r12 ¼ 0: ð13:58Þ
Putting the last two conditions for the tangential stresses (13.5100 ) and observing
that the functions ðF23  F12 Þ;2 and ðF31  F23 Þ;3 must be harmonic, it results

a0 ¼ b0 ¼ a0 ¼ 0:
13.1 Conditions for Stresses 569

Hence, one must have


1 1
ðF23  F12 Þ;2  b0 x3 ¼ 0; ðF31  F23 Þ;3  b0 x2 ¼ 0;
2 2
wherefrom
1
F23;2 ¼ b0 x3 þ e1;2 ; F12;2 ¼ e1;2
2
and
1
F31;3 ¼ e2;3 ; F23;3 ¼  b0 x2 þ e2;2 ;
2
the functions e1 ¼ e1 ðx2 Þ and e2 ¼ e2 ðx3 Þ being arbitrary.
But the last two groups of relations are not compatible in what concerns the
function F23 ; hence, the constant b0 must vanish. One obtains
F23 ¼ e1 þ e2 ; F31 ¼ e2 þ e3 ; F12 ¼ e1 þ e4 ;

where e3 ¼ e3 ðx1 Þ and e4 ¼ e4 ðx1 Þ are arbitrary functions. Observing that the above
functions are harmonic and taking into account the other relations in which they
are involved, one is lead to
1  
F23 ¼ c1 x22  x23 þ c4 x2 þ c5 x3 þ c6 ;
4
1   1
F31 ¼  c1 x21  x23 þ ðc2 þ c3 Þx1 þ c5 x3 þ c7 ;
4 2
1  2 2
 1
F12 ¼ c1 x1  x2 þ ðc2  c3 Þx1 þ c4 x2 þ c8 ;
4 2
where c1 ; c2 ; . . .; c8 are arbitrary constants.
Taking into account (13.510 ), (13.5100 ) and (13.52), (13.520 ), one obtains the
state of stress

r11 ¼ c0 x2 x3 þ c0 x2 þ d0 x3 þ d 0 ð13:59Þ
1
r23 ¼  c0 x2 þ c1 x1 þ c2 ð13:590 Þ
2ð1 þ mÞ 1
and the state of displacement

E½u1  ðx2 x03 þ x3 x02 þ u01 Þ ¼ c0 x1 x2 x3 þ c0 x1 x2 þ d0 x3 x1 þ d0 x1 ; ð13:60Þ


570 13 Particular Cases of States of Strain and Stress

1
E½u2  ðx3 x01 þ x1 x03 þ u02 Þ ¼  c0 x3 ½3x21 þ mð3x22  x23 Þ
6
1
 c0 ½x21 þ mðx22  x23 Þ  md0 x2 x3
2
1
 md 0 x2 þ ð1 þ mÞc1 x3 x1 þ ð1 þ mÞc2 x3 ;
2 ð13:600 Þ
0 0 0 1 2 2 2 0
E½u3  ðx1 x2 þ x2 x1 þ u3 Þ ¼  c0 x2 ½3x1 þ mðx2  3x3 Þ  mc x1 x2 x3
6
1
 d0 ½x21  mðx22  x23 Þ  md0 x3
2
1
þ ð1 þ mÞc1 x1 x2 þ ð1 þ mÞc2 x2 :
2

13.1.2.4 The Case of Only One Non-zero Normal Stress

In the previous case, one puts the supplementary condition


r23 ¼ 0: ð13:61Þ
It results
c0 ¼ c1 ¼ c2 ¼ 0:

One obtains the state of stress

r11 ¼ c0 x2 þ d0 x3 þ d 0 ð13:62Þ

and the state of displacement

E½u1  ðx2 x03 þ x3 x02 þ u01 Þ ¼ c0 x1 x2 þ d0 x3 x1 þ d 0 x1 ; ð13:63Þ


1
E½u2  ðx3 x01 þ x1 x03 þ u02 Þ ¼  c0 ½x21 þ mðx22  x23 Þ  md0 x2 x3  md 0 x2 ;
2 ð13:630 Þ
0 0 0 0 1 2 2 2 0
E½u3  ðx1 x2 þ x2 x1 þ u3 Þ ¼ mc x2 x3  d0 ½x1  mðx2  x3 Þ  md x3 :
2

13.1.2.5 The Case of only one Non-zero Tangential Stress

If one puts the supplementary condition


r11 ¼ 0: ð13:64Þ
in Sect. 13.2.3, then one obtains

c0 ¼ c0 ¼ d0 ¼ d0 ¼ 0:
13.1 Conditions for Stresses 571

It results the state of stress


r23 ¼ c1 x1 þ c2 ð13:65Þ
and the state of displacement

u1 ¼ x2 x03 þ x3 x02 þ u01 ; ð13:66Þ

1
u2 ¼ x3 ðc1 x1 þ 2c2 Þ  x3 x01 þ x1 x03 þ u02 ;
4l
ð13:660 Þ
1
u3 ¼ x2 ðc1 x1 þ 2c2 Þ  x1 x02 þ x2 x01 þ u03 :
4l

13.1.2.6 The Case of a State of Pure Shear

One obtains a state of pure shear if all the normal stresses vanish
r11 ¼ r22 ¼ r33 ¼ 0: ð13:67Þ
If one puts the first condition to (13.51) and one takes into account (13.510 )–
(13.520 ), one obtains the state of stress
r23 ¼ ðF12  F31 Þ;1 ; r31 ¼ ðF23  F12 Þ;2 ; r12 ¼ ðF31  F23 Þ;3 ; ð13:68Þ

and the state of displacement


1 0 0 0
u1 ¼ F 23  x2 x3 þ x3 x2 þ u1 ;
l
1 0 0 0
u2 ¼ F 31  x3 x1 þ x1 x3 þ u2 ; ð13:69Þ
l
1 0 0 0
u3 ¼ F 12  x1 x2 þ x2 x1 þ u3 :
l
The functions in two variables F23 ; F31 ; F12 as well as their harmonic conjugate
 23 ; F
functions F  31 ; F
 12 are of class C3 :

13.1.2.7 The Case of Two Non-zero Tangential Stresses

If in the preceding case one puts the supplementary condition


r12 ¼ 0; ð13:70Þ
then one obtains
F31;3 ¼ F23;3 :
572 13 Particular Cases of States of Strain and Stress

Taking into account the harmonicity of the functions F23 ; F31 ; one may write
1
F23 ¼  kðx22  x23 Þ þ d1 x3 þ d2 x2 þ d3 ;
2
1
F31 ¼  kðx21  x23 Þ þ d1 x3 þ d4 x1 þ d5 ;
2
where k; d1 ; d2 ; . . .; d5 are arbitrary constants. Observing that one may use a stress
function of the form
F12 ¼ d4 x1  d2 x2
and introducing the terms which depend on the constants d3 ; d5 in the function
F12 ; it results the state of stress
r23 ¼ F12;1 þ kx1 ; r31 ¼ F12;2  kx2 ð13:71Þ

and the state of displacement


k k
u1 ¼  x2 x3  x2 x03 þ x3 x02 þ u01 ; u2 ¼ x3 x1  x3 x01 þ x1 x03 þ u02 ; ð13:72Þ
l l
1 0 0 0
u3 ¼ F 12  x1 x2 þ x2 x1 þ u3 : ð13:720 Þ
l

13.1.2.8 The Case r33 ¼ 0; r12 ¼ 0

Let be the conditions


r33 ¼ r12 ¼ 0: ð13:73Þ
Taking into account the representation given in Sect. 13.1.1, one equates to zero
the tangential stress r12 ; the formula (13.3100 ) leads to
1
ðo11  o22 ÞX ¼  ðx1 o2 þ x2 o1 ÞH: ð13:730 Þ
2ð1 þ mÞ

Applying Laplace’s operator and observing that the function X is harmonic, one
obtains
H;12 ¼ 0;

where the relation (A.102) has been taken into account; it results

H ¼ ðf0 þ g0 Þx3 þ f 0 þ g0 ;
the functions of a single variable f0 ¼ f0 ðx1 Þ; g0 ¼ g0 ðx2 Þ; f 0 ¼ f 0 ðx1 Þ; g0 ¼ g0 ðx2 Þ
being specified by the harmonicity condition of the function H: One obtains
13.1 Conditions for Stresses 573


1 1
H¼ b1 ðx21  x22 Þ þ b2 x1 þ b3 x2 þ b4 x3 þ b01 ðx21  x22 Þ þ b02 x1 þ b03 x2 þ b04 ;
2 2

b1 ; b2 ; . . .; b04 being arbitrary constants.


The condition (13.73) becomes
1
ðo11  o22 ÞX ¼  ½ðb3 x1 þ b2 x2 Þx3 þ b03 x1 þ b02 x2 
2ð1 þ mÞ
and leads to
1
X ¼ u1 þ u2 þ 2b2 x2 x3 ðx22  x23 Þ  2b3 x3 x1 ðx21  x23 Þ
24ð1 þ mÞ

þb02 x2 ðx22  3x21 Þ  b03 x1 ðx21  3x22 Þ ; ð13:74Þ

the functions u1 ¼ u1 ðx1 þ x2 ; x3 Þ and u2 ¼ u2 ðx1  x2 ; x3 Þ must verify the


equation
Du1 þ Du2 ¼ 0: ð13:740 Þ
By means of the representation (13.31)–(13.3100 ) one obtains easily the state of
stress which corresponds to this particular case. The state of displacement may be
analogically obtained.

13.1.2.9 Plane State of Stress

Let be the case in which


r33 ¼ 0; r31 ¼ r32 ¼ 0: ð13:75Þ
Using the results obtained in Sect. 13.1.1, one equates to zero the tangential
stresses (13.3100 ); one obtains thus
(    )
m 1 2 2
X;13 ¼ x3 H;2 þ x3 Du þ 1  u ;
1þm 2 m ;1
(    ) ð13:76Þ
m 1 2 2
X;23 ¼ x3 H;1 þ x3 Du þ 1  u :
1þm 2 m ;2

Applying Laplace’s operator and taking into account (13.7), (13.70 ), (13.9),
(13.29), it results (equations of Cauchy-Riemann type)
mH0;2 þ ð1  mÞDu;1 ¼ 0; mH0;1  ð1  mÞDu;2 ¼ 0;

the Eqs. (13.800 ) allow to write


H0;1 ¼ H0;2 ¼ 0
574 13 Particular Cases of States of Strain and Stress

and
Du;1 ¼ Du;2 ¼ 0;

hence
H0 ¼ k; Du ¼ C;
where k and C are arbitrary constants.
The relations (13.76) become
1 h i
X;13 ¼  mx3 H0;2 þ ð2  mÞu;1 ;
1þm
1 h i
X;23 ¼ mx3 H0;1  ð2  mÞu;2 :
1þm
Integrating with respect to x3 ; one obtains

1 1 2 0
X;1 ¼  mx H þ ð2  mÞx3 u;1 þ p;1 ;
1 þ m 2 3 ;2
 ð13:77Þ
1 1 2 0
X;2 ¼ mx H  ð2  mÞx3 u;2 þ p;2 ;
1 þ m 2 3 ;1

where the function p ¼ pðx1 ; x2 Þ has been introduced, so that Schwartz’s theorem
be verified for mixed derivatives of second order of the function X:
One applies once more Laplace’s operator to the relations (13.77); there result
the Cauchy-Riemann equations
m m
Dp;1 ¼ H0 ; Dp;2 ¼  H0 ; ð13:78Þ
1 þ m ;2 1 þ m ;1
hence, the function p is biharmonic
DDp ¼ 0: ð13:780 Þ
Taking into account (13.78), one can integrate the relations (13.77); it results
1 2m
X ¼  x23 Dp  x3 u þ p; ð13:79Þ
2 1þm
where an arbitrary function of x3 has been neglected; indeed, the function X
appears in the expressions of the stresses only by its derivatives with respect to x1
and x2 . One may easily verify that the function (13.79) is harmonic.
Replacing in the representation (13.31), (13.3100 ), one can write
13.1 Conditions for Stresses 575

1 2m 1 1
r11 ¼  x23 Dp;12  x3 u;12 þ p;12  ðx1 o1  x2 o2 ÞH0 þ ðkx3 þ H0 Þ;
2 1þm 4ð1 þ mÞ 2
1 2 2m 1 1
r22 ¼ x3 Dp;12 þ x1 u;12  p;12 þ ðx1 o1  x2 o2 ÞH0 þ ðkx3 þ H0 Þ;
2 1þm 4ð1 þ mÞ 2
1 2 2m
r12 ¼ x3 ðo11  o22 ÞDp þ x3 ðo11  o22 Þu
4 2ð1 þ mÞ
1 1
 ðo11  o22 Þp  ðx1 o2 þ x2 o1 ÞH0 :
2 4ð1 þ mÞ

Let be F1 ¼ F1 ðx1 ; x2 Þ a harmonic function


DF1 ¼ 0;
harmonic conjugate to the function u by the Cauchy-Riemann equations
2m 2m
F1;1 ¼ u ; F1;2 ¼  u : ð13:80Þ
1 þ m ;2 1 þ m ;1
The constant C has been taken equal to zero; the influence of this constant may be
introduced in the function of x3 neglected above
Further, let be

DF0 ¼ H0 ; ð13:81Þ

hence, the function F0 ¼ F0 ðx1 ; x2 Þ is biharmonic


DDF0 ¼ 0: ð13:810 Þ
The general integral of the Eq. (13.81) may be written in the form
 0 þ F00 ;
F0 ¼ F
0 ¼ F
where F  0 ðx1 ; x2 Þ is a harmonic function

0 ¼ 0
DF

and F00 ¼ F00 ðx1 ; x2 Þ is a particular integral of this equation.


One observes also that the Eq. (13.810 ) is equivalent to the equations

DF0;11 ¼ H0;11 ; DF0;22 ¼ H0;22 ; DF0;12 ¼ H0;12 ;

neglecting a polynomial of first degree in x1 and x2 and an arbitrary constant,


which may be attached to the function DF0 .
Taking into account (13.78), (13.70 ) and (A.102), one may write
576 13 Particular Cases of States of Strain and Stress


1 1
D p;12  ðx1 o1  x2 o2 ÞH0 þ H0 ¼ H0;22 ;
4ð1 þ mÞ 2

1 1 0
D p;12  ðx1 o1  x2 o2 ÞH  H ¼  H0;11 ;
0
4ð1 þ mÞ 2

1 1
D ðo11  o22 Þp þ ðx1 o2 þ x2 o1 ÞH ¼ H0;12 :
0
2 4ð1 þ mÞ
The above results lead to
1 1
p;12  ðx1 o1  x2 o2 ÞH0 þ H0 ¼  F0;22 0
;
4ð1 þ mÞ 2
1 1
p;12  ðx1 o1  x2 o2 ÞH0  H0 ¼  F0;11 0
;
4ð1 þ mÞ 2
1 1
ðo11  o22 Þp þ ðx1 o2  x2 o1 ÞH0 ¼F0;12
0
;
2 4ð1 þ mÞ

neglecting an arbitrary harmonic function of F  0 type; such a function may be


0
attached to the function F0 ; introducing thus the function F0 :
Starting from (13.78) and (13.80), one can introduce now a biharmonic function
F ¼ Fðx1 ; x2 ; x3 Þ
DDF ¼ 0; ð13:82Þ
of the from
m 1
F¼ x2 DF0 þ x3 F1 þ F0 þ hðx21 þ x22 Þx3 ; ð13:820 Þ
2ð1 þ mÞ 3 4
the state of stress is thus expressed by (representation of Airy type)
r11 ¼ F;22 ; r22 ¼ F;11 ; ð13:83Þ

r12 ¼ F;12 : ð13:830 Þ


The state of displacement is given by
k
2l½u1  ðx3 x02 þ x2 x03 þ u01 Þ ¼ F;1 þ x1 x3 þ f1 ;
1þm
ð13:84Þ
k
2l½u2  ðx1 x03 þ x3 x01 þ u02 Þ ¼ F;2 þ x2 x3 þ f2 ;
1þm
k
2l½u3  ðx2 x01 þ x1 x02 þ u03 Þ ¼ F;3  ðx2 þ x22 þ mx23 Þ; ð13:840 Þ
2ð1 þ mÞ 1

the harmonic conjugate functions fi ¼ fi ðx1 ; x2 Þ; i ¼ 1; 2; being specified by


1
f1;1 ¼ f2;2 ¼ DF0 ; f1;2 þ f2;1 ¼ 0: ð13:85Þ
1þm
13.1 Conditions for Stresses 577

In the case of a state of incompressible deformation one has the relation (13.38),
which leads to
r11 þ r22 ¼ 0: ð13:86Þ

Taking into account the representation (13.83) and applying the operator of
Laplace to the function (13.820 ), one obtains the conditions
DF0 ¼ 0; DF1 ¼ 0; k ¼ 0: ð13:87Þ
The stress function F ¼ Fðx1 ; x2 ; x3 Þ is thus harmonic and of the form
F ¼ x 3 F1 þ F0 : ð13:88Þ

13.1.3 Case of Two Zero Tangential Stresses

Hereafter we suppose that


r23 ¼ r31 ¼ 0: ð13:89Þ

13.1.3.1 General Case

The equations of equilibrium become


r11;1 þ r12;2 ¼ 0; r12;1 þ r22;2 ¼ 0; ð13:90Þ

r33;3 ¼ 0: ð13:900 Þ
The Eq. (13.90) lead to the representation of Airy
r11 ¼ F;22 ; r22 ¼ F;11 ; r12 ¼ F;12 ð13:91Þ

where the function F ¼ Fðx1 ; x2 ; x3 Þ is arbitrary. We observe that

r11 þ r22 ¼ D0 F;

where D0 ¼ o11 þ o22 is Laplace’s operator in two variables. The Eq. (13.900 ) leads
to
r33 ¼ h; ð13:910 Þ
where h ¼ hðx1 ; x2 Þ is an arbitrary function too.
The Beltrami equations (5.420 ) become
1 1
Dr11 þ H;11 ¼ 0; Dr22 þ H;22 ¼ 0; ð13:92Þ
1þm 1þm
578 13 Particular Cases of States of Strain and Stress

1
D0 r33 þ H;33 ¼ 0; ð13:920 Þ
1þm
H;23 ¼ 0; H;31 ¼ 0; ð13:9200 Þ

1
Dr12 þ H;12 ¼ 0: ð13:92000 Þ
1þm
The Eq. (13.920 ) leads to
H;3 ¼ f;3 ;

where f ¼ f ðx3 Þ is an arbitrary function. Taking into account (13.900 ), it results

H ¼ D0 F þ h ¼ f þ g þ h; ð13:93Þ

where g ¼ gðx1 ; x2 Þ is an arbitrary function too.


Starting from the Eq. (13.920 ), one obtains (if there is no confusion, one uses the
symbol D)
1
Dh þ f;33 ¼ 0;
1þm
where
Dh ¼ C; f;33 ¼ ð1 þ mÞC;

C being an arbitrary constant; it results


1
f ¼  ð1 þ mÞCx23 þ kx3 þ C1 ; ð13:94Þ
2
where C1 ; k are arbitrary constants, and
1
h ¼ F2 þ Cðx21 þ x22 Þ;
4
F2 ¼ F2 ðx1 ; x2 Þ being a harmonic function
DF2 ¼ 0:
Taking into account (13.93), (13.94), the condition of harmonicity of the
function H leads to
Dg ¼ mC;

hence
13.1 Conditions for Stresses 579

1
g þ mCðx21 þ x22 Þ;
g¼
4
g¼
where  gðx1 ; x2 Þ is a harmonic function
g ¼ 0:
D
By means of the representation (13.90), (13.900 ) and of the formula (13.93), the
Eqs. (13.92), (13.92000 ) lead to
1 1
F;2233 þ g;22 þ ðg þ hÞ;11 ¼ 0; F;1133 þ g;11 þ ðg þ hÞ;22 ¼ 0;
1þm 1þm
ð13:95Þ
1
F;1233  g;12 þ ðg þ hÞ;12 ¼ 0: ð13:950 Þ
1þm
Integrating the Eq. (13.950 ) with respect to x1 and x2 ; one gets
1
F;33 ¼ ðg þ hÞ  g þ u þ w; ð13:96Þ
1þm
where u ¼ uðx1 ; x3 Þ; w ¼ wðx2 ; x3 Þ are arbitrary functions. One has
1
DF ¼ f þ ðg þ hÞ þ u þ w ð13:97Þ
1þm
too, by tooking into account (13.93).
Introducing (13.96) in (13.93), one may write
u;11 þ C ¼ 0; w;22 þ C ¼ 0;

wherefrom
1 1
u ¼  Cx21 þ x1 f1 þ f2 ; w ¼  Cx22 þ x2 f3 þ f4 ;
2 2
the functions fi ¼ fi ðx3 Þ; i ¼ 1; 2; 3; 4, being arbitrary. The relation (13.97)
becomes
1 1 1
DF ¼ ðg þ F2 Þ  Cðx21 þ x22 Þ  ð1 þ mÞCx23
1þm 4 2
þ x1 f1 þ x2 f3 þ f2 þ f4 þ kx3 þ C1 ; ð13:98Þ

hence
Z Z
1 1 1

F ¼ F x2 ðg þ F2 Þ  Cðx21 þ x22 Þ2  ð1 þ mÞCx43 þ x1 dx3 f1 dx3
2ð1 þ mÞ 3 64 24
Z Z Z Z
1 1
þ x2 dx3 f3 dx3 þ dx3 ðf2 þ f4 Þdx3 þ kx23 þ C1 x23 : ð13:99Þ
6 2
580 13 Particular Cases of States of Strain and Stress

 ¼ Fðx
where F  1 ; x2 ; x3 Þ is a harmonic function

 ¼ 0:
DF ð13:980 Þ
Differentiating the function F with respect to x3 ; one obtains

 ;33 þ 1 1
F;33 ¼ F g þ F2 Þ  ð1 þ mÞCx23 þ x1 f1 þ x2 f3 þ f2 þ f4 þ kx3 þ C1 :
ð
1þm 2
Subtracting the latter relation from (13.97), one obtains

  1 Cðx21 þ x22 Þ:
D0 F ¼ D0 F
4
Taking into account (13.980 ), it follows that

D0 F  ;33 ¼ g þ 1 ð1 þ mÞCðx21 þ x22 Þ  1 ð1 þ mÞCx23 þ kx3 þ C1 :


 ¼ F ð13:100Þ
4 2
Integrating with respect to x3 ; it results

 ¼  1 x23 g  1 ð1 þ mÞCðx21 þ x22 Þx23 þ 1 ð1 þ mÞCx43


F
2 8 24
1 3 1 2
 kx3  C1 x3 þ x3 v1 þ v2 ; ð13:101Þ
6 2
where v1 ¼ v1 ðx1 ; x2 Þ; k v2 ¼ v2 ðx1 ; x2 Þ are arbitrary functions.
One applies the Laplace operator D0 to the latter relation; comparing with
(13.100), one gets
1
g þ ð1 þ mÞCðx21 þ x22 Þ þ kx3 þ C1 ;
x3 Dv1 þ Dv2 ¼ 
4
hence
1
g þ ð1 þ mÞCðx21 þ x22 Þ þ C1 :
Dv1 ¼ k; Dv2 ¼  ð13:102Þ
4
It follows that
1
v1 ¼ F1 þ kðx21 þ x22 Þ;
4
where F1 ¼ F1 ðx1 ; x2 Þ is a harmonic function
DF1 ¼ 0:
Applying Laplace’s operator to the second relation (13.102), one may write
DDv2 ¼ ð1 þ mÞC;
wherefrom
13.1 Conditions for Stresses 581

1
v2 ¼ F 0 þ ð1 þ mÞCðx21 þ x22 Þ2 ;
64
with F0 ¼ F0 ðx1 ; x2 Þ biharmonic function
DDF0 ;
it also results
g ¼ DF0  C1 :
 ð13:1010 Þ
The function (13.101) becomes

 ¼  1 x23 DF0  1 ð1 þ mÞCðx21 þ x22 Þx23 þ 1 C1 x23 þ 1 ð1 þ mÞCx43  1 kx33


F
2 8 2 24 6
1 2 1
þ x3 F1 þ kðx1 þ x22 Þx3 þ F0 þ ð1 þ mÞCðx21 þ x22 Þ2 : ð13:10100 Þ
4 64
Introducing (13.10100 ) and (13.1010 ) in (13.99) and neglecting the functions in
x1 and x3 or in x2 and x3 ; linear with respect to x1 or x2 ; respectively, as well as the
functions which depend only on the variable x3 ; which lead to a vanishing state of
stress, one obtains the stress function
m 1 1
F ¼ x2 DF0 þ x2 F2 þ x3 F1 þ F0  ð1 þ mÞCðx21 þ x22 Þx23
2ð1 þ mÞ 3 2ð1 þ mÞ 3 8
1 2 1
þ kðx1 þ x22 Þx3 þ mCðx21 þ x22 Þ2 ; ð13:103Þ
4 64
where the functions F0 ; F1 ; F2 are of class C4 : The function F ¼ Fðx1 ; x2 ; x3 Þ
verifies the equation
DDF ¼ ð2 þ mÞC: ð13:1030 Þ
The state of stress is given by the representation (13.91), to which one adds
1
r33 ¼ F2 þ Cðx21 þ x22 Þ: ð13:9100 Þ
4
Starting from these results one may easily calculate the state of displacement,
similarly as in the case of a plane state of stress.

13.1.3.2 Case of a State of Incompressible Deformation

In the case of a state of incompressible deformation, the supplementary condition


H ¼ 0 is put (without to have m ¼ 1=2). Taking into account the formulation given
above, one obtains
1
D0 F þ F2 þ Cðx21 þ x22 Þ ¼ 0;
4
582 13 Particular Cases of States of Strain and Stress

which leads to
1 1
DF0 þ F2 þ ð1 þ mÞCðx21 þ x22 Þ  ð1 þ mÞCx23 þ kx3 ¼ 0:
4 2
The latter relation must be verified for any value of x3 ; one obtains
k ¼ C ¼ 0;

as well as
F2 þ DF0 ¼ 0:
The stress function (13.103) will be of the form
1
F ¼  x23 DF0 þ x3 F1 þ F0 ð13:104Þ
2
and the stress (13.9100 ) will be given by
r33 ¼ DF0 : ð13:1040 Þ
One remarks that, in this case, the stress function F is harmonic
DF ¼ 0: ð13:105Þ

13.1.4 Particular Cases

Starting from the results obtained above for the case of two zero tangential
stresses, one can put supplementary conditions to the other components of the
stress tensor, obtaining results which can be interesting for various problems.

13.1.4.1 The Plane State of Stress

Let be the supplementary condition


r33 ¼ 0: ð13:106Þ

Taking into account (13.9100 ), one obtains


1
F2 þ Cðx21 þ x22 Þ ¼ 0;
4
but the function x21 þ x22 is not harmonic, so that one must have
C ¼ 0; F2 ¼ 0

and the stress function (13.103) takes the form (13.820 ).


13.1 Conditions for Stresses 583

13.1.4.2 State of Stress Without Shear

Let be the conditions


r23 ¼ r31 ¼ r12 ¼ 0: ð13:107Þ
Equating to zero the tangential stress given by (13.91), one obtains
F;12 ¼ 0; ð13:108Þ

calculating now the mixed derivative of second order of the formula (13.103) and
observing that the expression must vanish for any x3 ; one obtains
 
1
ðF2  mDF0 Þ;12 ¼ 0; F1;12 ¼ 0; F0 þ mCx1 x2 ¼ 0: ð13:1080 Þ
8 ;12

Applying Laplace’s operator to the last relation (13.1080 ), one obtains


DF0;12 ¼ 0;

introducing now in the first relation (13.1080 ), one may write


F2;12 ¼ 0:

Now, it results
F1 ¼ u1 þ w1 : F2 ¼ u2 þ w2 :
where the functions of a single variable u1 ¼ u1 ðx1 Þ; u2 ¼ u2 ðx1 Þ; w1 ¼ w1 ðx2 Þ;
w2 ¼ w2 ðx2 Þ will be determined by the harmonicity condition of the functions F1
and F2 . It results
1
F1 ¼ Kðx21  x22 Þ þ K1 x1 þ K2 x2 þ K3 ;
2
1 2 2   
F2 ¼ Kðx 1  x2 Þ þ K1 x1 þ K2 x2 þ K3 ;
2
 . . .; K
where K; K;  3 are arbitrary constants.
The function F0 is given by
1
F0 ¼  mCðx21 þ x22 Þ þ u3 þ w3 ;
64
where u3 ¼ u3 ðx1 Þ and w3 ¼ w3 ðx2 Þ; the condition of biharmonicity of this
function leads to
   
1 1  x41  1 mCx21 x22 þ 1 1 mC  K  x42
F0 ¼ mC þ K
24 8 32 24 8
1  3  3 1  2  2  5 x1 þ K
 6 x2 þ K
7;
þ ðK1 x1 þ K2 x2 Þ þ ðK3 x1 þ K4 x2 Þ þ K
6 2
584 13 Particular Cases of States of Strain and Stress

 K
with K;  1 ; . . .; K
 7 arbitrary constants.
One obtains the stress function
1  2  x2 Þx2  1 ð1 þ mÞCðx2 þ x2 Þx2
  mKÞðx
F¼ ðK 1 2 3 1 2 3
4ð1 þ mÞ 8
   
1 1  x2 þ 1 mC  K  x2 þ 1 Kðx2  x2 Þx
þ mC þ K 1 2 1 2 3
24 2 2 2
1 1  3  3 1  2  2
þ kðx21 þ x22 Þx3 þ ðK 1 x1 þ K 2 x3 Þ þ ðK 3 x1 þ K 4 x2 Þ;
4 6 2
where a polynomial of first degree in x1 and x2 ; as well as a function which
depends only on x3 ; which lead to a zero state of stress have been neglected.
The representation (13.83) leads to
 
1  2  1 ð1 þ mÞCx2 þ 1 1 mC  K
  mKÞx  x2
r11 ¼  ðK 3 3 2
2ð1 þ mÞ 4 2 2
1  x þK  ;
 Kx3 þ kx3 þ K 2 2 4
2  
1  2  1 ð1 þ mÞCx2 þ 1 1 mC þ K
  mKÞx  x2
r22 ¼ ðK 3 3 1
2ð1 þ mÞ 4 2 2
1  x þK  ;
þ Kx3 þ kx3 þ K 1 1 3
2
and the formula (13.9100 ) allows to write
1 2 2 1 2 2   
r33 ¼ Kðx 1  x2 Þ þ Cðx1 þ x2 Þ þ K1 x1 þ K2 x2 þ K3 :
2 4
Introducing the notations
1   mKÞ  1 C; k ¼ 1  þ 1 C;
  mKÞ
k1 ¼ ðK 2 ðK
2ð1  m2 Þ 4 2ð1  m2 Þ 4
1 
  mKÞ;
k3 ¼ ðK
2ð1  m2 Þ
 ; a ¼ 1 k  K; a ¼ K
a1 ¼ K  ; b ¼ 1 k þ K; b ¼ K  ;b ¼K  ;
2 2 3 4 1 2 1 3 3
2 2
 1 ; c2 ¼ K
c1 ¼ K  2 ; c3 ¼ K
3;

one obtains the state of stress in the form

r11 ¼ ðk3  mk1 Þx22  ðk2  mk1 Þx23 þ a1 x2 þ a2 x3 þ a3 ;


r22 ¼ ðk1  mk2 Þx23  ðk3  mk2 Þx21 þ b1 x3 þ b2 x1 þ b3 ; ð13:109Þ
r33 ¼ ðk2  mk3 Þx21  ðk1  mk3 Þx22 þ c1 x1 þ c2 x2 þ c3 ;

where k1 ; k2 ; . . .; c3 are arbitrary constants.


13.1 Conditions for Stresses 585

By elementary calculations, one can represent the state of displacement in the


form
1
E½u1  ðx2 x03 þ x3 x02 þ u01 Þ ¼ ð1  m2 Þðk3 x22  k2 x23 Þx1 þ ð1 þ mÞðk2  k3 Þx31
3
1
 ½ðb2  mc1 Þx22 þ ðc1  mb2 Þx23 þ mðb2 þ c1 Þx21 
2
þ ½ða1  mc2 Þx2 þ ða2  mb1 Þx3 þ a3  mðb3 þ c3 Þx1 ;
1
E½u2  ðx3 x01 þ x1 x03 þ u02 Þ ¼ ð1  m2 Þðk1 x23  k3 x21 Þx2 þ ð1 þ mÞðk3  k1 Þx32
3
1
 ½ðc2  ma1 Þx23 þ ða1  mc2 Þx21 þ mðc2 þ a1 Þx22  ð13:110Þ
2
þ ½ðb1  ma2 Þx3 þ ðb2  mc1 Þx1 þ b3  mðc3 þ a3 Þx2 ;
1
E½u3  ðx1 x02 þ x2 x01 þ u03 Þ ¼ ð1  m2 Þðk2 x21  k1 x22 Þx3 þ ð1 þ mÞðk1  k2 Þx33
3
1
 ½ða2  mb1 Þx21 þ ðb1  ma2 Þx22 þ mða2 þ b1 Þx23 
2
þ ½ðc1  mb2 Þx1 þ ðc2  ma1 Þx2 þ c3  mða3 þ b3 Þx3 :

In the case of an incompressible state of deformation ðH ¼ 0Þ; we are led to

ð1þmÞ½ðk2  k3 Þx21 þ ðk3  k1 Þx22 þ ðk1  k2 Þx23 


þ ðb2 þ c1 Þx1 þ ðc2 þ a1 Þx2 þ ða2 þ b1 Þx3 þ a3 þ b3 þ c3 ¼ 0;

so that
1
k1 ¼ k2 ¼ k3 ¼ K;
1m
a2 ¼ b1 ; b2 ¼ c1 ; c2 ¼ a1 ; a3 ¼ b  c; b3 ¼ c  a; c3 ¼ a  b;

where K; a; b; c are new arbitrary constants.


The state of stress reads
r11 ¼ Kðx22  x23 Þ þ a1 x2  b1 x3 þ b  c;
r22 ¼ Kðx23  x21 Þ þ b1 x3  c1 x1 þ c  a; ð13:111Þ
r33 ¼ Kðx21  x22 Þ þ c1 x1  a1 x2 þ a  b

and the state of displacement is given by

2l½u1  ðx2 x03 þ x3 x02 þ u01 Þ


 
1
¼ Kx1  b2 ðx22  x23 Þ þ ða1 x2 þ a2 x3 þ b  cÞx1 ;
2
2l½u2  ðx3 x01 þ x1 x03 þ u02 Þ
 
1 ð13:112Þ
¼ Kx2  c2 ðx23  x21 Þ þ ðb1 x3 þ b2 x1 þ c  aÞx2 ;
2
2l½u3  ðx1 x02 þ x2 x01 þ u03 Þ
 
1
¼ Kx3  a2 ðx21  x22 Þ þ ðc1 x1 þ c2 x2 þ a  bÞx3 :
2
586 13 Particular Cases of States of Strain and Stress

13.1.4.3 Case of Two Non-zero Normal Stresses

Let be the conditions


r33 ¼ r23 ¼ r31 ¼ r12 ¼ 0: ð13:113Þ
If in the case in Sect. 13.4.2 one equates to zero the normal stress r33 ; then it results
m
k1 ¼ k2 ¼ mk3 ¼ c; c1 ¼ c2 ¼ c3 ¼ 0;
1m
where c is a new arbitrary constant.
The state of stress is expressed by

r11 ¼ c½ð1 þ mÞx22  mx23  þ a1 x2 þ a2 x3 þ a3 ;


ð13:114Þ
r22 ¼ c½ð1 þ mÞx21  mx23  þ b1 x3 þ b2 x1 þ b3

and the state of displacement is given by


 
1
E½u1  ðx2 x03 þ x3 x02 þ u01 Þ ¼ ð1 þ mÞc x22  mx23  x21 x1
3
1
 b2 ½x22  mðx23  x21 Þ
2
þ ½a2 x2 þ ða2  mb1 Þx3 þ a3  mb3 x1 ;
  ð13:115Þ
1
E½u2  ðx3 x01 þ x1 x03 þ u02 Þ ¼ ð1 þ mÞc x21  mx23 þ x22 x2
3
1 2 2 2
 a1 ½x1  mðx3  x1 Þ
2
þ ½b2 x1 þ ðb1  ma2 Þx3 þ b3  ma3 x2 ;

1
E½u3  ðx1 x02 þ x2 x01 þ u03 Þ ¼ mð1 þ mÞcðx21  x22 Þx3  ½ða2  mb1 Þx21
2
þ ðb1  ma2 Þx2 þ mða2 þ b1 Þx23 
2

 mðb2 x1 þ a1 x2 þ a3 þ b3 Þx3 : ð13:1150 Þ


In the case of an incompressible state of deformation, one obtains the condition

ð1 þ mÞcðx21  x22 Þ þ b2 x1 þ a1 x2 þ ða2 þ b1 Þx3 þ a3 þ b3 ¼ 0;

hence,
c ¼ 0; a1 ¼ b2 ¼ 0; a2 ¼ b1 ¼ a; a3 ¼ b3 ¼ b;

where a; b are new arbitrary constants. The state of stress becomes


r11 ¼ r22 ¼ ax3 þ b ð13:116Þ
13.1 Conditions for Stresses 587

and the state of displacement reads

2l½u1  ðx2 x03 þ x3 x02 þ u01 Þ ¼ðax3 þ bÞx1 ;


ð13:117Þ
2l½u2  ðx3 x01 þ x1 x03 þ u02 Þ ¼  ðax3 þ bÞx2 ;

1
2l½u3  ðx1 x02 þ x2 x01 þ u03 Þ ¼  aðx21  x22 Þ: ð13:1170 Þ
2

13.1.5 Problems of the Straight Cylinder. Discussion

Hereafter we make some considerations for a straight bar in form of a cylinder,


acted upon by superficial loads. A discussion concerning the results obtained
above is then made.

13.1.5.1 Problems of the Straight Cylinder

The results concerning the state of stress in case of an incompressible state of


deformation with the condition r33 ¼ 0 have been given by E. Almansi [1], using a
stress function X:
For a harmonic potential X of the form
1 0 þ 1 0
X¼ x3 ðo11  o22 ÞH x3 ðx1 o1  x2 o2 ÞH
12ð1 þ mÞ 3 4ð1 þ mÞ
 
1m 1 3
þ x3 Du  x3 u ; ð13:118Þ
1þm 6

the formulae (13.31), (13.310 ) lead to the state of stress


2m 1m
r11 ¼  x3 ðo11  o22 ÞH0 þ x3 ðx1 o1  x2 o2 ÞH0
12ð1 þ mÞ 3 4ð1 þ mÞ
1 1
 ðx1 o1  x2 o2 ÞH0 þ H;
4ð1 þ mÞ 2
ð13:119Þ
2m 1m
r22 ¼ x33 ðo11  o22 ÞH0  x3 ðx1 o1  x2 o2 ÞH0
12ð1 þ mÞ 4ð1 þ mÞ
1 1
þ ðx1 o1  x2 o2 ÞH0 þ H;
4ð1 þ mÞ 2
1  2 
r23 ¼  x3 H0 þ mx3 H0 :2 ;
2ð1 þ mÞ
 2  ð13:1190 Þ
1
r31 ¼ x H0 þ mx3 H0 :1 ;
2ð1 þ mÞ 3
588 13 Particular Cases of States of Strain and Stress

2m 3 1
r12 ¼  x3 H0;12 þ ðx1 o2 þ x2 o1 Þ½ð1  mÞx3 H0  H0 :
6ð1 þ mÞ 4ð1 þ mÞ
ð13:11900 Þ
The harmonic part of the biharmonic function u ¼ uðx1 ; x2 Þ which, as a matter
of fact, is arbitrary in the formulae (13.3100 ) will be taken equal to the harmonic
conjugate of the function a2 H0 =ð2  mÞ: Putting H0 ¼ 0 and introducing the
notation
1m  0;
w¼ ðx1 o2 þ x2 o1 ÞH
4ð1 þ mÞ
which leads to
1m
Dw ¼  H0 ;
1þm
one obtains the state of stress
2m 3 1
r11 ¼ x H0;22 þ x3 H0 þ x3 w;22 ;
6ð1 þ mÞ 3 1þm
ð13:120Þ
2m 2 1
r22 ¼ x3 H0;11 þ x3 H0 þ x3 w;11 ;
6ð1 þ mÞ 1þm
1
r23 ¼  ðx2  a2 ÞH0;2 ;
2ð1 þ mÞ 3
ð13:1200 Þ
1
r31 ¼ ðx2  a2 ÞH0;1 ;
2ð1 þ mÞ 3
2m 3
r12 ¼  x H0;12  x3 w;12 ; ð13:12000 Þ
6ð1 þ mÞ 3
given by J. H. Michell [6] and A.-E. H. Love [5] for a straight cylinder of finite
length 2a (or a thick plate of thickness 2a), the end sections x3 ¼ a of which are
free of loads (r33 ¼ r31 ¼ r32 ¼ 0 for x3 ¼ a).
One obtains thus the must important classic results. Moreover, one observes
that the formulae (13.31)–(13.3100 ) are much more simple that the previous ones
and have a larger field of applicability (as a matter of fact, it is a complete
representation, but the above representation has not this property). For example,
2ð1 þ mÞ
 F12 ðx1 ; x2 Þ; DF;12 ¼ 0;
2m
is the harmonic part of the biharmonic function u; if one puts
2
H0 ¼ H0 ¼ 0; X ¼ cx33  cx3 ðx21 þ x22 Þ;
3
13.1 Conditions for Stresses 589

then one obtains the state of stress


r11 ¼ r22 ¼ r33 ¼ r12 ¼ 0; ð13:121Þ

r23 ¼ F12;1 þ cx1 ; r31 ¼ F12;2  cx2 ; ð13:1210 Þ

which corresponds to the problem of torsion of a straight cylindrical bar of arbi-


trary cross section (particular case of an antiplane state of stress); this result cannot
be obtained by particularization from the Michell-Love formulae or from Al-
mansi’s results (because of the function u which is missing).
We considered in [19, 20] the Michell-Love problem mentioned above, with the
conditions
r31 ðx1 ; x2 ; aÞ ¼ r22 ðx1 ; x2 ; aÞ ¼ 0: ð13:122Þ
The representative (13.310 ) allows to write
     
1  1 2
1 þ X;31  ¼  aH;2  a2 Du  1  u ;
m x ¼a 2 m ;1
  3   
1  1 2
1 þ X;32  ¼  aH0;1  a2 Du  1  u :
m x3 ¼a 2 m ;2

Let be now

X ¼ Xþ þ X ; ð13:123Þ

where the functions Xþ ¼ Xþ ðx1 ; x2 ; x3 Þ; X ¼ X ðx1 ; x2 ; x3 Þ verify the conditions

Xþ ðx3 Þ ¼ Xþ ðx3 Þ; X ðx3 Þ ¼ X ðx3 Þ: ð13:1230 Þ


It results

DXþ ¼ 0; DX ¼ 0: ð13:12300 Þ


We introduce again the biharmonic function v ¼ vðx1 ; x2 Þ
DDv ¼ 0; ð13:124Þ

connected to the function H0 by the Cauchy-Riemann equations

H0;1 ¼ Dv;2 ; H0;2 ¼ Dv;1 : ð13:1240 Þ

On the frontier x3 ¼ a one obtains the conditions


  
1 þ 
1 þ X;3  ¼  aDv;
m x3 ¼a
    
1  a2 2
1 þ X  ¼  Du þ 1  u;
m ;3  x3 ¼a2 m
590 13 Particular Cases of States of Strain and Stress

which determine the harmonic functions Xþ and X :


Hence, one may write
m

;3 ¼ x3 Dv;
1þm
m 2m 1
X
;3 ¼  x2 Du  uþ ðx2  a2 ÞDu;
2ð1 þ mÞ 3 1þm 1þm 3

one may easily verify that these functions are harmonic.


Integrating, one obtains
m m
Xþ ¼ x23 Dv  v þ ,; ð13:125Þ
2ð1 þ mÞ 1þm
 
 m 3 2m 1 1 3 2
X ¼ x Du  x3 u þ x  a x3 Du; ð13:1250 Þ
6ð1 þ mÞ 3 1þm 1þm 3 3

where , ¼ ,ðx1 ; x2 Þ is a harmonic function


D, ¼ 0: ð13:12500 Þ
The conditions (13.123) are verified too.
One remarks that only the function (13.1250 ) corresponds to the Michell-Love
solution; the function (13.125) leads to new results.

13.1.5.2 Discussion

One has presented above the state of stress and the state of displacement for two
bidimensional (eventually, quasi-bidimensional) fundamental problems (the case
of a zero normal stress and the case of two zero tangential stresses), as well as for
many particular cases, by means of real stress potential functions. One remarks
thus that one can make a unitary study of a lot of problems the solutions of which
are known, as well as of new ones. The use of functions in two variables has a
great importance for the formulations of boundary value problems and each par-
ticular case may be put in connection to such a problem.
The case r33 ¼ 0 corresponds to the approximate study of a plate, the sepa-
ration planes (x3 ¼ const) of which are acted upon by tangential loads; in this case,
this stress may be neglected with respect to the other normal stresses, even for
plates of mean thickness. But G. R. Kirchhoff’s hypothesis must no more be used.
As well, at a point of the boundary one may put exact conditions on the contour of
the plate (three conditions at a point instead of two conditions as in the classic
theory of thin plates). The planes x3 ¼ const are sliding planes.
The case r31 ¼ r32 ¼ 0 corresponds to the study of superposed plates (stratified
body), the sliding being free between the faces x3 ¼ const: No other simplifying
hypothesis is introduced (linear element, zero normal stress r33 etc.). Such a case
13.1 Conditions for Stresses 591

may also be used for an initial study of the ‘‘sandwich’’ plates. The normal
directions to these plates are principal directions.
The representation in which one of the stress functions (e.g., the function H)
has a specified physical significance allows easily to obtain the results concerning
an incompressible state of deformation.
The case of a plane state of stress (Sects. 13.1.2.9 and 13.1.4.1) corresponds to a
plate free of loads on the parallel faces and acted upon on the contour by loads
parallel to the middle plane. Eventually, one must apply the principle of Saint-
Venant along the contour, obtaining a generalized plane state of stress.
The case of an antiplane state of stress (Sect. 13.22) corresponds to a cylindrical
body of finite length acted upon only on the end faces by tangential stresses and by
normal stresses which have a linear variation on these faces. This problem may be
considered as a complementary one to the preceding one. It corresponds to the
pure torsion superposed to the pure bending of the straight bars; it is the must
general case where the shear force does not pass through the centre of torsion.
The case in which two tangential stresses are non-zero (Sect. 13.1.2.7) corre-
sponds to the pure torsion of the straight bars; the shear force passes through the
centre of torsion and no bending effect appears.
The case of a state of stress without shear (Sect. 13.1.4.2) corresponds to a
triaxial state of stress; the principal directions are the same for any point of the
elastic body. In particular, if the normal stresses are equal, then one obtains a
uniform state of stress.
The case of a state of pure shear (Sect. 13.1.2.6) corresponds to the case of three
triorthogonal families of sliding planes.
The other considered particular cases have an interest for itself, either in what
concerns the possibility to specify sliding planes or principal directions for all the
body. For example, in the case in which one has r11 6¼ 0 and r23 6¼ 0 (Sect. 13.1.2.3)
the straight lines x1 ¼ const; x2 ¼ const are principal directions, while in the
planes parallel to these lines one has a sliding along the directions normal to them.
The given demonstrations allow to affirm that the representations thus obtained
are complete; it means that any state of stress which verifies the imposed condi-
tions may be expressed in this form. Moreover, one must remark that in these
representations one has used the minimal number of necessary stress functions.
In the case in which one has r33 ¼ 0; the stress vector on the planes x3 ¼ const
is tangent to these planes; it is a family of sliding planes (parallel planes). In the
case in which one has r31 ¼ r32 ¼ 0; the stress vector on the planes x3 ¼ const is
normal to these planes; it is a family of principal planes (parallel planes). The
directions normal to these planes are principal directions.
Proceeding analogically for the stresses in arbitrary curvilinear co-ordinates one
may obtain families of sliding surfaces or of principal surfaces.
592 13 Particular Cases of States of Strain and Stress

13.2 Conditions for Strains

Hereafter we shall put conditions for linear or for angular strains, obtaining general
potential functions for the states of strain and stress. The corresponding problems
are bidimensional problems, as in the preceding paragraph.

13.2.1 Case of a Zero Linear Strain

We deal, at the beginning, with the general problem of elastostatics which results
by putting the condition that one linear strain vanishes, namely
r33 ¼ 0: ð13:126Þ

13.2.1.1 General Case

Imposing the condition (13.162), Hooke’s law (4.3) leads to

r33 ¼ mðr11 þ r22 Þ;


hence

1þm
rkk ¼ ð1 þ mÞðr11 þ r22 Þ ¼ r33 ;
m
so that

1 1
e11 ¼ ðr11  m0 r22 Þ ¼ ðr11  r33 Þ;
E0 2l
ð13:127Þ
1 1
e22 ¼ ðr22  m0 r11 Þ ¼ ðr22  r33 Þ;
E0 2l
where one has introduced the generalized elastic constants
E m E E0
E0 ¼ ; m0 ¼ ;l¼ ¼ : ð13:1270 Þ
1  m2 1m 2ð1 þ mÞ 2ð1 þ m0 Þ
The equations of Cauchy (2.28) lead to
u3;3 ¼ 0;

hence

2lu3 ¼ f;12 ;
13.2 Conditions for Strains 593

where f ¼ f ðx1 ; x2 Þ is an arbitrary function of two variables of class C3 :


The Lamé equations (5.12) become
1
Du1 þ ðu1;1 þ u2;2 Þ;1 ¼ 0;
1  2m ð13:128Þ
1
Du2 þ ðu1;1 þ u2;2 Þ;2 ¼ 0;
1  2m
1
Du3 þ ðu1;1 þ u2;2 Þ;3 ¼ 0: ð13:1280 Þ
1  2m
Starting from the third of these equations, one obtains

2lðu1;1 þ u2;2 Þ;3 ¼ ð1  2mÞðx3 Df þ DgÞ;12 ;

g¼
where  gðx1 ; x2 Þ is an arbitrary function; it results that
1
2lu1;1 ¼ U;12 þ ð1  2mÞðx3 Df þ DgÞ;12 ;
2
1
2lu2;2 ¼  U;12 þ ð1  2mÞðx3 Df þ DgÞ;12 ;
2
hence
1
2lu1 ¼U;2 þ ð1  2mÞðx3 Df þ DgÞ;2 þ h1;2 ;
2 ð13:129Þ
1
2lu2 ¼  U;1 þ ð1  2mÞðx3 Df þ DgÞ;1 þ h2;1 ;
2
with U ¼ Uðx1 ; x2 ; x3 Þ; h1 ¼ h1 ðx2 ; x3 Þ; h2 ¼ h2 ðx1 ; x3 Þ:
The harmonic Eq. (5.15) leads to

2lDðu1;1 þ u2;2 Þ ¼ ð1  2mÞDDðx3f þ gÞ;12 ¼ 0;

hence

DDf;12 ¼ 0; DD
g;12 ¼ 0;

wherefrom
f ¼ f þ f1 þ f2 ; 
g ¼ g þ g1 þ g2 ; ð13:130Þ

with f1 ¼ f1 ðx1 Þ; f2 ¼ f2 ðx2 Þ; g1 ¼ g1 ðx1 Þ; g2 ¼ g2 ðx2 Þ, the functions in two


variables f ¼ f ðx1 ; x2 Þ; g ¼ gðx1 ; x2 Þ being biharmonic
DDf ¼ 0; DDg ¼ 0:
Taking into account the Eqs. (13.128), the expressions (13.129) of the
displacements, with (13.130), allow to write
594 13 Particular Cases of States of Strain and Stress

1
DU;2  ð1  2mÞðx3 f2 þ g2 Þ;22222 þ Dh1;2 þ Dðx3 f þ gÞ;112 ¼ 0;
2
1
DU;1 þ ð1  2mÞðx3 f1 þ g1 Þ;11111  Dh2;1  Dðx3 f þ gÞ;122 ¼ 0;
2
wherefrom
1
DU ¼ ð1  2mÞðx3 f2 þ g2 Þ;2222  Dh1  Dðx3 f þ gÞ;11 þ Dw2 ;
2
1
DU ¼  ð1  2mÞðx3 f1 þ g1 Þ;1111 þ Dh2 þ Dðx3 f þ gÞ;22  Dw1 ;
2
with w1 ¼ w1 ðx2 ; x3 Þ; w2 ¼ w2 ðx1 ; x3 Þ; one observes that one must have
1 h i
ð1  2mÞ ðx3 f1 þ g1 Þ;1111 þ ðx3 f2 þ g2 Þ;2222 þ Dðw1  h1 Þ þ Dðw2  h2 Þ ¼ 0
2
or
1
Dðw1  h1 Þ þ ð1  2mÞðx3 f2 þ g2 Þ;2222 ¼ w3;33 ;
2
1
Dðw2  h2 Þ þ ð1  2mÞðx3 f1 þ g1 Þ;1111 ¼  w3;33 ;
2
with w3 ¼ w3 ðx3 Þ: Hence,
1
Dw1 ¼ Dh1 þ w3;33  ð1  2mÞðx3 f2 þ g2 Þ;2222 ;
2
1
Dw2 ¼ Dh2  w3;33  ð1  2mÞðx3 f1 þ g1 Þ;1111
2
and one obtains
1 h i
DU ¼  ð1  2mÞ ðx3 f1 þ g1 Þ;1111  ðx3 f2 þ g2 Þ;2222
2
1
 Dðh1  h2 Þ  w3;33  ðo11  o22 ÞDðx3 f þ gÞ:
2
The function U will thus be of the form
1
U ¼ u  ðo11  o22 Þ½ð1  2mÞðx3 f1 þ g1 Þ  ðx3 f þ gÞ  ðh1  h2 Þ  w3 ;
2
where the function u ¼ uðx1 ; x2 ; x3 Þ must be harmonic
Du ¼ 0: ð13:131Þ
Introducing this function in (13.129) and taking into account (13.130), one
obtains
13.2 Conditions for Strains 595

2l½u1  ðx2 x03 þ x3 x02 þ u01 Þ ¼ u;2  ½o11  ð1  mÞo22 ðx3 f þ gÞ;2 ;
2l½u2  ðx3 x01 þ x1 x03 þ u02 Þ ¼  u;1 þ ½ð1  mÞo11  mo22 ðx3 f þ gÞ;1 ;
ð13:132Þ

2l½u3  ðx1 x02 þ x2 x01 þ u03 Þ ¼ f;12 : ð13:1320 Þ

Hence, one may express the state of displacement which verifies the condition
(13.131) by means of the representation (13.132), (13.1310 ), where u ¼
uðx1 ; x2 ; x3 Þ is a harmonic function of class C3 and where f ¼ f ðx1 ; x2 Þ; g ¼
gðx1 ; x2 Þ are biharmonic functions of class C5 in two variables; the corresponding
problem is thus a quasi-bidimensional problem.
By means of Cauchy relations (2.28), one obtains the state of strain

2le11 ¼u;12  ½mo11  ð1  mÞo22 ðx3 f þ gÞ;12 ;


ð13:133Þ
2le22 ¼  u;12 þ ½ð1  mÞo11  mo22 ðx3 f þ gÞ;12 ;

1 1
2le23 ¼  u;13 þ ½ð1  mÞo11  ð1 þ mÞo22  f;1 ;
2 2 ð13:1330 Þ
1 1
2le31 ¼ u;23  ½ð1 þ mÞo11  ð1  mÞo22  f;2 ;
2 2
1
2le12 ¼  ðo11  o22 Þu  ðx3 f þ gÞ;1122 ð13:13300 Þ
2
and the volume strain
2lem ¼ ð1  2mÞDðx3 f þ gÞ;12 : ð13:133000 Þ

The state of stress is given by

r11 ¼ u;12 þ ðx3 f þ gÞ;1222 ;


ð13:134Þ
r22 ¼ u;12 þ ðx3 f þ gÞ;1112 ;

r33 ¼ mDðx3 f þ gÞ;12 ; ð13:1340 Þ

1 1
r23 ¼  u;13 þ ½ð1  mÞo11  ð1 þ mÞo22  f;1 ;
2 2 ð13:13400 Þ
1 1
r31 ¼ u;23  ½ð1 þ mÞo11  ð1 þ mÞo22  f;2 ;
2 2
1
r12 ¼  ðo11  o22 Þu  ðx3 f þ gÞ;1122 : ð13:134000 Þ
2
596 13 Particular Cases of States of Strain and Stress

13.2.1.2 State of Incompressible Deformation

We will consider now that the volume strain vanishes, i.e.


1  2m
h ¼ v ¼ H ¼ 0; ð13:135Þ
E
in the case considered in the previous subsection. Equating to zero the relation
(13.13300 ) and observing that this condition is independent of x3 ; it results
Df;12 ¼ 0; Dg;12 ¼ 0;

so that
 þw þw ;
f ¼w g ¼ v þ v1 þ v2 ;
1 2

 ¼ wðx
with w1 ¼ w1 ðx1 Þ; w2 ¼ w2 ðx2 Þ; v1 ¼ v1 ðx1 Þ; v2 ¼ v2 ðx2 Þ and w  1 ; x2 Þ;
v ¼ vðx1 ; x2 Þ; the latter functions being harmonic ones
 ¼ 0;
Dw Dv ¼ 0:
But the functions f and g are biharmonic, wherefrom it results
ðo1111 þ o2222 Þðw1 þ w2 Þ ¼ 0; ðo1111 þ o2222 Þðv1 þ v2 Þ ¼ 0;

wherefrom
 1 þ d1 ;
w1;111 ¼ dx1 þ d1 ; v1;111 ¼ dx
w2;222 ¼ dx2 þ d2 ; v2;222 ¼ dx 2 þ d2 :

The formulae (13.131), (13.1310 ) allow to express the state of displacement


 þ vÞ
2l½u1  ðx2 x03 þ x3 x02 þ u01 Þ ¼ u;2 þ ðx3 w ;222
 2 þ d2 Þ;
þ ð1  mÞ½x3 ðdx2 þ d2 Þ þ ðdx
2l½u2  ðx3 x0 þ x1 x0 þ u0 Þ ¼  u þ ðx3 w þ vÞ
1 3 2 ;1 ;111
 1 þ d1 Þ;
þ ð1  mÞ½x3 ðdx1 þ d1 Þ þ ðdx
 :
2l½u3  ðx1 x0 þ x2 x0 þ u0 Þ ¼  w
2 1 3 ;12

One observes that one can add the harmonic function



1  
 ðo11  o22 Þðx3 w  þ vÞ  ð1  mÞ x3 1 d 3 x2 þ x2  2x2
1 2 3
2 6

1  2 2 2

 
þ d1 x1  d2 x2 þ d x1 þ x2  2x3 þ d1 x1  d2 x2
2
13.2 Conditions for Strains 597

to the arbitrary function u; obtaining thus a new arbitrary harmonic function; one
can thus neglect the polynomial terms in the above representation, as well as the
term in w and v in the representation of u1 and u2 ; without losing the generality.
With the notation
 ¼ w
w ;12

the function w ¼ wðx1 ; x2 Þ being harmonic too


Dw ¼ 0;
one obtains finally

2l½u1  ðx2 x03 þ x3 x02 þ u01 Þ ¼u;2 ;


ð13:136Þ
2l½u2  ðx3 x01 þ x1 x03 þ u02 Þ ¼  u;1 ;

2l½u3  ðx1 x02 þ x2 x01 þ u03 Þ ¼ w; ð13:1360 Þ

the state of displacement is thus expressed by means of two harmonic functions of


the class C3 ; one of which of three variables; the problem remains a quasi-bidi-
mensional one.
The state of stress (13.9)–(13.134000 ) becomes
r11 ¼ r22 ¼ u;12 ; r33 ¼ 0; ð13:137Þ

1 1
r23 ¼  ðu;13  w;2 Þ; r31 ¼ ðu;23  w;1 Þ; ð13:1370 Þ
2 2
1
r12 ¼  ðo11  o22 Þu: ð13:13700 Þ
2

13.2.1.3 Plane State of Strain

In particular, we will consider a plane state of strain, putting the supplementary


conditions
e22 ¼ e31 ¼ 0 ð13:138Þ

in the results in Sect. 13.2.1.1. It results


u;13 ¼ ½ð1  mÞo11  ð1 þ mÞo22  f;1 ; u;23 ¼ ½ð1 þ mÞo11  ð1  mÞo22  f;2 ;

wherefrom one obtains


u;3 ¼ ½ð1  mÞo11  ð1 þ mÞo22  f þ v1 ; u;3 ¼ ½ð1 þ mÞo11  ð1  mÞo22  f þ v2 ;
598 13 Particular Cases of States of Strain and Stress

with v1 ¼ v1 ðx2 ; x3 Þ; v2 ¼ v2 ðx1 ; x3 Þ:


Obviously, one must have
2mDf þ v2  v1 ¼ 0
or, differentiating with respect to x3 ;
v2;3 ¼ v1;3 ¼ ,3;33 ;

with ,3 ¼ ,3 ðx3 Þ; finally, it results


v1 ¼ ,3;3 þ ,2 ; ,2 ¼ ,3;3 þ ,1 ;

with ,1 ¼ ,1 ðx1 Þ; ,2 ¼ ,2 ðx2 Þ: One is led thus to


1
Df ¼ ð,2  ,1 Þ;
2m
then
1
u;3 ¼ ðo11  o22 Þf þ ð,1 þ ,2 Þþ,3;3 :
2
Observing that the function f is biharmonic, one can write
D,1 ¼ ,1;11  D,2 ¼ ,2;22 ¼ c;

hence
1 1
,1 ¼ cx21 þ a1 x1 þ b1 ; ,2 ¼ cx22 þ a2 x2 þ b2 :
2 2
One obtains thus
1 1
Df ¼  cðx21  x22 Þ  ða1 x1  a2 x2 þ b1  b2 Þ;
4m 2m
where
1 1 1
f ¼ f0  cðx4  x42 Þ  ða1 x31  a2 x32 Þ  ðb1 x21  b2 x22 Þ;
48m 1 12m 4m
the function f 0 ¼ f 0 ðx1 ; x2 Þ being harmonic

Df 0 ¼ 0: ð13:1310 Þ
Returning to the function u; it results
 
0 1 1 1 2 2
u;3 ¼ ðo11  o22 Þf þ v3;3 þ 1 cðx þ x2 Þ þ a1 x1 þ a2 x2 þ b1 þ b2 ;
2 m 2 1
13.2 Conditions for Strains 599

hence
 
0 1 1 1 2
u ¼ x3 ðo11  o22 Þ f þ v3 þ 1 cðx þ x22 Þ
2 m 2 1

þa1 x1 þ a2 x2 þ b1 þ b2 x3 þ v0 ;

with v0 ¼ v0 ðx1 ; x2 Þ: Taking into account the conditions (13.131), (13.1310 ), one
may write
 
1
Dv0 þ D,3 þ 1  cx3 ¼ 0;
m

differentiating with respect to x3 ; one obtains


 
1
,3;333 þ 1  c ¼ 0;
m

wherefrom
 
1 1
,3 ¼  1  cx33 þ c2 x23 þ c2 x3 þ c3 :
6 m

Finally,

Dv0 þ 2c1 ¼ 0;

so that
1
v0 ¼ 
v  c1 ðx21 þ x22 Þ;
2
v¼
the function  vðx1 ; x2 Þ satisfying the harmonic equation
v ¼ 0:
D
It results
 

0 1 1 1 2
u ¼ x3 ðo11  o22 Þf  
vþ 1 c 3ðx1 þ x22 Þ  2x23
2 m 6

1
þa1 x1 þ a2 x2 þ b1 þ b2 x3  c1 ðx21 þ x22  2x23 Þ þ c2 x3 þ c3 :
2
We introduce the notations
0
f;12 ¼ F0 ; 
g;12 ¼ F;

the new functions F 0 ¼ F 0 ðx1 ; x2 Þ; k  ¼ Fðx


F  1 ; x2 Þ verifying the equations
600 13 Particular Cases of States of Strain and Stress

DF 0 ¼ 0;  ¼ 0:
DDF ð13:1380 Þ

Similarly, let us introduce the harmonic function v ¼ vðx1 ; x2 Þ


Dv ¼ 0; ð13:13800 Þ

v by means of the Cauchy-Riemann equations


conjugate to the function 
v;1 ¼ 
v;2 ; v;2 ¼ v;1 :

Taking into account the formulae (13.134)–(13.134000 ) and the conditions


(13.13), (13.13800 ), one obtains the state of stress in the form
0
r11 ¼ x3 F;22  ;22 þv;22 ; r22 ¼ x3 F;11
þF 0  ;11 þv;11 ; r33 ¼ mDF;
þF 
0
r23 ¼ r31 ¼ 0; r12 ¼ x3 F;12  F;12  v;12 :

But, using Almansi’s formula (A.100), the arbitrary harmonic function v may be
 without any
included in the harmonic part of an arbitrary biharmonic function F;
lose of generality of the above representation (one may take v ¼ 0). With the
function

F ¼ x3 F 0 þ F; ð13:139Þ

which is biharmonic
DDF ¼ 0; ð13:1390 Þ
one obtains the formulae of Airy type
r11 ¼ F;22 ; r22 ¼ F;11 ; r33 ¼ mDF; ð13:140Þ

r23 ¼ r31 ¼ 0; r12 ¼ F;12 ; ð13:1400 Þ

corresponding to a plane state of strain. The function F 0 ¼ F 0 ðx1 ; x2 Þ is a har-


monic function of class C3 ; while the function F  ¼ Fðx
 1 ; x2 Þ is a biharmonic
4
function of class C ; hence, the corresponding problem is a bidimensional
problem.
Assuming that the state of stress does not depend on the variable x3 ; one has
F 0 ¼ 0; so that one gets the classic form of these formulae.
The formulae (13.132), (13.1320 ) allow to write the state of displacement in the form
 ;1  v;1 þ ð1  mÞDg;2  c1 x2 ;
2l½u1  ðx2 x03 þ x3 x02 þ u01 Þ ¼  x3 F;10  F
 ;2  v;2 þ ð1  mÞDg;1  c1 x1 ;
2l½u2  ðx3 x01 þ x1 x03 þ u02 Þ ¼  x3 F;20  F
2l½u3  ðx1 x02 þ x2 x01 þ u03 Þ ¼ F 0 :
13.2 Conditions for Strains 601

The displacements u1 and u2 which correspond to the constant c1 may be included


in the rigid body displacement. Taking into account the preceding observation
concerning the function v and using the same notation (13.139), one gets
Z
0 0 0
2l½u1  ðx2 x3 þ x3 x2 þ u1 Þ ¼ F;1 þ ð1  mÞ DFdx1 þ f2 ;
Z ð13:141Þ
2l½u2  ðx3 x01 þ x1 x03 þ u02 Þ ¼ F;2 þ ð1  mÞ DFdx2 þ f1 ;

2l½u3  ðx1 x02 þ x2 x01 þ u03 Þ ¼ F;3 ; ð13:1410 Þ

where f1 ¼ f1 ðx1 Þ; f2 ¼ f2 ðx2 Þ are functions of a single variable. Hooke’s law


2lðu1;2 þ u2;1 Þ ¼ 2r12

allows to specify the functions f1 and f2 by means of the equation


Z Z 
f1;1 þ f2;2 þ ð1  mÞ DF;2 dx1 þ DF;1 dx2 ¼ 0; ð13:142Þ

where DF ¼ DF  depends only on the variables x1 and x2 : Observing that the zero
sum (because of the biharmonicity of the function F and of the theorem of Fubini)
Z Z Z Z
dx2 DF;22 dx1 þ dx1 DF;11 dx2

differs of the sum between parentheses in the Eq. (13.142) only by a sum of two
functions of a single variable, it results that the latter equation is of the form
f1;1 þ f2;2 þ F1 þ F2 ¼ 0; ð13:1420 Þ

where F1 ¼ F1 ðx1 Þ; F2 ¼ F2 ðx2 Þ are known functions. The Eq. (13.142) is thus
with separate variables and one may take
Z Z
f1 ¼  F1 dx1 ; f2 ¼ F2 dx2 : ð13:14200 Þ

One supposed that f1 and f2 do not depend on the variable x3 ; indeed, this
hypothesis is justified by the form of the representations (13.16), (13.1410 ). Indeed,
assuming such a dependence and taking into account the relations
u1;2 þ u2;1 ¼ 0; u1;3 þ u3;1 ¼ 0;

one is led to the same conclusion.


If the displacements ui ; i ¼ 1; 2; 3; do not depend on the variable x3 ; one
obtains the classic results, with F 0 ¼ 0; one has thus u3 ¼ 0; neglecting the motion
of rigid body.
In the case of an incompressible plane state of strain, one may use the formulae
(13.140), (13.150 ) of Airy’s type. The condition (13.138) leads to
602 13 Particular Cases of States of Strain and Stress

DF ¼ 0; ð13:143Þ

the state of stress will be thus given by


r11 ¼ F;22 ; r22 ¼ F;11 ; r12 ¼ F;12 ð13:144Þ

r33 ¼ r23 ¼ r31 ¼ 0; ð13:1440 Þ


the function F being of the form (13.139), with

DF 0 ¼ 0;  ¼ 0:
DF ð13:145Þ
0
What concerns the state of displacement (13.141), (13.141 ), taking into
account (13.143), the functions f1 and f2 correspond to a motion of rigid body and
one obtains.

2lu1 ¼ F;1 ; 2lu2 ¼ F;2 ; 2lu3 ¼ F 0 ; ð13:146Þ

neglecting a motion of rigid body. The two harmonic functions of class C3 which
appear in this representation are both in two variables, the problem being a
bidimensional one.

13.2.2 Case of Two Zero Angular Strains

Hereafter, we deal with the general problem of elastostatics which results by


putting the condition that two angular strains vanish, namely
e23 ¼ e13 ¼ 0: ð13:147Þ

13.2.2.1 General Case

Taking into account Hooke’s law (4.30 ), it results


r23 ¼ r13 ¼ 0: ð13:148Þ

too. But it is just the case considered in Sect. 13.3. The state of stress is given by
Airy’s type formulae of the form (13.91), (13.9100 ), where the stress is given by the
formula (13.103).
It is obvious that particular cases may easily obtained starting from these
results.
The same considerations can be made for an incompressible state of strain.
13.2 Conditions for Strains 603

13.2.2.2 Discussion

All the representations given above, both for bidimensional and quasi-bidimen-
sional problems, are obtained by a demonstration step by step, with any lose of
generality. Hence, these representations are complete (any state of stress, any state
of strain and any state of displacement which verify the imposed conditions may
be represented in the given form) for a simply connected domain; in the case of a
multiply connected domain, one introduced certain supplementary terms, corre-
sponding to singularities.
What concerns an antiplane state of strain, excepting the normal stresses
r11 ; r22 which are non-zero, the results which one obtains are similar to those
obtained in the case of an antiplane state of stress (r11 ¼ r22 ¼ r12 ¼ 0).
These representations are useful for the study of a bar in the form of a straight
cylinder.

13.3 Plane and Antiplane Problems

Among the bidimensional and quasi-bidimensional problems, an important rôle is


played by the plane and antiplane ones. These problems have been considered
above in the static case; we will put hereafter the accent on the dynamic problems.

13.3.1 Plane and Antiplane States of Stress

First of all, we give some general results concerning the plane and antiplane states
of stress, problems which complete one another.

13.3.1.1 General Considerations

We remark that the stress tensor can be decomposed in the sum of two tensors, i.e.
Tr ¼ Tpr þ Tar ð13:149Þ

where the stress tensor is


2 3
r11 r12 r13
Tr  4 r21 r22 r23 5; ð13:1490 Þ
r31 r32 r33

the plane stress tensor reads


604 13 Particular Cases of States of Strain and Stress

2 3
r11 r12 0
Tpr  4 r21 r22 05 ð13:14900 Þ
0 0 0

and the antiplane stress tensor is of the form


2 3
0 0 r13
Tar  4 0 0 r23 5: ð13:149000 Þ
r31 r32 r33

13.3.1.2 Static and Dynamic Case

In the static case, the stresses depend on the three variables x1 ; x2 ; x3 ; while in the
dynamic case, one takes into account the time variable t too.
The plane state of stress corresponds to a plate acted upon by loads parallel to
the middle plane. If the loads are uniformly distributed on the thickness at the
frontier of the plate, than the problem is a bidimensional problem; otherwise, the
problem is a quasi-bidimensional one.
In the static case, one uses the equations of equilibrium
r11;1 þ r12;2 ¼ 0; r21;1 þ r22;2 ¼ 0; ð13:150Þ

which lead to a representation of the Airy type for the stresses, in the absence of
volume forces.
In the dynamic case, inertia forces are added, and the equations of motion read
r11;1 þ r12;2 ¼ q€
u1 ; r21;1 þ r22;2 ¼ q€u2 ; ð13:1500 Þ

the problem leads to partial differential equations of hyperbolic type.


This problem is called also the problem of the short cylinder.
The antiplane state of stress corresponds to a straight cylinder of finite length
(theoretically of infinite length), acted upon by loads normal to its axis.
In the static case, one uses the equations of equilibrium
r13;3 ¼ 0; r23;3 ¼ 0; r31;1 þ r32;2 þ r33;3 ¼ 0; ð13:151Þ

in the absence of volume forces.


In the dynamic case, adding inertia forces one obtains
r13;3 ¼ q€
u1 ; r23;3 ¼ q€
u2 ; r31;1 þ r32;2 þ r33;3 ¼ q€u3 : ð13:1510 Þ
The problem is called the problem of the long cylinder too.
13.3 Plane and Antiplane Problems 605

13.3.2 Plane and Antiplane States of Strain

Hereafter we give some general results concerning plane and antiplane states of
strain; then we deal with representation of such states of strain in the dynamic case.

13.3.2.1 General Considerations

As the stress tensor, the strain tensor may also be decomposed in the sum of two
tensors, i.e.

Te ¼ Tpe þ Tae ; ð13:152Þ

where the strain tensor is


2 3
e11 e12 e13
Te  4 e21 e22 e23 5; ð13:1520 Þ
e31 e32 e33

the plane strain tensor is of the form


2 3
e11 e12 0
Tpe  4 e21 e22 05 ð13:15200 Þ
0 0 0

and the antiplane strain tensor reads


2 3
0 0 e13
Tae  4 0 0 e23 5: ð13:152000 Þ
e31 e32 e33

13.3.2.2 Dynamic Plane State of Strain

In the following we assume that

e33 ¼ 0; e31 ¼ e32 ¼ 0; ð13:153Þ

which, taking into account Cauchy’s relations, lead to


u3;3 ¼ 0; u3;1 þ u1;3 ¼ 0; u2;3 þ u3;2 ¼ 0;

hence, one obtains the state of displacement


u1 ¼ x3 u;1 þ w1 ; u2 ¼ x3 u;2 þ w2 ; u3 ¼ u; ð13:154Þ
606 13 Particular Cases of States of Strain and Stress

where the potential functions u ¼ uðx1 ; x2 ; x3 ; tÞ; wi ¼ wi ðx1 ; x2 ; x3 ; tÞ; i ¼ 1; 2;


are introduced.
The non-vanishing strains are given by
e11 ¼ x3 u;11 þ w1;1 ; e22 ¼ x3 u;22 þ w2;2 ; ð13:155Þ

1
e12 ¼ x3 u;12 þ ðw1;2 þ w2;1 Þ: ð13:1550 Þ
2
Saint-Venant’s equations of continuity are satisfied by means of Lamé’s equations
1
h2 ui þ h;i ¼ 0; i ¼ 1; 2; 3; h ¼ ei;i ¼ x3 Du þ w1;1 þ w2;2 ; ð13:156Þ
1  2m
where the hyperbolic operator of d’Alembert has been introduced (see Sect. 5.3.1),
One obtains the equations
1
x3 h2 u;1 þ h2 w1 þ ðx3 Du;1 þ w1;11 þ w2;12 Þ ¼ 0;
1  2m
1
x3 h2 u;2 þ h2 w2 þ ðx3 Du;2 þ w1;12 þ w2;22 Þ ¼ 0;
1  2m
1
h2 u þ ðDuÞ ¼ 0;
1  2m
so that
ð1  2mÞh2 u;1 þ Du;1 ¼ 0; ð1  2mÞh2 u;2 þ Du;2 ¼ 0; ð13:1560 Þ

ð1  2mÞh2 u  Du ¼ 0 ð13:15600 Þ

and
ð1  2mÞh2 w1 þ ðw1;1 þ w2;2 Þ;1 ¼ 0; ð1  2mÞh2 w2 þ ðw1;1 þ w2;2 Þ;2 ¼ 0:

From the Eqs. (13.156) it results

ð1  2mÞh2 u þ Du ¼ 2€f ;
where f ¼ f ðtÞ; using also the Eq. (13.15600 ), it follows that

ð1  2mÞh2 u ¼ Du ¼ €f :
One obtains thus
2m 2 € m 2€
€¼
u c2 f ¼ c f;
1  2m 1m 1
so that
m 2
u¼ c f þu
 1t þ u
 2;
1m 1
13.3 Plane and Antiplane Problems 607

i ¼ u
where u  i ðx1 ; x2 Þ; i ¼ 1; 2; by applying the operator D; we have

u2 ¼ €f ;
u1 þ D
tD
so that
ui ¼ ai ;
D i ¼ 1; 2;
and
1 1
f ¼ a1 t 3 þ a2 t 2 þ a3 t þ a4 ; ð13:157Þ
6 2
where ak ; k ¼ 1; 2; 3; 4; are constants. One observes that
1
 i ¼ ui þ ai ðx21 þ x22 Þ;
u i ¼ 1; 2: ð13:158Þ
4
The above formulae lead to the representation of the state of displacement in
the form
 
1 1
u1 ¼  u1;1 þ a1 x1 t þ u2;1 þ a2 x1 x3 þ w1 ;
2 2
  ð13:159Þ
1 1
u2 ¼  u1;2 þ a1 x2 t þ u2;2 þ a2 x2 x3 þ w2 ;
2 2
 
m 2 1 3 1 2
u3 ¼  c 1 a1 t þ a2 t þ a3 t þ a4
1m 6 2

1 1
þ u1 þ a1 ðx21 þ x22 Þ t þ u2 þ a2 ðx21 þ x22 Þ; ð13:1590 Þ
4 4

where ui ¼ ui ðx1 ; x2 Þ; i ¼ 1; 2; are harmonic functions of class C3


Dui ¼ 0; i ¼ 1; 2; ð13:160Þ
so that the Lamé equations be satisfied.
Concerning the functions w1 and w2 ; they verify Lamé’s equations too, so that
1 1
w1 ¼  w ; w2 ¼ h2 w þ w ð13:161Þ
1  2m ;12 1  2m ;11
or in the form
1 1
w 1 ¼ h2 w þ w ; w2 ¼  w ; ð13:1610 Þ
1  2m ;22 1  2m ;12
where the function w ¼ wðx1 ; x2 ; tÞ of class C4 verifies the double waves equation
h1 h2 w ¼ 0: ð13:1600 Þ
608 13 Particular Cases of States of Strain and Stress

By the aid the formulae (13.155), (13.1550 ), one obtains the state of strain and
then the state of stress. It results
r31 ¼ 0; r32 ¼ 0 ð13:162Þ

and
2lm
r33 ¼ h2 w;2  x3 ða1 t þ a2 Þ ; ð13:163Þ
1  2m
in case of the representation (13.161), or
2lm
r33 ¼ h2 w;1  x3 ða1 t þ a2 Þ ; ð13:1630 Þ
1  2m
in case of the representation (13.1610 ).
Starting from the representation (13.155), (13.1550 ), which is polynomial in
t (excepting the functions w1 and w2 ), we obtain the static case if all the coeffi-
cients of t vanish; hence, we must have u1;1 ¼ u1;2 ¼ u1 ¼ 0; a1 ¼ a2 ¼ a3 ¼ 0:
At the same time, the velocity c1 vanishes. In this case, we use the formulae
(13.164), where the function u ¼ uðx1 ; x2 Þ of class C3 is a harmonic function
Du ¼ 0: ð13:164Þ
The functions w1 ; w2 are given by
1 1
w1 ¼  w ; w2 ¼ Dw þ w ð13:165Þ
1  2m ;12 1  2m ;11
or by
1 1
w1 ¼ Dw þ w ; w2 ¼  w ; ð13:1650 Þ
1  2m ;22 1  2m ;12
where the function w ¼ wðx1 ; x2 Þ of class C4 is biharmonic
DDw ¼ 0: ð13:1640 Þ
We obtain the state of stress in an analogous manner. We have
r31 ¼ r32 ¼ 0 ð13:166Þ

and
2Gm
r33 ¼ Dw;2 ; ð13:167Þ
1  2m
in case of the representation (13.156), or
2Gm
r33 ¼ Dw;1 ; ð13:1670 Þ
1  2m
in case of the representation (13.1560 ).
13.3 Plane and Antiplane Problems 609

Observing that the above representations (in the dynamic case, as well as in the
static one) have been obtained by a demonstration step by step, the respective
representations being thus complete.
The study puts into evidence the approximation made assuming that in the
plane state of strain the displacements depend only on the two variables x1 and x2 :

13.3.2.3 Dynamic Antiplane State of Strain

Hereafter, we assume that


e33 ¼ e22 ¼ 0; e12 ¼ 0; ð13:168Þ
which, taking into account Cauchy’s relations, lead to
u1;1 ¼ 0; u2;2 ¼ 0; u1;2 þ u2;1 ¼ 0;

hence,
u1 ¼ x 2 u þ u
 1;3 ; u2 ¼ x1 u þ u
 2;3 ; ð13:169Þ

where u ¼ uðx3 ; tÞ and ui ¼ u


 i ðx3 ; tÞ; i ¼ 1; 2, are potential functions.
To determine the potential functions, we use the general theory of Lamé’s
equations (see Sect. 5.3.1) and obtain

u3;31 ¼  ð1  2mÞðx2 h2 u þ h2 u
 1;3 Þ;
u3;32 ¼  ð1  2mÞðx1 h2 u þ h2 u  2;3 Þ;
 
1  2m 1
u3;33 ¼  u3;11 þ u3;22  2 €u3 :
2ð1  mÞ c3

u;3 ¼ u)
Integrating with respect to x3 ; it results (

u3;1 ¼  ð1  2mÞh2 ðx2 u


 þu 1 Þ þ w1 ;
ð13:170Þ
u3;2 ¼  ð1  2mÞh2 ðx1 u þu  2 Þ þ w2 ;

with wi ¼ wi ðx1 ; x2 ; tÞ; i ¼ 1; 2; we may then write


1  2m   1  2m 1
u3;33 ¼  w;11 þ w;22 þ €u3 :
2ð1  mÞ 2ð1  mÞ c22

Finally, we get
1  2m
h 1 u3 ¼ w þ w2;2 ; ð13:171Þ
2ð1  mÞ 1;1

where, in the operator h1 appears only the variable x3 :


610 13 Particular Cases of States of Strain and Stress

From (13.147), it results


u3;12 ¼ ð1  2mÞh2 u
 þ w1;2 ; u3;21 ¼ ð1  2mÞh2 u
 þ w2;1 ;

so that
w1;2  w2;1 ¼ 2ð1  2mÞh2 u
 ¼ 2ð1  2mÞh;

with h ¼ hðtÞ: One may write


 þ ð1  2mÞx2 hðtÞ;
w1 ¼ w   ð1  2mÞx1 hðtÞ;
w2 ¼ w
1 2

 ¼w
with the condition w  ; hence, w  ¼W  ;1 ; w2 ¼ W
 ;2 with W
 ¼ Wðx
 1 ; x2 ; tÞ:
1;2 2;1 1
Introducing in (13.170) and taking into account
 ¼ 0;
h2 u ð13:172Þ

we get
 ;1  ð1  2mÞh2 u
u3;1 ¼ W  ;2  ð1  2mÞh2 u
 1 ; u3;2 ¼ W  2;

so that
  ð1  2mÞðx1 h2 u
u3 ¼ W  1 þ x2 h2 u 
 2 þ UÞ; ð13:1690 Þ
 ¼ Uðx
with U  3 ; tÞ:
The Eq. (13.171) leads to
  ð1  2mÞðx1 h1 h2 u
h1 u3 ¼ h1 W  1 þ x2 h 1 h 2 u 
 2 þ h1 UÞ
1  2m   Þ ¼  1  2m DW: 
¼ ðw1;1 þ w2;2
2ð1  mÞ 2ð1  mÞ
Differentiating with respect to x3 ; one obtains

 1;3 þ x2 h1 h2 u
x1 h1 h2 u  ;3 ¼ 0;
 2;3 þ h1 U

this polynomial in x1 and x2 must be identical zero, so that

 1;3 ¼ 0; h1 h2 u
h1 h2 u  ;3 ¼ 0;
 2;3 ¼ 0; h1 U

wherefrom

 1 ¼ h1 ; h 1 h 2 u
h1 h2 u  ¼ H;
 2 ¼ h2 ; h 1 U ð13:173Þ

with hi ¼ hi ðtÞ; i ¼ 1; 2; H ¼ HðtÞ. Hence, it results


 ¼ 2ð1  2mÞðx1 h1 þ x2 h2 þ HÞ:
h2 W ð13:1730 Þ
13.3 Plane and Antiplane Problems 611

Starting from (13.173), one may write


Z Z Z Z
2 2
u1 ¼ u1 þ c1 c2 dt dt dt h1 dt;

Z Z Z Z ð13:174Þ
 2 ¼ u2 þ c21 c22 dt dt dt h2 dt;
u
Z Z
 ¼ U  c2
U dt h1 dt;
1

where the functions ui ¼ ui ðx3 ; tÞ; i ¼ 1; 2; U ¼ Uðx3 ; tÞ verify the equations


h1 h2 u1 ¼ 0; h1 h2 u2 ¼ 0; ð13:175Þ

h1 U ¼ 0; ð13:1750 Þ

as well, the Eq. (13.1730 ) leads to


 Z Z Z Z Z Z 
 ¼ W  2ð1  mÞc2 x1 dt h1 dt þ x2 dt h2 dt þ dt Hdt ;
W 2

ð13:1740 Þ
where the function W ¼ Wðx1 ; x2 ; tÞ verifies the equation
h2 W ¼ 0: ð13:176Þ
Differentiating the Eq. (14.135) with respect to x3 ; it results
h2 u ¼ 0: ð13:1760 Þ
Introducing the functions thus obtained in the relations (13.169) and (13.1690 )
and taking into account (13.174) and (13.1740 ), one may express the state of
displacement in the form
u1 ¼ x2 u þ u1;3 ; u2 ¼ x1 u þ u2;3 ; ð13:177Þ

u3 ¼ W  ð1  2mÞðx1 h2 u1 þ x2 h2 u2 þ UÞ; ð13:1770 Þ


where the potential functions U ¼ Uðx1 ; x2 ; tÞ and W ¼ Wðx1 ; x2 ; tÞ; u ¼ uðx3 ; tÞ
are of class C2 and verify the simple longitudinal wave equation (13.1750 ) and the
simple transverse wave equations (13.176), (13.1760 ), respectively, while the
potential functions ui ¼ ui ðx3 ; tÞ; i ¼ 1; 2; are of class C4 and verify the double
wave equations (13.175).
The state of stress is obtained in the form
m
r11 ¼ r22 ¼ r33 ¼ 2mlðx1 h2 u1;3 þ x2 h2 u2;3 þ U;3 Þ; ð13:178Þ
1m
612 13 Particular Cases of States of Strain and Stress

   
2m 2m
r31 ¼ l W;1 þ x2 u;3  2 u
€ ; r32 ¼ l W;2  x1 u;3  2 u
€ ; ð13:1780 Þ
c2 1 c2 2

obviously, r12 ¼ 0:
In the corresponding static case, if we start from the representation (13.1770 )
and equate to zero the dependence of the potential functions on the time variable t,
then we take the functions h1 ; h2 and H as constants. The Eq. (14.3.25) show that
the functions ui ¼ u
 i ðx3 Þ; i ¼ 1; 2; are polynomials of fourth degree, the func-
 
tion / ¼ /ðx3 Þ being a polynomial of second degree; as well, the particular
integral of (13.1730 ) will be a polynomial of third degree, while the function
W ¼ Wðx1 ; x2 Þ must be harmonic
DW ¼ 0: ð13:179Þ
We obtain thus the state of displacement

E½u1  ðx2 x03 þ x3 x02 þ u01 Þ ¼ ½ða1 x3 þ 3b1 Þx3 þ ax2 x3 ;
ð13:180Þ
E½u2  ðx3 x01 þ x1 x03 þ u02 Þ ¼ ½ða2 x3 þ 3b2 Þx3  ax1 x3 ;

E½u3  ðx1 x02 þ x2 x01 þ u03 Þ ¼ W þ ð1  mÞ½2ða1 x31 þ a2 x32 Þ  3bðx21 þ x22 Þ
 3ð1  2mÞ½ða1 x3 þ 2b1 Þx1
þ ða2 x3 þ 2b2 Þx2  ðbx3 þ cÞx3 ; ð13:1800 Þ

where ai ; bi ; i ¼ 1; 2; a; b; c are constants and where we have put in evidence the


rigid body motion.
The above representations are complete for a simply connected domain.
As well, the study puts in evidence the approximation made assuming that in an
antiplane state of deformation the displacements depend only on two space vari-
ables (x1 and x2 ).

References

A. Books

1. Clebsch, A.: Théorie de l’élasticité des corps solides. Paris (1883)


2. Love, A.E.H.: A Treatrise on the Mathematical Theory of Elasticity, 4th edn. Cambridge
University Press, Cambridge (1934)
3. Teodorescu, P.P., Ille, V.: Teoria elasticitătßii ßsi introducere în mecanica solidelor deformabile
(Theory of Elasticity and Introduction to Mechanics of Deformable Solids), II. Dacia, Cluj-
Napoca (1979)
References 613

B. Papers

4. Almansi, E.: Sulle deformationi delle piastre elastiche. Rend R. Accad. Lincei, Cl. Sci. fis.
mat. e nat. ser. 6. 17, 12 (1933)
5. Davidescu-Moisil, A.: Asupra unui caz de echilibru al corpurilor elastice izotrope (On a Case
of Equilibrium of Isotropic Elastic Bodies). Stud. cerc. mat. 15, 225 (1964)
6. Michell, J.H.: On the direct determination of stress in an elastic solid with application to the
theory of plates. Proc. Lond. Math. Soc. 31, 100 (1900)
7. Moisil, Gr.C.: Integrale ale ecuatßiilor echilibrului elastic. I. Integrale caracterizate prin
conditßii geometrice privind deplasările (Integrals of the equations of elastic equilibrium. I.
Integrals characterized by conditions concerning displacements). Bul. ßst. Acad. ser. mat. fiz.
chim. 2, 283 (1950)
8. Supino, G.: I sistemi elastici in due dimensioni e le loro relazioni con la deformazione
spaziale. Rend. R. Accad. Lincei Cl. Sci. fis. mat. e nat. ser. 6. 1, 116 (1925)
9. Supino, G.: Sul problema di Clebsch. Rend. R. Accad. Lincei Cl. Sci. fis. mat. e nat. ser. 6.
15, 366 (1932)
10. Supino, G.: Sopra la deformazione delle lastre. Rend. R. Accad. Lincei Cl. Sci. fis. mat. e nat.
ser. 6. 15, 448 (1932)
11. Supino, G.: Il problema elastico piano e la sua interpretazione nello spazio. Rend. R. Accad.
Lincei Cl. Sci. fis. mat. e nat. ser. 6. 22, 522, 581 (1935)
12. Supino, G.: Sopra la teoria delle lastre elastiche. Ann. mat. pura ed appl. ser. IV. 24, 39
(1945)
13. Supino, G.: Calcolo approxximato delle piatre elastiche. Atti V. Congresso Unione Mat.
Italiana, Pavia-Torino, 154 (1955)
14. Teodorescu, P.P.: Sur l’approximation de calcul bidimensionnel dans le cas d’un état de
tension plane. Mathematica (Cluj) 1(24), 345 (1959)
15. Teodorescu, P.P.: Considérations concernant la formulation mathématique du problème plane
de la théorie de l’élasticité. Rev. Roum. Math. Pures Appl. 9, 317 (1964)
16. Teodorescu, P.P.: Quelques problèmes bidimensionnels de la théorie de l’élasticité. Bull.
Math. Soc. Sci. Math. Roum. 10(58), 345 (1966)
17. Teodorescu, P.P.: Problèmes bidimensionnels de la théorie de l’élasticité. I. Une tension
normale nulle. Atti. Accad. Naz. Lincei. ser. VIII Rend. Cl. Sci. fis. mat. e nat. XLIV, 201
(1968)
18. Teodorescu, P.P.: Problèmes bidimensionnels de la théorie de l’élasticité. II. Deux tensions
tangentielles nulles. Atti. Accad. Naz. Lincei. ser. VIII, Rend. Cl. Sci. fis. mat. e nat. XLIV,
370 (1968)
19. Teodorescu, P.P. : Sur les problèmes bidimensionnels de la théorie de l’élasticité. Rend. Sem.
Matem. Torino, 27, 87 (1967–1968)
20. Teodorescu, P.P.: Quelques considérations sur les problèmes bidimensionnels de la théorie de
l’élasticité. Rev. Roum. Math. Pures Appl. XIII, 1467 (1968)
21. Teodorescu, P.P.: Quelques considérations sur le problème antiplan de la théorie de
l’élasticité. Rev. Roum. Sci. Techn. Méc. Appl. 22, 907 (1977)
22. Teodorescu, P.P.: La représentation par fonctions potentiel d’un état de déformation
antiplane. Rev. Roum. Sci. Techn., Méc. Appl. 23, 687 (1978)
23. Teodorescu, P.P.: Sur quelques problèmes bidimensionnels de la théorie de l’élasticité. Rev.
Roum. Sci. Techn. Méc. Appl. 25, 401 (1980)
24. Teodorescu, P.P.: On some quasibidimensional problems of the theory of elasticity.
SiSOM’99, C39 (1999)
25. Teodorescu, P.P.: On the Plane State of Strain of Elastodynamics. Bul. S ßt. Univ. Transilvania,
Brasßov, ser. A, Mec. 79 (2001)
26. Teodorescu, P.P.: Asupra stării de deformatßie antiplană în elastodinamică (On the Antiplane
State of Deformation in Elastodynamics). Bul. ßstiintß. Conf. natß. mec. sol. Brăila, 191 (2002)
Chapter 14
Anisotropic and Non-homogeneous Bodies

We remind that between the fundamental hypotheses of the theory of elasticity


presented in Sect. 2.1.2.2 are that of isotropy and homogeneity; the study made till
now has respected these hypotheses. Hereafter we will consider the cases in which
these hypotheses are no more respected.
If the mechanical (and physical) properties of the material, which intervene in
the constitutive law, are not the same in any direction in the neighbourhood of an
arbitrary point of the body, then the body is anisotropic (acolotropic).
As well, if the mechanical properties of the materials, corresponding to a
given direction, are not the same at any point of the body, then this one is
non-homogeneous.
In the following, we shall deal with these properties for elastic bodies, in the
general, as well in various particular cases.

14.1 Anisotropic Elastic Bodies

We have introduced in Sect. 4.1.3.1 Hooke’s tensor Hijkl , putting in evidence the
conditions which must be verified so that the quadratic form of the elastic potential
(4.54) be positive defined, using the criterion of Sylvester. Considering succes-
sively bodies with a plane of elastic symmetry (Sect. 4.1.3.2) and orthotropic
bodies (Sect. 4.1.3.3), one obtained Hooke’s law for an isotropic body (Sects.
4.1.3.4 and 4.1.3.5), which has been used in the study made till now.
Besides the bodies with intrinsic anisotropic properties, we may consider
structures which can be modelled as anisotropic ones, e.g., beams with transverse
ribs used for bridges and modelled as orthotropic plates. Other modelling as
anisotropic bodies is that of composite bodies, e.g., fiber-reinforced composites,
which—theoretically—are studied as such ones (see, for instance, the monograph
of Cristescu, Crăciun and Soós [3]).
Hereafter we shall consider various cases of anisotropy and we shall present
some elements of crystallography.

P. P. Teodorescu, Treatise on Classical Elasticity, Mathematical and Analytical 615


Techniques with Applications to Engineering, DOI: 10.1007/978-94-007-2616-1_14,
Ó Springer Science+Business Media Dordrecht 2013
616 14 Anisotropic and Non-homogeneous Bodies

14.1.1 Various Cases of Anisotropy

Starting from the general form of Hooke’s tensor (i.e., of Hooke’s law) we shall
give mechanical interpretations to the elastic coefficients which are involved, in
the general case as well as in particular cases. Thermoelastic and dynamic prob-
lems will be considered too.

14.1.1.1 General Case of Anisotropy

S. G. Lekhnitski [5] expresses the linear constitutive law in the form

e11 ¼ a11 r11 þ a12 r22 þ a13 r33 þ a14 r23 þ a15 r31 þ a16 r12 þ e011 ; . . .; ð14:1Þ

c23 ¼ a41 r11 þ a42 r22 þ a43 r33 þ a44 r23 þ a45 r31 þ a46 r12 þ c023 ; . . .; ð14:10 Þ

where the 36 coefficients aij ; i; j ¼ 1; 2; . . .; 6, are coefficients of deformation.


The relations may be written also in the form
r11 ¼ A11 e11 þ A12 e22 þ A13 e33 þ A14 c23 þ A15 c31 þ A16 c12 ; . . .; ð14:2Þ

r23 ¼ A41 e11 þ A42 e22 þ A43 e33 þ A44 c23 þ A45 c31 þ A46 c12 ; . . .; ð14:20 Þ

where the 36 coefficients Aij ; i; j ¼ 1; 2; . . .; 6, are elasticity moduli.


The quantities e011 ; e022 ; e033 ; c023 ; c031 ; c012 are initial strains. If these strains satisfy
Saint-Venant’s equations of continuity, then no initial state of stress will corre-
spond to them as has shown H. Reissner [20]; therefore, they have no more been
mentioned in the relations (14.2), (14.20 ).
Taking into account the Betti principle of reciprocity of the work, one may
show that the relations
aij ¼ aji ; Aij ¼ Aji ; i; j ¼ 1; 2; . . .; 6; ð14:3Þ

so that the 36 coefficients aij ; Aij are reduced to only 21.


Some researches tried to give a more technical form to these constants and to
the above relations, in general. So, A. L. Rabinovich [18] introduced elasticity
moduli for any direction as well as some coefficients of reciprocal influence; but
the coefficients thus introduced have not a specified physical significance and,
moreover, the results cannot be deduced by superposing the effects (principle
which is very important in the linear theory of elasticity). We shall try to obtain the
generalized Hooke law just by using this principle (see [24]).
To do this, we isolate an infinitesimal element of parallelepipedical form, with
the faces parallel to the planes of co-ordinates. Let us suppose that only the normal
stress r11 is acting. Due to this stress, appears a linear strain e11 ¼ r11 =E1 too,
where E1 is the longitudinal modulus of elasticity on the direction of Ox1 .
14.1 Anisotropic Elastic Bodies 617

If we suppose that r11 [ 0 corresponds to a traction, then in the directions Ox2


and Ox3 will appear the specific shortenings given by
m21 m31
e22 ¼  r11 ; e33 ¼  r11 :
E1 E1
Here, m21 and m31 are coefficients of transverse contraction (of Poisson’s type); the
first index indicates the direction of the deformation, while the second index
corresponds to the direction in which acts the cause which produces this
deformation.
Due to the normal stress r11 appear also the angular strains
g23;1 g31;1 g12;1
c23 ¼ r11 ; c31 ¼ r11 ; c12 ¼ r11 ;
E1 E1 E1
where g23;1 ; g31;1 ; g12;1 are coefficients of reciprocal influence, which will be called,
after Rabinovich, of second species. The first index (at the left of the comma)
shows the plane in which takes place the sliding; the second index gives the
direction in which acts the cause which produces the deformation.
If on the faces of the parallelepiped act only the tangential stresses r23 ¼ r32 ,
then appear the angular strains c23 ¼ r23 =G23 ; c32 ¼ r32 =G32 , where G23 ¼ G32
are the moduli of transverse elasticity in a plane normal to the Ox1 -axis. To the
tangential stress r23 correspond the linear strains
g1;23 g2;23 g3;23
e11 ¼ r23 ; e22 ¼ r23 ; e33 ¼ r23
G23 G23 G23
and the angular strains
m31;23 m12;23
c31 ¼ r23 ; c12 ¼ r23 ;
G23 G23
where g1;23 ; g2;23 ; g3;23 are coefficients of reciprocal influence of first species. The
first index shows the directions in which takes place the elongation (or the
shortening), while the second index indicates the plane in which acts the cause
which produces this deformation. As well, m31;23 and m12;23 are coefficients which
characterize the sliding in a plane normal to another one in which takes place a
sliding. We may call them the coefficients of Chentsov [9], being studied by this
researcher; as a matter of fact, Lekhnitski [5] gave them this name. The first index
indicates the plane in which takes place the sliding; the second index, shows the
plane in which takes place the cause which produces these slidings.
The stresses r11 ; r33 ; r31 ; r12 produce analogous strains. Using the principle of
superposition of effects, valid in the linear theory of elasticity, one may obtain the
components of the strain tensor in the most general form, i.e., one may write the
generalized law of Hooke as follows
618 14 Anisotropic and Non-homogeneous Bodies

1 m12 m13 g1;23 g1;31 g1;12


e11 ¼ r11  r22  r33 þ r23 þ r31 þ r12 ;
E1 E2 E3 G23 G31 G12
m21 1 m23 g2;23 g2;31 g2;12
e22 ¼ r11 þ r22  r33 þ r23 þ r31 þ r12 ; ð14:4Þ
E1 E2 E3 G23 G31 G12
m31 m32 1 g3;23 g3;31 g3;12
e33 ¼ r11  r22 þ r33 þ r23 þ r31 þ r12 ;
E1 E2 E3 G23 G31 G12
g23;1 g23;2 g23;3 1 m23;31 m23;12
c23 ¼ r11 þ r22 þ r33 þ r23 þ r31 þ r12 ;
E1 E2 E3 G23 G31 G12
g31;1 g23;2 g31;3 m31;23 1 m31;12
c31 ¼ r11 þ r22 þ r33 þ r23 þ r31 þ r12 ; ð14:40 Þ
E1 E2 E3 G23 G31 G12
g12;1 g12;2 g12;3 m12;23 m12;31 1
c12 ¼ r11 þ r22 þ r33 þ r23 þ r31 þ r12 :
E1 E2 E3 G23 G31 G12
In the constitutive law are involved 3 moduli of longitudinal elasticity
E1 ; E2 ; E3 , 3 moduli of transverse elasticity G23 ; G31 ; G12 (all these moduli have
the dimension of a stress, hence L1 MT2 ), 6 coefficients of transverse contraction
(of Poisson type) m23 6¼ m32 ; m31 6¼ m13 ; m12 6¼ m21 , 6 coefficients of transverse sliding
(Chentsov) m31;12 6¼ m12;31 ; m12;23 6¼ m23;12 ; m23;31 6¼ m31;23 , 9 coefficients of reciprocal
influence of first species g1;23 ; g1;31 ; g1;12 ; g2;23 ; g2;31 ; g2;12 ; g3;23 ; g3;31 ; g3;12 and 9
coefficients of reciprocal influence of second species g23;1 ; g23;2 ; g23;3 ; g31;1 ;
g31;2 ; g31;3 ; g12;1 ; g12;2 ; g12;3 ; hence, there are 3 þ 3 þ 6 þ 6 þ 9 þ 9 ¼ 36 non-
dimensional constants aij ; i; j ¼ 1; 2; . . .; 6.
Taking into account (14.3), one can state that these constants are not inde-
pendent. We may write following relations, in the order given by P. Bekhterev’s
classification [7].
I.
1 1 1
a11 ¼ ; a22 ¼ ; a33 ¼ ; ð14:5Þ
E1 E2 E3
II.
m12 m21 m23 m32 m31 m13
a12 ¼  ¼ ; a23 ¼  ¼ ; a13 ¼  ¼ ; ð14:50 Þ
E2 E1 E3 E2 E1 E3
III.
1 1 1
a44 ¼ ; a55 ¼ ; a66 ¼ ; ð14:500 Þ
G23 G31 G12
IV.
m23;31 m31;23 m31;12 m12;31 m12;23 m23;12
a45 ¼ ¼ ; a56 ¼ ¼ ; a46 ¼ ¼ ; ð14:6Þ
G31 G23 G12 G31 G23 G12
14.1 Anisotropic Elastic Bodies 619

V.
g1;23 g23;1 g2;31 g31;2 g3;12 g12;3
a14 ¼ ¼ ; a25 ¼ ¼ ; a36 ¼ ¼ ; ð14:60 Þ
G23 E1 G31 E2 G12 E3
VI.
g2;23 g23;2 g3;31 g31;3 g1;31 g31;1
a24 ¼ ¼ ; a35 ¼ ¼ ; a15 ¼ ¼ ;
G23 E2 G31 E3 G31 E1
g2;12 g12;2 g3;23 g23;3 g1;12 g12;1 ð14:600 Þ
a26 ¼ ¼ ; a34 ¼ ¼ ; a16 ¼ ¼ :
G12 E2 G23 E3 G12 E1

One has thus 3 þ 3 þ 3 þ 6 ¼ 15 relations between the 36 elastic constants;


hence, only 36  15 ¼ 21 elastic constants are independent, the other ones being
expressed by means of those ones.
The above results can be particularized for various important cases.

14.1.1.2 Bodies with Transverse Isotropy

We call body with transverse isotropy that one for which the planes normal to a
certain axis are isotropic planes (in an arbitrary point of the body one has the same
properties in all the directions contained in such a plane). This axis (called by
P. Bekhterev axis of monotropy) is an axis of total symmetry (of order n ¼ 1, as it
has been shown by S. G. Lekhnitski); this axis is identical to that with an order of
symmetry n ¼ 6.
Starting from the general relations (7.4), (6.1.40 ), one may write the generalized
Hooke law in the form (see [25])
1 m0 1 m0
e11 ¼ ðr11  mr22 Þ  0 r33 ; e22 ¼ ðr22  mr11 Þ  0 r33 ; ð14:7Þ
E E E E
m00 1
e33 ¼ ðr11 þ r22 Þ þ 0 r33 ; ð14:70 Þ
E E
1 1 1
c23 ¼ r23 ; c31 ¼ 0 r31 ; c12 ¼ 0 r12 ; ð14:700 Þ
G0 G G
where we have chosen the Ox3 -axis as axis of monotropy. Here, E and E0 are
moduli of longitudinal elasticity, G and G0 are moduli of transverse elasticity and
m; m0 and m00 are coefficients of transverse contraction of Poisson type.
One has thus 7 elastic constants, which verify the relations
E E
G¼ ; m00 ¼ 0 m0 ; ð14:8Þ
2ð1 þ mÞ E
620 14 Anisotropic and Non-homogeneous Bodies

hence, there remain 5 independent elastic constants.


There are two important cases which can be taken into consideration. Firstly,
we consider the case in which the monotropy axis Ox3 is normal to the Ox1 x2 -
plane, of a plane problem for which
r33 ¼ 0; r31 ¼ r32 ¼ 0: ð14:9Þ
Hooke’s law reads
1 1
e11 ¼ ðr11  mr22 Þ; e22 ¼ ðr22  mr11 Þ; ð14:10Þ
E E
2ð1 þ mÞ
c12 ¼ r12 ; ð14:100 Þ
E
where two elastic constants are involved as in the isotropic case. The linear strain
along the axis of monotropy Ox3 is given by
m0
e33 ¼  ðr11 þ r22 Þ; ð14:1000 Þ
E0
where the ratio m0 =E0 is a new elastic constant.
In the case of a plane state of strain, we assume that
e33 ¼ c31 ¼ c32 ¼ 0; ð14:90 Þ

we are lead to the same relations (14.10), (14.100 ), where we replace the elastic
constants E and m by the generalized elastic constants
E m þ m0 m00
E0 ¼ 0 00
; m0 ¼ ; ð14:11Þ
1mm 1  m0 m00
because the deformation is stopped in the direction of the Ox3 -axis, appears the
normal stress

r33 ¼ m0 ðr11 þ r22 Þ: ð14:12Þ

Excepting the determination of e33 in the case of a plane state of stress and of r33
in the case of a plane state of strain, the problem is identical, from the mathe-
matical point of view, in both cases; as well, the problem is identical to that of an
isotropic body.
The second important case is that in which the axis of monotropy is contained
in the plane in which appear the stresses and the strains. Some results in this
direction have been given by A. Moisil [17] and M. Iacovache [13], as well as
by M. Cristea [10].
We assume that the plane mentioned above is normal to the Ox2 -axis. In the
case of a plane state of stress one has
r22 ¼ 0; r21 ¼ 0; r23 ¼ 0: ð14:13Þ
14.1 Anisotropic Elastic Bodies 621

The relations (14.7)–(14.700 ) lead to


1 m0 1
e11 ¼ r11  0 r33 ¼ ðr11  m00 r33 Þ;
E E E ð14:14Þ
m00 1 1
e33 ¼  r11 þ 0 r33 ¼ 0 ðr33  m0 r11 Þ;
E E E
1
c31 ¼ r31 ; ð14:140 Þ
G0
where we used the relations (14.8) too. The linear strain along the Ox2 -axis is
given by
m m0 1
e22 ¼  r11  0 r33 ¼  ðmr11 þ m00 r33 Þ; ð14:1400 Þ
E E E
one has thus four distinct elastic constants.
In the case of a plane state of strain, we assume that
e22 ¼ 0; c21 ¼ c23 ¼ 0; ð14:130 Þ

obtaining the same relations (14.14), (14.1400 ), where we replace the elastic con-
stants E; E0 ; m0 ; m00 by the generalized elastic constants
E E0
E0 ¼ ; E00 ¼ ;
1m 2 1  m0 m00 ð14:16Þ
1 þ m 0 00 m00
m00 ¼ 0 00
m ; m0 ¼ :
1mm 1m
There appears also the normal stress

r22 ¼ mr11 þ m00 r33 : ð14:17Þ


As in the previous case, excepting the determination of e22 in the case of a plane
state of stress and of r22 for a plane state of strain, the problem is identical, in both
cases, from the mathematical point of view.
To the equations of equilibrium (in the absence of the volume forces)
r11;1 þ r13;3 ¼ 0; r31;1 þ r33;3 ¼ 0; ð14:18Þ

we add the equation of continuity of Saint-Venant


e11;33 þ e33;11 ¼ c13;13 : ð14:180 Þ

Taking into account the relations (14.14), (14.140 ), we may write the equation of
continuity in stresses
 
1 1 1 m00 m0
r11;33 þ 0 r33;11 ¼   r13;13 ; ð14:1800 Þ
E E G0 E E 0
622 14 Anisotropic and Non-homogeneous Bodies

where we took into account the equations of equilibrium (14.18) too.


We may use Airy’s representation
r11 ¼ F;33 ; r33 ¼ F;11 ; r13 ¼ F;13 ; ð14:19Þ

the stress function F ¼ Fðx1 ; x3 Þ must satisfy the partial differential equation
 
1 1 m00 m0 1
0
F;1111 þ   0 F;1133 þ F;3333 ¼ 0: ð14:20Þ
E G E E E
This equation may be written also in the form
ðs1 o1 þ io3 Þðs1 o1  io3 Þðs2 o1 þ io3 Þðs2 o1  io3 ÞF ¼ 0; ð14:200 Þ
where
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
s ffi
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u 2
u E E E
t
s1;2 ¼  m00   m00  0
2G0 2G0 E
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
rffiffiffiffiffiffi sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
rffiffiffiffiffiffi
1 E E 1 E E
¼  2m00 þ 2   2m00  2 ð14:2000 Þ
2 G0 E0 2 G0 E0

and where we used the relations (14.8). Thus, the general integral of this equation
may be represented, by means of certain functions of complex variable, in the form

Fðx1 ; x3 Þ ¼ uðx1 þ is1 x3 Þ þ wðx1  is1 x3 Þ


þ vðx1 þ is2 x3 Þ þ ,ðx1  is2 x3 Þ: ð14:21Þ

One may obtain then the state of strain and the state of displacement.

14.1.1.3 Orthotropic Bodies

We call orthotropic body that one through any point of which pass three planes of
elastic symmetry (see [26]).
Starting from the general results given in Sect. 14.1.1.1, we obtain the gen-
eralized Hooke law
1 m12 m13
e11 ¼ r11  r22  r33 ;
E1 E2 E3
m21 1 m23
e22 ¼ r11 þ r22  r33 ; ð14:22Þ
E1 E2 E3
m31 m32 1
e33 ¼ r11  r22 þ r33 ;
E1 E2 E3
1 1 1
c23 ¼ r23 ; c31 ¼ r31 ; c12 ¼ r12 ; ð14:220 Þ
G23 G31 G12
14.1 Anisotropic Elastic Bodies 623

there are involved 12 elastic constants of the material, between which take place
the relations
m12 m21 m23 m32 m31 m13
¼ ; ¼ ; ¼ ; ð14:23Þ
E2 E1 E3 E 2 E1 E3
so that only 9 elastic constants are independent.
In the case of a plane state of stress, we put the conditions
r33 ¼ 0; r31 ¼ r32 ¼ 0
and the relations (14.22), (14.220 ) take the form
1 m12 m21 1
e11 ¼ r11  r22 ; e22 ¼  r11 þ r22 ; ð14:24Þ
E1 E2 E1 E2
1
c12 ¼ r12 ; ð14:240 Þ
G12
to which one adds the linear strain
m31 m32
e33 ¼  r11  r22 ; ð14:2400 Þ
E1 E2
one has thus only 6 distinct elastic constants.
In the case of a state of plane strain one has e33 ¼ 0; c31 ¼ c32 ¼ 0; one is thus
led to the same relations (14.24), (14.240 ), where the elastic constants
E1 ; E2 ; m12 ; m21 are replaced by the generalized elastic constants
E1 E2
E10 ¼ ; E20 ¼ ;
1  m13 m31 1  m23 m32
ð14:25Þ
m12 þ m13 m32 0 m21 þ m23 m31
m012 ¼ ; m21 ¼ :
1  m23 m32 1  m13 m31
There appears also the stress
r33 ¼ m13 r11 þ m23 r22 ; ð14:26Þ
because the deformation in the Ox3 -direction vanishes. There appear only 6 elastic
constants too.
Excepting the determination of e33 for the state of plane stress and of r33 for the
state of plane strain, the problem is identical from the mathematical point of view,
involving only 4 distinct elastic constants (the other 2 constants are used to
determine the quantities mentioned above).
An interesting study is due to H. A. Lang [14] who, by an affine transformation,
replaces an orthotropic problem by an isotropic one. We remark also the results
given by E. Reissner [19] for some particular cases, especially if some constants
tend to zero. I .S. Sokolnikoff [23] gave an interesting approximate method of
624 14 Anisotropic and Non-homogeneous Bodies

calculation; starting from the corresponding isotropic problem, one introduces the
method of perturbations (a parameter of deviation from the isotropic solution to
the anisotropic one).
To the equations of equilibrium (without volume forces)
r11;1 þ r12;2 ¼ 0; r12;1 þ r22;2 ¼ 0

one adds the continuity conditions of Saint-Venant in the form


e11;22 þ e22;11 ¼ c12;12 :

Using the relations (14.24), (14.240 ), the equations of continuity in stresses take
the form
 
1 1 1 m21 m12
r11;22 þ r22;11 ¼   r12;12 : ð14:27Þ
E1 E2 G12 E1 E2

We introduce Airy’s representation (14.19), where the stress function F ¼


Fðx1 ; x2 Þ satisfies the equation
 
1 1 m21 m12 1
F;1111 þ   F;1122 þ F;2222 ¼ 0; ð14:28Þ
E2 G12 E1 E2 E1

which may be written in the form


ðs1 o1 þ io2 Þðs1 o1  io2 Þðs1 o1 þ io2 Þðs1 o1  io2 ÞF ¼ 0; ð14:280 Þ
where
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
rffiffiffiffiffi sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
rffiffiffiffiffi
1 E1 E 1 1 E1 E1
s1;2 ¼  2m21 þ 2   2m21  2 : ð14:2800 Þ
2 G12 E2 2 G12 E2

The general integral of this equation may be represented by means of the functions
of complex variable in the form

Fðx1 ; x2 Þ ¼ uðx1 þ is1 x2 Þ þ wðx1  is1 x2 Þ


ð14:29Þ
þ vðx1 þ is2 x2 Þ þ ,ðx1  is2 x2 Þ:

Solving the Eqs. (14.24), (14.240 ) with respect to stresses, one can write
E1 E2
r11 ¼ ðe11 þ m12 e22 Þ; r22 ¼ ðe22 þ m21 e11 Þ; ð14:30Þ
1  m12 m21 1  m12 m21
r12 ¼ G12 c12 : ð14:300 Þ
Replacing these expressions in the equations of equilibrium and taking into
account Cauchy’s relations, one obtains the equations which must be satisfied by
the displacements
14.1 Anisotropic Elastic Bodies 625

E1
ðu1;11 þ m12 u2;12 Þ þ G12 ðu1;11 þ u2;12 Þ ¼ 0;
1  m12 m21
ð14:31Þ
E2
G12 ðu1;12 þ u2;11 Þ þ ðu2;22 þ m21 u1;12 Þ ¼ 0:
1  m12 m21
Because the differential operators which act upon the functions u1 and u2 are
prime between them, it results that one may represent the displacements by means
of only one stress function F ¼ F ðx1 ; x2 Þ in the form
 
E1 m12 E1
u1 ¼  þ G12 F ; 12 ; u2 ¼ F ; 11 þ G12 F ; 22 ; ð14:32Þ
1  m12 m21 1  m12 m21

or in the form
 
E2 E1 m12
u1 ¼ F ; 22 þ G12 F ; 11 ; u2 ¼  þ G12 F ; 12 : ð14:320 Þ
1  m12 m21 1  m12 m21
If we put the condition that each of these representations, given by one of the
Eq. (14.31), verifies the other one, we observe that the function F must verify the
same Eq. (14.28) of elliptic type. This representation is analogue to that given by
K. Marguerre [15] for the plane isotropic problem. The solution is useful for the
first fundamental problem.

14.1.1.4 Bodies with a Plane of Elastic Symmetry

We call body with a plane of elastic symmetry that one for which there exists a plane
with the property that two directions symmetric with respect to it are equivalent from
the point of view of the elastic properties (see [27]). Starting from the general results
in Sect. 14.1.1.1, one can write Hooke’s generalized law in the form
1 m12 m13 g1;12
e11 ¼ r11  r22  r33 þ r12 ;
E1 E2 E3 G12
m21 1 m23 g2;12
e22 ¼ r11 þ r22  r33 þ r12 ; ð14:33Þ
E1 E2 E3 G12
m31 m32 1 g3;12
e33 ¼ r11  r22 þ r33 þ r12 ;
E1 E2 E3 G12
1 m23;31
c23 ¼ r23 þ r31 ;
G23 G31
ð14:330 Þ
m31;23 1
c31 ¼ r33 þ r31 ;
G23 G31
g12;1 g12;2 g12;3 1
c12 ¼ r11 þ r22 þ r33 þ r12 ; ð14:3300 Þ
E1 E2 E3 G12
where the Ox3 -axis is normal to the plane of elastic symmetry.
626 14 Anisotropic and Non-homogeneous Bodies

In these relations are involved 20 constants between which take place the
relations
m21 m12 m32 m23 m13 m31
¼ ; ¼ ; ¼ ; ð14:34Þ
E1 E2 E2 E 3 E3 E1
g12;1 g1;12 g12;2 g2;12 g12;3 g3;12
¼ ; ¼ ; ¼ ; ð14:340 Þ
E1 G12 E2 G12 E3 G12
m31;23 m23;31
¼ ; ð14:3400 Þ
G23 G31
there remain thus only 13 independent elastic constants.
The conditions
r33 ¼ 0; r31 ¼ 0; r32 ¼ 0

lead to a state of plane stress, for which Hooke’s law takes the form
1 m12 g1;12
e11 ¼ r11  r22 þ r12 ;
E1 E2 G12
ð14:35Þ
m21 1 g2;12
e22 ¼ r11 þ r22 þ r12 ;
E1 E2 G12
g12;1 g12;2 1
c12 ¼ r11 þ r22 þ r12 ð14:350 Þ
E1 E2 G12
to which one adds the linear strain
m31 m32 g3;12
e33 ¼  r11  r22 þ r12 ; ð14:3500 Þ
E1 E2 G12
there remain 9 distinct elastic constants.
If we assume that c33 ¼ 0; c31 ¼ c32 ¼ 0, then, for the case of a plane state of
strain, one can use the same constitutive law (14.35), (14.350 ), where the elastic
constants are replaced by the generalized elastic constants
E1 E2 G12
E10 ¼ ; E20 ¼ ; G012 ¼ ; ð14:36Þ
1  m13 m31 1  m23 m32 1  g12;3 g3;12

m12 þ m13 m32 0 m21 þ m23 m31


m012 ¼ ;m ¼ ; ð14:360 Þ
1  m23 m32 21 1  m13 m31
g1;12 þ m13 g3;12 0 g12;1 þ m31 g12;3
g01;12 ¼ ; g12;1 ¼ ;
1  g12;3 g3;12 1  m13 m31
ð14:3600 Þ
g2;12 þ m23 g3;12 0 g12;2 þ m32 g12;3
g02;12 ¼ ; g12;2 ¼ :
1  g12;3 g3;12 1  m23 m32

Because the deformation is stopped in the direction Ox3 . there appears the normal
stress
14.1 Anisotropic Elastic Bodies 627

r33 ¼ m13 r11 þ m23 r22 þ g12;3 r12 ; ð14:37Þ

there remain 9 arbitrary elastic constants.


Excepting the determination of e33 for the state of plane stress and of r33 for the
state of plane strain, the problem is identical from the mathematical point of view;
only 6 elastic constants are involved, the other 3 elastic constants appearing only
in the expressions of the remaining quantities.
We mention a study of Green [12] in this problem.
Taking into account the equations of continuity of Saint-Venant and the gen-
eralized law of Hooke, the first equations, written in stresses, read
 
1 1 1 m21 m12
r11;22 þ r22;11 ¼   r12;12
E1 E2 G12 E1 E2
g1;12 g2;12
2 r12;22  2 r12;11 : ð14:38Þ
G12 G12
The state of stress can be represented by Airy’s formulae (14.19), where the stress
function F ¼ Fðx1 ; x2 Þ must satisfy the partial differential equation
 
1 g2;12 1 m21 m12
F;1111  2 F;1112 þ   F;1122
E2 G12 G12 E1 E2
g1;12 1
2 F;1222 þ F;2222 ¼ 0: ð14:39Þ
G12 E1
One can also give a solution in displacements of the problem, analogous to that
given by K. Marguerre [15] for isotropic bodies and by us [26] for orthotropic
bodies. The potential function must verify the same Eq. (14.39).
One may obtain, as particular cases, the results corresponding to an orthotropic
body, as well as the results corresponding to a body with transverse isotropy, if the
axis of monotropy is normal to the plane of elastic symmetry.
We remark that in the case of a body with an axis of symmetry of order 2 (for
which two arbitrary directions are superposing after a rotation of 2p=2 ¼ p about
the axis are equivalent from the point of view of the elastic properties) one obtains
a constitutive law of the same form (14.33), (14.3300 ).

14.1.1.5 Bodies with an n-gonal Axis of Symmetry

We say that a body has an axis of symmetry of order n (an n-gonal axis) if two
arbitrary directions are superposing after a rotation of 2p=n about that axis are
equivalent from the point of view of the elastic properties. It has been stated that
one can have only n ¼ 2 (digonal axis), n ¼ 3 (trigonal axis), n ¼ 4 (tetragonal
axis), n ¼ 6 or 1 (hexagonal axis) (see, e.g., [6]); this corresponds to the law of
rational indices. The cases n ¼ 2 (plane of elastic symmetry) and n ¼ 6 (trans-
verse isotropy) have been considered above (see [28, 29]).
628 14 Anisotropic and Non-homogeneous Bodies

In the case of a trigonal axis, the generalized Hooke law takes the form
1 m12 m13 g1;23 g1;31
e11 ¼ r11  r22 þ r33 þ r23 þ r31 ;
E1 E2 E3 G23 G31
ð14:40Þ
m21 1 m23 g2;23 g2;31
e22 ¼ r11  r22  r33 þ r23 þ r31 ;
E1 E2 E3 G23 G31
m31 m32 1
e33 ¼  r11  r22 þ r33 ; ð14:400 Þ
E1 E2 E3
g23;1 g23;2 1 m23;12
c23 ¼ r11 þ r22 þ r23 þ r12 ;
E1 E2 G23 G12
ð14:4000 Þ
g31;1 g31;2 1 m31;12
c31 ¼ r11 þ r22 þ r31 þ r12 ;
E1 E2 G31 G12
m12;23 m12;31 1
c12 ¼ r23 þ r31 þ r12 ; ð14:40000 Þ
G23 G31 G12
where the Ox3 -axis has been chosen as axis of symmetry. There are 24 elastic
constants of the material, which verify the relations
m21 m12 m32 m23 m13 m31
¼ ; ¼ ; ¼ ;
E1 E2 E2 E3 E3 E1
g23;1 g1;23 g31;1 g1;31 g23;2 g2;23
¼ ; ¼ ; ¼ ; ð14:41Þ
E1 G23 E1 G31 E2 G23
g31;2 g2;31 m23;12 m12;23 m31;12 m12;31
¼ ; ¼ ; ¼ ;
E2 G31 G12 G23 G12 G31
because the Ox3 -axis is an axis of symmetry of third order, there appear following
relations between the elastic constants
 
1 1 m12
E1 ¼ E2 ; G23 ¼ G31 ; ¼2 þ ; m13 ¼ m23 ;
G12 E1 E 2
ð14:410 Þ
g23;1 g23;2 m31;12 g31;1 g31;2 m23;12
¼ ¼ ; ¼ ¼ :
E1 E2 G12 E1 E2 G12
Hence, there remain 24  ð9 þ 8Þ ¼ 7 independent elastic constants.
The relations (14.41), (14.410 ) allow to write Hooke’s law in the form
1 m0 1
e11 ¼ ðr11  mr22 Þ  0 r33 þ 0 ðgr23  g0 r31 Þ;
E E G ð14:42Þ
1 m0 1
e22 ¼ ðr22  mr11 Þ  0 r33  0 ðgr23  g0 r31 Þ;
E E G
m00 1
e33 ¼  ðr11 þ r22 Þ þ 0 r33 ; ð14:420 Þ
E E
14.1 Anisotropic Elastic Bodies 629

g00 1 m
c23 ¼ ðr11  r22 Þ þ 0 r23 þ r12 ;
E G G ð14:4200 Þ
g000 1 m0
c31 ¼  ðr11  r22 Þ þ 0 r31 þ r12 ;
E G G
1 00 1
c12 ¼ ðm r23 þ m000 r31 Þ þ r12 ; ð14:42000 Þ
G0 G
using thus 15 elastic constants; between these constants there exist 8 relations
m0 m00 m0 m000 m m00 1 2ð1 þ mÞ
0
¼ ; ¼ 0; ¼ 0; ¼ ;
E E G G G G G E ð14:43Þ
g g00 g0 g000 g00 m0 g000 m
¼ ; ¼ ; ¼ ; ¼ :
G0 E G0 E E G E G
If we put the conditions r33 ¼ 0; r31 ¼ r32 ¼ 0 for a state of plane stress (the
elastic axis has been considered to be normal to the plane in which take place the
relations)
1 1
e11 ¼ ðr11  mr22 Þ; e22 ¼ ðr22  mr11 Þ;
E E ð14:44Þ
2ð1 þ mÞ 1
c12 ¼ r12 ¼ r12 ;
E G
where appears the elastic constants of the isotropic bodies. The linear strain along
the elastic axis is given by
m00
e33 ¼  ðr11 þ r22 Þ; ð14:440 Þ
E
to which one adds the angular strains (which, in the isotropic case, are equal to
zero)
1 00
c23 ¼ ½m ðr11  r22 Þ þ mr12 ;
G ð14:4400 Þ
1
c31 ¼ ½mðr22  r11 Þ þ m0 r12 ;
G
where the relations (14.43) have been used.
In the case of a plane state of strain (e33 ¼ 0; c31 ¼ c32 ¼ 0) one uses the same
relations (14.44), where the generalized elastic constants
E G
E0 ¼ ; G0 ¼ ;
1 m0 m00 00
 ðgg þ g g Þ0 000 1  ðmm  m0 m000 Þ
00
ð14:45Þ
m þ m0 m00  ðgg00 þ g0 g000 Þ
m0 ¼
1  m0 m00  ðgg00 þ g0 g000 Þ
are introduced. We mention that these constants are independent.
630 14 Anisotropic and Non-homogeneous Bodies

Because the deformation is stopped in the direction of the axis, there appears
the normal stress
r33 ¼ m0 ðr11 þ r22 Þ ð14:46Þ

and the tangential stresses

r31 ¼ m00 ðr11  r22 Þ  m000 r12 ; r32 ¼ m000 ðr22  r11 Þ  m00 r12 : ð14:460 Þ
The problem of plane stress is the same as in the isotropic case; but in the
problem of plane strain, the equation of continuity in stresses takes the form
 
1 1 m0
ðr11;22  r22;11 Þ ¼ 2 r12;12 : ð14:47Þ
E0 G0 E0

One uses the elastic representation of Airy, where the potential function F ¼
Fðx1 ; x2 Þ verifies the equation
 
E0
F ;1111 þ  2m0 F;1122 þ F ;2222 ¼ 0; ð14:48Þ
G0

the general solution of which is of the form

Fðx1 ; x2 Þ ¼ uðx1 þ is1 x2 Þ þ wðx1  is1 x2 Þ


þ vðx1 þ is2 x2 Þ þ ,ðx1  is2 x2 Þ; ð14:480 Þ

where
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 E0 1 E0
s1;2 ¼ þ 2ð1  m0 Þ   2ð1  m0 Þ: ð14:4800 Þ
2 G0 2 G0
In displacements, the solution may be represented in the form
 
E0 m 0 E0
u1 ¼  þ G0 F ; 12 ; u2 ¼ F ; 11 þ G0 F ; 22 ð14:49Þ
1  m20 1  m20

or in the form
 
E0 E 0 m0
u1 ¼ F ; 22 þ G0 F ; 11 ; u 2 ¼  þ G0 F ; 12 ; ð14:490 Þ
1  m20 1  m20

where the function F ¼ F ðx1 ; x2 Þ satisfies the same Eq. (14.48).


Unlike the isotropic case, in the present case the tangential stresses r31 ; r32 are
non-zero.
In the case of a tetragonal axis, the generalized law of Hooke takes the form
1 m12 m13 g1;12
e11 ¼ r11  r22  r33 þ r12 ;
E1 E2 E3 G12
ð14:50Þ
m21 1 m23 g2;12
e22 ¼ r11 þ r22  r33 þ r12 ;
E1 E2 E3 G12
14.1 Anisotropic Elastic Bodies 631

m31 m32 1
e33 ¼  r11  r22 þ r33 ; ð14:500 Þ
E1 E2 E3
1 1
c23 ¼ r23 ; c31 ¼ r31 ; ð14:5000 Þ
G23 G31
g12;1 g12;2 1
c12 ¼ r11 þ r22 þ r12 ; ð14:50000 Þ
E1 E2 G12
where the Ox3 -axis is that of elastic symmetry.
There appear 16 elastic constants of the material, between which take place the
relations
m21 m12 m32 m23 m13 m31
¼ ; ¼ ; ¼ ;
E1 E 2 E2 E 3 E3 E1
g12;1 g1;12 g12;2 g2;12 ð14:51Þ
¼ ; ¼ :
E1 G12 E2 G12
Because the elastic axis is of fourth order, there take place the relations

E1 ¼ E2 ; G23 ¼ G31 ;
ð14:510 Þ
m13 ¼ m23 ; g1;12 ¼ g2;12

too. Thus, one has only 16  ð5 þ 4Þ ¼ 7 independent elastic constants.


One can write now Hooke’s law in the form
1 m0 g
e11 ¼ ðr11  mr22 Þ  0 r33 þ r12 ;
E E G ð14:52Þ
1 m0 g
e22 ¼ ðr22  mr11 Þ  0 r33  r12 ;
E E G
m00 1
e33 ¼  ðr11 þ r22 Þ þ 0 r33 ; ð14:520 Þ
E E
1 1
c23 ¼ r23 ; c31 ¼ 0 r31 ; ð14:5200 Þ
G0 G
g0 1
c12 ¼ ðr11  r22 Þ þ r12 ; ð14:52000 Þ
E G
using 9 elastic constants, which verify the relations
m00 m0 g0 g
¼ ; ¼ ; ð14:53Þ
E E0 E G
hence, there remain 7 distinct elastic constants.
The conditions r33 ¼ 0; r31 ¼ r32 ¼ 0 lead to a plane state of stress, the con-
stitutive law taking the form
632 14 Anisotropic and Non-homogeneous Bodies

1 g 1 g
e11 ¼ ðr11  mr22 Þ þ r12 ; e22 ¼ ðr22  mr11 Þ  r12 ; ð14:54Þ
E G E G
g0 1
c12 ¼ ðr11  r22 Þ þ r12 ; ð14:540 Þ
E G
where appear 5 elastic constants, connected by the second relation (14.53). The
linear strain along the elastic axis is given by
m00
e33 ¼  ðr11 þ r22 Þ: ð14:5400 Þ
E
For a plane state of strain, we use the same relations (14.54), (14.540 ), the
generalized elastic constants taking the form
E m þ m0 m00
E0 ¼ 0 00
; m0 ¼ ;
1mm 1  m0 m00 ð14:55Þ
g0
g00 ¼ ;
1  m0 m00
because the linear strain along the axis of symmetry is stopped, there appears the
normal stress

r33 ¼ m0 ðr11 þ r22 Þ: ð14:56Þ


Excepting the linear strain e33 in the case of a plane state of stress and the
normal stress r33 in the case of a plane state of strain, the problem is the same from
the mathematical point of view, 4 distinct elastic constants being involved.
The equation of continuity of Saint-Venant is expressed in stresses in the form
 0 
0 g
r11;22 þ r22;11 ¼ 2g ðo11  o22 Þr12 þ  2m r12;12 : ð14:57Þ
g
One can express the state of stress in the form given by G. B. Airy, where the
stress function F ¼ Fðx1 ; x2 Þ must verify the partial differential equation
 0 
0 g
F;1111 þ 2g F;1112 þ  2m F;1122  2g0 F;1222 þ F;2222 ¼ 0; ð14:58Þ
g

which may be studied similarly as the previous ones.


The state of displacement can be obtained analogically.

14.1.1.6 Case of Arbitrary Volume Forces

Let us consider the case of an orthotropic body (see Sect. 14.1.1.3). In a plane state
of stress, the equations of equilibrium read
r11;1 þ r12;2 þ F1 ¼ 0; r12;1 þ r22;2 þ F2 ¼ 0; ð14:59Þ
14.1 Anisotropic Elastic Bodies 633

where Fi ¼ Fi ðx1 ; x2 Þ; i ¼ 1; 2, are the volume forces. Using Airy’s representation,


one can express the stresses in the form
Z Z
r11 ¼ F;22  F1 dx1 ; r22 ¼ F;11  F2 dx2 ; ð14:60Þ

r12 ¼ F;12 ; ð14:600 Þ

where a system of particular integrals has been added (see [30]).


Denoting by DF the first member of the Eq. (14.28), one can write the equation
which determines the potential function F ¼ Fðx1 ; x2 Þ in the form
Z Z
1 1 m21 m12
DF ¼ F1;22 dx1 þ F2;11 dx2  F2;2  F1;1 : ð14:61Þ
E1 E2 E1 E2
Similarly, if one denotes by D1 u1 þ D2 u2 and by D01 u1 þ D02 u2 the first
members of the Eq. (14.31), then the system of partial differential equations which
is verified by the displacements reads

D1 u1 þ D2 u2 þ F1 ¼ 0; D01 u1 þ D02 u2 þ F2 ¼ 0: ð14:62Þ

If F 1 ¼ F 1 ðx1 ; x2 Þ and F 2 ¼ 0, then the displacements are given by (14.320 ), where


the potential function F ¼ F ðx1 ; x2 Þ verifies the equation
1  m12 m21
DF þ F1 ¼ 0; ð14:63Þ
E1 E2 G12
if F 1 ¼ 0 and F 2 ¼ F 2 ðx1 ; x2 Þ, then the displacements are given by (14.32), where
the potential function F ¼ F ðx1 ; x2 Þ verifies the equation
1  m12 m21
DF þ F2 ¼ 0: ð14:630 Þ
E1 E2 G12
In the general case, one superposes the effects.
Analogically, one can study the other cases of anisotropy.
An important problem is the uniqueness of the potential function for a given
state of stress or strain. In this direction, we mention the synthesis paper of M.
M. Fridman [11], as well as the monograph of S. G. Lekhnitski [5]. The problem
has been solved by S. G. Mikhlin [16] in 1936 and has been taken again by
G. N. Savin for bodies with a part of the frontier at infinite [21] or for simply
connected bodies of finite dimensions [22].
We may state (see [30]) that: In the case of the plane problem of elasticity of an
anisotropic body acted upon by arbitrary volume forces, the potential function, as
well as its partial derivatives are univocally determined in the interior of a simply
connected domain if the loads on the frontier are equilibrated by the volume
forces.
The equation which must be verified by the potential function (homogeneous
equation) is of the form
634 14 Anisotropic and Non-homogeneous Bodies

aF;1111 þ bF;1112 þ cF;1122 þ dF;1222 þ eF;2222 ¼ 0; ð14:64Þ

where the coefficients a; b; c; d; e depend on the elastic constants of the material. A


particular integral is of the form

Fðx1 ; x2 Þ ¼ ea1 x1 þa2 x2 ; ð14:65Þ


where ai ; i ¼ 1; 2, are related by the relation

aa41 þ ba31 a2 þ ca21 a22 þ da1 a32 þ ea42 ¼ 0; ð14:650 Þ

which may take a simpler form in important particular cases. By means of such
particular integrals, one may construct, e.g., expansions into Fourier series too.

14.1.1.7 Dynamic Problems

The problems considered above have been studied also in the dynamic case; we
mention, e.g., the results of R. E. D. Bishop [8] concerning both plane problems:
state of stress and state of strain.
In the case of an orthotropic elastic body, the equations corresponding to a
plane state of stress are
r11;1 þ r12;2 ¼ q€
u1 ; r21;1 þ r22;2 ¼ q€u2 ; ð14:66Þ

1 m12 m21 1
u1;1 ¼ r11  r22 ; u2;2 ¼  r11 þ r22 ; ð14:67Þ
E1 E2 E1 E2
1
u1;2 þ u2;1 ¼ r12 ; ð14:670 Þ
G12
where the elastic constants are related by
m12 m21
¼ : ð14:68Þ
E2 E1
In the case of a plane state of strain, the same equations are used, the elastic
constants being replaced by
E1 E2
E10 ¼ ; E0 ¼ ;
1  m13 m31 2 1  m23 m32
ð14:69Þ
m12 þ m13 m32 0 m23 þ m23 m31
m012 ¼ ; m23 ¼ ;
1  m23 m32 1  m13 m31
where m13 6¼ m31 ; m12 6¼ m21 are coefficients of Poisson type.
Eliminating the stresses between these equations, one obtains the equations of
Lamé type
14.1 Anisotropic Elastic Bodies 635

E1
ðu1;11 þ m12 u2;12 Þ þ G12 ðu1;22 þ u2;12 Þ ¼ q€u1 ;
1  m12 m21
ð14:70Þ
E2
G12 ðu1;12 þ u2;11 Þ þ ðu2;22 þ m21 u1;12 Þ ¼ q€u2 :
1  m12 m21
Because the differential operators which act on the displacements u1 and u2 are
prime between them, it results that the displacements may be represented by a
single potential function F in the form
 
E1 m12 E1 € ð14:71Þ
u1 ¼  þ G12 F ;12 ; u2 ¼ F ;11 þ G12 F ;22  qF
1  m12 m21 1  m12 m21

or in the form
 
E2 € E1 m12
u1 ¼ F ;22 þ G12 F ;11  qF ; u2 ¼  þ G12 F ;12 :
1  m12 m21 1  m12 m21
ð14:710 Þ
If we put the condition that each of these representations, given by one of the
equation (14.70), verify the other equation too, then we conclude that the function
F ¼ F ðx1 ; x2 ; tÞ is given by the equation
 
1 1 m21 m12 1
F ;1111 þ   F ;1122 þ F ;2222
E2 G12 E1 E2 E1
   
q 1  m12 m21 1 € q 1  m12 m21 1 €
 þ F ;11  þ F ;22
E2 E1 G12 E1 E2 G12
q2 ð1  m12 m21 Þ €
€ ¼ 0:
þ F ð14:72Þ
E1 E2 G12
Such a solution of the problem is useful in the case of the first problem of the
theory of elasticity.
The discriminant of the form (14.72), considered as a quadratic form, is given
by
    
 1 1 1
 2 m21
 q 1m12 m21
þ 1 
 G12 
  E2  2 G12 E1 2E2
 E1 
 
 1 1 m21
2 G12  2 E1
1
 2Eq 1 1mE122 m21 þ G112 : ð14:73Þ
    E1
 
 q 1m12 m21 1 q 1m12 m21 1 q2 ð1m12 m21 Þ 
  2E E þ G  2E E þ G E E G

2 1 12 1 2 12 1 2 12

One observes easily that this discriminant does not vanishes (if we consider it as a
polynomial in 1/G12 then we observe that the term of third degree of it has a non-
zero coefficient). Hence, unlike the isotropic case, the elastic waves can no more
be obtained by superposing two types of waves (irrotational and shearing waves,
for instance), but there exists only one wave front.
636 14 Anisotropic and Non-homogeneous Bodies

Thus, in the case of irrotational waves, the relation


u1;2  u2;1 ¼ 0

must be fulfilled; hence, taking into account (14.80), the displacements must verify
the equations

E1 u1;22 þ ½2G12 ð1  m12 m21 Þ þ E1 m12 u1;22  qð1  m12 m21 Þ€u1 ¼ 0;
ð14:74Þ
½2G12 ð1  m12 m21 Þ þ E2 m21 u2;11 u2;11 þ E2 u2;22  qð1  m12 m21 Þ€u2 ¼ 0:

On the other hand, these displacements are integrals of the equation (14.72). These
conditions may be fulfilled simultaneously only in particular cases.
For a solution in stresses of the problems, one eliminates u1 and u2 between the
equations of the problem; we get
 
1 m12
r11;11 þ r12;12 ¼ q €11 
r r
€22 ;
E1 E2
  ð14:75Þ
1 m21
r12;12 þ r22;22 ¼ q €22 
r r
€11 :
E2 E1
Subtracting these equations one of the other, we can state that the normal
stresses may be expressed in the form
q €
r11 ¼ F;22  ð1 þ m12 Þ F;
E2
q € ð14:76Þ
r22 ¼ F;11  ð1 þ m21 Þ F;
E1
while the tangential stress will be given by
ZZ  
o2 1 1 1  m12 m21 o2
r12 ¼  F;12 þ q 2 o11 þ o22  q 2 Fdx1 dx2
ot E2 E1 E1 E2 ot
2 
o 1 1
þq 2 f1 ðx2 ; tÞ þ f2 ðx1 ; tÞ : ð14:760 Þ
ot E2 E2

Then, one can express the components of displacements


Z Z
q € 1 þ f1;2 ;
E1 u1 ¼ F;22 dx1  m21 F;1  ð1  m12 m21 Þ Fdx
E2
Z Z ð14:77Þ
q € 2 þ f2;1 :
E2 u2 ¼ F;11 dx2  m12 F;2  ð1  m12 m21 Þ Fdx
E1
The equation of continuity, obtained from (14.67) and written in stresses, reads
1 1 1
ðo22  m21 o11 Þr11 þ ðo11  m21 o22 Þr22 ¼ r12;12 ; ð14:78Þ
E1 E2 G12
which shows that the potential function Fðx1 ; x2 ; tÞ must satisfy the same Eq. (14.72).
14.1 Anisotropic Elastic Bodies 637

The relation (14.670 ) allows to determine the functions f1 ðx2 ; tÞ and f2 ðx1 ; tÞ in
the form
Z Z  
1 1 1 m21 m12
F;22 dx1 þ F;11 dx2 þ   F;12
E1 E2 G12 E1 E2
 Z Z 
o2 1 1
 ð1  m12 m21 Þq 2 F;2 dx1 þ F;1 dx2
ot E2 E1
2 Z Z   ð14:79Þ
q o 1 1 1  m12 m21 o2
 o 11 þ o 22  q Fdx dx
1 2
G12 ot2 E2 E1 E 1 E2 ot2
 
1 1 q o2 1 1
þ f1;22 þ f2;11  f1 þ f2 ¼ 0:
E1 E2 G12 ot2 E1 E2
If we introduce everywhere double integrals with respect to x1 and x2 , we obtain
under the integral sign the expression in (14.72). Hence, Eq. (14.79) is with
separate variables and can be decomposed in two equations of the form
1€
f1;22  f1 þ F1 ðx2 ; tÞ þ E1 x0 ðtÞ ¼ 0;
a2 ð14:790 Þ
1
f2;11  2 €f2 þ F2 ðx1 ; tÞ  E2 x0 ðtÞ ¼ 0;
a
the function x0 ðtÞ being arbitrary, while a is a velocity of waves propagation,
given by
1 q
2
¼ : ð14:80Þ
a G12
By integration, one obtains

 1 2 0
f1 ðx2 ; tÞ  u1 ðx2  atÞ þ w1 ðx2 þ atÞ þ f1 ðx2 ; tÞ þ E  x0 ðtÞx2 þ u1 ðtÞx2 ;
2

 1 2 0
f2 ðx2 ; tÞ  u2 ðx1  atÞ þ w1 ðx1 þ atÞ þ f2 ðx2 ; tÞ þ E x0 ðtÞx1 þ u1 ðtÞx1 ;
2
ð14:81Þ

where u1 ; u2 ; w1 ; w2 are arbitrary functions; f1 ðx2 ; tÞ; f2 ðx1 ; tÞ are particular inte-
grals of the Eq. (14.790 ), while u01 ðtÞ; u02 ðtÞ and x0 ðtÞ are the displacements and the
rotation of the body, respectively, considered as a rigid. This constitutes a gen-
eralization for the dynamic case of the results obtained in the static one.
Choosing a function F of the form
Fðx1 ; x2 ; tÞ ¼ Fðx1 ; x2 ÞTðtÞ;
one may assume that the stress function is obtained by a superposition of harmonic
vibrations
638 14 Anisotropic and Non-homogeneous Bodies

X
1
2np
Fðx1 ; x2 ; tÞ ¼ Fn ðx1 ; x2 Þ cosðhn þ cn tÞ; cn ¼ ; ð14:82Þ
n¼1
T

where cn is the phase shift, while T is the period.


The Eq. (14.72) becomes
 
1 1 m21 m12 1
Fn;1111 þ   Fn;1122 þ Fn;2222
E2 G12 E1 E2 E1
   
q 1  m12 m21 1 q 1  m12 m21 1
þ þ Fn;11 þ þ Fn;22
E2 E1 G12 E1 E2 G12
q2 ð1  m12 m21 Þ
þ Fn ¼ 0; ð14:83Þ
E1 E2 G12
obtaining thus the state of strain and stress corresponding to stationary vibrations.
One can use Fourier methods, i.e., Fourier series in case of periodic loads or
Fourier integrals in case of bodies subjected to a shock (appearing the notion of
momentum).

14.1.2 Elements of Crystallography

Crystallography is a branch of science which studies the external form and the
internal structure of crystals, their development and growth and their properties.
Initially, crystallography was a part of mineralogy, becoming an independent
discipline only at the end of nineteenth century, when it has been found that there
exist non-mineral substances with crystalline structure and that the number of such
substances is sufficiently great. The bodies with such geometrical characteristics
have many specific properties. Nowadays, cristallography is developing in close
relationship with mineralogy and metallography, which use the results of crysta-
lography, as well as with chemistry, physics and mathematics, the results of which
are applied in crystallography.

14.1.2.1 Short History

The first references concerning the regular structure of the spatial arrangement of
the particles that make up a crystalline substance can be found in the works of
J. Kepler (1619), R. Hooke (1665) and C. Huygens (1690). In 1669, N. Stensen
discovered the law of constancy of the angle between the crystalline planes and, a
century later, in 1774 R. J. Ha} uy formulated the law of integers (of rational
indices), relative to the characterization of the positions of crystalline planes.
In the first decade of the nineteenth century, the systematic measurement of the
angles formed by crystalline planes began and W. H. Miller introduced the use of
14.1 Anisotropic Elastic Bodies 639

the methods of analytical geometry in crystallography. At the end of the century,


the fundamental principles of symmetry of crystals where established and the
classification of crystals on the basis of their properties of symmetry was com-
pleted. All possible crystalline classes had been first determined by Hessel, in
1830; independently, the 32 types of symmetry have also been specified geo-
metrically by A. W. Gadolin, in 1867.
The last stage in the development of crystallography began in 1912. Since
nobody had yet succeeded to prove experimentally the validity of the theory
asserting the reticular structure of crystals, at the end of nineteenth century there
were even doubts about the correctness of the theories which maintained the
discrete, atomic, molecular structure of crystals. However, the situation changed
radically, after W. K. R} ontgen had discovered, in 1895, a new type of radiation,
considered by some scientists as having a wave nature, a kind of electromagnetic
field characterized by a very short wavelength, whereas others asserted that it
consisted in an emission of corpuscles, because no reflexion, refraction or inter-
ference of these ‘X-rays’ could be experimentally observed.
Since even the finest diffraction gratings that could be produced artificially had
no effect on these rays, von Laue decided to use a natural diffraction grating, i.e., a
crystal; if the theory of reticular structure of crystals and the undulatory, elec-
tromagnetic theory of X-rays were correct, then the phenomenon of interference
must occur in interaction between R} ontgen’s radiation and a crystalline material.
Thus, Laue’s aim was to verify, through a single experiment, the validity of two
theories, which in fact he experimentally did. This result opened new possibilities
for the development of crystallography, physics and chemistry. The subsequent
successes of crystallography due to E. S. Fedorov, G. V. Wulff, W. H. Bragg and
W. L. Bragg, are based on the famous results obtained by Laue, which enabled a
determination of the geometrical structure of crystalline materials and the effects
of this specific structure on the physical and chemical properties of polycrystalline
substances.

14.1.2.2 Geometrical Properties of Crystals. Crystalline Lattices

The crystals are bounded by faces (planes), edges (straight lines) and vertices
(points); the exterior faces of perfect crystals have the following properties:
(i) the angles between the corresponding faces of the crystals of a given chemical
compound are constant;
(ii) if the positions of the faces are expressed in a definite co-ordinate system, then
the position of any face can always be specified by three integral numbers.
The study of physical properties of crystals shows that, in general, the prop-
erties are not the same in different directions; for instance, mechanical rigidity,
heat conductivity, dielectric permittivity vary relative to the direction in which
they are determined; for this reason, we say that the crystals are anisotropic.
However, there can exist several different directions along which some properties
640 14 Anisotropic and Non-homogeneous Bodies

of a given crystal are identical; the spatial arrangement of these directions reveals
the existence of a symmetry of the geometrical structure of crystalline bodies.
The observed characteristics can be partly expressed without taking into
account the nature and the form of the constituent particles of crystals, if we
suppose that the crystalline substances, more precisely the monocrystalline sub-
stances which form the real polycrystalline media, have a regular reticular struc-
ture; for this reason, the spatial arrangement of the particles which form a perfect
crystal can be represented mathematically through a rectilinear, three-dimensional,
regular net (lattice); formed by an enormous, but finite, number of identical par-
allelepipeds, called unit cells, without any gaps between them. The vertices of the
parallelepipeds constitute the space lattice; these points are called nodes, the lines
drawn through them are rows and the planes determined by two concurrent rows
are called plane nets. Using three concurrent non-coplanar edges of a unit cell, an
oblique co-ordinate system can be introduced, having as bases just the vectors
determined by the three edges and as origin their common vertex; in this system,
the co-ordinates of this nodes are represented by triplets of integers. We say that
the geometrical structure of a crystal is characterized by a simple Bravais net if
all the component particles of the crystal are mutually identical and their totality
forms the nodes of a regular net generated by a unit cell, having particles as its
nodes. By superposition of a finite number of simple Bravais nets, we obtain a
complex Bravais net. Real crystals are formed by particles of different types, hence
their geometrical structure is characterized by a complex Bravais net; every
component simple net is constituted by particles of a certain type. With these
specifications, we can attack one of the fundamental problems of crystallography:
the determination of all classes of symmetry of real crystals.

14.1.2.3 The Concept of Crystalline Symmetry

Let E3 be a three-dimensional Euclidean space, Ox1 x2 x3 be a Cartesian orthogonal


system in E3 ; xi ; i ¼ 1; 2; 3, be the co-ordinates, and r be the position vector of an
arbitrary point in E3 . A one-to-one mapping of E3 into E3 represents an orthogonal
transformation or a motion in E3 if and only if it preserves the distances between
the points of E3 . Any motion can be represented by the relation

r0 ¼ w þ Qr; QQT ¼ QT Q ¼ 1; ð14:84Þ


corresponding to the complete orthogonal group; we have denoted by r and r0 the
position vectors of an arbitrary point of E3 before and after the motion being
performed, respectively, by w a given vector and by Q a given orthogonal tensor,
the orthogonality condition being expressed the second relation (14.84). If the
origin O is fixed, w is uniquely determined through the motion considered, which
characterizes a translation of the points of the Euclidean space E3 . The orthogonal
tensor Q is also uniquely determined by the given motion and characterizes the
Euclidean motions which are not translations. The components of the vector w and
14.1 Anisotropic Elastic Bodies 641

of the tensor Q depend obviously on the co-ordinate system through which the
point transformation (14.84) is specified; in what follows, we shall choose co-
ordinate transformations such that these components have the simplest possible
form.
Symmetry is a special regularity observable in the spatial arrangement of objects
or of parts of objects. Any geometrical body is formed of systems of points. We
say that two geometrical figures (bodies) are mutually symmetric or congruent if
there exists a motion through which one of the geometrical figures can coincide
point by point with the other one. If a figure is formed of two congruent parts, we
say that it is self-symmetric.
The motions that realize the coincidence of two mutually symmetric figures or
of two parts of a self-symmetric figure are called operations of symmetry, corre-
sponding to the two geometrical figures and to the self-symmetric figure,
respectively, and the geometrical elements (points, lines, planes a.s.o.) charac-
teristic for such operations are called elements of symmetry. These motions are
defined by the relations (14.84).
Further, we shall deal only with those operations and elements of symmetry
which can occur with finite figures. The terminology and the notations used are
specific for crystallography.

(i) Reflection relative to a plane P. We suppose that the plane P contains the Ox3 -
axis and its intersection with the Ox1 x2 -plane forms the angle a with the Ox1 -
axis. We denote by P the orthogonal tensor which characterizes a reflection
relative to P. In the chosen co-ordinate system, the motion corresponding to a
reflection relative to P is described by the relations
2 3
cos 2a sin 2a 0 x01 ¼ x1 cos 2a þ x2 sin 2a;
0 4 5
r ¼ Pr; P  sin 2a  cos 2a 0 ; x02 ¼ x1 sin 2a  x2 cos 2a; ð14:85Þ
0 0 1 x03 ¼ x3 :

The element of symmetry in this motion is the plane P, called the plane of
symmetry.
(ii) Rotation about an axis L. Sometimes, by rotating a figure through a certain
angle about an axis L, it can come into a congruent figure (position). In this
case, L is called the axis of symmetry and it represents an element of sym-
metry in the considered motion. The angle of rotation must be an exact
divisor of 360 . Sometimes, the congruence can be obtained by rotating a
figure through different angles; the smallest of them, denoted by a, is called
the elementary angle of rotation. The number
360
n¼ ð14:86Þ
a
represents the order of the axis L, which we denoted further by Ln and will be
called the symmetry axis of nth order or n-fold axis of symmetry.
642 14 Anisotropic and Non-homogeneous Bodies

Let us suppose that Ln coincides with the Ox3 -axis and let us denote by Ln the
orthogonal tensor which characterizes the rotation about Ln . In this case, the
motion corresponding to the considered rotation is described by the relation
2 3
cos a  sin a 0 x01 ¼ x1 cos a  x2 sin a;
r ¼ L r; L  4 sin a cos a 0 5; x02 ¼ x1 sin a þ x2 cos a;
0 n n
ð14:87Þ
0 0 1 x03 ¼ x3 :

Obviously, any axis of symmetry of order n is also an axis of symmetry of


order n=k, where k 6¼ 0 is an integer.
(iii) Inversion relative to a point C. The inversion relative to a point is an
operation similar to that of reflection relative to a plane; this time, the
reflection is realized relative to a point C, which is an element of symmetry
in the considered motion and is called centre of symmetry. We suppose that
C coincides with the origin of the co-ordinate axes and we denote by C the
orthogonal tensor which characterizes the inversion relative to C. In this
case, the motion corresponding to the considered inversion is described
through the relations
2 3 0
1 0 0 x1 ¼ x1 ;
r0 ¼ Cr; C  4 0 1 0 5; x02 ¼ x2 ; ð14:88Þ
0 0 1 x03 ¼ x3 :
(iv) Roto-reflection. In this operation of symmetry, the congruent position of a
figure is realized through the simultaneous application of a rotation about an
Ln -axis and of a reflection relative to a plane P, perpendicular to Ln . The axis
Ln is called the axis of roto-reflection; it represents an element of symmetry.
We suppose that Ln coincides with the Ox3 -axis and the plane P to the Ox1 x2 -
plane; we denote by Ln2n the orthogonal tensor which characterizes the roto-
reflection. Then the motion corresponding to the considered operation of
symmetry is described by the relations
2 3 0
cos a  sin a 0 x1 ¼ x1 cos a  x2 sin a;
r0 ¼ Ln2n r; Ln2n  4 sin a cos a 0 5; x02 ¼ x1 sin a þ x2 cos a; ð14:89Þ
0 0 1 x03 ¼ x3 :

We stress the fact that, in general, a roto-reflection does not reduce to a rotation
with a separate reflection.
Applying a roto-reflection two fold, we obtain a motion which is described by
the relations 2 3
cos 2a  sin 2a 0
r0 ¼ Ln2n Ln2n r; Ln2n Ln2n ¼ 4 sin 2a cos 2a 0 5; ð14:90Þ
0 0 1
14.1 Anisotropic Elastic Bodies 643

what shows, on the basis of the formulae (14.86), (14.87) that

Ln2n Ln2n ¼ Ln=2 ; ð14:91Þ

from which we may state that a roto-reflection of an even order n is at the


same time a rotation axis of order n=2.
This result provides, at the same time, the justification for the rotation used
for the roto-reflection; the lower index characterizes the roto-reflection,
whereas the upper index characterizes the rotation.
(v) Roto-inversion. This operation of symmetry is somewhat similar to the pre-
vious one, but this time the rotation about an axis Ln is applied simultaneously
with an inversion relative to a point C situated on Ln . The axis Ln is called the
axis of inversion and it represents an element of symmetry. We suppose that
Ln coincides with the Ox3 -axis and that C coincides with the origin of the co-
ordinates system; we note by Lni the orthogonal tensor which characterizes the
roto-inversion. In this case, the motion corresponding to the considered
operation of symmetry is described by the relation
2 3
 cos a sin a 0
r0 ¼ Lni r; Lni ¼ 4  sin a  cos a 0 5: ð14:910 Þ
0 0 1

We stress that, in general, a roto-inversion does not reduce to the composition of a


rotation with an inversion.

14.1.2.4 The Fundamental Theory of Crystallography

We return to a further discussion of crystals. We specify, first, that the translations


cannot be operations of symmetry for real, finite crystals. It follows from this that
no combination of the five types of specified motions, which reduce to translations
of geometrical figures, can be present among the admissible symmetry operations
either. From this remark it follows, on the basis of certain elementary geometrical
considerations, that a real crystalline structure can possess several axes of sym-
metry only if these axes have a common point; it cannot have two parallel planes
of symmetry and it can possess at most one centre of symmetry.
While geometrical figures can possess symmetry axes of any order, the crystals,
owing to their discrete, reticular structure, can possess only certain axes of sym-
metry, of definite orders, specified by Theorem 14.1.
Theorem 14.1 In the case of crystals, only L1 ; L2 ; L3 ; L4 and L6 can be axes of
symmetry.
Let us suppose that L5 is an axis of symmetry for a crystal and let N1 be the
node closest to L5 . Because L5 is an axis of symmetry, there must exist a plane net
644 14 Anisotropic and Non-homogeneous Bodies

perpendicular on L5 , containing the node N1 . It also follows that the vertices


N2 ; N3 ; N1 ; N5 of the regular pentagon (Fig. 14.1) must be lattice nodes too.
Joining the nodes N1 and N3 , we obtain a line parallel to the side N4 N5 of the
pentagon. However, the rows situated in a plane net either are parallel or intersect
at points which are nodes of the space lattice. Therefore, the parallel drawn
through N4 to the lattice row N1 N5 is also a lattice row like the line N1 N3 ;
consequently, N, the intersection point of these two lines, must be a node of the
lattice. But this contradicts the initial assumption that N1 is the nearest node to the
axis L5 , which shows that the existence of an axis of symmetry of fifth order is not
possible for a crystal.
We can make sure of the validity of this assertion also in another way: there
must always exist a plane not perpendicular to the symmetry axis, so that the
‘‘loops’’ formed by the nodes contained in this plane have a symmetry corre-
sponding to the order of the axis. Hence, if L5 is a symmetry axis, these ‘‘loops’’
must be regular pentagons. However, a plane cannot be continuously cowered with
such figures, because the angle between two contiguous sides of a regular pentagon
is 108 , which is not a divisor of 360 ; hence, L5 cannot be a symmetry axis for a
crystal.
The same result is obtained for n ¼ 7; 8; . . .; this proves the theorem.
Consequently, in a crystal there can exist only the roto-reflection axes and the
inversion axes, respectively, characterized by the following orthogonal tensors

L12 ¼ P; L24 :L36 :L1i ¼ C; L2i ¼ P; L3i ; L4i ; L6i : ð14:92Þ

We leave to the reader the proof of the fact that L3i may be expressed through L3
and C; hence, among the inversion axes, only L4i and L6i have a proper meaning.

14.1.2.5 Combinations of Symmetry Operations

We shall state hereafter some theorems, necessary for deducing all possible
combinations of symmetry elements for crystals. Thus, we state

Fig. 14.1 Fundamental N5 N4


theorem of crystallography;
impossible case

L5

N1 N3
N

N2
14.1 Anisotropic Elastic Bodies 645

Theorem 14.2 If a geometrical figure has two or more symmetry planes P, then
their lines of intersection are symmetry axes L. The elementary angle of rotation
corresponding to such an axis is twice the angle formed by the two planes which
determine this axis.
Let P1 and P2 be two planes of symmetry. The co-ordinate axes can be always
be chosen in such a way that the line of intersection of the two planes coincide
with the Ox3 -axis and P1 coincide with the Ox1 x3 -plane. We denote by a the angle
formed by the Ox1 -axis with the intersection of the Ox1 x2 -plane and P2 . In this
case
2 3 2 3
1 0 0 cos 2a sin 2a 0
P1  4 0 1 0 5; P2  4 sin 2a  cos 2a 0 5;
0 0 12 0 30 1
ð14:93Þ
cos 2a  sin 2a 0
P1 P2 ¼ 4 sin 2a  cos 2a 0 5:
0 0 1

The comparison of the last equality (14.33) with the relation (14.27) shows that the
successive application of reflections relative to the planes P1 and P2 , respectively,
results in a rotation about the Ox3 -axis, through the angle 2a, which represents
twice the angle formed by the too planes. Since Ox3 is the line of intersection of
the two planes of symmetry, the theorem is proved.
There result two important consequences:
(i) The action of a symmetry axis can be replaced by the combined action of two
symmetry planes, intersecting along it.
(ii) If a figure has a symmetry plane P1 with a symmetry axis Ln contained in it,
then this figure must also have a second plane of symmetry P2 , intersecting P1
along the axis Ln , at an angle determined by the order n of the axis.
Let us state now
Theorem 14.3 (Euler) If a geometrical figure has two concurrent axes of sym-
metry Ln and Lm , with elementary angles of rotation a and b, respectively, then it
must also have a resultant axis of symmetry Lp .
According to the consequence (i), we can replace the axis Ln by two symmetry
planes P1 and P2 , which intersect along Ln , the angle between them being a=2. We
choose the plane P1 such that it passes through Lm . Analogically, we replace the
axis Lm by two planes of symmetry P3 and P4 , intersecting along Lm at an angle
b=2, such that P3 passes through Ln . The combined action of the planes P1 and P3
reduces to the identical transformation. Thus, the combined action of the axes Ln
and Lm is equivalent to two successive reflections through the symmetry planes P2
and P4 . But, according to the Theorem 14.2, the final result of these two motions
corresponding to a rotation through a known angle about the symmetry axis Lr ,
which coincides with the line of intersection of the two planes. The theorem is thus
proved.
646 14 Anisotropic and Non-homogeneous Bodies

We remark that, without changing the order of the axes Ln and Lm or the angle
between them, their replacement with couples of symmetry planes can be realized
in two different ways, which shows that there also exists a second resultant
0
symmetry axis Ln . Hence, Lm and Ln have two resultant axes of symmetry,
arranged simmetrically relative to the plane formed by Lm and Ln . We leave to the
reader the task of verifying this. We state now in Theorem 14.4.
Theorem 14.4 The point at which a symmetry axis of an even order pierces a
symmetry plane which is normal to it is a centre of symmetry.
We suppose that the symmetry axis L2n coincides with the Ox3 -axis and the
symmetry plane P coincides with the Ox1 x2 -plane. From (14.85)–(14.87) it results
that
2 3 2 3
cos 180 =n  sin 180 =n 0 1 0 0
L2n  4 sin 180 =n cos 180 =n 0 5; P  4 0 1 0 5 ð14:94Þ
0 0 1 0 0 1

and from (14.94), we obtain


2 3 2 3

2n n 1 0 0

1 0 0
n n
L ¼ 4 0 1 0 5; L2n P ¼ P L2n ¼ 4 0 1 0 5: ð14:95Þ
0 0 1 0 0 1

The last relation (14.95), compared with (14.88), shows that the origin of the co-
ordinates, hence the point of intersection of the axis L2n with the plane P, is indeed
a centre of symmetry. We state hereafter
Theorem 14.5 A centre of symmetry C is equivalent to an infinite number of
twofold roto-reflection axes, i.e., C ¼ 1L12 .
We suppose that the centre of symmetry C coincides with the origin of the
co-ordinate axes. From (14.86), (14.89) it follows that
2 3
1 0 0
L12 ¼ C ¼ 4 0 1 0 5: ð14:96Þ
0 0 1

Hence, the Ox3 -axis is a roto-reflection axis of second order. However, since the
relations (14.96) have the same structure in any orthogonal Cartesian co-ordinate
system, it follows that any line which passes through the centre of symmetry is
also a twofold roto-reflection axis. We state now
Theorem 14.6 The existence of a roto-reflection axis Ln2n , where n is an odd
number, always implies the existence of a centre of symmetry situated on this axis.
We suppose that the roto-reflection axis L2pþ1
2ð2pþ1Þ coincides with the Ox3 -axis.
From (14.86) and (14.89) it follows that
14.1 Anisotropic Elastic Bodies 647

2 3
cos 180 =ð2p þ 1Þ  sin 180 =ð2p þ 1Þ 0
6 7
L2pþ1 
2ð2pþ1Þ  4 sin 180 =ð2p þ 1Þ cos 180 =ð2p þ 1Þ 0 5;
0 0 1
2 3 ð14:97Þ
 2pþ1 1 0 0
6 7
L2pþ1
2ð2pþ1Þ ¼ 4 0 1 0 5:
0 0 1

The last relation shows that the origin of the co-ordinate axes is a centre of
symmetry. We may also state
Theorem 14.7 If a roto-reflection axis Ln2n , where n is an even number, is
combined with a twofold symmetry axis L2 , normal to L2n , then we get n symmetry
planes passing through Ln2n .
We suppose that the axis L2n ¼ L4p coincides with the Ox3 -axis and L2 coin-
cides with the Ox1 -axis; from (14.86), (14.87) and (14.89) it follows that
2 3 2 3
cos 90 =p  sin 90 =p 0 1 0 0
L2p 4 sin 90 =p  cos 90 =p 0 5; L2  4 0 1 0 5;
4p  ð14:98Þ
0 0 1 0 0 1

from which we obtain


2 3
cosð90 =pÞ  sinð90 =pÞ 0
L2 L2p
4p ¼ 4 sinð90 =pÞ  cosð90 =pÞ 0 5: ð14:99Þ
0 0 1

The last relation, compared with (14.85), shows that there exists a symmetry plane
which passes through the Ox3 -axis and the intersection of which with the Ox1 x2 -
plane forms an angle of 45 =p with the Ox1 -axis. Determining successively the
components of the orthogonal tensors ðL2p 2 2p 4p
4p Þ ; . . .; ðL4p Þ and combining the
2
results thus obtained with those relative to L , the reader can himself verify the fact
that there exist still n  1 symmetry planes which pass through the Ox3 -axis.

14.1.2.6 Derivation of all Possible Classes of Crystals

A crystal class is a combination of symmetry operations, in which all the resultant


elements of symmetry have been derived. For symmetries and elements of sym-
metry we use Flint’s notation.
From the results presented in previous subsections it follows that, in finite
geometrical figures specific for crystals, there can exist the following fundamental
elements of symmetry:
(i) the plane of symmetry P,
648 14 Anisotropic and Non-homogeneous Bodies

(ii) the symmetry axes L1 ; L2 ; L3 ; L4 and L6 ,


(iii) the roto-reflection axes L12 ¼ C; L24 ; L36 ,
(iv) the centre of symmetry C.
We shall divide the totality of possible combinations of these elements of
symmetry into two subsets A and B, where A includes the crystal classes in which,
after the addition of symmetry elements, there exists only one symmetry axis with
n greater than 2, whereas B includes the crystal classes corresponding to those
possible combinations which can contain several Ln axes of symmetry with orders
greater than 2. We shall deal, first, with the subset A, exhausting in several stages,
all possible cases.

(a) We assume that the entire symmetry of the figure is characterized by a single
symmetry axis Ln ; in this case, the following symmetries are possible in
crystals:

L1 ; L2 ; L3 ; L4 ; L6 : ð14:100Þ
(b) It is easy to see that, when the figure possesses a symmetry axis Ln , with
n [ 2, then, in order to avoid the occurrence of a new axis of order greater
than 2, only an L2 -axis that is normal to Ln can be added to Ln ; the addition of
an axis of symmetry of second order, that is not normal to Ln , would result in
the displacement of the Ln -axis in a new position, which is equivalent to the
existence of a new axis of symmetry of the same order n.
At the same time, we can see, on the basis of the Theorem 14.2 and on
Euler’s theorem concerning the composition of rotations, that the addition of
the L2 -axis to the Ln -axis, when L2 and Ln are mutually perpendicular, will
necessarily result in the occurrence of other n  1 axes of symmetry of
second order, perpendicular to Ln . This new type of symmetry will be denoted
by Ln nL2 . We shall therefore have the following symmetries:

L1 L2 ¼ L2 ; L2 2L2 ¼ 3L2 ; L3 3L2 ; L4 4L2 ; L6 6L2 : ð14:101Þ

On the basis of similar reasoning, we can conclude that for the crystal classes of
subset A, a symmetry plane can be added to Ln either perpendicularly or parallel to
Ln , hence passing through this axis.

(c) In the first case, according to the Theorem 14.4, the point of intersection of a
symmetry axis of an even order with the symmetry plane normal to it will
necessarily be a centre of symmetry. The new types of possible symmetries
are denoted by Ln P and Ln PC; their complete list looks as follows:

L1 P ¼ P; L2 PC; L3 P; L4 PC; L6 PC: ð14:102Þ


(d) In the second case, according to the Theorem 14.2, consequence (2), after the
addition of a symmetry plane containing Ln , there will necessarily occur
14.1 Anisotropic Elastic Bodies 649

another symmetry plane, which also contains the axis Ln and forms an angle
of a=2ða ¼ 360 =nÞ with the first plane. Repeating this reasoning, we can
conclude that n  1 such additional symmetry planes, passing through Ln , can
occur, with angles a=2 between each other. The new types of symmetry are
denoted by Ln nP; their complete list is:

L1 P ¼ P; L2 2P; L3 3P; L4 4P; L6 6P: ð14:103Þ


(e) Obviously, the possibility of adding simultaneously a symmetry plane normal
to Ln and another one parallel to Ln cannot be excluded. The types of sym-
metry possible for this case are obtained by combining the results (c) and (d);
taking also into account the fact that the addition of two symmetry planes,
with an angle of 90 between them, to the Ln -axis will result in the occurrence
of a twofold axis of symmetry, which coincides with the intersection line of
the two planes and hence is normal to L5 . The number of these axes of L2 type
will be equal to the number of symmetry planes which pass through Ln and
hence to n according to the result of (d). The types of symmetry obtained in
this way will be denoted by combining the notations used in (c) and (d); their
complete list is:

L1 L2 2P; 3L2 3PC; L3 3L2 4P; L4 4L2 5PC; L6 6L2 7PC: ð14:104Þ

To this point, we have exhausted all combinations of symmetry axes and


symmetry planes for the subset A. We shall now deal with roto-reflection axes.

(f) We suppose, first, that the entire symmetry of the figure is characterized by a
single roto-reflection axis Ln2n . According to the Theorem 14.6, if n is odd,
there necessarily occurs also a centre of symmetry C, situated on the axis Ln2n .
Thus, the complete list of possible symmetries is the following:

L12 ¼ C; L24 ; L36 C: ð14:105Þ

(g) Combining a roto-reflection axis Ln2n with a symmetry axis of order 2L2 ,
normal to Ln2n (see the reasoning of (b)), we necessarily obtain, according to
the Theorem 14.7, when n is even, n additional symmetry planes which pass
through Ln2n . Taking into account the Theorem 14.6 as well, we obtain the
following complete list of the symmetries possible in this case:

L12 L2 P ¼ L2 PC; L4 2L2 2P; L36 3L2 3PC: ð14:106Þ

According to the series (14.100)–(14.106), we have obtained, for the subset A,


31 types of possible symmetry; but it is easy to see that some of them are repeated.
These ones are marked with an asterisk in Table 14.1, from which we can see
immediately that the subset A contains, on the whole, 27 possible crystal classes.
650 14 Anisotropic and Non-homogeneous Bodies

Before attacking the general case, it is convenient to analyse a particular case of


the subset B. We suppose that the figure admits two symmetry axes L3 and L4 ,
which intersect at an arbitrary angle. Since L3 is of third order, it follows that,
besides L4 the figure admits two more symmetry axes of fourth order, which are
obtained from L4 by rotating it through 120 and 240 , respectively, about L3 .
Analogically, we conclude that the figure possesses necessarily three more sym-
metry axes of third order, obtained from L3 , by rotating it through angles of
90 ; 180 and 270 , respectively, about L4 . Thus, from the initial existence of the
two symmetry axes, it follows that the figure must have at least seven axes of
symmetry. However, the above considerations apply for any new symmetry axis;
therefore, in general, the number of the new additional axes of symmetry can
increase indefinitely.
In the case of crystals, characterized by a discrete structure, only a finite number
of symmetry axes can exist. We have thus to establish whether the occurrence of
new axes of symmetry can be stopped after a finite number of steps, and, if this is
so, to see in what conditions and how this happens. The first part of the answer can
be obviously formulated in the following way: either the number of axes will grow
indefinitely and shall arrive at a spherical symmetry of the type 1L1 PC or the
new axes and those which result from their composition coincide with the existing
ones and thus will be finite in number. As we have outlined, for crystals only the
second situation is admissible; therefore, we should establish in what conditions it
can occur.
To answer this question, we suppose that the problem has already been solved;
hence, starting from two symmetry axes of higher order, we have obtained a finite
number of symmetry axes, intersecting at the common point of the two initial axes.
In this situation, we have to establish the number, the order and the spatial
arrangement of these additional axes.
To this end, we describe a sphere R of unit radius around the common point of
the symmetry axes and we take into account the consequence (1) of the Theorem
14.2, according to which every axis can be considered as being the line of inter-
section of two planes of symmetry, the angle between the planes being half of the
elementary angle of rotation a, corresponding to the axis. Replacing the axes by
their corresponding planes, we reach the conclusion that R is divided into spherical
triangles determined by the intersections of these planes with R; at the points at
which the axes intersect the sphere, the symmetry planes corresponding to an axis,
hence also the sides of the spherical triangles determined by these planes, intersect
at an angle a=2. For crystals, these angles can be only 30 ; 45 ; 60 and 90 .
It is a well-known fact that the sum S of the angles of a spherical triangle must
satisfy the inequalities
180 \ S \ 3  180 ; ð14:107Þ
a fact which allows us to determine the admissible triangles for crystals. Since we
deal with the subset B, any admissible spherical triangle must have at least two
vertices corresponding to two axes of orders greater than 2; because S must satisfy
14.1 Anisotropic Elastic Bodies 651

Table 14.1 Types of geometrical symmetries


Series General formula Possible symmetries in the case A
n¼1 n¼2 n¼3 n¼4 n¼6
(a) Ln L1 L2 L3 L4 L6
n 2 * 1 2 2 2 2 2 3 2 4 2
(b) L nL L L ¼L L 2L ¼ 3L L 3L L 4L L6 6L2
(c) Ln P? ðCÞ 1
L P¼P 2
L PC 3
L P L PC 4
L6 PC
(d) Ln Pk * 1
L P¼P 2
L 2P 3
L 3P L 4P 4
L6 6P
n 2 * 1 2 2 2 2 2 3 2 4 2
(e) L nL ðn þ 1ÞPðCÞ L L 2P ¼ L 2P L 2L 3PC ¼ 3L 3PC L 3L 4P L 4L 5PC L6 6L2 7PC
(f) Ln2n ðCÞ L12 ¼ C L24 L36 C – –
*
(g) Ln2n nL2 nPðCÞ L2 PC L24 2L2 2P L36 2L2 3PC – –

the relations (14.107), the third vertex can possibly correspond to a symmetry axis
of second order, as a possible resultant axis. We give the list of all combinations
which are not excluded by the above considerations:

ðL6 ; L6 ; L2 Þ; S ¼ 150 ; ðL4 ; L4 ; L2 Þ; S ¼ 180 ; ðL6 ; L4 ; L2 Þ; S ¼ 165 ;


ð14:108Þ
ðL4 ; L3 ; L2 Þ; S ¼ 195 ; ðL6 ; L3 ; L2 Þ; S ¼ 180 ; ðL3 ; L3 ; L2 Þ; S ¼ 120 :

From (14.107) and (14.108) it follows that only the combinations ðL4 ; L3 ; L2 Þ
and ðL3 ; L3 ; L2 Þ are admissible. Knowing the angles of the two types of spherical
triangles corresponding to these combinations and using the formulae of spherical
trigonometry, we can compute the angles corresponding to the sides of the tri-
angles; we obtain:

L4 L3 = 54o 44′ 08″ , L4 L2 = 45o , L3 L2 = 35o15′ 32″ ,


ð14:109Þ
L3 L3 = 70o 31′ 44″ , L3 L2 = 54o 44′ 08″.

If the surface of the sphere R can be covered continuously with such triangles,
then their number and mutual arrangement determine the possible symmetries in
the case B. Using also Gauss’s formula
S ¼ p þ A; ð14:110Þ
which relates the sum S of the angles of a spherical triangle to the area A, it can be
shown that this covering is realizable and, in the first case, only the combination
3L4 4L3 6L2 can exist, whereas the second case permits only the combination
3L4 4L3 , the mutual position of the axes of symmetry being determined by the
relations (14.109).
Applying repeatedly the consequence (ii) of the Theorem 14.2, we conclude
that the addition of a plane of symmetry to the combination 3L4 4L3 6L2 , such that
no new axes of symmetry occur, will result in the occurrence of 8 additional planes
of symmetry, the positions of which can be determined from the fact that the plane
652 14 Anisotropic and Non-homogeneous Bodies

added initially must be normal to one of the L4 -axes. At the same time, the
Theorem 14.5 shows that the common point of the 13 symmetry axes is, in this
case, a centre of symmetry; therefore, the resultant combination has the structure
3L4 4L3 6L2 9PC.
Since the occurrence of the axes of symmetry is not permitted, the addition of a
plane of symmetry cannot be done arbitrarily in the second case, either. We leave
to the reader to verify that a plane of symmetry can be added to the combinations
3L2 4L3 only in two ways:
(i) the plane passes through the axes L3 and L4 ;
(ii) the plane passes through the L2 -axis, symmetrically relative to the L3 -axis.
In (i) results a symmetry of the type 3L2 4L3 6P, whereas in (ii) the symmetry of
the type 3L2 4L3 3PC. Hence, in the subset B there exist only five crystal classes

3L4 4L3 6L2 ; 3L4 4L3 6L2 9PC; 3L2 4L3 ; 3L2 4L3 6P; 3L2 4L3 3PC: ð14:111Þ
In the first and the fifth classes, the threefold symmetry axes are also sixfold
roto-reflection axes, and in the fourth class, the twofold symmetry axes become
roto-reflection axes of the fourth order.
Consequently, the totality of crystal classes can contain only 32 distinct ele-
ments; the Table 14.1 together with (14.111), exhaust all possible types of geo-
metrical symmetries for the external form of real crystals.

14.2 Non-homogeneous Elastic Bodies

Hereafter we treat problems of non homogeneous elastic bodies, considering both


continuous non-homogeneity and piecewise continuous non-homogeneity. We start
from some initial considerations given in Sect. 4.1.3.8.

14.2.1 Three-Dimensional Problems

We give first, in what follows, some general results concerning the static problem.
The dynamics case will be emphasized too.

14.2.1.1 General Results

In the case of continuous non-homogeneous elastic bodies, the elastic bodies


become functions of point; thus, E ¼ Eðx1 ; x2 ; x3 Þ; m ¼ mðx1 ; x2 ; x3 Þ and, conse-
quently, k ¼ kðx1 ; x2 ; x3 Þ; l ¼ lðx1 ; x2 ; x3 Þ; as well, q ¼ qðx1 ; x2 ; x3 Þ. Because
14.2 Non-homogeneous Elastic Bodies 653

Poisson’s ratio m has, in general, small variations, one may take m ¼ const,
remaining with only one elastic constant as function of point (see [34, 35]).
We remember that
1  i ¼ 1 Fi ; i; j ¼ 1; 2; 3;
ij ¼
r rij ; F ð14:112Þ
E E
are the reduced stresses and the reduced volume forces, respectively. Hooke’s law
becomes
eij ¼ ð1 þ mÞ
rij  m
rkk dij ; i; j ¼ 1; 2; 3; ð14:113Þ

being a constitutive law with constant coefficients, in the isotropic case. We


confine within the isotropic case, although the previous idea may be extended to
anisotropic bodies.
It may be convenient to represent the longitudinal modulus of elasticity in the
form

E ¼ E0 e f ; ð14:114Þ

where f ¼ f ðx1 ; x2 ; x3 Þ is a function continuous and differentiable whatever nec-


essary (in general, it is sufficient to be of class C4 ); E0 corresponds to a constant
rigidity and, in general, E ¼ const on a surface f ðx1 ; x2 ; x3 Þ ¼ const.
In particular, if the function f is of the form
f ðx1 ; x2 ; x3 Þ ¼ a  r þ const ¼ ai xi þ const; i ¼ 1; 2; 3; a ¼ const; ð14:115Þ
then we have to do with a constant rigidity in parallel planes; the equations of the
problem are with constant coefficients.
In the static case, the equations of equilibrium take the form

ij;i þ f;i r
r  j ¼ 0; j ¼ 1; 2; 3;
ij þ F ð14:116Þ

and Lamé’s equations read


1
uj;ii þ ðui;ij þ 2mf;j ui;i Þ þ f;i ðui;j þ uj;i Þ
1  2m
þ 2ð1 þ mÞF j ¼ 0; j ¼ 1; 2; 3: ð14:117Þ

In a vector form, these equations become (in the absence of the volume forces and
taking into account (14.115))
1
ðD þ a  rÞu þ ½rðr  uÞ þ 2maðr  uÞ þ rða  uÞ ¼ 0: ð14:1170 Þ
1  2m
Using an analogue C ¼ CðC1 ; C2 ; C3 Þ, of Galerkin’s vector, which verifies the
equation
n m h 2 io
ðD þ a  rÞ ðD þ a  rÞ2  a D  ða  rÞ2 C ¼ 0; ð14:118Þ
1m
654 14 Anisotropic and Non-homogeneous Bodies

one can express the displacement vector u in the form


1
u¼ fðD þ a  rÞ½2ð1  mÞðD þ a  rÞC  rðr  CÞ
1  2m
ð1  2mÞðr  CÞa  2mrða  CÞ þ 2mr  f½a  ðr  CÞagg: ð14:119Þ

By means of Boggio’s theorem (see Sect. A.1.2.7), one can express the
potential vector C in the form
C ¼ C1 þ C2 ; ð14:120Þ

where C1 ¼ C1 ðx1 ; x2 ; x3 Þ verifies the equation


ðD þ a  rÞC1 ¼ 0 ð14:1200 Þ
and C2 ¼ C2 ðx1 ; x2 ; x3 Þ verifies the equation
m
ðD þ a  rÞ2 C2 ¼ ½a2 D  ða  rÞ2 C2 : ð14:12000 Þ
1m
This observation may be reflected in the representation (14.119) too.
In the dynamic case (see [33]), the equations of motion read

ij;i þ f;i r
r j ¼ q
ij þ F €uj ; j ¼ 1; 2; 3; ð14:121Þ

where
1

q q ð14:1160 Þ
E
is the reduced unit mass (the reduced density). The equations of Lamé in the
dynamic case take the form (in the absence of the volume forces)
1
uj;ii þ ðui;ij þ 2mf;j ui;i Þ þ f;i ðui;j þ uj;i Þ
1  2m
þ 2ð1 þ mÞðF j  quvj Þ ¼ 0: ð14:1170 Þ

Also in this case, one can give a representation of Somigliana-Iacovache type too.

14.2.1.2 Approximate Methods of Calculation

Hereafter, we shall indicate a method of successive approximations, valid in the


general case of non-homogeneity. In the case of the second fundamental problem
(in stresses), we introduce the reduced superficial loading
n 1 n

 p; i ¼ 1; 2; 3; ð14:122Þ
i E i
so that the boundary conditions take the form
14.2 Non-homogeneous Elastic Bodies 655

n n
pi ¼ r
  ij nj ; i ¼ 1; 2; 3: ð14:123Þ
We express the stresses in the form
X
n
ðkÞ
r 0ij þ
ij ¼ r r
ij : ð14:124Þ
k¼1

Neglecting the terms with variable coefficients (if a  r 6¼ const), the equations in
stresses take a form analogue to that which corresponds to homogeneous bodies,
one obtains an approximation of zero order, solving the problem for an homo-
geneous body acted upon by given volume forces and with the boundary condi-
tions (14.123).
To obtain the approximation of k-th order, assuming that one knows the
approximation of (k  1)-th order, which leads to the conventional volume forces

 jðkÞ ¼ f;i r
F
ðk1Þ
ij ; j ¼ 1; 2; 3; ð14:125Þ

one solves the problem for a homogeneous body, acted upon by the volume forces
 jðkÞ , with zero loads on the frontier.
F
The solution in stresses of the static elastic problem of a non-homogeneous
body may be thus reduced to the solution in stresses of a succession of elastic
problems for the same body, considered as homogeneous, acted upon by con-
ventional volume forces and (except the approximation of zero-th order) with
vanishing boundary conditions.
The convergence of the iterative process may be practically appreciated by a
mechanical interpretation (the conventional volume forces F  iðkÞ must give a state of
stress negligible with respect to the state of stress given by F  iðk1Þ ).
In the case of the first fundamental problem (in displacements), one may use an
analogue procedure, using the equations of Lamé type (14.117).
The corresponding dynamic problem can be treated similarly. Obviously, one
must also put initial conditions for the time, i.e.:
n n
t t0 : 
pi ¼ r
 ji nj ; i ¼ 1; 2; 3; ð14:126Þ

or
n
t t0 : ui ¼ u i ; i ¼ 1; 2; 3; ð14:1260 Þ

as well as

t ¼ t0 : ui ¼ u0i ðx1 ; x2 ; x3 Þ; u_ i ¼ u_ 0i ðx1 ; x2 ; x3 Þ: ð14:127Þ


656 14 Anisotropic and Non-homogeneous Bodies

14.2.2 Two-Dimensional Problems

Hereafter we deal with the two-dimensional problems of non-homogeneous bod-


ies, i.e., both the plane state of strain and the plane state of stress. After the
formulations in displacements and in stresses, we present an application for the
elastic half-plane.

14.2.2.1 General Results

One assumes that in the case of a plane state of strain the longitudinal modulus of
elasticity does not depend on x3 , i.e., E ¼ Eðx1 ; x2 Þ. In the case of a plane state of
stress one uses an analogous hypothesis, but with a mean elasticity moduli on the
thickness h of a plate, i.e.,
Z Z
1 h=2 1 h=2
Eðx1 ; x2 Þ ¼ Eðx3 Þdx3 ; Gðx1 ; x2 Þ ¼ Gðx3 Þdx3 ; ð14:128Þ
h h=2 h h=2

defining thus a generalized plane state of stress.


We shall assume to be in the case of a plane state of stress; in the case of a plane
state of strain, one uses the same equations, introducing the generalized elastic
coefficients
E m
E0 ¼ ; m0 ¼ ; ð14:1280 Þ
1  m2 1m
which depend on the co-ordinates of the point too.
Introducing the reduced quantities (14.112), one uses the Hooke law
e11 ¼ r
11  m
r22 ; e22 ¼ r
22  m
r11 ; ð14:129Þ

c12 ¼ 2ð1 þ mÞ


r12 : ð14:1290 Þ

The equations of equilibrium take the form


1

11;1 þ r
r 12;2 þ 11 þ E;2 r
E;1 r 12 ¼ 0;
E ð14:130Þ
1

12;1 þ r
r 22;2 þ E;1 r 22 ¼ 0:
12 þ E;2 r
E
being with variable coefficients.
Representing the longitudinal elasticity modulus in the form

E ¼ ea1 x1 þa2 x2 þa3 ¼ Eð0; 0Þea1 x1 þa2 x2 ; a1 ; a2 ; a3 ¼ const; ð14:131Þ


where Eð0; 0Þ is the elasticity modulus at the origin of the co-ordinate axes. One
observes that the relations do not change if one multiplies the modulus E by an
arbitrary constant (i.e. if the rigidity of the body is enlarged or diminished).
14.2 Non-homogeneous Elastic Bodies 657

Eq. (14.130) become


ðo1 þ a1 Þ
r11 þ ðo2 þ a2 Þ
r12 ¼ 0; ðo1 þ a1 Þ
r12 þ ðo2 þ a2 Þ
r22 ¼ 0: ð14:1300 Þ
The equations of continuity in stresses take the form


11;22 þ r
r 22;11  2
r12;12 ¼ m r11;11 þ r
22;22 þ 2
r12;12 : ð14:132Þ
Taking into account the equations of equilibrium (14.1300 ), one may write

22 Þ ¼ ð1 þ mÞða21 r
r11 þ r
Dð 11 þ a22 r
22 þ 2a1 a2 r
12 Þ: ð14:1320 Þ

Hence, the partial differential equations of the problem are (14.1300 ) and (14.1320 ).
The Eqs. (14.1300 ) lead to

11 ¼ ðo2 þ a2 Þu; r


r 12 ¼ ðo1 þ a1 Þu;
22 ¼ ðo1 þ a1 Þw; r
r 12 ¼ ðo2 þ a2 Þw;

where u ¼ uðx1 ; x2 Þ; w ¼ wðx1 ; x2 Þ are arbitrary functions. Equating the two


expressions of r
12 , one obtains
u ¼ ðo2 þ a2 ÞF; w ¼ ðo1 þ a1 ÞF;

so that

11 ¼ ðo2 þ a2 Þ2 F; r
r 22 ¼ ðo1 þ a1 Þ2 F;
ð14:133Þ
12 ¼ ðo1 þ a1 Þðo2 þ a2 ÞF;
r

where the stress function F ¼ Fðx1 ; x2 Þ is a stress function of Airy type.


Introducing this representation in (14.1320 ), we obtain the equation
h i h
D ðo1 þ a1 Þ2 þ ðo2 þ a2 Þ2 F  ð1 þ mÞ a21 ðo2 þ a2 Þ2
i
þa22 ðo1 þ a1 Þ2  2a1 a2 ðo1 þ a1 Þðo2 þ a2 Þ F ¼ 0; ð14:134Þ

which must be satisfied by the stress function F; after elementary calculations, one
obtains
h i
D ðo1 þ a1 Þ2 þ ðo2 þ a2 Þ2 F  ð1 þ mÞða1 o2  a2 o1 Þ2 F ¼ 0 ð14:1340 Þ

or

ðD þ a1 o1 þ a2 o2 Þ2 F  mða1 o2  a2 o1 Þ2 F ¼ 0; ð14:13400 Þ

wherefrom
pffiffiffi pffiffiffi pffiffiffi
D þ ða1  a2 mÞo1 þ ða2 þ a1 mÞo2 D þ ða1 þ a2 mÞo1
pffiffiffi
þða2  a1 mÞo2 F ¼ 0; ð14:134000 Þ
658 14 Anisotropic and Non-homogeneous Bodies

using again Boggio’s theorem, one may write


Fðx1 ; x2 Þ ¼ Uðx1 ; x2 Þ þ Wðx1 ; x2 Þ; ð14:135Þ
where the new potential functions must verify the partial differential equations
pffiffiffi pffiffiffi
D þ ða1  a2 mÞo1 þ ða2 þ a1 mÞo2 U ¼ 0;
pffiffiffi pffiffiffi ð14:1350 Þ
D þ ða1 þ a2 mÞo1 þ ða2  a1 mÞo2 W ¼ 0:
To solve the boundary value problem, we put the conditions
n n
p 1 ¼ r11 n1 þ r12 n2 ; p 2 ¼ r21 n1 þ r22 n2 ð14:136Þ
on the frontier; using the reduced loading (14.122) on the frontier, one may write
these conditions in the form
n
p 1 ds ¼ ðo2 þ a2 Þ2 Fdx2 þ ðo1 þ a1 Þðo2 þ a2 ÞFdx1 ;

n
ð14:1360 Þ
2
p 2 ds ¼ ðo1 þ a1 Þðo2 þ a2 ÞFdx2  ðo1 þ a1 Þ Fdx1 :


By a change of function, one may obtain a total differential in the right member
n
p 1 ea1 x1 þa2 x2 ds ¼ d½o2 ðFea1 x1 þa2 x2 Þ;

n
p 2 ea1 x1 þa2 x2 ds ¼ d½o1 ðFea1 x1 þa2 x2 Þ;


wherefrom, by integration,
 Z sn
 p 2 ds ¼ V0  V;
ðEFÞ;1 ¼ ðEFÞ;1  
0 0
 Z sn ð14:137Þ
 p 1 ds ¼ H0 þ H;
ðEFÞ;2 ¼ ðEFÞ;2  
0 0

where s is the curvilinear co-ordinate along the frontier of the plane domain.
In the case of a simply connected domain, one can take the constants H0 and V0
equal to zero, because a stress function of the form
F ¼ ðH0 x2 þ V0 x1 þ constÞea1 x1 a2 x2 ð14:138Þ
leads to no any stress, taking into account (14.133). The quantities H and V (which
play the rôle of the functions u and w) correspond to the components along the co-
ordinate axes of the external load acting on a bar along the frontier of the con-
sidered domain, starting from an arbitrary point. One observes that
o dx1 dx2
ðEFÞ ¼ ðEFÞ;1 þ ðEFÞ;2 ;
os ds ds ð14:139Þ
o dx1 dx2
ðEFÞ ¼ ðEFÞ;1 þ ðEFÞ;2 :
on dn dn
14.2 Non-homogeneous Elastic Bodies 659

In the case of a simply connected domain, one obtains


o
ðEFÞ ¼ V cosðn; x2 Þ þ H cosðn; x1 Þ ¼ T;
os ð14:140Þ
o
ðEFÞ ¼ V cosðs; x2 Þ  H cosðs; x1 Þ ¼ N:
on
The tangential and the normal derivative of the function EF are thus the shear
force (with a changed sign) and the axial force, respectively, in a bar which is
along the frontier of the domain, counterclockwise, the domain remaining at the
left side, beginning from an arbitrary fixed point.
On observes that
dðEFÞ ¼ ðEFÞ;1 dx1 þ ðEFÞ;2 dx2 ;

integrating by parts
Z s Z s
ðHdx2  Vdx1 Þ ¼ ðHx1 Þs  ðVx1 Þs  ðx2 dH  x1 dVÞ
0 0
Z s Z s
n n
¼ p 1 (x2 js x2 Þds þ p 2 (x1  x1 js Þds;
0 0

one obtains, finally,


EF ¼ M0 þ V0 x1 þ H0 x2 þ M: ð14:1400 Þ
Neglecting a function of the form (14.138), which leads to no efforts in the bar,
one may write
1
Fðx1 ; x2 Þ ¼ Mðx1 ; x2 Þ; ð14:14000 Þ
E
where M is the bending moment in the bar mentioned above.
By changing the sense of the passing through along the bar (in the clockwise
sense), one has
o
ðEF Þ ¼ T; ð14:140000 Þ
os
the other effort remaining with the same sign.
It follows that to solve the boundary value problem one must know on the
boundary: either F and oðEFÞ=on (M and N), i.e. F and oF=on or oðEFÞ=os and
oðEFÞ=on(T and N), i.e. oF=os and oF=on.
For a solution in displacements, one uses the relations of Cauchy, obtaining the
equations of Lamé in the form
660 14 Anisotropic and Non-homogeneous Bodies

h i
1
Du1 þ 1m ð1 þ mÞðu1;1 þ u2;2 Þ;1 þ 2a1 ðu1;1 þ mu2;2 Þ
h þa2 ðu1;2 þ u2;1 Þ ¼ 0; i ð14:141Þ
1
Du2 þ 1m ð1 þ mÞðu1;1 þ u2;2 Þ;2 þ 2a2 ðu2;2 þ mu1;1 Þ
þa1 ðu1;2 þ u2;1 Þ ¼ 0;

which may be written also in the form



2
o1 ðo1 þ a1 Þ þ o2 ðo2 þ a2 Þ u1
1m

1þm
þ o2 ðo1 þ a1 Þ  ða1 o2  a2 o1 Þ u2 ¼ 0;
1m
 ð14:1410 Þ
1þm
o1 ðo2 þ a2 Þ þ ða1 o2  a2 o1 Þ u1
1m

2
þ o2 ðo2 þ a2 Þ þ o1 ðo1 þ a1 Þ u2 ¼ 0:
1m
The displacements may be now represented by means of a potential function
F ¼ Fðx1 ; x2 Þ in the form
1þm
u1 ¼  o2 ðo1 þ a1 ÞF þ ða1 o2  a2 o1 ÞF;
1m ð14:142Þ
2
u2 ¼ o1 ðo1 þ a1 ÞF þ o2 ðo2 þ a2 ÞF
1m
or in the form
2
u1 ¼ o2 ðo2 þ a2 ÞF þ o1 ðo1 þ a1 ÞF;
1m ð14:1420 Þ
1þm
u2 ¼  o1 ðo2 þ a2 ÞF þ ða2 o1  a1 o2 ÞF;
1m
we observe that the potential function F must verify the same Eq. (14.134000 ) as in
the solution in stresses. Obviously, on the frontier the displacements must be
given.
As one can see, the Eq. (14.1350 ) have the form
ðD þ c1 o1 þ c2 o2 ÞXðx1 ; x2 Þ ¼ 0; ð14:143Þ

by a change of function

X ¼ eðc1 x1 þc2 x2 Þ=2 X; ð14:144Þ
one obtains the equation

  1 ðc2 þ c2 ÞX
DX  ¼ 0; ð14:1440 Þ
4 1 2
14.2 Non-homogeneous Elastic Bodies 661

which correspond to the solving both in stresses and in displacements of the


problem.
If one takes, for the first Eq. (14.1350 ),
pffiffiffi pffiffiffi
c 1 ¼ a1  a2 m ; c 2 ¼ a2 þ a1 m ð14:145Þ

and, for the second equation (14.1350 ),


pffiffiffi pffiffiffi
c 1 ¼ a1 þ a2 m ; c 2 ¼ a2  a1 m ; ð14:1450 Þ
one obtains the equation of elliptic type

  1 ð1 þ mÞða2 þ a2 ÞX
DX  ¼ 0; ð14:146Þ
1 2
4
which is obtained by two different changes of function.
If the modulus of longitudinal elasticity is not taken in the form (14.132), but
has a more general form

E ¼ ef ðx1 ;x2 Þ ; ð14:147Þ


where f ðx1 ; x2 Þ is an arbitrary function, then the equations of equilibrium take the
form
ðo1 þ f;1 Þ
r11 þ ðo2 þ f;2 Þ
r12 ¼ 0; ðo1 þ f;1 Þ
r12 þ ðo2 þ f;2 Þ
r22 ¼ 0: ð14:148Þ
Because the operators in the parentheses are invertible, i.e.,

ðo1 þ f;1 Þðo2 þ f;2 Þ ¼ ðo2 þ f;2 Þðo1 þ f;1 Þ


¼ o12 þ f;2 o1 þ f;1 o2 þ f;12 þ f;1 f;2 ; ð14:149Þ

one may continue on the same way as before, obtaining a representation of the
stresses in a form analogous to the Airy one, i.e.,

11 ¼ ðo2 þ f;2 Þ2 F; r


r 22 ¼ ðo1 þ f;1 Þ2 F; r
12 ¼ ðo1 þ f;1 Þðo2 þ f;2 ÞF; ð14:150Þ

where the function Fðx1 ; x2 Þ verifies the partial differential equation


h i nh i
D ðo1 þ f;1 Þ2 þ ðo2 þ f;2 Þ2 F ¼ ð1 þ mÞ ðf;1 Þ2  f;22 ðo2 þ f;2 Þ2
h i o
þ ðf;2 Þ2  f;22 ðo1 þ f;1 Þ2  2ðf;1 f;2  f;12 Þðo1 þ f;1 Þðo2 þ f;2 Þ F ð14:151Þ

or the equation
h i
D ðo1 þ f;1 Þ2 þ ðo2 þ f;2 Þ2 F  ð1 þ mÞðf;1 o2  f;2 o1 Þ2 F

¼ ð1 þ mÞ f;12 ½o1 ðo2 þ f;2 Þ þ o2 ðo1 þ f;1 Þ
 ð14:1510 Þ
f;11 o2 ðo2 þ f;2 Þ  f;22 o1 ðo1 þ f;1 Þ F:
662 14 Anisotropic and Non-homogeneous Bodies

We remark that one may obtain formulae similar to (14.137), (14.140)–


(14.14000 ) in solving the boundary value problem.
But, in the case of a solution in displacements, one can no more get analogous
results.

14.2.2.2 Particular Integrals

Let us search now some particular integrals of the Eqs. (14.143) and (14.146).
First of all, let us search a particular integral in form of a polynomial of nth
degree (except a polynomial of first degree). But only one polynomial is not
sufficient because, by applying it, one obtains a polynomial of ðn  1Þth degree
and a polynomial of ðn  2Þth degree and each of them must identically vanish.
Hence, the polynomial of nth degree must verify each of the equations
c1 v;1 þ c2 v;2 ¼ 0; Dv ¼ 0: ð14:152Þ

We observe that the first Eq. (14.152) has an integral of the form
v ¼ vðc2 x1  c1 x2 Þ;

the polynomial of nth degree must thus be of the form

v ¼ ðc2 x1  c1 x2 Þn ;
but it is not a harmonic one. Hence, one introduces a polynomial of ðn  1Þth
degree of the form

X ¼ nðn  1Þðc1 x1 þ c2 x2 Þðc2 x1  c1 x2 Þn2 ;


but neither this one is harmonic. One tries a polynomial of ðn  2Þth degree a.s.o.
One searches a polynomial of the form
Xn ¼ P0 ðx1 ; x2 Þ þ P1 ðx1 ; x2 Þ þ    þ Pn ðx1 ; x2 Þ; ð14:153Þ

where Pn ðx1 ; x2 Þ is a homogeneous polynomial of nth degree. Applying the


operator of the Eq. (14.143), one obtains a polynomial which must identically
vanish. Between the arbitrary constants of the polynomial Xn ðx1 ; x2 Þ, i.e.: 1 þ
2 þ . . . þ n ¼ nðn þ 1Þ=2 constants, there must take place 1 þ 2 þ    þ n  1 ¼
ðn  1Þn=2 relations, so that there remain nðn þ 1Þ=2  ðn  1Þn=2 ¼ n arbitrary
constants. Hence, there are n independent polynomials of the form (14.153) which
verify the Eq. (14.143).
14.2 Non-homogeneous Elastic Bodies 663

We may thus write

Xn ðx1 ; x2 Þ ¼ C0 þ C1 ðc2 x1  c1 x2 Þ þ C2 ½ðc2 x1  c1 x2 Þ2  2ðc1 x1  c2 x2 Þ


þ C3 ½ðc2 x1  c1 x2 Þ3  6ðc1 x1 þ c2 x2 Þðc2 x1  c1 x2 Þ þ . . .

þ Cn ðc2 x1  c1 x2 Þn  nðn  1Þðc1 x1 þ c2 x2 Þðc2 x1  c1 x2 Þn2

nðn  1Þðn  2Þðn  3Þ
þ ðc1 x1 þ c2 x2 Þ2 ðc2 x1  c1 x2 Þn1 þ    :
2!
ð14:154Þ
These polynomials allow to solve some elementary particular cases of loading.
Other interesting particular integrals are of the form
 1 ; x2 Þ ¼ X1 ðx1 ÞX2 ðx2 Þ;
Xðx ð14:155Þ
the Eq. (14.146) leads to the condition
1
X1 X2;22 þ X2 X1;11  ð1 þ mÞða21 þ a22 ÞX1 X2 ¼ 0 ð14:156Þ
4
or
1 1 1
X1;11 ¼  X2;22 þ ð1 þ mÞða21 þ a22 Þ ¼  k2 : ð14:1560 Þ
X1 X2 4
One can chose a function of the form
 ¼ ec1 x1 þc2 x2 ;
X ð14:157Þ
which leads to the condition
1 1
c21 þ c22 ¼ ð1 þ mÞða21 þ a22 Þ ¼ ðc21 þ c22 Þ: ð14:1570 Þ
4 4
Taking into account (14.144), one obtains the integral

Xðx1 ; x2 Þ ¼ eðc1 c1 =2Þx1 eðc2 c2 =2Þx2 ; ð14:158Þ


with which one may construct important expansions, useful in various boundary
value problems.
Let be
c1 ¼ a1  ib1 ; c2 ¼ a2  ib2 ; ð14:159Þ
pffiffiffiffiffiffiffi
where i ¼ 1; there result the relations
1
a21 þ a22  ðb21 þ b22 Þ ¼ ðc21 þ c22 Þ; a1 b1  a2 b2 ¼ 0: ð14:1590 Þ
4
664 14 Anisotropic and Non-homogeneous Bodies

To have only trigonometric lines for the variable x1 , one chooses


1
a1 ¼ c1 ; b1 arbitrary: ð14:15900 Þ
2
The conditions (14.1590 ) lead to
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffi u
u 2
 2
2 t 2 c2 2 c22
a2 ¼  b1 þ þ b1 þ þb21 c21 ;
2 4 4
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð14:160Þ
pffiffiffi u
u  2
  2
2t c 2 c2
b2 ¼   b21 þ þ b21 þ 2 þb21 c21 :
2 4 4

One obtains thus particular integrals of the form

Xðx1 ; x2 Þ ¼ eðc2 =2a2 Þx2 eib1 x1 eib2 x2 ; ð14:161Þ


there result trigonometric lines in both directions. We remark that, if a2 has the
sign þ, then the quantities b1 and b2 have the same sign if c1 [ 0 and an opposite
sign if c1 \0. If a2 has the sign , then the signs of the above mentioned quantities
are just opposite.
Choosing integrals of the form (14.154), one can—taking into account
(14.135)—construct stress functions of the form
F ¼ C ¼ const; F ¼ Cx1 ; F ¼ Cx2 ; . . . ð14:162Þ
The first function leads to the state of state

11 ¼ b2 C; r
r 22 ¼ a2 C; r
12 ¼ abC; ð14:163Þ

which corresponds to a traction in both directions, superposed with a simple shear


(in reduced stresses).
Similarly, one may consider the other functions (14.162) which lead to other
results and may be used in more complicated cases.
Integrals of the form (14.160), (14.161) can be used for domains the frontier of
which is formed by co-ordinate lines (for which one of the co-ordinates is known).
One can thus use a stress function of the form
Fðx1 ; x2 Þ ¼ K0 þ K1 x1 þ K2 x2
X
þ ekn x1 ½An sinðcn x1  an x2 Þ þ Bn cosðcn x1  an x2 Þ
n
X
þ eln x2 ½Cn sinðcn x1  bn x2 Þ þ Dn cosðcn x1  bn x2 Þ
n
X
þ evn x2 A0n sinðcn x1  dn x2 Þ þ B0n cosðcn x1  dn x2 Þ
n
X
þ e,n x2 Cn0 sinðcn x1  en x2 Þ þ D0n cosðcn x1  en x2 Þ ; ð14:164Þ
n
14.2 Non-homogeneous Elastic Bodies 665

where
np
cn ¼ ; n ¼ 1; 2; 3. . . ð14:1640 Þ
a
and 2a is the period; the quantities an ; bn ; dn ; en ; kn ; ln ; vn ; ,n may be expressed as
functions of cn by means of the relations (14.160). Under the trigonometric lines,
one takes the sign þ or  as one has c1 [ 0 or c1 \0, hence after the sign of the
pffiffiffi
expression a  b m. One obtains another stress function by interchanging x1 and
x2 .

14.2.2.3 Elastic Half-Plane

Let be the elastic half-plane x2 0 acted upon on the separation line x2 ¼ 0 by the
reduced periodic load
X X
pðx1 Þ ¼
 an sin cn x1 þ bn cos cn x1 ; ð14:165Þ
n n

we assume that the longitudinal modulus of elasticity is of the form

E ¼ Eð0; 0Þebx2 ; ð14:166Þ


where the rigidity of the half-plane is constant in the direction of Ox1 . The external
load is thus given by
pðx1 Þ ¼ Eð0; 0Þ
pðxÞ: ð14:1650 Þ
One may take a stress function of the form
X
Fðx1 ; x2 Þ ¼ ekn x2 ½An sinðcn x1 þ an x2 Þ þ Bn cosðcn x1 þ an x2 Þ
n
X
þ ekn x2 A0n sinðcn x1  an x2 Þ þ B0n cosðcn x1  an x2 Þ ; ð14:167Þ
n

where
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 1 1 2 1
kn ¼ b þ 2cn þ b þ 4c4n þ 2ð1 þ 2mÞb2 c2n þ b4 ;
2
2 2 2 4
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð14:168Þ
  rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 1 1
an ¼  2c2n þ b2 þ 4c4n þ 2ð1 þ 2mÞb2 c2n þ b4 :
2 2 4

The first sum above satisfies the second Eq. (14.1350 ) and the second sum
verifies the first of these equations.
666 14 Anisotropic and Non-homogeneous Bodies

By a linear relation, one obtains an integral of the form


X
Fðx1 ; x2 Þ ¼ ekn x2 ðA
 n sin an x2 þ B
 n cos an x2 Þ sin cn x1
n
X
0
þ ekn x2 A n
 0n cos an x2 cos cn x1 ;
 sin an x2 þ B ð14:1670 Þ
n

which does no more verify each of the equations (14.1350 ), but the Eq. (14.134000 ).
One also remarks that the functions (14.1670 ) are valid only for elasticity
moduli of the form (14.166).
By means of the formulae (14.133), one obtains the reduced stresses
X
22 ¼ 
r c2n ekn x2 ðA
 n sin an x2 þ B
 n cos an x2 Þ sin cn x1
n
X
0
 c2n ekn x2 A  0n cos an x2 cos cn x1 ;
 sin an x2 þ B
n
n
X
0
12 ¼
r cn ekn x2 ðb  kn Þ A n
 0n cos an x2
 sin an x2 þ B
n ð14:168Þ

0
 cos an x2  B
þan A n
 0n sin an x2 sin cn x1
X
 cn ekn x2 ½ðb  kn ÞðA  n sin an x2 þ B
 n cos an x2 Þ
n
 n cos an x2  B
þan ðA  n sin an x2 Þ cos cn x1 :

If we put the boundary conditions


x2 ¼ 0 : r
2 ¼ ~
pðx1 Þ; r
21 ¼ 0; ð14:169Þ
21 vanish for x2 ! 1.
2 and r
then the stresses r
The constants An; B 0 ; B
n; A  0n become
n

 n ¼ ðb  kn Þan ; A
A  0 ¼ ðb  kn Þbn ;
n
an c2n an c2n
ð14:170Þ
 n ¼  an ; B
B  0 ¼  bn :
c2n n c2n
The stress function (14.168) becomes
X1  
kn x2 b  kn
Fðx1 ; x2 Þ ¼ e sin an x2  cos an x2 ðan sin cn x1 þ bn cos cn x1 Þ
n
c2n an
ð14:171Þ

and leads to the state of stress


14.2 Non-homogeneous Elastic Bodies 667

X1 h ib  k 
kn x2 2 2 n
r
11 ¼ e an þ ðb  kn Þ sin an x2 þ cos an x2 ðan sin cn x1 þ bn cos cn x1 Þ;
n
c2n an
X  
b  kn
r
22 ¼ ekn x2 sin an x2 : cos an x2 ðan sin cn x1 þ bn cos cn x1 Þ;
n
an
ð14:172Þ
X 1 h i
12 ¼ 
r ekn x2 a2n þ ðb  kn Þ2 sin an x2 ðan cos cn x1  bn sin cn x1 Þ:
n
an c n
ð14:1720 Þ

One observes that the normal stresses on the separation line are given by
X a2 þ ðb  kn Þ2
n
11 ðx1 ; 0Þ ¼
r ðan sin cn x1 þ bn cos cn x1 Þ: ð14:173Þ
n
c2n

In the case of local normal loads acting upon the separation line, one may study
the problem on the same way, replacing the Fourier series by Fourier integrals. The
final results are thus similar to (14.172), (14.173), where the parameters kn ; an ; cn
become functions and the sums integrals.

References

A. Books

1. Beju, I., Soós, E., Teodorescu, P.P.: Euclidean Tensor Calculus with Applications. Abacus
Press, Tunbridge Wells (1983). (Ed. Tehnică, Bucuresti)
2. Brilla, J.: Anizotropické steny (Anisotropic Plates). Slov. Akad. Vied., Bratislava (1958)
3. Cristescu, N.D., Crăciun, E.M., Soós, E.: Mechanics of Elastic Composites. Chapman and
Hall/CRC, Boca Raton (2004)
4. Hearmon, R.F.S.: An Introduction to Applied Anisotropic Elasticity. Clarendon Press, Oxford
(1961)
5. Lekhnitskiǚ, S.G.: Teoriya uprugosti anizotropnogo tela (Theory of Elasticity of Anisotropis
Bodies). Ogiz, Moskva-Leningrad (1947)
6. Love, A.-E.-H.: A Treatise on the Mathematical Theory of Elasticity, 4th edn. Cambridge
University Press, Cambridge (1934)

B. Papers

7. Bekhterev, P.: Analiticheskoe issledovanie obobshchenovo zakona Guka (Analytical study of


the generalized hooke law). J. Russk. fiz.-khim. obshchestva. VII (1925)
8. Bishop, R.E.D.: Dynamical problems of plane stress and plane strain. Q.J. Mech. Appl. Math.
VI, 250 (1953)
668 14 Anisotropic and Non-homogeneous Bodies

9. Chentsov, N.G.: Issledovanie faneri, kak ortotropnoi plastinki (Application of bodies as


orthotropic plates). Tekhn. Zametki Tsagi. (1936)
10. Cristea, M.: Legea lui Hooke plană neizotropă (Plane anisotropic hooke’s law). Com. Acad.
Rom. I, 1007 (1951)
11. Fridman, M.M.: Matematicheskaya teoriya anizotropnykh sred (Mathematical theory of
anisotropic media). Prikl. Mat. Mekh. XIV, 321 (1950)
12. Green, A.E.: A note on stress systems in aelotropic materials. Proc. Roy. Soc. 162, 173
(1939). (173, 416, 420 (1920))
13. Iacovache, M. Aplicarea functiilor monogene în sensul lui Feodorov în teoria elasticitătii
corpurilor cu izotropie transversă (Application of the monogenic functions in the sense of
feodorov in the theory of elasticity of the bodies with transverse isotropy). Rev. Univ. si
Polit., Bucuresti. 58 (1952)
14. Lang, H.A.: The affine transformation for orthotropic plane-stress and plane-strain problems.
J. Appl. Mech. 23, 1 (1956)
15. Marguerre, K.: Ebenes und achsensymmetrisches Problem der Elastizitätstheorie. Z.A.M.M.
13, 437 (1933)
16. Mikhlin, S. G.: Ploskaya deformatsya v anizotropnoŭ srede (Plane deformation in anisotropic
media). Tr. seism. inst., ANSSSR. t (1936)
17. Moisil, A.: Asupra ecuatiilor echilibrului elastic plan pentru corpurile cu izotropie transversă
(On the equations of plane elastic equilibrium for bodies with transverse isotropy). Lucr. Ses.
Acad. Rom. 308 (1950)
18. Rabinovich, A. L.: Ob uprugikh postoiannykh i prochnosti anizotropnykh materialov (On
elastic strength of anisotropic materials). Trudy. Tsagi. (1946)
19. Reissner, E.: A contribution to the problem of elasticity of nonisotropic materials. Phil. Mag.
(1940)
20. Reissner, H.: Z.A.M.M. 11, 1 (1931)
21. Savin, G. N.: Osnovnaya ploskaya zadacha teorii uprugosti dlya anizotropnoŭ sredy
(odnosvyaznaya becknechnaya oblast’) (Fundamental plane problem of the theory of
elasticity for an isotropic body (Simple connected infinite domain)). Tr. Inst. Stroit. Mekh.,
ANSSSR (1938)
22. Savin, G.N.: Ob odnom metode resheniya osnovnoǚ ploskoǚ staticheskoǚ zadachi teorii
uprugosti anizotropnoǚ sredy (On a method of solving the fundamental plane statical problem
of the theory of elasticity of the anisotropic media). Tr. Inst. Matem. ANSSSR (1939)
23. Sokolnicoff, I.S.: Approximate methods of solution of twodimensional problems in
anisotropic elasticity. Proc. Symp. Appl. Math. (1950)
24. Teodorescu, P. P.: Asupra problemei plane a elasticitătii unor corpuri anizotrope. I. Relatii
între eforturi unitare si deformatii specifice (On the plane problem of elasticity of some
anisotropic bodies. I. Relations between stresses and strains). Com. Acad. Rom. VII, 395
(1957)
25. Teodorescu, P. P.: Asupra problemei plane a elasticitătii unor corpuri anizotrope. II. Corpuri
cu izotropie transversă (On the plane problem of elasticity of some anisotropic bodies. II.
Bodies with transverse isotropy). Com. Acad. Rom. VII, 401 (1957)
26. Teodorescu, P. P.: Asupra problemei plane a elasticitătii unor corpuri anizotrope. III. Corpuri
ortotrope (On the plane problem of elasticity of some anisotropic bodies. III. Orthotropie
bodies). Com. Acad. Rom. VII, 503 (1957)
27. Teodorescu, P. P.: Asupra problemei plane a elasticitătii unor corpuri anizotrope. IV. Corpuri
cu un plan de simetrie elastică (On the plane problem of elasticity of some anisotropic bodies.
IV. Bodies with a plane of elastic symmetry). Com. Acad. Rom. VII, 509 (1957)
28. Teodorescu, P. P: Asupra problemei plane a elasticitătii unor corpuri anizotrope. V. Corpuri
cu o axă de simetrie elastică de ordinul al treilea (On the plane problem of elasticity of some
anisotropic bodies. V. Bodies with an axis of elastic symmetry of third order). Bul. Acad.
Rom. VII, 641 (1957)
29. Teodorescu, P. P.: Asupra problemei plane a elasticitătii unor corpuri anizotrope. VI. Corpuri
cu o axă de simetrie elastică de ordinul patru (On the plane problem of elasticity of some
References 669

anisotropic bodies. VI. Bodies with an axis of elastic symmetry of fourth order). Bul. Acad.
Rom. VII, 753 (1957)
30. Teodorescu, P. P.: Asupra problemei plane a elasticitătii unor corpuri anizotrope. VII.
Corpuri actionate de forte masice oarecare (On the plane problem of elasticity of some
anisotropic bodies. VII. Bodies acted upon by arbitrary volume forces). Com. Acad. Rom.
VIII, 887 (1958)
31. Teodorescu, P.P.: Asupra problemei plane a elasticitătii unor corpuri anizotrope. VIII.
Influenta variatiei de temperatură (On the plane problem of elasticity of some anisotropic
bodies. VIII. Influence of the temperature variation). Com. Acad. Rom. VIII, 1119 (1958)
32. Teodorescu, P. P.: Asupra problemei plane a elasticitătii unor corpuri anizotrope. IX. Cazul
micilor miscări elastice (On the plane problem of elasticity of some anisotropic bodies. IX.
Case of small elastic motions). Com. Acad. Rom. VIII, 1243 (1958)
33. Teodorescu, P. P.: Ůber das Kinetische Problem nichthomogener elasticher Körper. Bull.
Acad. Pol. Sci., sér. Sci. Techn. XII, 595 (1964)
34. Teodorescu, P. P., Predeleanu, M.: Quelques considérations sur le problème des corps
élastiques hétérogènes (Non-homogeneity in Elasticity and Plasticity). In: IUTAM-
Symposium, pp. 31 (1958)
35. Teodorescu, P. P., Predeleanu, M.: Quelques considérations sur le problème des corps
élastiques hétérogènes. Bull. Acad. Pol. Sci., sér. Sci. Techn. VII, 81 (1959)
36. Teodorescu, P. P., Predeleanu, M.: Ůber das ebene Problem nichthomogeneuer elastischer
Körper. Acta Techn. Acad. Sci. Hung. XXVII, 349 (1959)
Chapter 15
Introduction to Thermoelasticity

Thermoelasticity contains the theory of heat conduction and the theory of strains
and stresses due to the flow of heat, when coupling of temperature and strain fields
occurs. An important case is that in which the coupling of temperature and strain is
neglected. In what follows some quasi-static and dynamic problems are dealt with.
Most of the monographs and treatises on the theory of elasticity contain smaller
of larger chapters on thermoelasticity. We mention some monographs dedicated to
heat conduction, i.e. those of H. C. Carlslaw [1], H. C. Carlslaw and J. C. Jaeger
[2] and Hamburger [3]; as to monographs dedicated to thermoelasticity, we
mention those of B. A. Boley and J. H. Weiner [4], E. Melan and H. Parkns [5],
W. Nowacki [6], H. Parkus [7] and O. Tedone [8].

15.1 Basic Relations and Equations

First of all, using also ideas from [6], we deal with some problems concerning heat
conduction and the relations to the equations of equilibrium and motion of elastic
bodies. Stationary and dynamic quasi-static problems will be then considered, as
well as dynamic ones. Some general principles and theorems will be presented too.

15.1.1 Heat Conduction. Equations of Thermodynamics

Hereafter, we introduce the partial differential equation of parabolic type of heat


propagation; we make then the connection to the elastic solids.

P. P. Teodorescu, Treatise on Classical Elasticity, Mathematical and Analytical 671


Techniques with Applications to Engineering, DOI: 10.1007/978-94-007-2616-1_15,
Ó Springer Science+Business Media Dordrecht 2013
672 15 Introduction to Thermoelasticity

15.1.1.1 Problems of Heat Conduction

Let us consider a deformable solid contained in a domain D of volume V; bounded by


the frontier S: We denote by Tðr; tÞ the temperature (T ¼ T0 þ #; see Sect. 4.10) at
the point of position vector r ¼ rðx1 ; x2 ; x3 Þ; at the time t: Because of the temperature
differences between various points of the domain D, it results the flow of heat
q ¼ kgradT ¼ krT; ð15:1Þ

where k is the coefficient of internal heat conduction.


In the time interval Dt; a quantity of heat
oT
DQ ¼ k dSDt ð15:2Þ
on
flows across a surface element dS at the point ðx1 ; x2 ; x3 Þ in the direction of the
external normal n; this represents the law of Fourier.
In case of a domain D1 of frontier S1 ; the quantity of heat is given by
ZZ
oT
DQ0 ¼ k dSDt;
S1 on

if inside of the considered domain there are heat sources too, the quantity of heat is
completed by the quantity
ZZZ
DQ00 ¼ WdVDt;
V1

where W is the quantity of heat generated per unit volume and per unit time. The
sum DQ ¼ DQ0 þ DQ00 may be determined on the basis of the temperature dif-
ferences in the domain D1 in the time interval Dt; thus, we obtain
ZZZ
DQ ¼ _
cqTdVDT;
V1

where q is the unit mass and c is the specific heat, i.e. the quantity of heat required
for a unit increase of temperature of unit mass. It results the equation
ZZZ ZZ
  oT
cqT_  W dV  k dS ¼ 0:
V1 S1 on

The Gauss-Ostrogradskiı̆ formula (A.1.96) leads now to


ZZZ
 
cqT_  W  kDT dV ¼ 0;
V1

which holds for an arbitrary domain D1 hence


1 Q
\T ¼ DT  T_ ¼  ; ð15:3Þ
, ,
15.1 Basic Relations and Equations 673

where \ is Nicolescu’s operator (see Sect. A.3.4.5) and


k W
,¼ ; Q¼ : ð15:30 Þ
qc qc
One obtains thus a partial differential equation of parabolic type (see
Sect. A.3.4.5), corresponding to non-stationary temperature fields. If the temper-
ature does not depend on time (T_ ¼ 0), then the corresponding field is stationary;
one obtains thus a Poisson equation of elliptic type
Q
DT ¼  : ð15:4Þ
,
If there are not heat sources inside the domain D, then the equation is a Laplace
equation
DT ¼ 0 ð15:5Þ

and the temperature field is a harmonic one.


There have been constructed also other models of heat propagation, which led
to a partial differential equation of hyperbolic type.

15.1.1.2 Boundary and Initial Condition

The temperature field is completely determined by the Eq. (15.3), to which one
must add appropriate boundary and initial conditions.
The boundary conditions are put for t [ t0 on the surface S and correspond to
the interaction between the body and the surrounding medium. They may be of
three kinds:
(i) The temperature is prescribed at all points of the surface S:
(ii) The temperature gradient is prescribed at all points of the surface S:
(iii) The function oT=on þ aT ¼ b; a; b ¼ const, is prescribed at all points of the
surface S:
The boundary condition (ii) occurs if the intensity of the heat flow from outside
into inside of the body is known. If there is no heat exchange with the surrounding
medium across the surface S (the body is thermally insulated), then oT=on ¼ 0 on
S: The condition (iii) corresponds to a free heat exchange on the surface S:
The initial condition determines the distribution of temperature T ¼ TðrÞ at the
initial moment t ¼ t0 :
If a part of the frontier is thrown to infinite, the boundary conditions are
replaced by the requirement that the temperature is bounded or takes a given value
on this part. If the conduction of heat is stationary, then the initial condition is
irrelevant; one puts only boundary conditions for the Eq. (15.4).
In what follows, we admit that the heat equation is solved, so that the tem-
perature Tðr; tÞ is known.
674 15 Introduction to Thermoelasticity

15.1.1.3 Equations of Equilibrium and Motion

Hereafter, we deal with an elastic, isotropic, homogeneous body with respect to


both its mechanical and thermal properties. We assume that for an unloaded body,
when the stresses are zero, the temperature is equal to zero too, so that Tðr; tÞ
represents the increment of temperature, beginning from this moment. By a we
denote a coefficient of linear expansion, which is independent on point and
direction.
Hooke’s law takes now the form
 
1  1 m
eij  aTdij ¼ ð1 þ mÞrij  mrkk dij ¼ rij  Hdij ; ð15:6Þ
E 2G 1þm

which represents the Duhamel-Neumann law. Making i ¼ j and summing, one has
1  2m
h  3at ¼ H; ð15:7Þ
E
the law mentioned above takes also the form
rij ¼ ðkh  bTÞdij þ 2leij ; ð15:8Þ

where we have introduced the notation


E
b ¼ ð3k þ 2lÞa ¼ a: ð15:9Þ
1  2m
In the particular case in which rij ¼ 0; one obtains

eij ¼ aTdij ; ð15:10Þ

the conditions of continuity of Saint-Venant (2.68) lead to T;ij ¼ 0; i; j ¼ 1; 2; 3;


hence to a linear distribution of temperature.
As in the classical static case, we use the Cauchy relations between strains and
displacements (5.2), the equations of Saint-Venant (2.68), the equations of equi-
librium (5.1) and, in the dynamic case, the equations of motion (5.6).

15.1.2 Stationary and Quasi-Static Problems

First of all, we deal with problems in which the temperature has not a decisive
importance in the variation of strains and stresses, i.e. with stationary and quasi-
static problems; we will consider thus both displacement and stress equations.
15.1 Basic Relations and Equations 675

15.1.2.1 Displacement Equations

Eliminating the strains and the stresses, as in the classical case, the equation of
Lamé type which must be verified by the displacement vector reads
lDu þ ðk þ lÞgrad divu þ F ¼ bgrad T ð15:11Þ

or
1 1
Du þ grad divu þ ðF  bgrad TÞ ¼ 0; ð15:110 Þ
1  2m l
We observe that the volume force F is replaced by F  bgrad T; obviously, in
the absence of volume forces one remains only with the quantity bgradT. As
usual, the boundary conditions are put in displacements (the first fundamental
problem).
J. N. Goodier has introduced the thermoelastic displacement potential Uðr; tÞ by
the relation
u ¼ grad U: ð15:12Þ

Introducing in (15.11), in the absence of the volume forces, one obtains


(grad div ¼ D)
lDgradU þ ðk þ lÞgradDU ¼ bgradT;
wherefrom the Poisson equation
1þm
DU ¼ cT; c¼ a; ð15:13Þ
1m
neglecting an arbitrary constant.
The state of stress is thus given by
rij ¼ 2lðU;ij  DUdij Þ; i; j ¼ 1; 2; 3: ð15:14Þ
For the elastic space (unbounded body), by means of the Poisson integral, one
may write
Z
c TðnÞ
UðrÞ ¼  dVðnÞ; ð15:15Þ
4p V Rðr; nÞ

where Rðr; nÞ is the distance between the points of vectors r and n:


Taking into account (15.12), we obtain
ZZZ
c o TðnÞ
ui ¼  dVðnÞ
4p oxi V Rðr; nÞ
ZZZ
¼b TðnÞUi ðr; nÞdVðnÞ; i ¼ 1; 2; 3; ð15:16Þ
V
676 15 Introduction to Thermoelasticity

where
 
1 1
Ui ðr; nÞ ¼  ; i ¼ 1; 2; 3: ð15:160 Þ
4pðk þ 2lÞ R ;i

This function can be regarded as the displacement of the point r in the direction of
the Oxi -axis, due to the action of a centre of pressure situated at the point n of the
elastic space, i.e. the dilatation at the point n; due to the action of a concentrated
unit force situated at the point r and having the direction of the Oxi -axis.
Denoting
Hi ¼ ð3k þ 2lÞUi ; i ¼ 1; 2; 3;
we get
ZZZ
ui ðrÞ ¼ a TðnÞHi ðn; rÞdVðnÞ; i ¼ 1; 2; 3: ð15:17Þ
V

The formulae (15.15) and (15.17) are obtained for the stationary case. If one
replaces the function TðnÞ by the function Tðn; tÞ; it follows that the results remain
valid for quasi-static problems.
Taking into account the homogeneous equation of that conduction (corre-
sponding to (15.3))
1
\T ¼ DT  T_ ¼ 0; ð15:300 Þ
,
_ it results
differentiating (15.13) with respect to time and eliminating T,

DU_ ¼ c,DT;
hence
Z t
U ¼ c, Tdt þ U0 þ U1 t; ð15:18Þ
0

where U1 is a harmonic function and U0 ¼ Uðr; 0Þ is the thermoelastic displace-


ment potential at the moment t ¼ 0; corresponding to the temperature Tðr; 0Þ ¼
T0 ðrÞ and satisfying the equation

DU_ 0 ¼ cT0 : ð15:180 Þ


We mention that the function Uðr; tÞ obtained for an infinite domain, satisfies
only a part of the boundary conditions, leading to the displacements
ui ¼ U;i ;
 i ¼ 1; 2; 3: ð15:19Þ
Hence, one has
ui þ 
ui ¼  ui ; i ¼ 1; 2; 3; ð15:20Þ
15.1 Basic Relations and Equations 677

where
lD
ui þ ðk þ lÞD
u;i ¼ 0; i ¼ 1; 2; 3: ð15:190 Þ
The vector of components 
ui may be expressed in the form
1 

ui ¼ ðk þ 2lÞDdij  ðk þ lÞoij uj ; i ¼ 1; 2; 3; ð15:21Þ
l
where the displacement functions uj ; j ¼ 1; 2; 3, are biharmonic

DDuj ; j ¼ 1; 2; 3: ð15:210 Þ

The representation (15.21) is called Galerkin’s representation and u is


Galerkin’s vector. The corresponding state of stress is given by
2l h i
ij ¼
r ðmDdij  oij Þuk;k þ 2ð1  mÞDuði;jÞ ; i; j ¼ 1; 2; 3: ð15:22Þ
1  2m

15.1.2.2 Stress Equations

For the second fundamental problem of thermoelasticity it is useful to introduce


the stress equations.
By means of the Duhamel-Neumann law (15.6), the equations of Saint-Venant
take the form
 
1 1þm
Drij þ H;ij þ 2al T;ij þ DTdij ¼ 0; i; j ¼ 1; 2; 3; ð15:23Þ
1þm 1m

one obtains thus the Beltrami type equations (in the absence of the volume forces).
The sum of the normal stresses is given by
2aE
DH þ DT ¼ 0: ð15:24Þ
1m
 
In the particular case of a stationary temperature field T_ ¼ 0 and no heat
sources acting inside the body ðQ ¼ 0Þ, the temperature function becomes a
harmonic one ðDT ¼ 0Þ; the sum of the normal stresses is a harmonic function too
ðDH ¼ 0Þ: Hence,
1
Drij þ ðH þ aET Þ;ij ¼ 0: ð15:25Þ
1þm
Starting from these equations, one has obtained various solutions for the elastic
space, for the elastic half-space, for axisymmetric problems a.s.o.
678 15 Introduction to Thermoelasticity

15.1.3 Dynamic Problems

As in the previous section, one performs a study both in displacements and


stresses. Any non-stationary problem of thermoelasticity is a dynamic problem. If
the changes in temperature are slow, then the problem can be regarded as a quasi-
static one; but if the changes of temperature may be sudden, then the inertia terms
cannot be neglected and the problem is a true dynamic one.

15.1.3.1 Displacement Equations

Eliminating the stresses and the strains between the equations of thermoelasticity
in the dynamic case, one obtains the equations of Lamé type in the form (see §4.3)
lh2 u þ ðk þ lÞgrad divu ¼ bgradT ð15:26Þ

or in the form
1 b
h2 u þ grad divu ¼ gradT; ð15:260 Þ
ð1  2mÞ l
in the absence of the volume forces.
Another form of these equations is
1 b
h1 u þ curlcurlu ¼ grad T: ð15:27Þ
2ð1  mÞ k þ 2l
By means of a scalar potential U ¼ Uðr; tÞ and of a vector potential W ¼
Wðr; tÞ; the displacement uðr; tÞ can be represented in the form
u ¼ gradU þ curlW; ð15:28Þ

obtaining a representation of Lamé-Clebsch type. The potential functions verify


simple wave equations of the form
h1 U ¼ cT; h2 W ¼ 0: ð15:29Þ
Eliminating the temperature between the parabolic Eq. (15.3) and the first
Eq. (15.29), one obtains the equation
cQ
h1 \U ¼  ; ð15:30Þ
,
which must be satisfied by the potential function U:
A particular solution of the first Eq. (15.29) may be obtained by the retarded
potential
15.1 Basic Relations and Equations 679

ZZZ
c Tðn; t  R=c1 Þ
Uðr; tÞ ¼  dVðnÞ; R  c1 t; ð15:31Þ
4p V Rðr; nÞ

where Rðr; nÞ is the distance between the points r and n: The integration is carried
over a volume of the body situated inside the sphere of radius R ¼ c1 t and of
centre at the point r: The formula corresponds for a body initially at rest, in the
natural state.
The displacement u can be obtained also as a sum u ¼ u þu ; where u
 satisfies

the non-homogeneous system (15.26). while u satisfies the homogeneous system
(of Lamé type) and is given by
1
 ¼ h2 u þ
u ðDu  graddivuÞ; ð15:32Þ
1  2m
where the vector potential u satisfies the double wave equation
h1 h2 u ¼ 0; ð15:320 Þ

one obtains thus the Somigliana-Iacovache representation (Sect. 5.3.2.2), which


generalizes the Galerkin one.

15.1.3.2 Stress Equations

In case of the second fundamental problem we try to find a solution in stresses, by


an extension of the Beltrami equations. Starting from the equations of continuity of
Saint-Venant and using the Duhamel-Neumann law, one obtains the continuity
equations in stresses in the form
2ðk þ lÞ k
Drij þ H;ij  DHdij
3k þ 2l 3k þ 2l
1  
 2r €ij  2bT;ij þ 2a lDTdij þ 3ðk þ lÞT;ij
c2
 
1 k
¼ 2 €kk  2alT€ dij ; i; j ¼ 1; 2; 3:
r ð15:33Þ
c2 3k þ 2l
By contraction, one may write
€ ¼ 0;
h1 H þ cð4lDT  3qTÞ ð15:34Þ

finally, the Beltrami type equations read


2ðk þ lÞ c2  c2 k € ij
h2 rij þ H;ij  1 2 2 2 Hd
3k þ 2l c1 c2 3k þ 2l
n 1 h
þ a 2lT;ij þ ð3k þ 2lÞDT
k þ 2l
i o
 ð5k þ 4lÞqT€ dij ¼ 0; i; j ¼ 1; 2; 3: ð15:35Þ
680 15 Introduction to Thermoelasticity

Applying the operator h1 ; one obtains also the interesting relation


  
€ ij ¼ 0; i; j ¼ 1; 2; 3:
h2 h1 rij  c 2lðT;ij  DTdij Þ þ qTd ð15:36Þ
In particular, one gets the relation
 
D Drij  2clðT;ij  DTdij Þ ¼ 0; i; j ¼ 1; 2; 3; ð15:37Þ

for the static case.


These results are valid in the absence of the volume forces; if these forces are
non-zero, one adds the terms
k
gij ¼ Fi;j þ Fj;i þ Fk;k dij ; i; j ¼ 1; 2; 3; ð15:38Þ
k þ 2l
in the left side of the Eq. (15.35), obtaining thus the equations of Beltrami-Michell
type.

15.1.4 General Considerations

We deal hereafter with some general considerations concerning the elastic bodies
subjected to thermal actions too, i.e. d’Alembert’s and Hamilton’s principles,
reciprocal theorems and coupling problems.

15.1.4.1 The d’Alembert-Lagrange Principle. Hamilton’s Principle

Let us consider a body in an isothermal state subjected at the instant t to a virtual


displacement dui (infinitesimal arbitrary displacement, independent on time and
compatible with the geometrical conditions on the surface of the body); we are
thus led to the d’Alembert-Lagrange principle (6.24), where the variation of the
external work is given by (6.23) and the variation of the internal work is given by
(6.230 ).
Introducing the influence of the temperature variation, one can express the
external work (6.23) in the form
ZZZ ZZ

n
dWe ¼ ðFi  bT;i Þdui dV þ pi þ bTni dui dS; ð15:39Þ
V S

using the Gauss-Ostrogradskiı̆ formula specify (A.1.960 ), we may write


ZZZ ZZ
ðTdui Þ;i dV ¼ Tni dui dS;
V S
15.1 Basic Relations and Equations 681

so that the internal work becomes


ZZZ
dWi ¼ dWi  b Tdekk dV; ð15:40Þ
V

representing an internal work generalized to the thermoelastic problem, i.e.


ZZZ

Wi ¼ W  dV; ð15:41Þ
V

where
1 Ea
W  ¼ rij eij  ekk T: ð15:410 Þ
2 1  2m
Hence, the principle takes the form
ZZZ
dWe  ui dui dV ¼ dWi ;
q€ ð15:42Þ
V

in dynamical thermoelasticity; it states that in a virtual displacement of a body


from the instantaneous state to a neighbouring one, the work performed by the
superficial forces and the volume forces is equal to the increment of the energy of
deformation.
In the case of a static loading and a stationary temperature field, one obtains
d’Alembert’s generalized principle
dWe ¼ dWi : ð15:43Þ
Let us now consider that the states of the elastic body vary continuously
between the times t0 and t1 : Integrating the relation (15.42) on the assumed time
interval, we get
Z t1 Z t1 Z t1 ZZZ
dWe dt ¼ dWi dt þ dt q€ui dui dV; ð15:44Þ
t0 t0 t0 V

if we assume that one has synchronous virtual displacements (the virtual dis-
placements vanish at t0 and t1 ), then
Z t1 Z t1 ZZZ
dTdt ¼  dt q€ui dui dV:
t0 t0 V

It follows that
Z t1 Z t1
d ðWi  TÞdt ¼ dWe dt:
t0 t0
682 15 Introduction to Thermoelasticity

Assuming that the external loads do not vary, we introduce the potential energy
P ¼ Wi  2We ; ð15:45Þ

where We is the external work (6.12), corresponding to the static case; we obtain
thus Hamilton’s principle in the form
Z t1
d ðP  TÞdt ¼ 0; ð15:46Þ
t0

where the integral is called action. Hamilton’s principle states that the action is
steady, while the motion of the elastic body corresponds to the extremals of this
functional in the synchronous case.
If the external load is static and the temperature field is steady, the relation
(15.46) yields the principle of minimum of the potential energy.
dP ¼ 0; ð15:47Þ

because the second variation of P is positive.

15.1.4.2 Principle of Reciprocity

Let us consider two states of strain and stress, due to two cases of loading, marked
by (‘) and (’’), respectively; we may state the principle (theorem) of Betti in the
form (6.36).
n
Replacing the volume forces Fi00 by bT;i and the superficial loads p 00i by bTni
and neglecting the volume forces Fi ; one may express this principle as follows
ZZZ ZZ ZZZ ZZ
n
Fi0 u00i dV þ p 0i u00i dS ¼ b T;i u0i dV þ b Tni u0i dS: ð15:48Þ
V S V S

Applying the Gauss-Ostrogradskiı̆ formula specify (A.1.960 ), we get


ZZZ ZZ
 0 
ui T ;i dV ¼ Tu0i ni dS;
V S

so that the principle of reciprocity is expressed in the form


ZZZ ZZ ZZZ ZZZ
n
Fi0 u00i dV þ p 0i u00i dS ¼ b Th0 dV ¼ a TH0 dV: ð15:49Þ
V S V V

Let be the state of strain and state denoted by the dash, due to a concentrated
unit force acting at the point r in the direction of the Oxi -axis; the relation (15.19)
leads to
15.1 Basic Relations and Equations 683

ZZZ
ui ðrÞ ¼ a TðnÞHðiÞ ðn; rÞdVðnÞ
V
ZZZ
¼b TðnÞhðiÞ ðn; rÞdVðnÞ; i ¼ 1; 2; 3: ð15:50Þ
V

One obtains thus Maysel’s formulae, which specify the displacements in a body,
due to the action of a steady temperature field. Obviously, these formulae may be
used also in the case of quasi-static problems, in which the time t is a parameter,
ZZZ
ui ðr; tÞ ¼ a Tðn; tÞHðiÞ ðn; rÞdVðnÞ; i ¼ 1; 2; 3: ð15:500 Þ
V

If we construct now the equation of virtual work, we may write the relation of
reciprocity in the form
ZZZ ZZ ZZZ
n
Fi0 du00i dV þ p 0i du00i dS þ a T 0 dH00 dV
V S V
ZZZ ZZ ZZZ
00 0 n 00 0
¼ dFi ui dV þ d p i ui dS þ a dT 00 H0 dV: ð15:51Þ
V S V

The above results are valid for stationary and quasi-static thermoelastic prob-
lems. If we wish to include also dynamic effects, then we must introduce inertia
terms in the relation (15.51); we obtain thus
ZZZ ZZ ZZZ
n
ðFi0  q€u0i Þdu00i dV þ p 0i du00i dS þ a T 0 dH00 dV
V S V
ZZZ ZZ ZZZ
00 00 0 n 00 0
¼ ðdFi  qd€ ui Þui dV þ d p i ui dS þ a dT 00 H0 dV: ð15:52Þ
V S V

If the thermal stresses are produced only by the temperature field T 0 ; i.e. if Fi0 ¼ 0;
n
p 0i ¼ 0; i ¼ 1; 2; 3; and dT 00 ¼ 0; then the formula (15.52) takes the form
ZZZ ZZ
00 00 0 n
ðdFi  qd€ ui Þui dV þ d p 00i u0i dS
V
ZZZ ZZZ S
0 00
þ q€ui dui dV ¼ a T 0 dH00 dV: ð15:53Þ
V V

15.1.4.3 Coupling of Temperature and Strain Fields

Heretofore we have assumed that the temperature field is independent of the


corresponding state of strain. Obviously, this is an approximation which is strictly
valid only in the case of a stationary temperature field. In fact, a change of the
amount of heat leads to a change of the state of strain and, conversely, a change in
the deformation of the body leads to a change of the temperature field.
684 15 Introduction to Thermoelasticity

The coupling of the strain and temperature fields represents a more correct
approach of the problem, which becomes thus more intricate too. The equation of
heat propagation (15.3), (15.30 ) is completed as follows
1 1
\T ¼ DT  T_ ¼  Q þ gdivu;
_ ð15:54Þ
, ,
where one takes use of the notation (15.30 ), as well as of
bT0
g¼ ; ð15:540 Þ
h
where h is the heat conductivity. Obviously, the coupling is due to this latter term.
We remember now Lamé’s type equation
lh2 u þ ðk þ lÞgraddivu þ F ¼ bgradT ð15:55Þ

and assume that


u ¼ gradU þ curl W; ð15:56Þ

F ¼ qðgradu þ curl wÞ: ð15:560 Þ


We obtain thus the system of equations
1
\T ¼ gDU_  Q; ð15:57Þ
,
1
h1 U ¼ cT  u: ð15:570 Þ
c21

1
h2 W ¼  w: ð15:5700 Þ
c22

Eliminating the temperature from the first two equations, we get


   
1o _ c 1 1o
D h1 U  cgDU ¼  Q  2 D  u: ð15:58Þ
, ot , c1 , ot
If the temperature varies slowly in time, then the inertia terms may be neglected
and the problem may be regarded as quasi-static.
The coupling of the temperature and strain fields is important in certain specific
technological applications (e.g., electronics). But in problems of thermal stresses
occuring in machine structures the coupling may be neglected.
15.2 Applications 685

15.2 Applications

Hereafter we will present some elementary examples and then some applications
concerning problems with axial symmetry using results due to S. Timoshenko and
J. N. Goodier [9]; as well, we deal with some plane problems of thermoelasticity,
using our results [10].

15.2.1 Elementary Examples

First of all, we deal with some simple problems concerning thin plates and then we
deal with a sphere of large radius.

15.2.1.1 Case of a Thin Plate

Let us consider a thin rectangular plate referred to the Ox1 and Ox2 axes, situated
in the middle plane; the frontier is specified by the sides xi ¼ ai ; i ¼ 1; 2: We
suppose firstly that T ¼ Tðx2 Þ is an even function. The longitudinal thermal
expansion aT is suppressed by applying the longitudinal stress r011 ¼ aET; lat-
erally we assume that the expansion is free, but at the ends of the plate there must
be forces of compression which must suppress any expansion in the direction of
the Ox1 -axis. If the plate is free from external forces, then we must superpose
distributed at the ends x1 ¼ a1 of resultant
Z a2
aETdx2 ; ð15:59Þ
a2

at a certain distance from the ends will appear approximate uniformly distributed
stresses (principle of Saint-Venant), of magnitude
Z a2
1
aETdx2 :
2a2 a2
The thermal stresses in the plate with free ends, at a considerable distance from the
ends, becomes
Z a2
1
r11 ¼ aETdx2  aET: ð15:60Þ
2a2 a2
686 15 Introduction to Thermoelasticity

If, e.g., the temperature is distributed parabolically


 
x22
T ¼ T0 1  2 ; ð15:61Þ
a2

then we obtain
 
2 x2
r11 ¼ aET0  aET0 1  22 ; ð15:610 Þ
3 a2

result valid except near the ends.


If the temperature T ¼ Tðx2 Þ is not symmetric with respect to the Ox1 -axis, to
the resultant (15.59) one must add a resultant moment
Z a2
aETx2 dx2 ; ð15:590 Þ
a2

hence, to the stresses r011 one must add the bending stresses r0011 ¼ rx2 =a2 ; which
have to be determined by equating to zero the total moment over the cross section.
It results
Z a2 Z a2
1 2
rx2 dx2  aETx2 dx2 ¼ 0:
a2 a2 a2

Integrating, one obtains,


Z a2
3
r0011 ¼ x2 aETx2 dx2 ;
2a32 a2

so that
Z a2 Z a2
1 3
r11 ¼ aET þ aETdx2 þ 3 x2 aETx2 dx2 : ð15:62Þ
2a2 a2 2a 2 a2

If the plate is not thin in the Ox3 -direction, i.e. if the plate is large in this
direction, we get a plate with the Ox1 x3 -plane as middle plane; we assume, as
before, that the temperature T ¼ Tðx2 Þ depends, only on one variable. Putting
e11 ¼ e33 ¼ aT; r22 ¼ 0; it results
aET
r11 ¼ r33 ¼  : ð15:63Þ
1m
By a similar reasoning, one obtains
Z a2
aET 1
r11 ¼ r33 ¼  þ aETdx2
1  m 2ð1  mÞa2 a2
Z a2
3x2
þ aETx2 dx2 ; ð15:64Þ
2ð1  mÞa32 a2
15.2 Applications 687

a result analogous to the previous one. One may thus easily calculate the state of
stress in the plate if the distribution of the temperature T over the thickness of the
plate is known.
If the faces x2 ¼ a2 are maintained at two different temperatures Tþ and T ;
then, after a certain time, the temperature is given by
1 1 x2
T ¼ ðTþ þ T Þ þ ðTþ  T Þ : ð15:65Þ
2 2 a2
If the plate is not restrained, then the thermal stresses vanish, but if the edges
are perfectly restrained against expansion and rotation, then the stress thus induced
is given by the relations (15.63). For instance, if T2 ¼ T1 ; we get
x2
T ¼ T1 ; ð15:650 Þ
a2
so that
aET1 x2
r11 ¼ r33 ¼  ; ð15:66Þ
1  m a2
the maximum stress is
aET1
r11max ¼ r33max ¼ : ð15:660 Þ
1m

15.2.1.2 Case of a Sphere

Let be now the case of a sphere of large radius; we assume that in a small spherical
element of radius q; situated at the centre of the sphere, there occurs a temperature
rise T: The expansion of this element being not possible, a pressure p appears on
its surface. Assuming that the radius of the sphere is much greater than q; we
obtain for r [ q the normal stress rrr and the stress rt (in the plane normal to r; the
centre of the sphere being taken as origin)

q 3 p
q 3
rrr ¼ p ; rt ¼ ; ð15:67Þ
r 2 r
for r ¼ q; we get
p
rrr ¼ p; rt ¼ : ð15:670 Þ
2
The increase of the radius r ¼ q; due to the pressure p; is given by
q pq pq
Dr ¼ ðqet Þr¼q ¼ ½rt  mðrrr þ rt Þr¼q ¼ ð1 þ mÞ ¼ :
E 2E 4l
688 15 Introduction to Thermoelasticity

It must equate the increase of the radius of the heated spherical element produced
by the rise of temperature and pressure; it follows that
pq pq
aqT  ð1  2mÞ ¼ ð1 þ mÞ ;
E 2E
wherefrom
2 aET
p¼ : ð15:68Þ
31  m
Introducing in the relation (15.67), we get the formulae
2 aET
q 3 1 aET
q 3
rrr ¼  ; rt ¼ : ð15:69Þ
31  m r 31m r

15.2.2 Problems with Axial Symmetry

Hereafter we deal with two problems with axial symmetry, i.e.: the case of a
circular disk and the case of a circular cylinder.

15.2.2.1 Thin Circular Disk

Let be a thin circular disk for which the temperature T ¼ TðrÞ does not vary over
its thickness; we assume that the stresses and the displacements have the same
property. The temperature, as well as the displacements and the stresses are
symmetric about the centre and do not depend on h:
The Duhamel-Neumann law takes the form
1 1
err  aT ¼ ðrrr  mrhh Þ; ehh  aT ¼ ðrhh  mrrr Þ: ð15:70Þ
E E
and corresponds to a plane state of stress.
Solving with respect to the stresses, one obtains,
E
rrr ¼ ½err þ mehh  ð1 þ mÞaT ;
1  m2 ð15:700 Þ
E
rhh ¼ ½ehh þ merr  ð1 þ mÞaT :
1  m2
The equation of equilibrium
drrr 1
þ ðrrr  rhh Þ ¼ 0 ð15:71Þ
dr r
15.2 Applications 689

takes the form


d dT
r ðerr þ mehh Þ þ ð1  mÞðerr  ehh Þ ¼ ð1 þ mÞar : ð15:710 Þ
dr dr
The Cauchy equations are
dur ur
err ¼ ; ehh ¼ : ð15:72Þ
dr r
Introducing in (15.710 ), we get

d2 ur 1 dur ur dT
2
þ  2 ¼ ð1 þ mÞa ;
dr r dr r dr
wherefrom

d 1 dðrur Þ dT
¼ ð1 þ mÞa : ð15:73Þ
dr r dr dr

By integration, we get
Z r
1 C2
ur ¼ ð1 þ mÞa TðrÞrdr þ C1 r þ ; ð15:74Þ
r r0 r

where r0 may be chosen arbitrarily.


For a solid disk, one takes r0 ¼ 0; in the case of a circular hole, r0 is the inner
radius.
Using the relations (15.72) and then (15.700 ), the state of stress reads
Z
aE r E C2
rrr ¼  2 TðrÞrdr þ ð1 þ mÞC1  ð1  mÞ 2 ;
r r0 1  m2 r
Z r ð15:75Þ
aE E C2
rhh ¼ 2 TðrÞrdr þ ð1 þ mÞC1  ð1  mÞ 2 ;
r r0 1  m2 r

where the arbitrary constants C1 and C2 are determined by the boundary


conditions.
In the case of a solid disk ðr0 ¼ 0Þ one takes C2 ¼ 0; because
Z
1 r
lim TðrÞrdr ¼ 0
r!0 r 0

and ur must be finite at the centre of the disk.


On the frontier r ¼ r1 one must have rrr ¼ 0; therefore
Z
a r1
C1 ¼ ð1  mÞ 2 TðrÞrdr;
r1 0
690 15 Introduction to Thermoelasticity

so that
Z r1 Z
1 1 r
rrr ¼ aE 2 TðrÞrdr  2 TðrÞrdr ;
r 0 r 0
1 Z r1 Z ð15:76Þ
1 1 r
rhh ¼ aE 2 TðrÞrdr þ 2 TðrÞrdr  TðrÞ :
r1 0 r 0
At the centre, one has
Z r
1 1
lim TðrÞrdr ¼ T0 ; ð15:77Þ
r!0 r 2 0 2
where T0 is the temperature at the centre.

15.2.2.2 Circular Cylinder

Let be a circular cylinder for which the Oz-axis is a geometrical axis of symmetry,
as well as for the temperature TðrÞ: We suppose that all the geometrical quantities
do not depend on z: We deal with a long cylinder (theoretically infinite, practically
finite) and suppose that the displacement uz vanishes.
The angular strains and the tangential stresses vanish and the Duhamel-Neumann
law takes the form
1
err  aT ¼ ½rrr  mðrhh þ rzz Þ;
E
1
ehh  aT ¼ ½rrr  mðrzz þ rrr Þ; ð15:78Þ
E
1
ezz  aT ¼ ½rzz  mðrrr þ rhh Þ
E
in cylindrical co-ordinates. Since ur ¼ 0; so that ezz ¼ 0; the third Eq. (15.78)
leads to
rzz ¼ mðrrr þ rhh Þ  aET; ð15:79Þ

wherefrom

1  m2
m
err  ð1 þ mÞaT ¼ rrr  rhh ;
E 1m ð15:790 Þ
1  m2
m
ehh  ð1 þ mÞaT ¼ rhh  rrr :
E 1m
These equations correspond for the case of a plane state of strain and may be
obtained from the equations (15.70) (a plane state of stress) if one replaces E by
E=ð1  m2 Þ; m by m=ð1  mÞ and a by ð1 þ mÞa: Therefore, we may use the formulae
(15.74) and (15.75) obtained in the preceding subsection; it results
15.2 Applications 691

Z r
1 þ ma C2
ur ¼ TðrÞrdr þ C1 r þ
; ð15:80Þ
1mr r0 r
Z  
aE 1 r E C1 C2
rrr ¼  TðrÞrdr þ  ;
1  m r 2 r0 1 þ m 1  2m r 2
Z   ð15:81Þ
aE 1 r aET E C1 C2
rhh ¼ TðrÞrdr  þ þ
1  m r 2 r0 1  m 1 þ m 1  2m r 2

and, taking into account (15.79), one has


aET 2mEC1
rzz ¼  þ : ð15:810 Þ
1  m ð1 þ mÞð1  2mÞ
In order to have uz ¼ 0; we superpose a uniform axial stress rzz ¼ C3 ; we
choose C3 so as the resultant forces at the end cross sections be zero. Conse-
quently, one must add the term mC3 r=E to ur :
In the case of a solid cylinder we take r0 ¼ 0; because the displacement ur must
vanish on the Oz-axis, we take C2 ¼ 0:
If the boundary of the cylinder r ¼ r1 is free of superficial loads, we obtain
Z r1
C1 a
¼ TðrÞrdr:
ð1 þ mÞð1  2mÞ ð1  mÞr12 0
The resultant of the axial stresses (15.810 ) is
Z r1 Z
2paE r1 2mEC1
rzz  2prdr ¼  TðrÞrdr þ pr 2
0 1m 0 ð1 þ mÞð1  2mÞ 1

and the resultant of the uniform axial stress C3 is C3  pr12 ; the total axial force
must vanish, so that
Z
2 2paE r1 2mEC1
C3  2pr1  TðrÞrdr þ pr 2 :
1m 0 ð1 þ mÞð1  2mÞ 1
Finally, the state of stress reads
Z Z
aE 1 r1 1 r
rrr ¼ TðrÞrdr  2 TðrÞrdr ;
1  m r12 0 r 0
Z Z
aE 1 r1 1 r
rhh ¼ TðrÞrdr þ TðrÞrdr ; ð15:82Þ
1  m r12 0 r2 0
Z
aE 2m r1
rzz ¼ TðrÞrdr  T
1  m r12 0

and the radial displacement is given by


692 15 Introduction to Thermoelasticity

Z Z
ð1 þ mÞa ð1  2mÞr r1 1 r
ur ¼ TðrÞrdr þ TðrÞrdr : ð15:820 Þ
1m r12 0 r2 0
Let us assume now that the initial temperature is T0 and that, beginning from
t ¼ 0; the boundary is maintained at the temperature T1 : The distribution of
temperature is given by
X1  
r pn t
T ¼ ðT0  T1 Þ An J0 bn e ; ð15:83Þ
n¼1
r1

where J0 ðbn r=r1 Þ is the Bessel function of zeroth order and where bn are the roots
of the equation J0 ðbÞ ¼ 0; as well, one has
2 k b2n
An ¼ ; pn ¼ ; ð15:830 Þ
bn J1 ðbn Þ cq r12

where J1 ðbn Þ is the Bessel function of first order.


Taking into account the relation
Z r    
r r1 r r
J0 bn rdr ¼ J1 b n ;
0 r1 bn r1

one obtains finally



2aEðT0  T1 Þ X1
1 r1 J1 ðbn r=r1 Þ pn t
rrr ¼ 2
1 e ;
1m n¼1 bn
r J1 ðbn Þ
ð15:84Þ
2aEðT0  T1 Þ X1
1 1 r1 J1 ðbn r=r1 Þ J0 ðbn r=r1 Þ pn t
rhh ¼ þ  e ;
1m b bn bn r J1 ðbn Þ
n¼1 n
J1 ðbn Þ

as well as

2aEðT0  T1 Þ X1
1 2 J0 ðbn r=r1 Þ pn t
rzz ¼  e : ð15:840 Þ
1m b bn
n¼1 n
J1 ðbn Þ

If the cylinder has a circular cylindrical hole of radius r0 ; which is symmetrical


with respect to the Oz-axis, then one can use the same (15.81), (15.810 ) formulae.
We put the conditions that the stress rrr vanishes for r ¼ r0 and r ¼ r1 : The
constants C1 and C2 are thus given by
C1 C2
 ¼ 0;
1  2m r02
Z r1  
aE E C1 C2
 TðrÞrdr þ  ¼ 0;
ð1  mÞr12 r0 1 þ m 1  2m r22

one obtains
15.2 Applications 693

Z r1
EC1 aE
¼  2  TðrÞrdr;
ð1 þ mÞð1  2mÞ ð1  mÞ r1  r02 r0
Z r1
EC2 aEr02
¼  2  TðrÞrdr:
1 þ m ð1  mÞ r1  r02 r0

Substituting in (15.81), (15.810 ), the state of stress reads


2 Z Z r
aE r  r02 r1
rrr ¼ TðrÞrdr  TðrÞrdr ; ð15:85Þ
ð1  mÞr 2 r12  r02 r0 r0
2 Z Z r
aE r þ r02 r1 2
rhh ¼ TðrÞrdr þ TðrÞrdr  Tr ;
ð1  mÞr 2 r12  r02 r0 r0
Z ð15:850 Þ
r1
aE 2
rzz ¼ TðrÞrdr  TðrÞ :
1  m r1  r02
2
r0

In the case of a steady heat flow, we assume that T0 is the temperature on the
inner surface; if the temperature on the outer surface is T1 ¼ 0; then the temper-
ature is given by
T0 r1
TðrÞ ¼ ln : ð15:86Þ
lnðr1 =r0 Þ r
In this particular case, the state of stress becomes
 
aET0 r1 r2 r2 r1
rrr ¼  ln þ 2 0 2 1  12 ln ;
2ð1  mÞ lnðr1 =r0 Þ r r1  r0 r r0
  ð15:87Þ
aET0 r1 r2 r2 r1
rhh ¼ 1  ln  2 0 2 1 þ 12 ln ;
2ð1  mÞ lnðr1 =r0 Þ r r1  r0 r r0

aET0 r1 2r 2 r1
rzz ¼ 1  2 ln  2 0 2 ln : ð15:870 Þ
2ð1  mÞ lnðr1 =r0 Þ r r1  r0 r0

15.2.3 Plane Problems of Thermoelasticity

The problems considered in the previous section are plane problems. Hereafter
we deal with a more general case in Cartesian co-ordinates; we will thus give
formulations in stresses and displacements.
694 15 Introduction to Thermoelasticity

15.2.3.1 General Results

We will consider in the following orthotropic homogeneous bodies, using the


results given in Sect. 14.1.1.3. Assuming a temperature given by T ¼ Tðx1 ; x2 Þ,
the Duhamel-Neumann law takes the form
1 m12 m13
e11  a1 T ¼ r11  r22  r33 ;
E1 E2 E3
m21 1 m23
e22  a2 T ¼  r11  r22  r33 ; ð15:88Þ
E1 E2 E3
m31 m32 1
e33  a3 T ¼  r11  r22 þ r33 ;
E1 E2 E3
1 1 1
e23 ¼ r23 ; e31 ¼ r31 ; e12 ¼ r12 : ð15:880 Þ
2G23 2G31 2G12
Between the 12 elastic constants E1 ; E2 ; E3 ; G23 ; G31 ; G12 ; m23 6¼ m32 ; m31 6¼
m13 ; m12 6¼ m21 take place the relations
m12 m21 m23 m32 m31 m13
¼ ; ¼ ; ¼ ; ð15:89Þ
E2 E 1 E3 E 2 E1 E3
so that 9 elastic constants are independent; a1 ; a2 ; a3 are the coefficients of linear
expansion in the three directions.
In the case of a plane state of stress, one has
r33 ¼ 0; r31 ¼ r32 ¼ 0 ð15:90Þ

and the constitutive law becomes

e11  a1 T ¼ E11 r11  mE122 r22 ;


ð15:91Þ
e22  a2 T ¼  mE211 r11 þ E12 r22 ;

1
e12 ¼ r12 ð15:910 Þ
2G12
to which we add
m31 m32
e33  a3 T ¼  r11  r22 : ð15:9100 Þ
E1 E2
Between the elastic constants takes place the first relation (15.89), so that one has
only 6 independent elastic constants.
In the case of a plane state of strain we assume that
e33 ¼ 0; e31 ¼ e32 ¼ 0; ð15:92Þ
we are thus led to the same relations (15.91), (15.910 ) in which the elastic constants
E1 ; E2 ; m12 6¼ m21 are replaced by the generalized elastic constants
15.2 Applications 695

E1 E2
E10 ¼ ; E20 ¼ ;
1  m13 m31 1  m23 m32
ð15:93Þ
m12 þ m13 m32 m21 þ m23 m31
m012 ¼ ; m021 ¼
1  m23 m32 1  m13 m31
and the coefficients of linear dilatation a1 ; a2 by the generalized coefficients of
linear dilatation

a01 ¼ a1 þ m13 a3 ; a02 ¼ a2 þ m23 a3 : ð15:94Þ

One can verify that between these generalized constants remains valid the first
relation (15.89). In this case, one obtains also
r33 ¼ m13 r11 þ m23 r22  E3 a3 T; ð15:95Þ
because the linear strain e33 vanishes and we took into consideration the relation
(15.89). In this case we have only 6 elastic constants too.
Excepting the determination of e33 for the state of plane stress and the deter-
mination of r33 for the state of plane strain, the problems are identical from the
mathematical point of view; there are involved only 4 distinct elastic constants (the
other two constants are used to determine the two quantities mentioned above).

15.2.3.2 Formulation in Stresses

To the equations of equilibrium (without volume forces)


r11;1 þ r12;2 ¼ 0; r21;1 þ r22;2 ¼ 0

we add Saint-Venant’s condition of continuity


e11;22 þ e22;11 ¼ 2e12;12 :

Using the relations (15.91), (15.910 ) and the equations of equilibrium mentioned
above, the equations of continuity in stresses read
     
1 1 1 m21 m12
r11 þ a1 T þ r22 þ a2 T ¼   r12;12 : ð15:96Þ
E1 ;22 E2 ;11 l12 E1 E2

This leads to a solution in stresses of the problem. In the case of arbitrary volume
forces, one must add some particular integrals.
We introduce a representation of Airy type of the form
r11 ¼ F;22 ; r22 ¼ F;11 ; r12 ¼ F;12 ; ð15:97Þ

where the function F is of the form


 þ F1 ;
F¼F ð15:98Þ
696 15 Introduction to Thermoelasticity

 is the general integral of the equation


the function F
DF ¼ 0 ð15:99Þ

and F1 is a particular integral of the equation


 
DF1 ¼  a1 T;11 þ a2 T;22 ; ð15:990 Þ

where the operator D is given by


 
1 1 m21 m12 1
D ¼ o1111 þ   o1122 þ o2222 : ð15:100Þ
E2 l12 E1 E2 E1

15.2.3.3 The problem of Temperature Variation

If the temperature variation is not known, then one must obtain it from the
equation of heat conduction in orthotropic media

T_ ¼ ,1 T;11 þ ,2 T;22 þ Q; ð15:101Þ

where
k1 k2
,1 ¼ ; ,2 ¼ ; ð15:1010 Þ
cq cq
one supposes that Q ¼ Qðx1 ; x2 ; tÞ; T ¼ Tðx1 ; x2 ; tÞ.
Taking into account (15.990 ), we obtain
T ¼ DF ; ð15:102Þ
where we have introduced a new potential function F ¼ F ðx1 ; x2 ; tÞ; which ver-
ifies the equation
Dð,1 o11 þ ,2 o22  ot ÞF ¼ Q: ð15:103Þ
The connection between the two potential functions is made by
F ¼ ða1 o11 þ a2 o22 ÞF ; ð15:104Þ

so that the function F verifies the equation


Dð,1 o11 þ ,2 o22  ot Þ F ¼ a2 M;11 þ a1 M;22 : ð15:105Þ
Applying Boggio’s theorem, we may assert that the function F is of the form
 þ F1 þ F 0 ;
F¼F ð15:106Þ
 is the general integral of the Eq. (15.99), F1 is a particular integral of the
where F
complete Eq. (15.105), while F 0 is the general integral of the heat equation
15.2 Applications 697

0 0
,1 F;11 þ ,2 F;22 ¼ F_ 0 : ð15:107Þ

If the flow of heat is stationary, one has T_ ¼ 0 and if one has not heat sources one
has Q ¼ 0 too; in this case, the potential function is given by
Dð,1 o11 þ ,2 o22 Þ F ¼ 0: ð15:108Þ
It is interesting to remark that, in the case of an anisotropic (in particular
orthotropic) body, a variation of the temperature leads not only to a variation of the
strain tensor but also to a variation of the stress tensor.
Using Cauchy’s equations, one may easily calculate the components of the
displacement vector.
To determine the potential function F; one puts three conditions on the
boundary: two for the stresses and one for the temperature.
In conclusion, to solve the thermoelastic problem one must integrate a partial
differential equation of the form
Dða1 o11 þ a2 o22  ot Þ F ¼ 0; ð15:109Þ

on the boundary being given the function F; the normal derivative oF=on and DF:
One may also put an initial condition concerning the distribution of temperature
Tðx1 ; x2 ; tÞ for t ¼ t0 :

15.2.3.4 Formulation in Displacements

In a solution in displacements of the problem, we eliminate the strains and the


stresses, obtaining Lamé’s type equations in the form

E1 ðu1;11 þ m12 u2;12 Þ þ l12 ð1  m12 m21 Þðu1;22 þ u2;12 Þ


 E1 ða1 þ m12 a2 ÞT;1 ¼ 0;
ð15:110Þ
l12 ð1  m12 m21 Þðu1;12 þ u2;11 Þ þ E2 ðu2;22 þ m12 u1;12 Þ
 E2 ða2 þ m21 a1 ÞT;2 ¼ 0:

In the general case (the function T is not known), we eliminate the temperature
function between the Eq. (15.110), obtaining an equation verified by the dis-
placements u1 and u2 : Because the differential operators which act on these
functions are prime each other, it results that we can express the displacements by
means of only one function in the form
 
a1 a2 þ m21 a1 a1 þ m12 a2
u1 ¼  F ;22 þ F ;11 ;
G12 E1 E2 ;1
  ð15:111Þ
a2 a1 þ m12 a2 a2 þ m21 a1
u2 ¼  F ;11 þ F ;22 :
G12 E2 E1 ;2
698 15 Introduction to Thermoelasticity

The results is no more valid in the particular case of the isotropic body, because
the above operators are no more prime each other.

References

A. Books

1. Carlslaw, H.S.: Introduction to the Mathematical Theory of Conduction of Heat in Solids.


2nd edn. Macmillan Publisher, London (1921)
2. Carlslaw, H.S., Jaeger, J.C.: Conduction of Heat in Solids. Clarendon Press, Oxford (1959)
3. Hamburger, L.: Introducere în teoria propagării căldurii (Introduction to the Theory of Heat
Propagation). I. Conductßia în solide (Conduction in Solids). Ed. Acad. Rom., Bucuresßti
(1936)
4. Boley, B.A., Weiner, J.H.: Theory of Thermal Stresses. Willey, New York (1960)
5. Melan, E., Parkus, H.: Wärmespannungen infolge strationärer Temperaturfelder. Springer,
Wien (1953)
6. Nowacki, W.: Thermoelasticity. Pergamon Press/PWN—Polish Science Publisher, London/
Warszawa (1962)
7. Parkus, H.: Instationäre Wärmespannungen. Springer, Wien (1959)
8. Tedone, O.: Allgemeine Theoreme der mathematischen Elastizitätstheorie. Thermische
Deformation, in Encykl., math. Wiss (1910)
9. Timoschenko, S., Goodier, J.N.: Theory of Elasticity. 2nd edn. McGraw-Hill, New York
(1951)

B. Papers

10. Teodorescu, P.P.: Asupra problemei plane a elasticitătßii unor corpuri anizotrope. VIII.
Influentßa variatßiei de temperatură (On the plane problem of elasticity of certain anisotropic
bodies. VIII. Influence of the temperature variation). Com. Acad. Rom., VIII (1958)
Chapter 16
Introduction to Linear Viscoelasticity

We have assumed in the classical theory of elasticity that the constitutive law (the
strain-stress relations) is linear and independent on time. As well, we have
assumed the hypothesis of small deformations with respect to unity, so that the
principle of superposition of effects may be applied. On the other hand, many
bodies do not respect the above hypotheses, appearing the influence of the time
too. A constitutive law has a more general form, i.e.
f ðr; e; tÞ ¼ 0;
where appears also the variable t.
To take into consideration the variation in time of the mechanical properties of
the body, we will assume that the medium is a viscoelastic one, i.e. it is formed of
two media: a perfectly elastic medium and a medium with properties of viscous
fluid, described by Hooke’s law and by Newton’s law, respectively.
In this order of ideas, we assume the superposition principle of Boltzmann as a
basis of the mathematical theory of linear viscoelastic bodies. In the case of a
Boltzmannian body, if the stress r1 ðtÞ leads to the strain e1 ðtÞ and the stress r2 ðtÞ
leads to the strain e2 ðtÞ, then the sum r1 ðtÞ þ r2 ðtÞ leads to the strain e1 ðtÞ þ e2 ðtÞ;
in particular, if r2 ¼ kr1 , then e2 ¼ ke1 (k = const) and if r2 ¼ r1 , then e2 ¼ e1 .
Thus, the above constitutive law is a linear differential or integral equation.

16.1 Linear Viscoelastic Solids

In viscoelasticity, an important rôle is played by the creep and relaxation func-


tions, which represent a measure of the mechanical properties of the body.
Hereafter we will put in evidence constitutive laws for one-dimensional and three-
dimensional models, using differential and integro-differential representations.

P. P. Teodorescu, Treatise on Classical Elasticity, Mathematical and Analytical 699


Techniques with Applications to Engineering, DOI: 10.1007/978-94-007-2616-1_16,
Ó Springer Science+Business Media Dordrecht 2013
700 16 Introduction to Linear Viscoelasticity

16.1.1 Constitutive Laws

By means of the creep and relaxation functions we will construct one- and three-
dimensional laws, putting in evidence their differential and integro-differential
form too. Complex moduli of creep and relaxation will also be put in evidence.

16.1.1.1 Creep and Relaxation Phenomena

In Sect. 4.2.2.1 has been considered a cylindrical sample of viscoelastic material,


subjected to a simple axial loading. Using this model, the phenomena of creep and
relaxation have been put into evidence.
One has thus introduced the creep function uðt; r0 Þ given by (4.164), repre-
sented in Fig. 4.20.
The relaxation function wðt; e0 Þ is given by (4.167) and is represented in
Fig. 4.21. Because r0  0; r_ 0  0 and r€0  0, from (4.167) it results that the
relaxation function has these properties too, i.e.

w  0; w_  0; w
€  0: ð16:1Þ
Moreover, the experiments show that the curvature of the relaxation function w is a
decreasing function, hence
v  0:
w ð16:10 Þ
Indeed, because the curvature 1=q corresponding to the function rðtÞ is given
by

1 rv
¼ ;
q ð1 þ r_ 2 Þ3=2

the condition put to the relaxation function and the relation (16.1) lead to

_ r2
3r€
rv  0;
1 þ r_ 2
wherefrom the relation (16.10 ). A study of the properties of the linear viscoelastic
bodies by means of the properties (16.1), (16.10 ) is due to M. Alain [8].

16.1.1.2 Constitutive Law of the Isotropic Viscoelastic Solid

The viscoelastic solids are bodies with memory, because the state of stress at the
moment t depends on the whole history of the deformations, i.e. on the state of
deformation on the interval ð1; t: Hereafter, we will characterize the linear
viscoelastic solid, deducing the corresponding constitutive law by means of the
methods used by M. Gurtin and E. Sternberg [10] and by Yu. N. Rabotnov [5].
16.1 Linear Viscoelastic Solids 701

Let be rðr; tÞ and eðr; tÞ; ðr; tÞ 2 X  R  R3  R; the tensor fields of stress
and strain, the components of which are functions of class C0;0 ðX  RÞ; which
satisfy the conditions

rij ¼ 0; eij ¼ 0; i; j ¼ 1; 2; 3; for t  0; 8rðx1 ; x2 ; x3 Þ 2 X  R3 ; ð16:2Þ

which correspond to the natural state before the initial moment t = 0. We remark
that, for a fixed r in X; the component rij ðr; tÞ of the stress tensor at the moment
t represents a real functional of the components ekl ðr; sÞ; k; l ¼ 1; 2; 3; s 2
ð1; s of the strain tensor. Consequently, the stress rij ðr; tÞ may be represented
as a functional in the form
rij ðr; tÞ ¼ Lij ðekl Þ; i; j; k; l ¼ 1; 2; 3; ð16:3Þ

where Lij represents a real functional defined on the set of functions ekl 2
C0;0 ðX  RÞ, equal to zero for t \ 0: For r 2 X, variable, Lij represents an oper-
ator. In a concentrated form, the relation (16.3) may be written in the form
r ¼ LðeÞ; ð16:4Þ

where L is a real functional

L : C0;0 ðX; RÞ ! R; ð16:40 Þ

with r fixed and ekl ¼ 0; k; l ¼ 1; 2; 3 for t \ 0; ekl 2 C0;0 ðX  RÞ.


Thus, we call viscoelastic linear solid a deformable solid with linearized
deformations (ui;j  1), the constitutive law of which is of the form (16.4) and
satisfies following conditions:

8e0 ; e00 2 C0;0 ðX  RÞ; e0 ¼ e00 ¼ 0 for t\0;

(i) 8a0 ; a00 2 R one has

Lða0 e0 þ a00 e00 Þ ¼ a0 Lðe0 Þ þ a00 Lðe00 Þ; ð16:5Þ

(ii) 8t; s 2 R one has

e0 ðr; tÞ ¼ eðr; t  sÞ ) r0 ðr; tÞ ¼ Lðe0 ðr; tÞÞ ¼ rðr; t  sÞ; ð16:50 Þ

(iii) 8t 2 R; 8r 2 X fixed, 9lt [ 0 so that

jLðeðr; tÞÞj  lt keðr; tÞk; ð16:500 Þ


702 16 Introduction to Linear Viscoelasticity

 
where keðr; tÞk ¼ max eij ðr; tÞ :
s2½0;t
The condition (i) expresses the linearity of the functional L and, from the
mechanical point of view, ensures the application of the principle of superposition
of effects of Boltzmann. The condition (ii) puts in evidence the fact that the
functional L is invariant with respect to the translation operation in time, hence
independent on the choice of the origin of time measure; on the other hand, this
condition shows that the constitution of the body does not depend on time and,
consequently, the correspondence between the tensors r and e defined by (16.4) is
independent on the choice of their origin, Finally, the condition (iii) expresses the
fact that the functional L is bounded and, together with (i), ensures its continuity.
Hence, it results that eðr; sÞ ¼ 0 for s 2 ½0; t implies rðr; tÞ ¼ 0 on ½0; t:
To write the law (16.4) in components, let us observe that the functional L is
equivalent to the lot of functionals Lijkl so that the constitutive law may be written
in the form
rij ðr; tÞ ¼ Lijkl ekl ðr; tÞ; i; j; k; l ¼ 1; 2; 3; ð16:6Þ

one may say that the functional L is a fourth order tensor functional of components
Lijkl ; i; j; k; l ¼ 1; 2; 3: Because the tensors r and e are symmetric, it follows the
symmetry with respect to the pairs of indices ði; jÞ, ðk; lÞ; i.e.:
Lijkl ¼ Ljikl ¼ Lijlk ; i; j; k; l ¼ 1; 2; 3: ð16:7Þ

It results that from the 34 ¼ 81 components only 36 are distinct; they satisfy the
conditions (i), (ii), (iii) and characterize the general anisotropic linear viscoelastic
solid. If the functional is invertible, then there exists the tensor functional p of
components pijkl ; i; j; k; l ¼ 1; 2; 3; so that

eij ðr; tÞ ¼ pijkl rkl ðr; tÞ; i; j; k; l ¼ 1; 2; 3; ð16:8Þ

obviously, one may write the relations


pijkl ¼ pjikl ¼ pijlk ; i; j; k; l ¼ 1; 2; 3: ð16:9Þ
We will express the constitutive law (16.6) in an integral form, using the
theorem of F. Riesz [7] for the representation of the linear functionals on the space
of continuous functions. Because, for the linear viscoelastic solids, the state of
strain and stress at the moment t is determined by the whole history of this state,
then taking into account the condition (ii) and the relations (16.2), rij is a linear
and continuous functional of the history of deformations, i.e. ekl ðr; t  sÞ; s 2
½0; t; k; l ¼ 1; 2; 3: The constitutive law may be thus represented, by means of the
Stieltjes integral, in the form
Z t
rij ðr; tÞ ¼ ekl ðr; t  sÞdwijkl ðr; tÞ; i; j ¼ 1; 2; 3; ð16:10Þ
0
16.1 Linear Viscoelastic Solids 703

where wijkl ðr; tÞ are functions with bounded variation with respect to t 2 R on any
interval and equal to zero for t  0: These functions are called relaxation functions
and are the components of the relaxation tensor w, which is in complete con-
cordance with the function introduced in the one-dimensional case.
Denoting
8 8
< f ðaÞ; x\a; < gðaÞ; x\a;
f ðxÞ ¼ f ðxÞ; x 2 ½a; b;  gðxÞ ¼ gðxÞ; x 2 ½a; b; ð16:11Þ
: :
f ðbÞ; x [ b; gðbÞ; x [ b;

the formula (A.278) allows to write


!
X
n
ðf 0 ðxÞ; 
gðxÞÞ ¼ ~
0
f ðxÞ þ si dðx  xi Þ; gðxÞ
i¼0
  X n
~f 0 ; g þ
¼  si ðf Þgðxi Þ; ð16:12Þ
i¼0

because ðdðx  xi Þ; 
gðxÞÞ ¼ gðxi Þ ¼ gðxi Þ. Taking into account (16.11), one
obtains
  Z b
~
f 0 ; 
g ¼ gðxÞ~f 0 ðxÞdx;
a

so that
Z b X
n Z b
ðf 0 ; 
gÞ ¼ gðxÞ~f 0 ðxÞdx þ si ðf Þgðxi Þ ¼ gðxÞdf ðxÞ: ð16:13Þ
a i¼0 a

If we assume that wijkl ðr; tÞ 2 C0;1 ðX  ½0; 1ÞÞ and take into account that
wijkl ¼ 0 for t  0; then it results that wijkl ðr; tÞ 2 C0;1 ðX  Rnf0gÞ; while t ¼ 0
represents a point of discontinuity of the first species, with the jump s0 ðwijkl Þ
¼ wijkl ðr; 0 þ 0Þ  wijkl ðr; 0  0Þ ¼ wijkl ðr; 0þÞ:
Assuming that wijkl ðr; tÞ has, at the origin, a discontinuity of the first species
with respect to t 2 R and that rij ðr; tÞ 2 C0;0 ðX  RÞ, while wijkl and rij vanish for
t  0; then, taking into account (16.13), the constitutive law (16.10) takes the form
Z t ~o
rij ðr; tÞ ¼ ekl ðr; tÞwijkl ðr; 0þÞ þ ekl ðr; t  sÞ wijkl ðr; sÞds
0 os
Z t
¼ ekl ðr; t  sÞdwijkl ðr; sÞ; i; j ¼ 1; 2; 3; ð16:14Þ
0

where ~o=os is the derivative in the usual sense; the derivative in the sense of the
theory of distributions is denoted by o=os. Using the convolution product in
the space of distributions with regard to t; we may write the constitutive law in the
form
704 16 Introduction to Linear Viscoelasticity

o o
rij ðr; tÞ ¼ ekl ðr; tÞ wijkl ðr; tÞ ¼ ekl ðr; tÞ wijkl ðr; tÞ; i; j ¼ 1; 2; 3: ð16:15Þ
ðtÞ ot ot ðtÞ

This relation may be considered in the distribution space K0þ with the supports
in ½0; 1Þ if we assume that rij ; eij ; wijkl ; with respect to t 2 R; are distributions in
K0þ , while r 2 X is a parameter.
Because of the relations
wijkl ¼ wjikl ¼ wijlk ; i; j; k; l ¼ 1; 2; 3; ð16:16Þ

in the case of an anisotropic linear viscoelastic solid, the relaxation tensor w has 36
distinct components.
In the same space K0þ ; the constitutive law (16.15) may be written in the
equivalent form
o o
eij ðr; tÞ ¼ rkl ðr; tÞ uijkl ðr; tÞ ¼ rkl ðr; tÞ uijkl ðr; tÞ; i; j ¼ 1; 2; 3: ð16:17Þ
ðtÞ ot ot ðtÞ

where u represents the creep tensor of components uijkl in the distribution space
K0þ ; hence, uijkl ¼ 0 for t  0. As in the case of the relaxation tensor, one has

uijkl ¼ ujikl ¼ uijlk ; i; j; k; l ¼ 1; 2; 3; ð16:160 Þ

remaining 36 distinct components.


Supposing that the components of the tensors r; e; w; u admit the Laplace
transform in the distribution space, we apply this transformation in the equivalent
constitutive law (16.15) and (16.17); one obtains

r ~ ðr; pÞ; i; j ¼ 1; 2; 3;
~kl ðr; pÞ ¼ ~eij ðr; pÞpw ð16:18Þ
ijkl

~eij ðr; pÞ ¼ r
~kl ðr; pÞp~
uijkl ðr; pÞ; i; j ¼ 1; 2; 3; ð16:180 Þ

where p is a complex variable.


We remark that the above relations are similar to Hooke’s law, expressed either
in the form (4.56) or in the form (4.5600 ); we may thus establish following
correspondences

rij ðrÞ $ r ~ ðr; pÞ; i; j; k; l ¼ 1; 2; 3;


~ij ðr; pÞ; Hijkl $ pw ð16:19Þ
ijkl

eij ðrÞ $ ~eij ðr; pÞ; Cijkl $ p~


uijkl ðr; pÞ; i; j; k; l ¼ 1; 2; 3: ð16:190 Þ

Thus, a dual of Hooke’s law is obtained for the linear anisotropic viscoelastic
bodies; hence, the results obtained for the classical linear elastic bodies may be
transposed to the linear viscoelastic bodies.
In the case of the isotropic elastic solid, Hooke’s tensor is given by (4.89);
taking into account (16.19), one may write
16.1 Linear Viscoelastic Solids 705

~ ðr; pÞ ¼ ~kðr; pÞdij dkl þ l


pw ~ðr; pÞðdik djl þ dil djk Þ; i; j; k; l ¼ 1; 2; 3; ð16:20Þ
ijkl

where kðr; tÞ and lðr; tÞ are distribution in K0þ . Applying the inverse Laplace
transform, one obtains
o
w ðr; tÞ ¼ kðr; tÞdij dkl þ lðr; tÞðdik djl þ dil djk Þ; i; j; k; l ¼ 1; 2; 3: ð16:21Þ
ot ijkl
Substituting in the constitutive law (16.15), one obtains the constitutive law for an
isotropic linear viscoelastic solid in the form
rij ðr; tÞ ¼ kðr; tÞ ekk ðr; tÞdij þ 2lðr; tÞ eij ðr; tÞ; i; j ¼ 1; 2; 3: ð16:22Þ
ðtÞ ðtÞ

Applying Laplace’s transform, the relations (16.22) read

~ij ðr; tÞ ¼ ~
r kðr; pÞ~ekk ðr; pÞdij þ 2~
lðr; pÞ~eij ðr; pÞ; i; j ¼ 1; 2; 3:

Introducing the deviators of stresses and strains r0ij ; e0ij and the spheric tensors r0 dij
and e0 dij ; respectively, one obtains
~0ij ðr; pÞ ¼ 2~
r lðr; pÞ~e0ij ðr; pÞ;
ð16:23Þ
~ pÞ~e0 ðr; pÞ;
~0 ðr; pÞ ¼ 3Kðr;
r

where

~ pÞ ¼ ~ 2
Kðr; kðr; pÞ þ l~ðr; pÞ; ð16:24Þ
3
the inverse Laplace transform leads to

r0ij ðr; tÞ ¼ 2lðr; tÞ e0ij ðr; tÞ;


ðtÞ
ð16:230 Þ
r0 ðr; tÞ ¼ 3Kðr; tÞ e0 ðr; tÞ;
ðtÞ

where
2
Kðr; tÞ ¼ kðr; tÞ þ lðr; tÞ: ð16:240 Þ
3
The distribution l is the relaxation modulus to sliding, while the distribution K is
the relaxation modulus to volume dilatation.
The relaxation distributions wi ðr; tÞ 2 K0þ ; i ¼ 1; 2; are defined by the relations
o o
w1 ðr; tÞ ¼ 2lðr; tÞ; w2 ðr; tÞ ¼ 3Kðr; tÞ; ð16:25Þ
ot ot
wherefrom
w1 ðr; tÞ ¼ 2hðtÞ lðr; tÞ; w2 ðr; tÞ ¼ 3hðtÞ Kðr; tÞ; ð16:250 Þ
ðtÞ ðtÞ
706 16 Introduction to Linear Viscoelasticity

where hðtÞ 2 K0þ is Heaviside’s distribution. Hence

o o
r0ij ðr; tÞ ¼ w1 ðr; tÞ ~e0ij ðr; tÞ ¼ w1 ðr; tÞ ~e0ij ðr; tÞ; i; j ¼ 1; 2; 3; ð16:26Þ
ot ðtÞ ðtÞ ot

o o
r0 ðr; tÞ ¼ w ðr; tÞ e0 ðr; tÞ ¼ w2 ðr; tÞ e0 ðr; tÞ: ð16:260 Þ
ot 2 ðtÞ ðtÞ ot

From (16.21), (16.230 ) and (16.25), one obtains (by integrating with respect to t)
the components of the relaxation tensor
1
wijkl ðr; tÞ ¼ ½w ðr; tÞ  w1 ðr; tÞdij dkl
3 2
1
þ w1 ðr; tÞðdik djl þ dil djk Þ; i; j; k; l ¼ 1; 2; 3: ð16:27Þ
2
Hence, the constitutive law (16.22) becomes
1 o
rij ðr; tÞ ¼ ½w2 ðr; tÞ  w1 ðr; tÞ ekk ðr; tÞdij
3 ðtÞ ot
o
þ w1 ðr; tÞ eij ðr; tÞ; i; j ¼ 1; 2; 3: ð16:28Þ
ðtÞ ot

In general, the relaxation tensor W depends on r and t; in the case of a dependence


only on time, the viscoelastic solid is homogeneous; as a consequence, the dis-
tributions k; l; K; w1 ; w2 depend only on t 2 R.
In the case in which
w1 ðtÞ ¼ 2l0 hðtÞ; w2 ðtÞ ¼ 3K0 hðtÞ ð16:29Þ
one obtains, as a particular case, the constitutive law of a linear elastic body.
Let us now consider the constitutive law of the form (16.17). Using an anal-
ogous way, one obtains

eij ðr; tÞ ¼  lðr; tÞ rij ðr; tÞ;


kðr; tÞdij rkk ðr; tÞ þ 2 ð16:30Þ
ðtÞ ðtÞ

where   2 K0þ for an isotropic linear viscoelastic body. By means of the same
k; l
notations, one may write

e0ij ðr; tÞ ¼ 2
lðr; tÞ r0ij ðr; tÞ;
ðtÞ
ð16:31Þ
 tÞ r0 ðr; tÞ;
e0 ðr; tÞ ¼ 3Kðr;
ðtÞ

where

 tÞ ¼  2  tÞ 2 K0þ ðRÞ:
Kðr; kðr; tÞ þ lðr; tÞ; Kðr; ð16:310 Þ
3
16.1 Linear Viscoelastic Solids 707

One obtains the creep distributions in the form

u1 ðr; tÞ ¼ 2hðtÞ l  tÞ;


ðr; tÞ; u2 ðr; tÞ ¼ 3hðtÞ Kðr; ð16:32Þ
ðtÞ ðtÞ

obtaining, finally,
1 o
eij ðr; tÞ ¼ ½u2 ðr; tÞ  u1 ðr; tÞ rkk ðr; tÞdij
3 ðtÞ ot
o
þ u1 ðr; tÞ rij ðr; tÞ; i; j ¼ 1; 2; 3: ð16:33Þ
ðtÞ ot

In the homogeneous one-dimensional and isotropic case, one obtains the


relations
o o
eðx1 ; tÞ ¼ uðtÞ rðx1 ; tÞ ¼ uðtÞ rðx1 ; tÞ; ð16:34Þ
ðtÞ ot ot ðtÞ

with the notation u2 ðtÞ ¼ uðtÞ: For rðtÞ ¼ r0 hðtÞ one obtains uðtÞ ¼ e=r0 ; the
denomination of creep distribution being justified.
If

u1 ðtÞ ¼ 2  0 hðtÞ;
l0 hðtÞ; u2 ðtÞ ¼ 3K ð16:35Þ
then one obtains, as a particular case, the constitutive law of a linear elastic body.
Thus, the creep and relaxation distributions completely characterize the
mechanical properties with respect to time of the linear viscoelastic solids.
Because the constitutive laws (16.28), (16.33) are equivalent, it results that the
creep and relaxation distributions, as well as the distributions k; l; k; l are not
independent. To obtain the corresponding relations, we apply the Laplace trans-
form to the above mentioned laws; it follows

~0ij ðr; pÞ ¼ 2~
r lðr; pÞ~e0ij ðr; pÞ; i; j ¼ 1; 2; 3; r ~ pÞ~e0 ðr; pÞ;
~0 ðr; pÞ ¼ 3Kðr; ð16:36Þ

~e0ij ðr; pÞ ¼ 2l
~ ~
ðr; pÞr ~ pÞ~
0ij ðr; pÞ; i; j ¼ 1; 2; 3; ~e0 ðr; pÞ ¼ 3Kðr; r0 ðr; pÞ; ð16:360 Þ

wherefrom

~
lðr; pÞl
4~ ðr; pÞ ¼ 9Kðr; ~
 pÞKðr;
 pÞ ¼ 1: ð16:37Þ
Applying the inverse Laplace transform and taking into account (16.25) and
(16.32), one obtains
o o
lðr; tÞ 2lðr; tÞ ¼
2 u1 ðr; tÞ w1 ðr; tÞ ¼ dðtÞ;
ðtÞ ot ðtÞ ot
ð16:38Þ
 tÞ 3Kðr; tÞ ¼ u2 ðr; tÞ o w2 ðr; tÞ ¼ dðtÞ;
3Kðr;
o
ðtÞ ot ðtÞ ot
708 16 Introduction to Linear Viscoelasticity

hence, the distributions oui =ot; i ¼ 1; 2 are inverse in the convolution algebra K0þ ;
i.e.
 1  1
oui owi owi ou
¼ ; ¼ i ; i ¼ 1; 2: ð16:39Þ
ot ot ot ot
One may remark from (16.38) that

o2
ðu wi Þ ¼ dðtÞ; i ¼ 1; 2;
ot2 i
wherefrom
ui ðr; tÞ wi ðr; tÞ ¼ thðtÞ ¼ tþ ; i ¼ 1; 2: ð16:40Þ
ðtÞ

Using these results, one may construct differential and integro-differential


representations of the constitutive law of the isotropic viscoelastic solid, as it has
been shown in Sect. 4.2.2.3; the creep and relaxation distributions may be found
again.
As well, the general theory allows to construct unidimensional viscoelastic
models; the grouping in series or in parallel leads to such models, e.g.: Kelvin-
Voigt, Maxwell etc. Models of this kinds have been considered in Sect. 4.2.2.2.
We mention that, starting from a unidimensional model, one may construct bi-
and tridimensional models. To do this, it is convenient to use representations by
means of deviators and spherical tensors. Thus, if the constitutive law is given in
an integro-differential form, we replace the stress r by r0ij and the strain e by e0ij ;
hence, starting from the one-dimensional law
ðPðDÞdðtÞÞ r ¼ ðQðDÞdðtÞ þ gÞ e;
ðtÞ ðtÞ

one obtains the constitutive law


 
ðPðDÞdðtÞ þ f Þ r0ij ¼ QðDÞdðtÞ g e0ij ; i; j ¼ 1; 2; 3; ð16:41Þ
ðtÞ ðtÞ ðtÞ

which corresponds to a spatial linear viscoelastic solid. What concerns the relation
between the spherical tensors, that one may be taken arbitrarily, i.e.
o
r0 ¼ 3KðtÞ e0 ¼ w2 ðtÞ e0 : ð16:410 Þ
ðtÞ ðtÞ ot

If by a hydrostatic compression or by a uniform dilatation, the viscoelastic solid


behaves as an elastic one, then one has
w2 ¼ 3K0 hðtÞ; K0 [ 0;
16.1 Linear Viscoelastic Solids 709

wherefrom
r0 ¼ 3K0 e0 :
In the case of a Kelvin-Voigt model, we replace the constitutive law (4.176) by
o 0
r0ij ðr; tÞ ¼ Ee0ij ðr; tÞ þ g e ðr; tÞ; i; j ¼ 1; 2; 3; ð16:42Þ
ot ij
r0 ðr; tÞ ¼ 3K0 e0 ðr; tÞ: ð16:43Þ
The first equation may be written in the form

r0ij ðr; tÞ ¼ ðEdðtÞ þ gd0 ðtÞÞ e0ij ðr; tÞ; i; j ¼ 1; 2; 3; ð16:420 Þ


ðtÞ

too, while the creep and relaxation distributions are given by


hðtÞ h i
u1 ðtÞ ¼ 1  eðE=gÞt ; w1 ðtÞ ¼ EhðtÞ þ gdðtÞ; ð16:44Þ
E

1 1
w2 ðtÞ ¼ 3K0 hðtÞ; u2 ðtÞ ¼ w002 ¼ 3ðK0 d0 Þ1 ¼ hðtÞ: ð16:440 Þ
3K0
The constitutive law (4.182) of the Maxwell model leads, in the three–
dimensional case, to

o 0 1 0 1
eij ¼ d ðtÞ þ dðtÞ r0ij ð16:45Þ
ot E g ðtÞ

r0 ¼ 3K0 e0 ; ð16:450 Þ
the creep and relaxation distributions are
1 1 1
u1 ðtÞ ¼ hðtÞ þ thðtÞ; u2 ðtÞ ¼ hðtÞ; ð16:46Þ
E g 3K0

w1 ðtÞ ¼ EhðtÞeðE=gÞt ; w2 ðtÞ ¼ 3K0 hðtÞ: ð16:460 Þ

16.1.2 The Complex Moduli of Relaxation and Creep

Hereafter we will consider a linear viscoelastic solid subjected to the action of


periodic loads with a harmonic variation in time; we will thus consider unidi-
mensional and tridimensional solids to obtain the complex moduli of relaxation
and creep.
710 16 Introduction to Linear Viscoelasticity

16.1.2.1 One-Dimensional Case

Let us consider a linear unidimensional viscoelastic solid, the constitutive law of


which is given in the differential form
PðDÞr1 ¼ QðDÞe1 ; ð16:47Þ

where PðDÞ; QðDÞ are differential operators with constant coefficients in the space
of distributions. We suppose that the solid is subjected to a periodic stress with a
harmonic variation in time, of the form
r1 ðtÞ ¼ r0 hðtÞ cosðxt þ bÞ; ð16:48Þ
where r0 ¼ r0 ðxÞ is the amplitude, function of the pulsation x, while T ¼ 2p=x
is the period of the motion. One may associate a complex stress rðtÞ given by

rðtÞ ¼ r hðtÞeixt ; ð16:49Þ


where

r ¼ r0 eib : ð16:490 Þ
The corresponding strain is of the form
e1 ðtÞ ¼ e0 hðtÞ cosðxt þ aÞ; ð16:480 Þ

to which we associate in complex

eðtÞ ¼ e hðtÞeixt ; ð16:50Þ

e ¼ e0 eia : ð16:500 Þ
Hence, r1 ¼ Re rðtÞ; e1 ¼ Re eðtÞ; so that the constitutive law (16.47) takes the
form PðDÞRe rðtÞ ¼ QðDÞRe eðtÞ; i.e.
Re½PðDÞrðtÞ ¼ Re½QðDÞeðtÞ: ð16:470 Þ
Hence, instead of the equivalent Eqs. (16.47), (16.470 ), one may consider the
equation
PðDÞr ¼ QðDÞe; ð16:4700 Þ
a complex constitutive equation of the unidimensional viscoelastic solid.
Because r and e are distributions of function type of class C1 ðRÞ; excepting the
origin, where they have discontinuities of the first species; one has

P ðDÞr
~ ¼ QðDÞe;
~ t [ 0; ð16:51Þ

where D ~ ¼~ d=dt represents the derivative in the usual sense. Hence, substituting
0
(16.49 ), (16.50) in (16.51), we get
16.1 Linear Viscoelastic Solids 711

r ðxÞPðDÞe
~ ixt ¼ e ðxÞQðDÞe
~ ixt ;

wherefrom
r ðxÞPðixÞ ¼ e ðxÞQðixÞ; ð16:52Þ

because
~ ixt ¼ eixt PðixÞ; QðDÞe
PðDÞe ~ ixt ¼ eixt QðixÞ:

It results
r r ðxÞ QðixÞ
¼ ¼ ; t [ 0: ð16:520 Þ
e e ðxÞ PðixÞ
We call complex modulus of relaxation corresponding to the viscoelastic solid
the complex number
QðixÞ
E ðixÞ ¼ : ð16:53Þ
PðixÞ
By means of the complex modulus E ; the relation (16.520 ) becomes
r ¼ E e ð16:54Þ

or

r ¼ E e ; r ¼ r0 eib ; e ¼ e0 eia : ð16:540 Þ


We remark that the formulae above are similar to Hooke’s law, wherefrom it
results that to the viscoelastic solid acted upon by loads which have a harmonic
variation in time corresponds an elastic solid with the constitutive law (15.54). The
algebraic form of the complex modulus E is
E ðixÞ ¼ E1 ðxÞ þ iE2 ðxÞ; ð16:55Þ
where E1 ¼ Re E ; E2 ¼ Im E .
We call complex creep modulus the inverse J(ix) of the complex modulus E*, i.e.

JðixÞ ¼ ðE ðixÞÞ1 ; ð16:56Þ


one has
e e
JðixÞ ¼ J1 ðxÞ  iJ2 ðxÞ ¼ ¼ : ð16:560 Þ
r r
The complex relaxation modulus E constitutes a characteristic of the visco-
elastic solid, which may be determined experimentally too.
Let be the solid of Kelvin-Voigt type for which
PðDÞr1 ¼ QðDÞe1 ; ð16:57Þ
712 16 Introduction to Linear Viscoelasticity

where
d
PðDÞ ¼ 1; QðDÞ ¼ gD þ E; D ¼ : ð16:570 Þ
dt
We have PðixÞ ¼ 1; QðixÞ ¼ E þ ixg; consequently,
QðixÞ
E ðixÞ ¼ ¼ E þ ixg ¼ E1 ðxÞ þ iE2 ðxÞ;
PðixÞ

hence
E1 ðxÞ ¼ E; E2 ðxÞ ¼ xg: ð16:58Þ
The complex creep modulus is
1 E  ixg
JðixÞ ¼ ¼ ¼ J1 ðxÞ  iJ2 ðxÞ;
E ðixÞ E2 þ x2 g2
wherefrom
E xg
J1 ðxÞ ¼ ; J2 ðxÞ ¼ 2 : ð16:580 Þ
E2 2
þx g 2 E þ x 2 g2
Hence, the constitutive law of the Kelvin-Voigt model in the complex form reads

r ¼ E e; r ¼ r hðtÞeixt ; e ¼ e hðtÞeixt : ð16:59Þ


In the case of Maxwell’s model, in the constitutive law (16.57) one has
1 1
PðDÞ ¼ D þ ; QðDÞ ¼ D; ð16:60Þ
E g
hence
QðixÞ ix 1 1 i
E ðixÞ ¼ ¼ ¼ ; JðixÞ ¼  : ð16:61Þ
PðixÞ ix=E þ 1=g 1=E  i=xg E xg

16.1.2.2 Three-Dimensional Case

Let be a three-dimensional linear viscoelastic solid; the constitutive law of integro-


differential form is given by

ðP1 ðDÞdðtÞ þ f1 ðtÞÞ r0ij ðr; tÞ ¼ ðQ1 ðDÞdðtÞ þ g1 ðtÞÞ e0ij ðr; tÞ; ð16:62Þ
ðtÞ ðtÞ

ðP2 ðDÞdðtÞ þ f2 ðtÞÞ r0 ðr; tÞ ¼ ðQ2 ðDÞdðtÞ þ g2 ðtÞÞ e0 ðr; tÞ; ð16:620 Þ
ðtÞ ðtÞ
16.1 Linear Viscoelastic Solids 713

where fi ; gi ; i ¼ 1; 2; are distributions of function type in K0þ ; generated by con-


tinuous functions on [0,?) with discontinuities of the first species at the origin and
which admit the Laplace transform.
We assume that the solid is acted upon by loads with a harmonic variation, i.e.

r0ij ðr; tÞ ¼ r ij ðr; xÞhðtÞeixt ; e0ij ðr; tÞ ¼ e ij ðr; xÞhðtÞeixt ; ð16:63Þ

r0 ðr; tÞ ¼ r 0 ðr; xÞhðtÞeixt ; e0 ðr; tÞ ¼ e 0 ðr; xÞhðtÞeixt ; ð16:630 Þ

where r ij ; e ij ; r 0 ; e 0 are functions of r 2 X and x 2 R: Taking into account the


one-dimensional case, one may write
r0ij r ij Q1 ðixÞ þ ~
g1 ðixÞ
0 ¼ ¼ ; ð16:64Þ
eij eij P1 ðixÞ þ ~f1 ðixÞ

r0 r 0 Q2 ðixÞ þ g ~2 ðixÞ
¼ ¼ ; ð16:640 Þ
e0 e0 P2 ðixÞ þ ~f2 ðixÞ

where ~fk ðixÞ and ~


gk ðixÞ represent the Laplace transforms of the functions fk(t) and
gk(t), k = 1,2, calculated at the point p = ix.
The complex moduli of deformation and compressibility are given by
Q1 ðixÞ þ ~
g1 ðixÞ
2l ðixÞ ¼ ; ð16:65Þ
P1 ðixÞ þ ~f1 ðixÞ

Q2 ðixÞ þ ~
g2 ðixÞ
3K ðixÞ ¼ ; ð16:650 Þ
P2 ðixÞ þ ~f2 ðixÞ
thus, the constitutive law takes the form

r0ij ðr; tÞ ¼ 2l ðixÞe0ij ðr; tÞ; ð16:66Þ

r0 ðr; tÞ ¼ 3K ðixÞe0 ðr; tÞ: ð16:660 Þ


Denoting
2
k ðixÞ ¼ K ðixÞ  l ðixÞ;
3
one obtains the constitutive law
rij ¼ k ðixÞekk dij þ 2l ðixÞeij ; i; j ¼ 1; 2; 3; ð16:67Þ

where

rij ¼ r0ij þ r0 dij ¼ hðtÞeixt ðr ij þ r 0 dij Þ ¼ hðtÞeixt r


ij ; i; j ¼ 1; 2; 3; ð16:68Þ

eij ¼ e0ij þ e0 dij ¼ hðtÞeixt ðe ij þ e 0 dij Þ ¼ hðtÞeixt e


ij ; i; j ¼ 1; 2; 3; ð16:680 Þ
714 16 Introduction to Linear Viscoelasticity

e0 ¼ e 0 ðr; xÞeixt ; ð16:69Þ

where r
ij ; eij ; e0 are the amplitudes of the stresses, strains and volume strain,
respectively, when the loads have an harmonic variation. One obtains thus the
relations between the amplitudes

r
ij ðr; xÞ ¼ k ðixÞe0 ðr; xÞdij þ 2l ðixÞeij ðr; xÞ; i; j ¼ 1; 2; 3: ð16:70Þ

To obtain the creep and relaxation distributions, we take account of


o o
r0ij ¼ w1 ðtÞ eij ; r0 ¼ w2 ðtÞ e0 ;
ðtÞ ot ðtÞ ot

applying the Laplace transforms, it results


~ ðpÞ~e0 ðr; pÞ; i; j ¼ 1; 2; 3; r
~0ij ðr; pÞ ¼ pw
r ~ ðpÞ~e0 ðr; pÞ;
~0 ðr; pÞ ¼ pw
1 ij 2

so that
r ij e ij r 0 e 0
~0ij ¼
r ; ~e0ij ¼ ~0 ¼
;r ; ~e0 ¼ : ð16:71Þ
p  ix p  ix p  ix p  ix
One obtains thus
r ij ðr; ixÞ
~ ðixÞ ¼ 2l ðixÞ;
¼ ixw1
e ij ðr; ixÞ
r 0 ðr; ixÞ ~ ðixÞ ¼ 3K ðixÞ:
¼ ixw2
e 0 ðr; ixÞ
Because
w0k u0k ¼ dðtÞ; k ¼ 1; 2;
ðtÞ

by applying the Laplace transform and taking p = ix, we get

ðixÞ2 u ~ ðixÞ ¼ 1; k ¼ 1; 2;
~ k ðixÞwk

the Laplace images of the creep distributions are


1 1
~ k ðixÞ ¼ 
u ; k ¼ 1; 2: ð16:72Þ
2 ~
x wk ðixÞ

16.2 Limit Problems

We deal now with the formulation of the problems of the linear theory of visco-
elasticity, presenting then some limit problems.
16.2 Limit Problems 715

16.2.1 Formulation of the Problems of the Linear Theory


of Viscoelasticity

Hereafter we shall give a formulation in displacements of the problems of linear


viscoelasticity, the fundamental problems being similar with the problems of the
linear elasticity. We consider then the correspondence principle too.

16.2.1.1 Formulation in Displacements

To solve in distributions the fundamental problems of linear viscoelasicity,


one must transcribe in the distribution space K0þ all the quantities which occur.
We introduce thus the distributions of function type

ij ðr; tÞ ¼ hðtÞrij ðr; tÞ; eij ðr; tÞ ¼ hðtÞeij ðr; tÞ; i; j ¼ 1; 2; 3;
r
ð16:73Þ
i ðr; tÞ ¼ hðtÞui ðr; tÞ; F
u  i ðr; tÞ ¼ hðtÞFi ðr; tÞ; i ¼ 1; 2; 3:

These quantities are distributions of function type in K0þ with respect to t 2 R;


depending on the parameter r 2 X and having discontinuities of the first species at
the origin (t = 0). The derivatives in the usual sense with respect to the space
variables coincide with the same derivatives in the sense of distributions. What
concerns the derivatives with respect to time, one must take into account the
relation between the derivative in the usual sense ~o=ot and the derivative in the
sense of distributions o=ot:
Taking into account the initial conditions

~ui ðr; tÞ
0 o
ui ðr; tÞjt¼þ0 ¼ ui ;
  ¼ u_ 0i ; i ¼ 1; 2; 3; r 2 X; ð16:74Þ
ot 
t¼þ0

the equations of motion take the form

r  i ¼ q€
ij;j þ F  _
ui  q½u_ 0i dðtÞ þ u0i dðtÞ; i ¼ 1; 2; 3: ð16:75Þ
The constitutive law reads
ij ¼ kðtÞ ekk ðr; tÞdij þ 2lðtÞ eij ðr; tÞ; i; j ¼ 1; 2; 3;
r ð16:76Þ
ðtÞ ðtÞ

where
dw1 ðtÞ dw ðtÞ
2lðtÞ ¼ ; 3KðtÞ ¼ 2 ; l; K; w1 ; w2 2 K0þ : ð16:760 Þ
dt dt
The equations of Cauchy are
1

eij ¼ ui;j þ 
 uj;i ; i; j ¼ 1; 2; 3; ð16:77Þ
2
716 16 Introduction to Linear Viscoelasticity

and the equations of continuity read


eij;kl þ ekl;ij ¼ eik;jl þ ejl;ik ; i; j; k; l ¼ 1; 2; 3: ð16:78Þ

Eliminating the stresses and the strains, one obtains the equations in displacements
of Lamé’s type, i.e.

lðtÞ l
i;jj þ ½kðtÞ þ lðtÞ  i
uj;ji þ F
ðtÞ ðtÞ

¼ q€
ui 
 q½u_ 0i dðtÞ þ _
u0i dðtÞ; i ¼ 1; 2; 3: ð16:79Þ
Applying Laplace’s transforms in distributions with respect to t 2 R; one
obtains the equations

l ui;jj ðr; pÞ þ ½~
~ðpÞ~
 kðpÞ þ l ~ðpÞ~ ~
uj;ji ðr; pÞ þ F
 i

¼ qp  2~ 0 0
ui ðr; pÞ  qðu_ i þ pui Þ; i ¼ 1; 2; 3; ð16:790 Þ

similar to Lamé’s equations.


Taking into account the relations (16.73), one may write these equations in the
form

dw1 ðtÞ 1 dw1 ðtÞ dw2 ðtÞ i
ui;jj þ þ2 uj;ji þ 2F
dt ðtÞ 3 dt dt ðtÞ
h i
¼ 2q € _
ui  u_ 0i dðtÞ  u0i dðtÞ : ð16:7900 Þ

Differentiating with respect to xi and summing, one has


 i;i ¼ q€
ðk þ 2lÞ Dekk þ F _
ekk  q½_e0kk dðtÞ þ e0kk dðtÞ; ð16:80Þ
ðtÞ

where the initial conditions have been put too for the volume strain.
In particular, in the case of the quasistatic problem, the volume forces and the
initial conditions being equal to zero, one obtains
ðk þ 2lÞ Dekk ¼ 0: ð16:81Þ
ðtÞ

Because k þ 2l 6¼ 0; it follows that


Dekk ¼ 0; ð16:810 Þ
hence ekk is a harmonic distribution. In the same case, applying the operator D to
the Eq. (16.79), it results
ui ¼ 0; i ¼ 1; 2; 3;
DD ð16:82Þ
so that the displacements are biharmonic distributions.
16.2 Limit Problems 717

If we introduce the operator of d’Alembert type

o2
h 1 ¼ D  qðk þ lÞ1 ; ð16:83Þ
ðtÞ ot2

then the Eq. (16.80) takes the form

h 1ekk þ ½kðtÞ þ 2lðtÞ1 F  i;i


ðtÞ

o
¼ q ½kðtÞ þ 2lðtÞ1 e_ 0kk þ ½kðtÞ þ 2lðtÞ1 e0kk ; ð16:84Þ
ot

where we assumed that kðtÞ þ 2lðtÞ 2 K0þ is invertible in the convolution algebra
K0þ :

16.2.1.2 Methods of Solutions of the Problems of Linear


Viscoelasticity. Principle of Correspondence

We have seen that in solving the limit problems of the linear viscoelasticity the
Laplace transform plays an important rôle. Comparing the formulation of the
problems of linear viscoelasticity to the similar ones of linear elasticity, one
observes that the difference consists in the constitutive law; but the Laplace image
of the constitutive law of linear viscoelastic bodies has the same mathematical
structure as the corresponding linear elastic law of Hooke. This remark led
T. Alfrey [9] and E. H. Lee [15] to the formulation of the principle of corre-
spondence as a method of solving the above mentioned problems. Vito Volterra
[18] dealt with this principle and Yu. N. Rabotnov and A. A. Ilyushin [6] called it
the principle of Volterra. This principle has been generalized and used by
W. T. Read [16], H. S. Tsien [17] and J. Mandel [4].
At it has been shown in [9], [16], [17], to a quasistatic problem of viscoelas-
ticity one may associate an elastostatic problem. W. T. Read [16] showed that,
using the Fourier transform, one can associate, in general, to a dynamic visco-
elastic problem an elastostatic problem too, but with a special character. Thus, to
the inertial forces in the viscoelastodynamics correspond volume forces direct
proportional to the components of the displacement in the elastostatic problem.
Let us thus remember the common equations (motion, Cauchy, continuity) of
both isotropic elastic and viscoelastic solids, i.e.:

r  i ðr; tÞ ¼ q€
ij;j ðr; tÞ þ F  _
ui  q½u_ 0i dðtÞ þ u0i dðtÞ; i ¼ 1; 2; 3; ð16:85Þ

eij ¼ ui;j þ 
 uj;i ; i; j ¼ 1; 2; 3; ð16:850 Þ
2
eij;kl þ ekl;ij ¼ eil;jk þ ejk;il ; i; j; k; l ¼ 1; 2; 3; ð16:8500 Þ
718 16 Introduction to Linear Viscoelasticity

where rij ; eij ;   i ; i = 1,2,3, are distributions in K0þ ; while r 2 X  R3 is a


ui ; F
parameter and the initial conditions are included.
For the mixed fundamental problem of elasticity or viscoelasticity, we intro-
duce the boundary conditions (S ¼ Su [ Sr ; t 2 R)
n
ui ðr; tÞ ¼ 
 ui ðr; tÞ on Su ; i ¼ 1; 2; 3; ð16:86Þ
n
ij ðr; tÞnj ðrÞ ¼ 
r pi ðr; tÞ on Sr ; i; j ¼ 1; 2; 3; ð16:860 Þ
n n
where  pi ; i ¼ 1; 2; 3 are distributions of function type in K0þ ; while the frontier
ui ; 
S is fixed.
If the considered solid is an elastic one, then one must add Hooke’s law

0ij ðr; tÞ ¼ 2le0ij ðr; tÞ; i; j ¼ 1; 2; 3;


r ð16:87Þ

0 ðr; tÞ ¼ 3Ke0 ðr; tÞ;


r ð16:870 Þ
where l(r), K(r) are the elastic constants of the isotropic, non-homogeneous linear
elastic body. Thus, we have the complete system of equations of the mixed fun-
damental problem of the linear elastodynamics.
If we consider the constitutive law of the isotropic, non-linear viscoelastic body

o
0ij ðr; tÞ ¼ 2l ðr; tÞ e0ij ðr; tÞ ¼
r w ðr; tÞ e0ij ðr; tÞ; i; j ¼ 1; 2; 3; ð16:88Þ
ðtÞ ot 1 ðtÞ

o
0 ðr; tÞ ¼ 3K ðr; tÞ e0 ðr; tÞ ¼
r w ðr; tÞ e0 ðr; tÞ ð16:880 Þ
ðtÞ ot 2 ðtÞ

too, then one has the complete system of equations of the mixed fundamental
problem of the linear elastoviscodynamics.
Obviously, for both solids we assume that the geometric form is the same, as
well as the volume forces, the boundary and the initial conditions. We mention
also that l, K and l*, K* are the same constants but with different notations, to
may distinguish the two solids.
Applying the Laplace transform in distributions to the constitutive laws (16.87),
(16.870 ), (16.88), (16.880 ) one obtains

0ij ðr; pÞ ¼ 2lðrÞ~


~
r e0ij ðr; pÞ; i; j ¼ 1; 2; 3; ð16:89Þ

0 ðr; pÞ ¼ 3KðrÞ~
~
r e0 ðr; pÞ; ð16:890 Þ

0ij ðr; pÞ ¼ 2l ðr; pÞ~


~
r e0ij ðr; pÞ; i; j ¼ 1; 2; 3; ð16:90Þ

0 ðr; pÞ ¼ 3K ðr; pÞ~


~
r e0 ðr; pÞ: ð16:900 Þ
16.2 Limit Problems 719

Comparing the Eqs. (16.89) and (16.890 ) with (16.90) and (16.900 ), respectively,
we see that they have the same structure, which leads to the bijection
~ ðr; pÞ;
lðrÞ $ l ð16:91Þ
~ ðr; pÞ
KðrÞ $ K ð16:910 Þ

between the quantities which characterize the two media; the complex variable
p plays the rôle of a parameter.
The other relations are the same for both media. It results that, knowing the
solution of a problem in elastodynamics, one may obtain the solution for the
analogue problem in elastoviscodynamics.
Hence, in conformity with the principle of correspondence, one must solve
firstly the correspondent problem of elastodynamics, considering then the Laplace
image of the solution thus obtained. Then, the elastic constants l and K are
replaced by Laplace’s images l ~ and K ~ ; respectively; one obtains thus the
Laplace image of the solution of the mixed problem of viscoelastodynamics.
Applying now the inverse Laplace transform, one gets the solution of the searched
problem.
A class of problems to which the principle of correspondence may be suc-
cessfully applied is that of quasi-static problems.

16.2.1.3 Quasi-Static Problems

Neglecting the inertial forces q€


u and the initial conditions, the equations of motion
are simplified and the problems thus formulated are called quasi-static. In this
case, to a quasi-static problem of viscoelasticity corresponds, by using the Laplace
transform, a static problem in elasticity.
In the quasi-static case, the complete system of equations of an isotropic linear
viscoelastic solid is

r  i ðr; tÞ ¼ 0; i ¼ 1; 2; 3;
ij;j ðr; tÞ þ F ð16:92Þ

1 
eij ðr; tÞ ¼ ui;j ðr; tÞ þ 
 uj;i ðr; tÞ ; i; j ¼ 1; 2; 3; ð16:920 Þ
2
eij;kl ðr; tÞ þ ekl;ij ðr; tÞ ¼ eil;jk ðr; tÞ þ ejk;il ðr; tÞ; i; j; k; l ¼ 1; 2; 3; ð16:9200 Þ

0ij ðr; tÞ ¼ 2l ðr; tÞ e0ij ðr; tÞ; i; j ¼ 1; 2; 3;


r ð16:93Þ
ðtÞ

0 ðr; tÞ ¼ 3K ðr; tÞ e0 ðr; tÞ;


r ð16:930 Þ
ðtÞ

n
 ui ðr; tÞ on Su ; i ¼ 1; 2; 3; t 2 R;
ui ðr; tÞ ¼  ð16:94Þ
720 16 Introduction to Linear Viscoelasticity

n
ij ðr; tÞnj ðrÞ ¼ 
r pi ðr; tÞ on Sr ; i; j ¼ 1; 2; 3: ð16:940 Þ
Applying the Laplace transform in distributions, one obtains

~ ~
 i ðr; pÞ ¼ 0; i ¼ 1; 2; 3;
ij;j ðr; pÞ þ F
r ð16:95Þ

1 
~eij ðr; pÞ ¼ ~
ui;j ðr; pÞ þ ~
 uj;i ðr; pÞ ; i; j ¼ 1; 2; 3;
 ð16:950 Þ
2
eij;kl ðr; pÞ þ ~
~ ekl;ij ðr; pÞ ¼ ~
eil;jk ðr; pÞ þ ~
ejk;il ðr; pÞ; i; j; k; l ¼ 1; 2; 3; ð16:9500 Þ

0ij ðr; pÞ ¼ 2~
~
r l ðr; pÞ~
e0ij ðr; pÞ; i; j ¼ 1; 2; 3; ð16:96Þ

~ ~ ðr; pÞ~
0 ðr; pÞ ¼ 3K
r e0 ðr; pÞ; ð16:960 Þ
n
~ u~i ðr; pÞ on Su ; i ¼ 1; 2; 3;
ui ðr; pÞ ¼ 
 ð16:97Þ
n
~
ij ðr; pÞnj ðrÞ ¼ ~
r pi ðr; pÞ on Sr ; i; j ¼ 1; 2; 3:
 ð16:970 Þ
Let us consider now following bijections between the quantities in elastostatics
and the Laplace images of the quantities in viscoelasticity

~
rij ðrÞ $ r ~
 i ðr; pÞ; i; j ¼ 1; 2; 3;
ij ðr; pÞ; Fi ðrÞ $ F ð16:98Þ

ui ðr; pÞ; eij ðrÞ $ ~


ui ðrÞ $ ~
 eij ðr; pÞ; i; j ¼ 1; 2; 3; ð16:980 Þ
n n
n n
u~i ðr; pÞ on Su ; pi ðrÞ $ 
ui ðrÞ $  p~i ðr; pÞ on Sr ; i; j ¼ 1; 2; 3; ð16:9800 Þ

lðrÞ $ l ~ ðr; pÞ:


~ ðr; pÞ; KðrÞ $ K ð16:99Þ
Hence, in conformity with the principle of correspondence, one must solve
firstly the associated static problem; then the mechanical quantities rij ; ui ; eij ; Fi ;
n n
n n ~
ui ; pi are replaced by the quantities r~ i ; ~
~
ij ; u  i; u
eij ; F ~i ; p~i and the elastic constants l,
K by the quantities l ~ ; respectively. One obtains thus the Laplace images of
~ ; K
the solution of the quasi-static problem of viscoelasticity; by applying the inverse
Laplace transform one determines the searched solution.

16.2.2 Applications

We shall apply now the above results in some particular cases, i.e.: the viscoelastic
disk in uniform circular motion, the viscoelastic tube with thick walls and the
viscoelastic cylindrical bar subjected to its own weight and to an axial force.
16.2 Limit Problems 721

16.2.2.1 Viscoelastic Disk in Uniform Circular Motion

Let us consider a circular disk of radius re of a homogeneous and isotropic linear


viscoelastic material in uniform motion of rotation with an angular velocity x
about an axis normal to the plane of the disk at its centre. The volume forces which
act upon the disk along the vector radius are the centrifugal forces Fr ¼ qx2 rhðtÞ;
where q is the density (Fig. 16.1). Neglecting the forces of inertia, one obtains a
quasi-static problem with the conditions
rrr jr¼re ¼ 0; ur jr¼0 ¼ 0; ð16:100Þ

If the disk is a linear elastic one, then it is subjected to a plane state of stress; using
Airy’s formulae, one obtains
" #
2
1 2 k þ 2l ðk þ 2lÞ
ur ¼ qx r ð2re2  r 2 Þ þ re2 : ð16:101Þ
8 4lðk þ lÞ 4lðk þ lÞð3k þ 2lÞ

Applying the principle of correspondence, the Laplace image of the displace-


ment ur of the linear viscoelastic disk is
" #
1 2 ~
k þ 2~
l ð~k þ 2~
lÞ2
2 2 2
ur ðr; pÞ ¼ qx r ð2re  r Þ
~ þ re ;
8 lðk
4~ ~þl~Þ lð~k þ l
4~ ~Þð3~k þ 2~

where

2~ ~ ðpÞ; 3~
l ¼ pw k þ 2~
l ¼ 3K ~ ðpÞ;
~ ¼ pw ð16:102Þ
1 2

w1 ; w2 being the relaxation distributions of the viscoelastic solid of which is made


the disk. Taking into account (16.102), one obtains

Fig. 16.1 Viscoelastic disk


in uniform circular motion x2

Fr

P
r θ
O
re x1
722 16 Introduction to Linear Viscoelasticity

ð~ lÞ2
k þ 2~ ~ l ~k þ 2~
k þ 2~ l
¼
4~ ~
l ðk þ l ~
~Þð3k þ 2~ ~
lðk þ l
lÞ 4~ ~
~Þ ð3k þ 2~

~ þw
ð2w ~ Þ2
¼ 1 2 ~
¼ GðpÞ; ð16:103Þ
~ w
3pw ~ ðw~ þ 2w ~ Þ
1 2 1 2

~
k þ 2~
l 2w~ þw ~
1 2 ~
¼ ¼ HðpÞ: ð16:1030 Þ
lð~
4~ kþl ~ ðw
~ Þ pw 1
~ þ 2w
1
~ Þ
2

~
Denoting GðtÞ ¼ L1 ½GðpÞ; ~
HðtÞ ¼ L1 ½HðpÞ; it results
1  
~ ~
ur ¼ L½ur  ¼ qx2 r ð2re2  r 2 ÞHðpÞ ~
þ re2 GðpÞ ;
8
wherefrom
1  
ur ðr; tÞ ¼ qx2 r ð2re2  r 2 ÞHðtÞ þ re2 GðtÞ ; ð16:104Þ
8
i.e. the state of displacement corresponding to the viscoelastic solid.
Let us suppose now that the viscoelastic body is of Kelvin-Voigt type, which
behaves as an elastic one at simple compression or traction; one has thus w2 ¼
3K0 hðtÞ; K0 [ 0, hence w ~ ðpÞ ¼ 3K0 =p: The relaxation distribution w ¼ gdðtÞ þ
2 1
~
EhðtÞ; leads to w1 ðpÞ ¼ ðgp þ EÞ=p:
Substituting in (16.103), (16.1030 ), one obtains

~ ½2ðgp þ 4EÞ þ 3K0 2


GðpÞ ¼ ;
9K0 ðgp þ EÞðgp þ E þ 6K0 Þ
ð16:105Þ
~ 2ðgp þ EÞ þ 3K0
HðpÞ ¼ ;
ðgp þ EÞðgp þ E þ 6K0 Þ

applying the inverse Laplace transform, we get


4 hðtÞ  ðE=gÞt 
GðtÞ ¼ dðtÞ þ e  9e½ðEþ6K0 Þ=gt ;
9K0 6g
ð16:1050 Þ
hðtÞ  ðE=gÞt 
HðtÞ ¼ e þ 3e½ðEþ6K0 Þ=gt :
2g
Substituting in (16.104), we may write
 
qx2 re2 r qx2 r 7 2
ur ðr; tÞ ¼ dðtÞ þ hðtÞ r  r eðE=gÞt
2
18K0 16g 3 e

2 2 ½ðEþ6K0 Þ=gt
þ 3ðre  r Þe ;
16.2 Limit Problems 723

but Dirac’s distribution dðtÞ puts in evidence only that the circular motion of the
disk is due at t ¼ 0 to a sudden cause. So that, the radial displacement of the linear
viscoelastic disk of Kelvin-Voigt type is given by

qx2 r hg  i
ur ðr; tÞ ¼ hðtÞ re2  r 2 eðE=gÞt þ 3ðre2  r 2 Þe½ðEþ6K0 Þ=gt ð16:106Þ
16g 3
for t [ 0:

16.2.2.2 Viscoelastic Tube with Thick Walls

Let us consider now the Lamé problem for a homogeneous and isotropic visco-
elastic tube subjected to an uniformly distributed internal and external compres-
sion of variable intensities pi ðtÞ and pe ðtÞ on the faces of radii ri and rc ;
respectively (Fig. 16.2). We neglect the volume forces, as well the forces of
inertia. The intensities pi ðtÞ and pe ðtÞ are distributions in K0þ :
We assume that the tube is subjected to a plane state of strain; because of the
axial symmetry, one has rrr ¼ rrr ðr; tÞ; rhh ¼ rhh ðr; tÞ; rrh ¼¼ 0: The equation
of motion becomes
o 1
rrr ðr; tÞ þ ½rrr ðr; tÞ  rhh ðr; tÞ ¼ 0 ð16:107Þ
or r
and the boundary conditions are of the form
rrr ðr; tÞjr¼ri ¼ pi ðtÞ; rrr ðr; tÞjr¼re ¼ pe ðtÞ: ð16:108Þ

In the elastic case, the biharmonic function of Airy is of the form

Fðr; tÞ ¼ AðtÞ ln r þ BðtÞr 2 þ CðtÞ;

Fig. 16.2 Viscoelastic tube pe(t)


with thick walls

P re
ri
O
pi (t)
724 16 Introduction to Linear Viscoelasticity

the stresses are given by

1 oF o2 F
rrr ¼ ; rhh ¼ 2 :
r or or
One obtains the stresses

1 ri2 re2 ri2 pi ðtÞ  re2 pe ðtÞ


rrr ¼ ½ p e ðtÞ  p i ðtÞ  þ ;
r 2 re2  ri2 re2  ri2
ð16:109Þ
1 r2 r2 r 2 pi ðtÞ  re2 pe ðtÞ
rhh ¼  2 2 i e 2 ½pe ðtÞ  pi ðtÞ þ i :
r re  ri re2  ri2
Finally, the Laplace image of the radial displacement becomes
~
k ~k þ 2~
l
ur ¼ r~
~ rrr ðr; pÞ þ r~
rhh ðr; pÞ ;
4~ ~
lð k þ l
~Þ ~
lðk þ l
4~ ~Þ

one has

3K ~ ðpÞ ¼ 2~
~ ¼ pw l þ 3~ ~ ðpÞ:
l ¼ pw
k; 2~ ð16:110Þ
2 1

We observe that
~
k ~ w
w ~
~ 1 ðpÞ ¼
G ¼ 2 1
;
~þl
lðk
4~ ~ ðw
~ Þ pw ~ þ 2w ~ Þ
1 1 2

as well, one uses the Laplace image HðpÞ~ given by (16.105). Applying the inverse
Laplace transform, one obtains

ur ðr; tÞ ¼ r rrr ðtÞ G1 ðtÞ  rhh ðtÞ HðtÞ : ð16:111Þ
ðtÞ ðtÞ

Considering now a Kelvin-Voigt model, as in the previous subsection one


obtains

~ 1 ðpÞ ¼ ðgp þ EÞ þ 3K0


G ;
ðgp þ EÞðgp þ E þ 6K0 Þ
wherefrom
hðtÞ h ðE=gÞt i
G1 ðtÞ ¼ e  3e½ðEþ6K0 Þ=gt ; ð16:112Þ
2g
while HðtÞ is given by (16.1050 ). Substituting now in (16.111), one gets the radial
displacement of the linear viscoelastic tube of Kelvin-Voigt type.
16.2 Limit Problems 725

16.2.2.3 Viscoelastic Cylindrical Bar Subjected to its Own Weight


and to an Axial Force

Let be a homogeneous, isotropic linear viscoelastic cylinder of arbitrary cross


section of length l, in a vertical position (Fig. 16.3). The origin O of the
co–ordinate axes coincides with the centre of gravity of the superior cross section
(x3 = 0), which is built in; the inferior cross section (x3 = l) is acted upon by a
uniform distributed load of resultant P(t) along the Ox3 -axis. One takes into
account the own weight c ¼ qghðtÞ too. Neglecting the inertial forces, one is led to
a quasi-static viscoelastic problem.
To may apply the principle of correspondence, we consider firstly the case of
the elastic cylinder. Finally, one has
~
11 ¼ r
r ~
22 ¼ r~
23 ¼ r~
31 ¼ r ~12 ¼ 0;
qg ~
PðpÞ
~
33 ðr; pÞ ¼
r ðl  x3 Þ þ ;
p A
where A is the area of the cross section of the cylindrical bar. It results that the
corresponding stresses are

11 ¼ r
r 22 ¼ r
23 ¼ r
31 ¼ r
12 ¼ 0;
PðtÞ ð16:113Þ
33 ðr; tÞ ¼ qghðtÞðl  x3 Þ þ
r ;
A
wherefrom one obtains the state of displacement.

Fig. 16.3 Viscoelastic


cylindrical bar subjected to its x1
O
own weight and to an axial
force x2
l

ρg

P(t)
x3
726 16 Introduction to Linear Viscoelasticity

~
kðpÞx1 qg ~
PðtÞ
~
u1 ðr; pÞ ¼ 
 ðl  x3 Þ þ ;
lð3~
2~ lÞ p
k þ 2~ A
ð16:114Þ
~
kðpÞx2 qg ~
PðtÞ
~
u2 ðr; pÞ ¼ 
 ðl  x3 Þ þ ;
lð3~
2~ lÞ p
k þ 2~ A

~
kðpÞqg x21 þ x22
~
u3 ðr; pÞ ¼ 

lð3~
2~ lÞ p
k þ 2~
~
kþl ~ qg 2 ~
PðtÞ
þ ðl  x23 Þ þ 2x3 ; ð16:1140 Þ
lð3~
2~ lÞ p
k þ 2~ A

where
~ ; 3K
l ¼ pw
2~ ~ :
~ ¼ pw ð16:115Þ
1 2

Because
dui ðtÞ dwi ðtÞ
¼ dðtÞ; i ¼ 1; 2; ð16:116Þ
dt ðtÞ dt

one has

~ ¼ 1
u
~ iwi : ð16:1160 Þ
p2
Taking into account (16.116), (16.1160 ), one obtains
~
k 2
¼ p½u
~ 1 ðpÞ  u
~ 2 ðpÞ;
~
~ð3k þ 2~
l lÞ 3
~
kþl ~ p
¼ ½2~
u1 ðpÞ þ u~ 2 ðpÞ;
~
~ð3k þ 2~
l lÞ 3

wherefrom
" #
~
k 2 du1 ðtÞ du2 ðtÞ
1
L ¼  ;
~ð3~
l k þ 2~
lÞ 3 dt dt
" #
~
kþl ~ 1 du1 ðtÞ du2 ðtÞ
1
L ¼ 2 þ :
~ð3~
l k þ 2~
lÞ 3 dt dt

Applying the inverse Laplace transform, the relations (16.73), (16.1160 ) lead to
16.2 Limit Problems 727


1 du1 ðtÞ du2 ðtÞ PðtÞ
u1 ðr; tÞ ¼ 
  qghðtÞðl  x3 Þ þ x1
3 dt dt ðtÞ A

qg x1 du1 ðtÞ du2 ðtÞ PðtÞ
¼  x1 ðl  x3 Þðu1  u2 Þ   ;
3 3 dt dt ðtÞ A

qg x2 du1 ðtÞ du2 ðtÞ PðtÞ ð16:117Þ
u2 ðr; tÞ ¼  x2 ðl  x3 Þðu1  u2 Þ 
  ;
3 3 dt dt ðtÞ A
qg 2

u3 ðr; tÞ ¼ 
 x1 þ x22 ðu1  u2 Þ
3
x3 du1 ðtÞ du2 ðtÞ PðtÞ
þ 2 þ qghðtÞðl  x3 Þ þ 2 :
6 dt dt ðtÞ A
In particular, if PðtÞ is a force constant in time, then PðtÞ ¼ P0 hðtÞ and the
formulae (16.117) become

1 P0
u1 ðr; tÞ ¼  ðu1  u2 Þ qgx1 ðl  x3 Þ þ
 ;
3 A

1 P0
u2 ðr; tÞ ¼  ðu1  u2 Þ qgx2 ðl  x3 Þ þ
 ;
3 A ð16:118Þ
qg 2 2

u3 ðr; tÞ ¼ 
 x þ x2 ðu1  u2 Þ
3 1
x3 P0
þ ð2u1 þ u2 Þ qgðl  x3 Þ þ 2 :
6 A

References

A. Books

1. Bland, D.R.: The Theory of Linear Viscoelasticity. Pergamon Press, London (1960)
2. Kecs, W.:Elasticitate si vâscoelasticitate (Elasticity and Viscoelasticity). Ed. Tehnică,
Bucuresti (1986)
3. Kolski, H.: Volny napryazheniya v tverdykh telakl (Stress Waves in Solid Bodies). Inostr.
Lit., Moskva (1955)
4. Mandel, J.: Cours de mécanique des milieux continus. I, II. Gauthier–Vilars, Paris (1966)
5. Rabotnov, Yu. N.: Elements of Hereditary Solid Mechanics. Mir Publ, Moecow (1980)
6. Rabotnov, Yu. N., A, Ilyushin: Methoden der Viskoelastizitätstheorie. VEB Fachbuchverlag,
Leipzig (1970)
7. Riesz, F., Sz Nagy, B.: Leçons d’analyse fonctionnelle. Budapest (1972)
728 16 Introduction to Linear Viscoelasticity

B. Papers

8. Alain, M.: Viscoélasticité linéaire et fonctions complètement monotones. J. de Mécanique 4,


12 (1978)
9. Alfrey, T.: Non-homogeneous stresses in viscoelastic media. Quart. Appl. Math. 2, 113
(1944)
10. Gurtin, M., Sternberg, E.: On the linear theory of viscoelasticity. Arch. Rat. Mech. Anal. 11,
291 (1962)
11. Kecs, W.: On the equations on the one-dimensional problem of the linear viscoelastic solid.
Letters Appl. Engng. Sci. Pergamon Press, 5, 389 (1977)
12. Kecs, W.: On models of one-dimensional linear viscoelastic solids. Rev. Roum. Sci. Techn.;
Méc. Appl. 24, 3 (1979)
13. Kecs, W.: Sur les équations des solides linéaires viscoélastiques. Rev. Roum. Sci. Techn.;
Méc. Appl., 25, 3 (1980)
14. Kecs, W.: On grouping models of viscoelastic solids. Bull. Math. Soc. Sci. Math. Roum.
28(76) (1984)
15. Lee, E.H.: Stress analysis in viscoelastic bodies. Quart. Appl. Math. 13, 183 (1955)
16. Read, W.T.: Stress analysis for compressible viscoelastic materials. J. Appl. Phys. 21, 671
(1950)
17. Tsien, H.S.: A generalization of Alfrey’s theorem for viscoelastic media. Quart. Appl. Math.
8, 104 (1950)
18. Volterra, V.: Sulle equazioni integro-differenziali della teoria dell’elasticità. Atti. R. Accad.
Lincei. 18, 295 (1909)
Appendix

In what follows, we shall deal with elements of tensor calculus, with the introduction
of curvilinear co-ordinates and with some notions of the theory of distributions;
as well, we shall introduce certain notations and integrals.

A.1 Elements of Tensor Calculus

We shall make some considerations concerning the algebra and the analysis
of tensors, dealing successively with scalars, tensors of first order and tensors of
nth order. Various particular tensors will be put into evidence. As well, we will
express in a tensor form the algebraic and analytic operations effected with
vectors. We consider only orthogonal affine tensors in R3 , the corresponding
notions being sufficient for the mathematical representation of the mechanical
phenomena with which we deal.

A.1.1 Orthogonal Affine Tensors

We shall state hereafter some considerations concerning the changes of


co-ordinate axes, the introduction of scalars, of tensors of first order, as well as
of tensors of nth order. We will mention some particular cases too. The Einstein’s
summation convention, in conformity with which the twice existence of an index
(called dummy index) in a monomial indicates the summation with respect to this
index, is used.

P. P. Teodorescu, Treatise on Classical Elasticity, Mathematical and Analytical 729


Techniques with Applications to Engineering, DOI: 10.1007/978-94-007-2616-1,
Ó Springer Science+Business Media Dordrecht 2013
730 Appendix

Fig. A.1 Change of


co-ordinate axes x3
x'3
a

P
i3 r i'2
x'2
i'3
i1 x2
O i2
i1'

x1
x'1

A.1.1.1 Changes of Co-ordinate Axes

Let be a positive orthonormed basis B of unit vectors (versors) ij ; j ¼ 1; 2; 3, hence


a system of orthogonal Cartesian axes Oxj ; j ¼ 1; 2; 3. Let, as well, be a second
positive orthonormed basis B0 of unit vectors i0k ; k ¼ 1; 2; 3, to which corresponds
a second system of orthogonal Cartesian axes Ox0k ; k ¼ 1; 2; 3 (Fig. A.1). We thus
have
i0k ¼ akj ij ; k ¼ 1; 2; 3; ðA:1Þ

where we have introduced the cosines


akj ¼ i0k  ij ¼ cosði0k ; ij Þ; j; k ¼ 1; 2; 3; ðA:2Þ

while ‘‘’’ indicates the scalar (dot) product of two vectors, represented by bold
letters; analogically, we may write
ij ¼ akj i0k ; j ¼ 1; 2; 3: ðA:10 Þ
A scalar product of the relation (A.1) by i0l and, analogically, a scalar product of
the relation (A.10 ) by il ; l ¼ 1; 2; 3, leads to the relations between the cosines
introduced above
aij aik ¼ aji aki ¼ djk ; ðA:3Þ

where djk is Kronecker’s symbol, defined in the form



1; j ¼ k;
djk ¼ ðA:4Þ
0; j 6¼ k:
Let be the position vector r of a point P, with respect to the origin O; we may
write, in the two systems of co-ordinates,
r ¼ xj ij ¼ x0k i0k ; ðA:5Þ
A.1 Elements of Tensor Calculus 731

where xj ; x0k ; j; k ¼ 1; 2; 3, are the co-ordinates of the point P in the two systems of
co-ordinates, respectively. Taking into account (A.1) and (A.10 ), respectively, we
get the relations which allow to pass from a system of co-ordinates to another one,
in the form
x0k ¼ xj akj ; xj ¼ x0k akj ; ðA:6Þ

these linear transformations are orthogonal, their coefficients verifying the relations
(A.3).
Hence, it results easily
ox0k oxj
¼ ¼ akj : ðA:7Þ
oxj ox0k

A.1.1.2 Scalar. Tensor of First Order

Let be a function U ¼ Uðx1 ; x2 ; x3 Þ; by a change of co-ordinates of the form (A.6),


one obtains the function U 0 ¼ U 0 ðx01 ; x02 ; x03 Þ. If the condition
U0 ¼ U ðA:8Þ

is fulfilled, hence if the considered function is invariant to a change of orthogonal


Cartesian co-ordinates, then we say that U is a tensor of zeroth order (or a scalar).
In the case of a constant function, one obtains a mathematical entity characterized
by sense (sign) and modulus.
Let now be the vector a ¼ aðrÞ (Fig. A.1); taking into account the two systems
of co-ordinates, we may write
a ¼ aj ij ¼ a0k i0k ; ðA:9Þ

where aj ¼ aj ðx1 ; x2 ; x3 Þ; a0k ¼ a0k ðx01 ; x02 ; x03 Þ; j; k ¼ 1; 2; 3. Taking into account
(A.1) and (A.10 ), respectively, we get the relations which allow passing from the
components of the vector a, with respect to the basis B, to the components of the
same vector with respect to the basis B0 and inversely, in the form
a0k ¼ aj akj ; aj ¼ a0k akj ; k; j ¼ 1; 2; 3: ðA:10Þ

In general, the sets of three functions, in each basis, which are transformed
according to the formulae (A.10), by a change of orthogonal Cartesian
co-ordinates, are the components of a tensor of first order (or the components of
a vector). A tensor of first order may be represented by its components with regard
to a certain basis B in the form of a column matrix or by means of its transpose
(a row matrix)
732 Appendix

32
a1
a  ½ai   4 a2 5  ½ a1 a2 a3  T : ðA:11Þ
a3
Taking into account (A.7), we may express the transformation relations in the
form
ox0k oxj
a0k ¼ aj ; aj ¼ a0k 0 ; j; k ¼ 1; 2; 3: ðA:100 Þ
oxj oxk

A.1.1.3 Tensor of Order n

Extending the result obtained in the previous section, we call tensor of second
order the totality of the sets of 32 ¼ 9 magnitudes aij ; i; j ¼ 1; 2; 3, which, by a
change of orthogonal Cartesian co-ordinates, become
a0kl ¼ aij aki alj ; aij ¼ a0kl aki alj ; i; j; k ¼ 1; 2; 3; ðA:12Þ

hence, the tensor of second order is a mathematical entity a which is expressed by


its components with respect to a positive orthonormed basis B in the form of a
square matrix
2 3
  a11 a12 a13
a  aij  4 a21 a22 a23 5: ðA:13Þ
a31 a32 a33

Similarly, a tensor of third order is the totality of the sets of 33 ¼ 27 magnitudes


aijk ; i; j; k ¼ 1; 2; 3, which are transformed by a change of orthogonal Cartesian
co-ordinates by means of the relations
a0lmn ¼ aijk ali amj ank ; aijk ¼ a0lmn ali amj ank ; i; j; k; l; m; n ¼ 1; 2; 3; ðA:14Þ

as well, it is a mathematical entity a, represented by its components aijk with


respect to an arbitrary positive orthonormed basis B.
In general, a tensor of nth order is the totality of the sets of 3n magnitudes
ai1 i2 ...in ; ik ¼ 1; 2; 3; k ¼ 1; 2; . . .; n, which are transformed by a change of
orthogonal Cartesian co-ordinates in the form
a0j1 j2 ...jn ¼ ai1 i2 ...in aj1 i1 aj2 i2 . . .ajn in ; ðA:15Þ

ai1 i2 ...in ¼ a0j1 j2 ...jn aj1 i1 aj2 i2 . . .ajn in ; ðA:150 Þ

it is obvious that such a tensor is a mathematical entity which is represented by its


components ai1 i2 ...in with respect to an arbitrary orthonormed Cartesian basis B.
A.1 Elements of Tensor Calculus 733

The tensors thus defined are Euclidean tensors.


Let be a  ½ai1 i2 ...in T a tensor of nth order. If one has the relation
ai1 i2 ...ij ...ik ...in ¼ ai1 i2 ...ik ...ij ...in ðA:16Þ

in a basis B, then we say that the tensor is symmetric with respect, to the indices ij
and ik ; if the property (A.16) holds for any indices ij and ik , then the tensor a is
totally symmetric.
If the components of the tensor a verify the relation
ai1 i2 ...ij ...ik ...in ¼ ai1 i2 ...ik ...ij ...in ; ðA:17Þ

in a basis B, then we say that the tensor is antisymmetric (skewsymmetric) with


respect to the indices ij and ik ; if the property (A.17) takes place for all the indices
ij and ik , then the tensor is totally antisymmetric.
A tensor a which has not one of the properties mentioned above is an
asymmetric tensor.
A tensor is, in general, defined at a point of the space; as well, we may consider
a tensor mapping ðx1 ; x2 ; x3 Þ ! ai1 i2 ...in ðx1 ; x2 ; x3 Þ or a tensor function
ai1 i2 ...in ¼ ai1 i2 ...in ðx1 ; x2 ; x3 ; tÞ; ðA:18Þ

defining thus a tensor field.

A.1.1.4 Particular Cases

If we use the upper index ‘‘ 0 ’’ for the quantities which are obtained from the
relation (A.2), by a change of co-ordinate axes, then it follows that
a0lm ¼ i0l  im ; l; m ¼ 1; 2; 3:

the relation of definition of these cosines remaining the same; indeed, the scalar
products (A.2) are invariant with respect to any frame of reference, in particular
with respect to the frames B and B0 . We see that a relation of the form (A.12)
holds, because
a0lm ¼ akj alk amj ¼ alk dkm ¼ alm ; l; m ¼ 1; 2; 3;

where we have introduced Kronecker’s symbol. Hence, aij ; i; j ¼ 1; 2; 3, are the


components of a tensor of second order
2 3
  a11 a12 a13
aij  4 a21 a22 a23 5; ðA:19Þ
a31 a32 a33

which makes clear the position of the orthonormed basis B0 with respect to the
basis B and inversely.
734 Appendix

We can define Kronecker’s symbol with respect to the basis B0 in the form

0 0; k 6¼ l;
dkl ¼ ; k:l ¼ 1; 2; 3;
1; k ¼ l:

also in this case, a relation of the form (A.12) is verified, because


d0kl ¼ dij aki alj ¼ dkl ; k; l ¼ 1; 2; 3:

Hence, Kronecker’s symbols dij ; i; j ¼ 1; 2; 3, are the components of a tensor of


second order too
2 3
  1 0 0
d  dij  4 0 1 0 5; ðA:20Þ
0 0 1

which is just the unit tensor (d ¼ 1); this tensor is symmetric and we may write
dji ¼ dij ; i; j ¼ 1; 2; 3: ðA:200 Þ
Let us consider also Ricci’s permutation symbol, defined by the formula
8
< 1; ði; j; kÞ ¼ ð1; 2; 3Þ;
ijk ¼ 1; ði; j; kÞ ¼ ð2; 1; 3Þ; ðA:21Þ
:
0; i ¼ j or j ¼ k or k ¼ j;

where by ði; j; kÞ ¼ ð1; 2; 3Þ we understand that the indices i; j; k take the distinct
values 1; 2; 3 or a cyclic permutation of them. We notice that
2 3
al1 al2 al3
cijk ali amj ank  4 am1 am2 am3 5; l; m; n ¼ 1; 2; 3; ðA:22Þ
an1 an2 an3

taking into account (A.1) and introducing the scalar triple product of three vectors
ðii ; ij ; ik Þ ¼ cijk ; ði0l ; i0m ; i0n Þ ¼ c 0lmn ; i; j; k; l; m; n ¼ 1; 2; 3; ðA:23Þ

it results
c0lmn ¼ cijk ali amj ank ; l; m; n ¼ 1; 2; 3:
Hence, the permutation symbol is a tensor of third order; this tensor is totally
antisymmetric and we may write
cijk ¼ cjki ¼ ckij ¼ cjik ¼ cikj ¼ ckji ; i; j; k ¼ 1; 2; 3: ðA:24Þ
Let be now the determinant
2 3
  a11 a12 a13
det apq  4 a21 a22 a23 5; ðA:25Þ
a31 a32 a33
A.1 Elements of Tensor Calculus 735

the permutation tensor allows to express this determinant in the form


det½apq  ¼ cijk a1i a2j a3k ¼ cijk ai1 aj2 ak3 : ðA:250 Þ
We notice that we may write
clmn det½apq  ¼ cijk ali amj ank ¼ cijk ail ajm akn ; l; m; n ¼ 1; 2; 3; ðA:2500 Þ

too or
1
det½apq  ¼ cijk clmn ail ajm akn ; ðA:25000 Þ
6
which constitutes an extension of the above results.

A.1.2 Operations with Tensors

In what follows we shall deal with algebraic or analytic operations effected with
tensors, as well as with the tensor expression of such operations effected with
vectors. By the way, we shall give some results of the field theory.

A.1.2.1 Algebraic Operations

Two tensors a and b of the same order n are equal (a ¼ b) if they have the same
components
ai1 i2 ...in ¼ bi1 i2 ...in ; ij ¼ 1; 2; 3; j ¼ 1; 2; . . .; n; ðA:26Þ

in an arbitrary frame B; this relation has the well known properties of reflexivity,
symmetry and transitivity.
The sum of two tensors a and b of the same order n is a tensor c = a ? b of the
same order; hence, in a frame B, we have
ai1 i2 ...in þ bi1 i2 ...in ¼ ci1 i2 ...in : ðA:27Þ

The addition of tensors is commutative and associative.


Multiplying all the component of a tensor a of nth order in a frame B by the
same scalar k, one obtains a new tensor b ¼ ka of the same order and of
components kai1 i2 ...in ; this operation is distributive with respect to the addition of
tensors, as well as of scalars. In particular, we may have k ¼ 1; hence, by
subtracting two tensors of nth order one obtains a tensor of the same order. If
b ¼ 0 (all the components of a null tensor in an arbitrary basis B vanish), then we
have k ¼ 0 or a ¼ 0.
Starting from the observations made at the previous section, one can show that
an asymmetric tensor of nth order can be univocally decomposed in a sum of two
tensors, the first one symmetric with respect to the indices ij and ik and the second
736 Appendix

one antisymmetric with respect to the same indices; we can thus write
1 
ai1 i2 ...ij ...ik ...in ¼ ai1 i2 ...ij ...ik ...in þ ai1 i2 ...ik ...ij n
2
1 
þ ai1 i2 ...ij ...ik ...in  ai1 i2 ...ik ...ij ...in : ðA:28Þ
2
The tensor product (external product) of two tensors a and b of nth and mth
order, respectively, is a tensor c ¼ a  b of (n þ m)th order; in a frame B we can
write
ai1 i2 ...in bj1 j2 ...jm ¼ ck1 k2 ...knþm ; ðA:29Þ

where the indices kl ; l ¼ 1; 2; . . .; n þ m, are the indices ip ; p ¼ 1; 2; . . .; n, and


jq ; q ¼ 1; 2; . . .; m.
For instance, by the dyadic product of two tensors of first order one obtains a
tensor of second order
ai bj ¼ cij ; i; j ¼ 1; 2; 3: ðA:30Þ
The external product of tensors is not commutative but is associative and
distributive with respect to their addition.
If we make jk ¼ jl in the relation of definition (A.15) and take into account the
formulae (A.3), then we obtain, in a basis B,
a0j1 j2 ...jk1 jk jkþ1 ...jl1 jk jlþ1 ...jn ¼ ai1 i2 ...ik1 ik ikþ1 ...il1 ik ilþ1 ...in
 aj1 i1 aj2 i2 . . .ajk1 ;ik1 ajkþ1 ;ikþ1 . . .ajl1 ;il1 ajlþ1 ;ilþ1 . . .ajn ;in ; ðA:31Þ

this operation is called the contraction of the tensor. Hence, by the contraction of
two indices of a tensor of nth order one obtains a tensor of (n  2)th order.
For instance, by the contraction of a tensor of second order, of components aij
in a basis B one obtains a scalar aii , called the trace of the tensor a and denoted
tr a ¼ aii : ðA:310 Þ
In particular, by the contraction of Kronecker’s tensor one obtains tr1 ¼ dii ¼ 3.
The internal product (the contracted tensor product) of two tensors a and b of
nth and mth order, respectively, is a tensor ab of (n þ m  2p)th order, where p is
the number of effected contractions.
For instance, the scalar product of two vectors a and b, of components ai and
bi , respectively, will be given by the contracted product (Fig. A.2a)
a  b ¼ ab cosða; bÞ ¼ ai bi ; ðA:32Þ
where a; b are the moduli of the respective vectors. If this product vanishes, then
the two vectors are orthogonal one to the other (or one of them is null).
If c ¼ a  b is the vector product of the two vectors, then we may write
(Fig. A.2b)
ci ¼ cijk aj bk ; i ¼ 1; 2; 3; ðA:33Þ
A.1 Elements of Tensor Calculus 737

b c c b
(a,b) b

a a a
(a) (b) (c)

Fig. A.2 Products of vectors: a Scalar. b Vector. c Triple scalar

we notice that the vector c is normal to the plane formed by the vectors a and b,
while its modulus is equal to the area of the parallelogram formed by these two
vectors. If this product vanishes, then the two vectors are collinear.
The triple scalar product of three vectors a; b and c can be expressed by means
of a contracted product too, in the form
ða; b; cÞ ¼ cijk ai bj ck ; ðA:34Þ

representing the volume of the parallelepiped constructed on the three vectors


(Fig. A.2c). If the product vanishes, then the vectors are coplanar. If a; b; c are the
unit vectors of an orthonormed basis, then one obtains relations of the form (A.23);
if the mixed product is positive, then the orthonormed basis is positive.
We have seen that by algebraic operations of addition and of external or internal
products one obtains new tensors. To identify if a quantity is a tensor, the quotient
law is frequently used; so, if we have a relation of the form
f ði1 ; i2 ; . . .; in Þbj1 j2 ...jm ¼ ck1 k2 ...knþm ; ðA:35Þ

where the indices kl ; l ¼ 1; 2; . . .; n þ m, are just the indices ip ; p ¼ 1; 2; . . .; n, and


jq ; q ¼ 1; 2; . . .; m, then the function f is of the form
f ði1 ; i2 ; . . .; in Þ ¼ ai1 i2 ...in ; ðA:36Þ

obtaining thus the components of a tensor of nth order in a basis B if b and c are
tensors.
With two permutation tensors one may effect the external product
 
 dil dim din 
 
ijk lmn ¼  djl djm djn ; i; j; k; l; m; n ¼ 1; 2; 3: ðA:37Þ
 dkl dkm dkn 

To prove this identity, one considers all the possible values which may be given to
the indices i; j; k and l; m; n, respectively; if two indices are equal, then one obtains
zero in both members, while if the indices are different, then one considers
successively the cases in which they form a permutation of the numbers 1; 2; 3 or a
permutation of the numbers 2; 1; 3.
738 Appendix

If k ¼ n, then one obtains


     
 dil dim  d dim  d djm 

cijk clmk ¼ dkk    djk  il þ dik  jl ;
djl djm  dkl dkm  dkl dkm 

which leads to the contracted product


 
 dil dim 

cijk clmk ¼   ¼ dil djm  dim djl ; i; j; l; m ¼ 1; 2; 3: ðA:38Þ
djl djm 

If one has also j ¼ m, then it results


cijk cljk ¼ djj dil  dij djl ;
wherefrom
cijk cljk ¼ 2dil ; i; l ¼ 1; 2; 3: ðA:39Þ
Finally, in the case in which one has i ¼ l too, then we obtain the scalar
cijk cijk ¼ 6: ðA:40Þ
Taking into account (A.38), one may verify the identity
cijk aj ðcklm bl cm Þ ¼ ðaj cj Þbi  ðaj bj Þci ;

while the relations (A.33) allows to write the triple vector product of three vectors
a, b and c (Fig. A.3) in the form
a  ðb  cÞ ¼ ða  cÞb  ða  bÞc: ðA:41Þ
In the case of a continuous system of vectors W ¼ WðrÞ, definite on the
domain D of volume V, one obtains the resultant and the resultant moment
ZZZ
R¼ WðrÞ dV;
V
ZZZ ðA:42Þ
MO ¼ r  WðrÞ dV;
V

of components
ZZZ
Ri ¼ Wi ðrÞ dV;
V
ZZZ i ¼ 1; 2; 3: ðA:420 Þ
MOi ¼ cijk xj Wk ðrÞ dV;
V

Fig. A.3 Triple vector


product
a c
b
A.1 Elements of Tensor Calculus 739

Let be an asymmetric tensor of second order, expressed by its components aij . If


we denote the symmetric part of this tensor by
1 
aðijÞ ¼ aij þ aji ; i; j ¼ 1; 2; 3; ðA:43Þ
2
and its antisymmetric part by
1 
a½ij ¼ aij  aji ; i; j ¼ 1; 2; 3; ðA:430 Þ
2
then one may write, univocally,
aij ¼ aðijÞ þ a½ij ; i; j ¼ 1; 2; 3: ðA:4300 Þ

What concerns the symmetric tensor of second order, its canonic decomposition
in the spheric tensor and the stress deviator, as well as the study of the variation of
its components along various directions, the determination of the principal
directions, the extreme values, the corresponding graphic representations etc.,
these are the problems which have been considered in the Chaps. 2 and 3, in
connection with the particular tensors Te and Tr .
The antisymmetric tensor a of second order will be represented in the form
 
0 a12 a31 

½aij  ¼  a12 0 a23 ; ðA:44Þ
 a31 a23 0 

having only three distinct components. This is a degenerate tensor, equivalent to a


tensor of first order; we may thus introduce an equivalent vector a of components
1
ak ¼ cijk aij ; k ¼ 1; 2; 3: ðA:45Þ
2
Conversely, multiplying both members by cklm and taking into account (A.38),
we get
1 1 1
cklm ak ¼ cklm cijk aij ¼ ðdli dmj  dlj dmi Þaij ¼ ðalm  aml Þ;
2 2 2
hence, we may write
aij ¼ cijk ak ; i; j ¼ 1; 2; 3: ðA:450 Þ
For instance, the vector product of two vectors, given by the formula (A.33),
leads to an antisymmetric tensor of second order, of the form
ai bj ¼ 2cijk ck ; i; j ¼ 1; 2; 3: ðA:330 Þ
We remark that, unlike the vectors a and b, which are polar vectors, the vector c,
which is their vector product, is of another nature; it is an axial vector, corresponding
to an antisymmetric tensor of second order. One cannot add axial vectors to polar
740 Appendix

ones, hence one cannot add a vector product to an axial vector. Instead, the scalar
product of a polar vector by an axial one (which leads to a scalar, i.e. the mixed
product of three vectors), as well as the vector product of two such vectors (which
leads to a polar vector, i.e. the double vector product of three vectors) are operations,
which have sense.

A.1.2.2 Analytic Operations

Let be a scalar field U ¼ Uðx1 ; x2 ; x3 Þ, defined on a domain D, with U 2 C1 ðDÞ


(the function U admits continuous derivatives of first order on the domain D) and
let be oU=oxi the components of the conservative vector field thus obtained; by a
change of co-ordinates, one obtains U 0 ¼ U 0 ðx01 ; x02 ; x03 Þ and one can write
oU oU oxi oU
¼ ¼ aji ; j ¼ 1; 2; 3; ðA:46Þ
ox0j oxi ox0j oxi

where we have taken into account the relations (A.7) and the relation (A.8), which
defines a scalar field. Therefore, the derivative of a scalar with respect to an
independent variable leads to a tensor of first order; we denote
oU
¼ oi U ¼ U;i ; i ¼ 1; 2; 3; ðA:47Þ
oxi
where the indices at the right of the comma indicate the differentiation with respect
to the corresponding variable.
Similarly, starting from the vector field Vi ¼ Vi ðx1 ; x2 ; x3 Þ with Vi 2 C1 ðDÞ,
i ¼ 1; 2; 3, and making a change of co-ordinates of the form (A.6), we may write
oVl0 oðVk alk Þ oxi oVk
¼ ¼ alk aji ; l; j ¼ 1; 2; 3; ðA:48Þ
ox0j oxi ox0j oxi

hence, the derivatives of first order of a tensor of first order are the components of
a tensor of second order. We may write
oVi
¼ oj Vi ¼ Vi;j ; i; j ¼ 1; 2; 3: ðA:49Þ
oxj
In general, the derivatives of first order of tensor of nth order are the
components of a tensor of (n þ 1)th order.
Similarly, one can define derivatives of higher order of a tensor.
If ai 2 C2 ðDÞ, then we may write
ai;jk ¼ ai;kj ; i; j; k ¼ 1; 2; 3; ðA:50Þ

hence, the mixed derivatives of second order do not depend on the order of
differentiation (Schwarz’s theorem). We notice thus that the tensor ai;jk is
symmetric with respect to the indices j and k.
A.1 Elements of Tensor Calculus 741

A.1.2.3 Differential Operators of First Order

Let us introduce the vector differential operator nabla (del) of Hamilton, denoted
by r, in the form
o
r ¼ ij ¼ ij oj : ðA:51Þ
oxj
In this case, the conservative vector field (A.47) will define the gradient of the
scalar field U by the relation
gradU ¼ rU ¼ oj Uij ¼ U;j ij : ðA:52Þ
As well, we assume that Vi;j are the components of the tensor of second order,
denoted by GradV; in general, the gradient of a tensor of nth order will be a tensor
of (n þ 1)th order.
In this case, the differential of the scalar field U ¼ Uðx1 ; x2 ; x3 Þ may be written
in the form
dU ¼ gradU  dr: ðA:53Þ
Let us consider the field of vectors V ¼ Vðx1 ; x2 ; x3 Þ defined by the relations
Vi ¼ U;i ; ðA:54Þ

considering an arbitrary unit vector n, we observe that


oU
V  n ¼ U;i ni ¼ ; ðA:55Þ
on
hence the definition given to the vector field does not depend on the chosen system
of co-ordinate axes.
The vector field thus defined is called conservative field (field which derives
from the potential U); the corresponding vectors are called conservative vectors,
while the scalar field U is called potential field.
If U ¼ Uðr; tÞ and r ¼ rðtÞ, then we obtain a vector field defined by the
formulae (A.54) too; this is a quasi-conservative field, while the corresponding
vectors are quasi-conservative vectors, the function U being a quasi-potential. The
differential is written in the form
oU
dU ¼ gradU  dr þ dt; ðA:56Þ
ot
while the total (substantial) derivative is given by
dU _
¼ gradU  r_ þ U; ðA:560 Þ
dt
being the sum of the space derivative and the temporal derivative.
742 Appendix

Fig. A.4 Work of a vector


along a curve. a Open. V V dr C
b Closed dr
P 1
P
P0 P r
C
r O

O
(a) (b)

We denote by
dW ¼ VðrÞ  dr ¼ Vi ðrÞdxi ðA:57Þ
the elementary work of the vector V ¼ VðrÞ; this one is, in general, not an exact
differential. We notice that the work of the sum of n vectors applied at the same
point is equal to the sum of the works of each vector; this result is obvious, taking
into account the property of distributivity of the scalar product with respect to the
addition of vectors.
The work of a vector V along a curve C, between the points P0 and P1
(Fig. A.4a), is given by

ðA:58Þ

the travelling sense along this curve being from P0 to P1 . As well, we notice that
the work of a vector is a scalar quantity.
In the case of a closed curve C (Fig. A.4b), we consider the curvilinear vector
integral
I I
WC ðVÞ ¼ VðrÞ  dr ¼ Vi dxi ; ðA:59Þ
C C

the travelling sense being the counterclockwise one; this work is called the
circulation of the vector V along the closed curve C.
We mention that the curvilinear vector integrals along a closed curve do not
depend on the initial point on the curve.
In the case of a conservative vector field, the elementary work of a conservative
vector is given by
dW ¼ gradU  dr ¼ dU; ðA:60Þ
so that it is an exact differential; it results that
A.1 Elements of Tensor Calculus 743

ðA:600 Þ

where we used the formula (A.58). Hence, in the case of a conservative vector, the
work between two points does not depend on the path, bat only on the values of
the potential at the ends; similarly, starting from the formula (A.59), we notice that
the work of a conservative vector along a closed curve vanishes (if the curve
belongs to a simply connected domain).
Let be
UðrÞ ¼ C; C ¼ const; ðA:61Þ

the equation of a surface having the property that at each of its points the scalar
potential is constant; we assume that U 2 C1 . This surface is called an
equipotential surface; in the case in which U ¼ Uðr; tÞ, we have an equiquasi-
potential surface
Uðr; tÞ ¼ C; C ¼ const; ðA:610 Þ

which is variable in time.


Through a point r0 passes only one equipotential surface.
The variation of a scalar function U ¼ UðrÞ may be appreciate by means of the
gradient, defined by the formula (A.52).
We notice that gradU is normal to the equipotential surface (A.61). The relation
(A.55) may be written in the form
oU
n  gradU ¼ ; ðA:550 Þ
on
where oU=on is the derivative of the scalar field U on the direction of the unit
vector n; if we take n ¼ vers gradU in the relation (A.550 ), then it results that the
gradient of the scalar function U is a vector directed in the sense in which the value
of this function at a point is increasing. The congruence of gradient lines will be
thus normal to the family of corresponding equipotential surfaces, the travelling
sense being that in which the value of the scalar function U is increasing. These
properties are valid in the case of a quasi-conservative scalar field too.
We mention the properties
gradðU1 þ U2 Þ ¼ grad U1 þ grad U2 ; ðA:62Þ
gradCU ¼ CgradU; C ¼ const; ðA:620 Þ
gradC ¼ 0; C ¼ const: ðA:6200 Þ
Introducing the function r ¼ rðsÞ, where s is the curvilinear co-ordinate along a
curve C, it results that U ¼ UðsÞ and we may write
oU dU dr
¼ ¼ gradU  ¼ gradU  s; ðA:63Þ
os ds ds
744 Appendix

obtaining thus the derivative on the direction of unit vector s of the tangent to this
curve.
We mention the properties:
gradðU1 U2 Þ ¼ U1 gradU2 þ U2 gradU1 ; ðA:64Þ

gradf ðUÞ ¼ f 0 ðUÞgradU; f 2 C1 ; ðA:640 Þ


Z
f ðUÞ gradU ¼ grad f ðUÞ dU; f integrable; ðA:6400 Þ

gradf ðU1 ; U2 Þ ¼ fU0 1 gradU1 þ fU0 2 gradU2 ; ðA:64000 Þ

valid in the case of a quasi-potential scalar field.


Concerning the position vector r, we also notice that:
1 f 0 ðrÞ
gradr ¼ r; gradf ðrÞ ¼ r; f 2 C1 ; ðA:65Þ
r r

!
gradðC  rÞ ¼ C; C ¼ const: ðA:650 Þ
The formula (A.53) shows that the operator r which has been introduced
allows to conceive the symbol d of the total differential as an operator, in the form
of a scalar product
d ¼ dr  r ¼ dr  grad: ðA:66Þ
As well, the formula (A.55) leads to the introduction of the operator derivative
on the direction of the unit vector n, in the form
o
¼ n  r ¼ n  grad; ðA:67Þ
on
in the case in which r ¼ rðsÞ and n ¼ s, it results the operator
o d
¼ ¼ r0 ðsÞ  r ¼ s  grad: ðA:670 Þ
os ds
Thus, one introduces the linear differential operator
o
A  grad ¼ A  r ¼ Ai ¼ Ai o i ; ðA:68Þ
oxi
where A is a given vector, constant or variable. By applying this operator to the
scalar field U, one obtains the scalar
A  gradU ¼ Ai U;i ; ðA:69Þ

as well, by applying this operator to the vector field V, one gets the vector
ðA  gradÞV ¼ Aj Vk;j ik : ðA:70Þ
A.1 Elements of Tensor Calculus 745

In particular, we obtain the total differential of a vector field V ¼ VðrÞ in the


form
dV ¼ ðdr  gradÞV; ðA:71Þ

the derivative on the direction of unit vector n being given by


oV
¼ ðn  gradÞV; ðA:72Þ
on
for r ¼ rðsÞ and n ¼ s, we get
oV dV
¼ ¼ ðs  gradÞV: ðA:720 Þ
os ds
If
V;i ¼ Vj;i ij ; ðA:73Þ

then it results
dV ¼ GradV  dr ¼ V;i dxi : ðA:730 Þ
The curves for which the tangents to each point of them are directed along the
vectors V ¼ VðrÞ of the field are called vector lines; these lines form a congruence
of curves. Because the differential dr is tangent to these lines, their vector equation
will be of the form
VðrÞ  dr ¼ 0; ðA:74Þ
equivalent to
cijk Vj dxk ¼ 0; i ¼ 1; 2; 3: ðA:740 Þ
As well, we introduce the vector differential operator
o
A  r ¼ cjkl Aj il ¼ cjkl Aj il ok : ðA:75Þ
oxk
The differential operator ‘‘r’’ if applied to a vector V, then it defines its
divergence in the form
oVi
divV ¼ r  V ¼ ¼ oi Vi ¼ Vi;i ; ðA:76Þ
oxi
this scalar quantity is invariant to a change of co-ordinate axes.
The differential operator ‘‘r’’ if applied to a vector V, then it leads to the curl
of this vector; we may write
oVk
curlV ¼ r  V ¼ cjkl il ¼ cjkl oj Vk il ¼ cjkl Vk;j il : ðA:77Þ
oxj
746 Appendix

Similarly, one can introduce the operators div and curl for tensors of higher
order.
A field of vectors for which
curlV ¼ r  V ¼ 0 ðA:78Þ
is called irrotational. One easily observer that a field of gradients
V ¼ gradU ¼ rU ðA:780 Þ

is irrotational (curl gradU ¼ 0); hence, the fields of conservative vectors are
irrotational, these fields being the only ones which have this property. The property
is maintained in the case of a quasi-conservative field too; as we observe, this is
also the condition that the integral in (A.58) be an exact differential.
A vector field for which
divV ¼ r  V ¼ 0 ðA:79Þ

is called solenoidal. One can easily notice that a field of curls


V ¼ curlW ¼ r  W ðA:790 Þ
is solenoidal (div curlW ¼ 0); one may show that these fields are the only ones
with this property.

A.1.2.4 Differential Operators of Second Order

We introduce the differential operator of second order


o o
D ¼ r2 ¼ div grad ¼ ¼ oi oi ; ðA:80Þ
oxi oxi
known as Laplace’s operator. If D is applied to a scalar U ¼ UðrÞ, then one
obtains

o2 U o 2 U o2 U o2 U
DU ¼ ¼ 2 þ 2 þ 2 ðA:81Þ
oxi oxi ox1 ox2 ox3

and if it is applied to a vector V ¼ VðrÞ, then we get


DV ¼ DVj ij : ðA:82Þ

A function U 2 C2 ðDÞ which satisfies the equation


DU ¼ 0; ðA:83Þ
in the domain D, is called harmonic in this domain; similarly, a scalar function
U 2 C4 ðDÞ, which verifies the equation

D2 U ¼ 0; ðA:84Þ
A.1 Elements of Tensor Calculus 747

in the domain D, is called biharmonic in this domain. On the same way one may
introduce polyharmonic functions.
Laplace’s operator is of elliptic type. Analogically, we introduce d’Alembert’s
operator of hyperbolic type, in the form

1 o2
hc ¼ D  ; c ¼ const; ðA:85Þ
c2 ot2
assuming that U ¼ Uðr; tÞ, the equation
hc U ¼ 0; ðA:86Þ

where U 2 C2 ðDÞ, is called the waves equation, while c is the propagation velocity
of the waves.
If U 2 C4 ðDÞ, then one may introduce the double waves equation
h1 h2 U ¼ 0; ðA:87Þ

too, where

1 o2
hi ¼ D  ; ci ¼ const, i ¼ 1; 2; ðA:870 Þ
c2i ot2

corresponding to two simple waves equations. Similarly, one may consider


functions which verify a polywaves equation.
We introduce also the caloric operator of parabolic type
1o
\¼D ; a ¼ const, ðA:88Þ
a ot
where a is the thermic diffusivity; assuming that U ¼ Uðr; tÞ, the equation
\ U ¼ 0; ðA:89Þ

where U 2 C2 ðDÞ will be called caloric equation; similarly, one may introduce
polycaloric functions.

A.1.2.5 Relations Between Differential Operators

Concerning the divergence operator, we mention the relations


divðkVÞ ¼ kdivV þ V  gradk; k scalar; ðA:90Þ
divðV  WÞ ¼ W  curlV  V  curlW: ðA:900 Þ
Concerning the curl operator, we may write
curlðkVÞ ¼ kcurlV  V  gradk; k scalar; ðA:91Þ
curl curlV ¼ grad divV  DV: ðA:910 Þ
748 Appendix

We mention, as well, the formulae


gradðV  WÞ ¼ ðW  rÞV þ ðV  rÞW
ðA:92Þ
þ V  curlW þ W  curlV;

curlðV  WÞ ¼ ðW  rÞV  ðV  rÞW


ðA:920 Þ
þ VdivW  WdivV:
In particular, from (A.93) one obtains
1
gradV2 ¼ ðV  rÞV þ V  curlV: ðA:9200 Þ
2
We notice also that the vector differential operator (A.75) is applied to the
vector V by the formula
ðV  rÞ  W ¼ V  curlW: ðA:93Þ
The above results are easily obtained by using the methods of the vector algebra
and by formal calculations with the operator r.

A.1.2.6 Integral Formulae

Let be a closed curve C, situated on the sufficiently smooth surface R, limiting on


it a simply connected domain S (reducible, by continuous deformation, to a point,
without leaving the surface R) (Fig. A.5). The surface R is oriented by means of
the unit vector n of the normal to it; as well, we assume a sense of travelling
through the curve C.
Let us consider the function W ¼ WðrÞ with W 2 C1 ðDÞ, where D is a domain
which includes S þ C. One may prove Stokes’s formula
I ZZ
V dr ¼ n  curlV dS; ðA:94Þ
C S

because the left member of the formula depends only on the curve C, we may
replace the surface S by any other surface SO  D, which satisfies analogous
conditions.

Fig. A.5 Stokes’s formula


x3
n

P S Σ

r C

O x2

x1
A.1 Elements of Tensor Calculus 749

The above curvilinear integral is called circulation (it represents the work of a
field of vectors), while the surface integral represents the flux of a field of curls; in
the above mentioned conditions, it results that the circulation of a field of vectors
along a closed curve is equal to the flux of the curl of the very same field of vectors
through a sufficiently smooth arbitrary surface, bounded by a given curve. We also
mention that the circulation of a field of irrotational vectors, hence of a field of
conservative vectors, vanishes.
With respect to an orthonormed frame of reference Ox1 x2 x3 , we may write
I ZZ
Wi dxi ¼ ijk ni Wk;i dS: ðA:940 Þ
C S

In particular, if V1 ¼ F 2 C1 ðDÞ; V2 ¼ V3 ¼ 0, then it results


I ZZ
Fdx1 ¼ ðn2 F;3  n3 F;2 ÞdS: ðA:9400 Þ
C S

If in Stokes’s formula (A.94) we concentrate all the surface at a point of


position vector r, then we get
H
V  dr
prn curlV ¼ n  curlV ¼ lim C : ðA:94000 Þ
S!0 S
Hence, we may determine the projection of the vector curlV on an arbitrary axis of
unit vector n (hence the very same vector), without any reference to a frame; thus,
the definition given to the curl in Sect. A.1.2.3 has an intrinsic value, being
immaterial on the frame.
Let be now a sufficiently smooth closed surface S, limiting a domain D and a
field of vectors W ¼ WðrÞ (Fig. A.6); we assume that W 2 C1 ðD þ SÞ. We may
write the Gauss–Ostrogradskiı̆ formula in the form
ZZ ZZZ
W  ndS ¼ divWds; ðA:95Þ
S D

Fig. A.6 The Gauss–Ostro-


gradskiı̆ formula x3
n
S

P
D

r x2
O
x1
750 Appendix

where n is the unit vector of the external normal to the surface; the surface integral
represents the flux of the field of vectors through the surface S, so that the formula
is called the flux-divergent formula too.
In components, we may write
ZZ ZZZ
Wi ni dS ¼ Wi;i ds: ðA:950 Þ
S D

In particular, if we consider the component Vj ¼ F 2 C1 ðD þ SÞ, the other


components vanishing, we may write
ZZ ZZZ
Fnj dS ¼ F;j ds; j ¼ 1; 2; 3; ðA:9500 Þ
S D

multiplying by the unit vector ij and summing, we get


ZZ ZZZ
Fn dS ¼ gradF ds: ðA:95000 Þ
S D

If we take F of the form cijk Vk in (A.9500 ), then we obtain


ZZ ZZZ
cijk nj Vk dS ¼ cijk Vk;j ds;
S D

so that
ZZ ZZZ
n  V dS ¼ curlV ds: ðA:950000 Þ
S D

Let be a domain D0  D, reducible by continuous deformation to a point of


position vector r and S0 the surface (sufficiently smooth) which bounds it. Starting
from the Gauss-Ostrogradskiı̆ formula (A.95), we may represent the divergence of
the vector V in the form
RR 0
S0 V  n dS
divV ¼ lim 0
; ðA:9500000 Þ
0
D !0 D
hence, the definition given to the divergence in Sect. A.1.2.3 is immaterial of the
frame and has an intrinsic value.
The formula (A.95) leads to
ZZ ZZZ
oU
dS ¼ DUds ðA:96Þ
S on D

for a field of conservative vectors W ¼ gradU, where we took into account the
formula (A.550 ) and the definition of Laplace’s operator. If W ¼ wU; w scalar,
then we may write
ZZ ZZZ
wU  n dS ¼ ðwdivU þ U  gradwÞds;
S D
A.1 Elements of Tensor Calculus 751

where we took into account the formula (A.90). If, after this, we take U ¼ gradu,
u scalar, then we get
ZZ ZZZ
ou
w dS ¼ ðwDu þ gradu  gradwÞds; ðA:97Þ
S on D

inverting u and w and subtracting the relation thus obtained from (A.97). it results
ZZ
ZZZ
ou ow
w u dS ¼ ðwDu  uDwÞds: ðA:98Þ
S on on D

The formulae (A.96–A.98) are known as Green’s formulae.

A.1.2.7 Theorems of Almansi and Boggio Type

Often, one can reduce the study of certain partial differential equations to the study
of several differential equations of the same type, but of a lower order.
Thus, let be the biharmonic equation (A.84) and two functions U1 , and U2 ,
harmonic in the domain D and verifying the Eq. (A.83); Almansi showed that the
biharmonic function U may be expressed univocally in the form
U ¼ U 1 þ x1 U 2 ; ðA:99Þ
hence, a biharmonic function is, in a certain way, equivalent to two harmonic
functions. We notice that we can replace the variable x1 by anyone of the other
variables; as well, we may write

U ¼ U1 þ r 2 U2 : ðA:990 Þ
To verify the above formulae, we mention the relations
Dðxi UÞ ¼ 2U;i þ xi DU; i ¼ 1; 2; 3; ðA:100Þ

Dðr 2 UÞ ¼ 8U þ 4r  gradU þ r 2 DU; ðA:1000 Þ

which, if U is a harmonic function, become


Dðxi UÞ ¼ 2U;i ; i ¼ 1; 2; 3; ðA:101Þ

Dðr 2 UÞ ¼ 8U þ 4r  gradU; ðA:1010 Þ


we mention also the formula
Dðr  VÞ ¼ 2divV þ r  DV; ðA:102Þ

which, in case of a harmonic vector, becomes


Dðr  VÞ ¼ 2divV: ðA:1020 Þ
In general, we may state following theorem of Almansi type: If D is a
differential operator of order m; in s variables q1 ; q2 ; . . .; qs ; and if we assume that
U ¼ Uðq1 ; q2 ; . . .; qs Þ, then the solution of the equation
752 Appendix

Dn U ¼ 0; ðA:103Þ
where U 2 Cmn ðDÞ, may be written in the form

U ¼ U0 þ q1 U1 þ q21 U2 þ    þ qn1
1
Un1 ; ðA:1030 Þ

where U0 ; U1 ; U2 ; . . .; Un1 are functions of the same variables, verifying the


equations
DUi ¼ 0; i ¼ 0; 1; 2; ; . . .; n  1: ðA:10300 Þ
Obviously, one may replace the variable q1 by anyone of the other s  1
independent variables.
As well, let be D i ; i ¼ 1; 2; . . .; p, differential operators of order mi in s
variables q1 ; q2 ; . . .; qs , and a function U ¼ Uðq1 ; q2 ; . . .; qs Þ. We may state
following theorem of Boggio type: If D i are relatively prime and permutable
differential operators.
D iD j ¼ D jD i; i; j ¼ 1; 2; ; . . .; p; ðA:104Þ

then the solution of the equation


D 1 D 2    D p U ¼ 0; ðA:105Þ

where U 2 Cm1 þm2 þþmp ðDÞ may be written in the form


U ¼ U1 þ U2 þ    þ Up ; ðA:1050 Þ

where U1 ; U2 ; . . .; Up are functions of the same variables, which verify the


equations
D i Ui ¼ 0; i ¼ 1; 2; . . .; p: ðA:10500 Þ
The condition (A.104) is fulfilled, e.g., in the case of operators with constant
coefficients.
In particular, in the case of the double waves equations (A.87), we obtain
U ¼ U1 þ U2 ; ðA:106Þ

where U1 and U2 satisfy the equations


h1 U1 ¼ 0; h2 U2 ¼ 0; ðA:1060 Þ
d’Alembert’s operators being given by (A.870 ).
We mention also the formula
hi ðr  VÞ ¼ 2divV þ r  hi V; i ¼ 1; 2; ðA:107Þ

analogue to the formula (A.102); if the vector V satisfies a simple waves equation,
then it becomes
hi ðr  VÞ ¼ 2divV; i ¼ 1; 2; ðA:1070 Þ
A.1 Elements of Tensor Calculus 753

If we take
V ¼ curl W; ðA:108Þ

then the formulae (A.102) and (A.107) take the form


Dðr  curl WÞ ¼ r  Dcurl W ¼ r  curl DW; ðA:109Þ
hi ðr  curl WÞ ¼ r  hi curl W ¼ r  curl hi W; i ¼ 1; 2: ðA:1090 Þ

A.2 Curvilinear Co-ordinates

In what follows we shall consider orthogonal curvilinear co-ordinates, in general,


as well as cylindrical and spherical co-ordinates, in particular. We will then put in
evidence some important differential operators.

A.2.1 Orthogonal Curvilinear Co-ordinates

First of all, we deal with some general results concerning orthogonal curvilinear
co-ordinates, particularizing then the results thus obtained to cylindrical and
spherical co-ordinates.

A.2.1.1 General Results

Let be a point M of position vector r, specified by the orthonormed Cartesian


co-ordinates xi ; i ¼ 1; 2; 3 (Fig. A.7). The relations
xi ¼ xi ðq1 ; q2 ; q3 Þ; i ¼ 1; 2; 3; ðA:110Þ

where xi are functions at least of class C1 , put in evidence three curvilinear lines
passing through the point M, as we take q2 ; q3 ¼ const or q3 ; q1 ¼ const or

Fig. A.7 Curvilinear


co-ordinates x3
q3
q2 q1
i s3
i s2 i s1

r M ( x1, x 2, x3)
i3

i1 O x2
i2

x1
754 Appendix

q1 ; q2 ¼ const. These three lines are the curvilinear co-ordinate lines q1 ; q2 ; q3 ,


which pass through the point M; the co-ordinates along these lines are called
curvilinear co-ordinates.
Giving various values to these constants, we find three families of co-ordinate
lines, upon which lean three families of co-ordinate surfaces, obtained making
qa ¼ const, a ¼ 1; 2; 3 (we use Greek indices for curvilinear co-ordinates).
If the Jacobian

oxi
J  det ðA:111Þ
oqa

is nonzero, one may find, univocally, relations of the form


qa ¼ qa ðx1 ; x2 ; x3 Þ; a ¼ 1; 2; 3; ðA:1100 Þ

where qa are functions at least of class C1 too.


The unit vectors of the system of curvilinear lines are specified by
1 1 oqa
isa ¼ grad qa ¼ ij ð!Þ; a ¼ 1; 2; 3; ðA:112Þ
ha ha oxj

where we have introduced Lamé’s differential parameters of first order


sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
oqa oqa
ha ¼ jgrad qa j ¼ ð!Þ; a ¼ 1; 2; 3: ðA:113Þ
oxi oxi

In the Sect. A.2, the sign (!) indicates ‘‘without summation with respect to
Greek indices’’.
In case of orthogonal curvilinear co-ordinates, one has
isb  isc ¼ dbc ; b; c ¼ 1; 2; 3; ðA:114Þ

wherefrom, taking into account (A.112) and that ij has an analogous property, we
get
oqb oqc
¼ hb hc dbc ð!Þ; b; c ¼ 1; 2; 3; ðA:115Þ
oxi oxi
in particular,
oqb oqc
¼ 0; b 6¼ c; b; c ¼ 1; 2; 3: ðA:1150 Þ
oxi oxi
These are necessary and sufficient conditions for the transformation (A.1100 ) to be
orthogonal.
The direction cosines of the directions qa will be given by the scalar products
cosðqa ; xi Þ ¼ isa  ii ;
A.2 Curvilinear Co-ordinates 755

which leads to
1 oqa
cosðqa ; xi Þ ¼ ; i; a ¼ 1; 2; 3; ðA:116Þ
ha oxi
where we took into account (A.112).
We notice that
oqb oqb oxi
¼ ¼ dbc ð!Þ; b; c ¼ 1; 2; 3:
oqc oxi oqc

To solve the system of nine linear equations thus obtained, we assume that
oxi oqc
 f ðcÞ ð!Þ;
oqc oxi

introducing in the previous relations, we get


oqb oqc
f ðcÞ ¼ dbc ð!Þ:
oxi oxi
Taking into account (A.115), it results
f ðcÞhb hc dbc ¼ dbc ð!Þ;

wherefrom, for b ¼ c;
1
f ðcÞ ¼ ;
h2c

hence,
oxi 1 oqa
¼ 2 ð!Þ; i; a ¼ 1; 2; 3: ðA:117Þ
oqa ha oxi

Replacing in (A.115), it results


oxi oxi dbc
¼ ð!Þ; b; c ¼ 1; 2; 3; ðA:118Þ
oqb oqc hb hc

in particular,
oxi oxi
¼ 0; b 6¼ c; ðA:1180 Þ
oqb oqc

obtaining thus the necessary and sufficient conditions for the transformation
(A.110) to be orthogonal.
Starting from (A.116), we get the direction cosines in the form
oxi
cosðqa ; xi Þ ¼ ha ð!Þ; i; a ¼ 1; 2; 3; ðA:1160 Þ
oqa
756 Appendix

Taking into account (A.117), we may express the differential parameters of first
order also in the form

oxi oxi 1=2


ha ¼ ð!Þ; a ¼ 1; 2; 3:
oqa oqa

A.2.1.2 Cylindrical and Spherical Co-ordinates

In the case of cylindrical co-ordinates (Fig. A.8a), we may write the relations
(orthonormed basis ir ; ih ; iz )
x1 ¼ r cos h; x2 ¼ r sin h; x3 ¼ z; r 0; 0
h\2p; z 2 R; ðA:119Þ
wherefrom qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
x2
r¼ x21 þ x22 ; h ¼ arg tan ; z ¼ x3 : ðA:1190 Þ
x1
The differential parameters of first order are
1
hr ¼ hz ¼ 1; hh ¼ : ðA:120Þ
r
In the case of spherical co-ordinates (space polar co-ordinates) (Fig. A.8b), we
have (orthonormed frame of unit vectors iR ; iu ; ih )

x1 ¼ R sin u cos h; x2 ¼ R sin u sin h; x3 ¼ R cos u;


ðA:121Þ
R 0; 0
u
p; 0
h\2p;
and
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffi x2 x2
R¼ xi xi ¼ x21 þ x22 þ x23 ; u ¼ arg cos ; h ¼ arg tan : ðA:1210 Þ
R x1

Fig. A.8 Cylindrical (a) and


x3 iz x3 iR
spherical (b) co-ordinates iθ

M M
ir iϕ
R
ϕ
z
i3 i3
O i2 O i2
i1 x2 i1 x2
r r
θ θ
x1 x1
M' M'
(a) (b)
A.2 Curvilinear Co-ordinates 757

One may easily pass from cylindrical co-ordinates to spherical ones, observing
that
r ¼ R cos u: ðA:122Þ
The differential parameters of first order are given by
1 1
hR ¼ 1; hu ¼ ; hh ¼ : ðA:123Þ
R R sin u

A.2.2 Differential Operators in Curvilinear Co-ordinates

In what follows, we give some results concerning the determination of certain


differential operators in orthogonal curvilinear co-ordinates, in general, and in
cylindrical and spherical co-ordinates, in particular.

A.2.2.1 Orthogonal Curvilinear Co-ordinates

Starting from
oxi
dxi ¼ dqa ; i ¼ 1; 2; 3; ðA:124Þ
oqa
the arc element in orthogonal curvilinear co-ordinates is given by (Fig. A.9)
X3
1 2
ds2 ¼ dxi dxi ¼ dq ; ðA:125Þ
h2 a
a¼1 a

where we took into account (A.1130 ) and (A.118).


Making successively q2 ; q3 ¼ const or q3 ; q1 ¼ const or q1 ; q2 ¼ const, we
obtain the arc elements along the three co-ordinates lines

Fig. A.9 Element of arc in


curvilinear orthogonal
co-ordinates
ds

i s3
ds3

i s2 i s1
ds1
ds2
M ( x1, x 2, x3)
758 Appendix

1
dsa ¼ dqa ð!Þ; a ¼ 1; 2; 3; ðA:126Þ
ha
and

ds2 ¼ dsa dsa : ðA:127Þ


We notice that, along the curvilinear lines, one can write
o dqa o o
¼ ¼ ha ð!Þ; a ¼ 1; 2; 3; ðA:128Þ
osa dsa oqa oqa
where we took into account (A.126).
The volume element, i.e. the volume of the curvilinear parallelepiped built with
the vectors isa dsa ð!Þ; a ¼ 1; 2; 3, is given by
dV ¼ ds1 ds2 ds3 ; ðA:129Þ
assuming that we deal with a positive basis.
One, easily, observes that we can introduce the differential operators of first
order in the form
o oqa o oxi o
¼ ¼ ; i ¼ 1; 2; 3; ðA:130Þ
oxi oxi oqa osa osa
where we took into account (A.117) and (A.128).
Using the formula (A.64000 ), we may write the gradient of a scalar field U ¼
Uðx1 ; x2 ; x3 Þ in the form
oU
gradU ¼ rU ¼ gradqa ;
oqa
taking into account (A.112) and (A.128), one obtains the operator r in the form
o
r ¼ grad ¼ isa : ðA:131Þ
osa
By means of the formula (A.117) and taking into account the operator (A.128),
the unit vectors (A.112) of the curvilinear system of axes may be written in the
form
oxj
isa ¼ ij ; a ¼ 1; 2; 3; ðA:132Þ
osa
too; in this case, the operator r takes the form
oxj o
r ¼ grad ¼ ij : ðA:1310 Þ
osa osa
If we apply the operator o=osc , of the form (A.128), to the operator o=osb , given
by the same formula, we get
A.2 Curvilinear Co-ordinates 759

o2 ohb o o2
¼ þ hc hb ð!Þ;
osc osb osc oqb oqc oqb

similarly

o2 ohc o o2
¼ þ hb hc ð!Þ:
osb osc osb oqc oqb oqc

Subtracting one relation of the other, taking into account the symmetry of the
mixed derivative operators with respect to qb and qc and using the formula
(A.128), we get the relation between the mixed derivative operators with respect to
sb and sc in the form

o2 o2 1 ohc o 1 ohb o
 ¼  ð!Þ; b; c ¼ 1; 2; 3: ðA:133Þ
osb osc osc osb hc osb osc hb osc osb
Taking into account the relation (A.132), the condition (A.114) leads to
oxi oxi
¼ dbc ; b; c ¼ 1; 2; 3; ðA:134Þ
osb osc

one can obtain this result also starting from the relation (A.118) and from the
operator (A.128).
Differentiating the relation (A.134) with respect to sa , we get

o2 xi oxi o2 xi oxi
þ ¼ 0; a; b; c ¼ 1; 2; 3; ðA:135Þ
osa osb osc osa osc osb

making b ¼ c, we obtain, in particular,

o2 xi oxi
¼ 0; a ¼ 1; 2; 3; ðA:136Þ
osa osb osb

relation which is true if one sums with respect to b, as well as if one does not sum
with respect to this index.
Taking into account the relations (A.133, A.134), the relation (A.135) also
allows to write

o2 xi oxi o2 xi oxi 1 oha oha


þ ¼ dab þ dac
osb osa osc osc osa osb ha osc osb

1 ohb 1 ohc
dbc þ ð!Þ; a; b; c ¼ 1; 2; 3: ðA:1350 Þ
hb osa hc osa
Making b ¼ c, we get, in particular,

o2 xi oxi dab oha 1 ohb


¼  ð!Þ; a; b; c ¼ 1; 2; 3; ðA:1360 Þ
osb osa osb ha osb hb osa
760 Appendix

summing with respect to b, one has

o2 xi oxi 1 ohb 1 ohc


¼ 
osd osa osd hb osa hc osa

1 oðhb hc Þ o 1
¼ ¼ hb hc ð!Þ;
hb hc osa osa hb hc
a 6¼ b 6¼ c 6¼ a; with summation with respect to d: ðA:13600 Þ

Similarly, we establish the relation


2

o o xi o 2 xi 1 oha oha
þ ¼ dab þ dac
osa osb osc osc osb ha osc osb

1 ohb 1 ohc
þdbc þ ð!Þ; a; b; c ¼ 1; 2; 3: ðA:136000 Þ
hb osa hc osa
The divergence of a vector V is given by
o   oVsb ois
divV ¼ r  V ¼ isa  Vsb isb ¼ isa  isb þ Vsb isa  b ;
osa osa osa
where we used the operator (A.131); taking into account (A.132), we may write
oisb oxi o2 xi
isa  ¼ ; ðA:137Þ
osa osa osa osb

while the formula (A.13600 ) leads to


oisb o 1
isa  ¼ ha hc ð!Þ: ðA:1370 Þ
osa osb ha hc

Finally, we get

oVsa 1 ohb 1 ohc


divV ¼  þ Vs a
osb hb osa hc osa

oVsa 1 oðhb hc Þ o Vsa


¼  Vs a ¼ hb h c ð!Þ;
osa hb hc osa osa hb hc
b; c ¼ 1; 2; 3; with summation with respect to a: ðA:138Þ
Similarly, the curl of a vector V is expressed in the form
o   oVsb ois
curlV ¼ r  V ¼ isa  Vsb isb ¼ isa  isb þ Vsb isa  b ;
osa osa osa
we may write
isa  isb ¼ cabc isc ; ðA:139Þ

where from
A.2 Curvilinear Co-ordinates 761

oxi oxj oxk


cijk ¼ cabc : ðA:140Þ
osa osb osc

Then
oisb oxi o2 xj oxk
isa  ¼ cijk is ;
osa osa osa osb osc c

where we used the relation


oxi
ij ¼ is ; ðA:1320 Þ
osa a
analogue to the relation (A.132); by means of the formula (A.140), it follows that
oisb oxi o2 xi
isa  ¼ cacd is : ðA:141Þ
osa osd osa osb c

Observing that cacd is antisymmetric with respect to a and d, we may write


ois 1 oxi o2 xi oxi o2 xi


isa  b ¼  cacd  is ;
osa 2 osd osa osb osa osd osb c

but

1 oxi o2 xi oxi o2 xi 1 oxi o2 xi oxi o2 xi


  ¼ þ
2 osd osa osb osa osd osb 2 osa osd osb osb osa osd

1 oxi 1 ohd oxi 1 oha oxi 1 1 ohd 1 oha


¼  ¼ dbd  dab ð!Þ;
2 osb hb osa osd ha osd osa 2 hd osa ha osd

where we took into account (A.129) and (A.133). The relation (A.141) becomes
oisb 1 ohb
isa  ¼ cabc is ð!Þ;
osa hb osa c
with summation with respect to a and c: ðA:1410 Þ

Finally, we get
X
3

oVsb 1 ohb
curlV ¼ cabc  Vs isc
b¼1
osa hb osa b
X3

o oVsb
¼ cabc hb isc : ðA:142Þ
b¼1
osa hb

Taking into account the definition formula (A.80) and the differential operators
(A.131) and (A.138), we may write Laplace’s operator in the form
762 Appendix

o2 1 ohb 1 ohc o o2
D¼  þ ¼
osa osa hb osa hc osa osa osa osa

ðA:143Þ
1 oðhb hc Þ o o 1 o
 ¼ hb hc ð!Þ:
hb hc osa osa osa hb hc osa

A.2.2.2 Cylindrical and Spherical Co-ordinates

In cylindrical co-ordinates, the arc element is given by

ds2 ¼ dr 2 þ r 2 dh2 þ dz2 ; ðA:144Þ


while, along the three co-ordinate lines, we have
dsr ¼ dr; dsh ¼ rdh; dsz ¼ dz: ðA:1440 Þ
As well, we may write
o o o 1 o o o
¼ ; ¼ ; ¼ : ðA:14400 Þ
osr or osh r oh osz oz
The volume element is expressed in the form
dV ¼ dsr dsh dsz ¼ rdrdhdz: ðA:145Þ
The operator r becomes
o 1 o o
r ¼ grad ¼ ir þ ih þ iz : ðA:146Þ
or r oh oz
The divergence of the vector V is given by
1o 1 oVh oVz
divV ¼ ðrVr Þ þ þ ; ðA:147Þ
r or r oh oz
while the curl of the same vector reads

1 oVz oVh oVr oVz


curlV ¼  ir þ  ih
r oh oz oz or

1 o oVr
þ ðrVh Þ  iz : ðA:148Þ
r or oh

For Laplace’s operator one may write


1o o 1 o2 o2
D¼ r þ 2 2þ 2: ðA:149Þ
r or or r oh oz
A.2 Curvilinear Co-ordinates 763

In spherical co-ordinates, the arc element is given by

ds2 ¼ dR2 þ R2 du2 þ R2 sin2 udh2 ; ðA:150Þ

while, along the three co-ordinate lines, we have


dsR ¼ dR; dsu ¼ Rdu; dsh ¼ R sin u dh: ðA:1500 Þ
As well, we may write
o o o 1 o o 1 o
¼ ; ¼ ; ¼ : ðA:15000 Þ
osR oR osu R ou osh R sin u oh
The volume element is expressed in the form

dV ¼ dsR dsu dsh ¼ R2 sinudRdudh: ðA:151Þ


The operator r becomes
o 1 o 1 o
r ¼ grad ¼ iR þ iu þ ih : ðA:152Þ
oR R ou R sin u oh
The divergence of the vector V reads
1 o 2 1 o 1 oVh
divV ¼ ðR VR Þ þ ðsin uVu Þ þ ; ðA:153Þ
R2 oR R sin u ou R sin u oh
while the curl of the same vector is given by

1 o oVu
curlV ¼ ðsin uVh Þ  iR
R sin u ou oh

1 1 oVR o
þ  ðRVh Þ iu
R sin u oh oR

1 o oVR
þ ðRVu Þ  ih : ðA:154Þ
R oR ou
For Laplace’s operator it results

1 o 2 o 1 o o 1 o2
D¼ 2 R þ 2 sin u þ 2 2 : ðA:155Þ
R oR oR R sin u ou ou R sin u oh2

A.3 Elements of the Theory of Distributions

In the study of discontinuous phenomena and for their representation in a unitary


form, together with the continuous ones, it is necessary to use some notions of the
theory of distributions. In what follows, we give some results concerning the
composition of distributions and the integral transforms in distributions; as well,
we introduce the notion of basic solution of a differential equation in the sense of
the theory of distributions. These notions acquire thus a larger interest.
764 Appendix

A.3.1 Fundamental Results of the Theory of Distributions

We shall state hereafter the notion of distribution, giving a classification of them


too; some fundamental formulae will be emphasized and the derivative of a
distribution will be introduced.

A.3.1.1 The Notion of Distribution. Classification of Distributions

We call functional a mapping of a vector space X (with respect to C ) into C. If C is


the corpus of real numbers R, then we say that we have to do with a real
functional. We call distribution a continuous linear functional, defined on a
topological vector space X (fundamental space).
By definition, the fundamental space K is constituted of the functions of real
variables uðxÞ (x  ðx1 ; x2 ; . . .; xn Þ represents a point in Rn ), indefinite
differentiable (of class C1 ) and vanishing with all their derivatives in the
exterior of certain bounded domains; these domains, together with their
bounderies, determine the support of these functions, called fundamental
functions (by support of a function uðxÞ we mean the smallest closed set which
contains the set of points x for which uðxÞ 6¼ 0). By an extension of the space K we
reach another class of functions, which determine the fundamental space S. The
functions belonging to this class have also the property to be indefinite
differentiable; for jxj ! 1 these functions tend to zero together with their
derivatives of any order, more rapidly than any power of 1=jxj. We introduce also
the space K m which includes the functions with compact support, having
continuous derivatives up to and including the mth order (of class Cm ). The
distributions defined on the spaces K, S and K m are called of infinite order,
temperate and of finite order p
m, respectively.
Let f ðxÞ be a function defined on the real axis R; we say that this function is
absolutely integrable in a closed interval ½a; b of R if the integral
Zb
jf ðxÞjdx\1 ðA:156Þ
a

exists.
If the function f ðxÞ is absolutely integrable in any finite interval of R, then we
say that f ðxÞ is a locally integrable function. We mention that an absolutely
integrable function is integrable too, i.e. the integral
Zb
f ðxÞdx ðA:1560 Þ
a

exists.
A.3 Elements of the Theory of Distributions 765

The locally integrable functions generate an important class of distributions; we


assume thus that to any fundamental functions uðxÞ 2 K there corresponds a real
number
Z Zb
ðf ; uÞ ¼ f ðxÞuðxÞdx ¼ f ðxÞuðxÞdx; ðA:157Þ
R a

where f ðxÞ is a locally integrable function and ½a; b is the support of uðxÞ. It is
easy to see that the functional thus defined is linear and continuous. The functional
defined on the space K by means of the locally integrable function f ðxÞ represents
a distribution on this space which will be denoted by f ðxÞ too, like the generating
function. Such distributions are called regular distributions (distributions of
functional type). Similarly, one can define temperate regular distributions on the
space S.
The distributions which are not regular distributions are called singular
distributions.
If to any function uðxÞ 2 K we attach its value at the origin (the value uð0Þ),
then we see that the respective functional is linear and continuous, hence it is a
distribution which is not regular; this is the Dirac distribution, which will be
denoted by the symbol dðxÞ. We may write
ðdðxÞ; uðxÞÞ ¼ uð0Þ: ðA:158Þ
We can define the Dirac distribution on the space K 0 too; it will be called, in
this case, the Dirac measure.
If the fundamental functions u 2 K are defined in Rn , then we have
ðdðx1 ; x2 ; . . .; xn Þ; uðx1 ; x2 ; . . .; xn ÞÞ ¼ uð0; 0; . . .; 0Þ: ðA:1580 Þ
The relation
ðf ðx  x0 Þ; uðxÞÞ ¼ ðf ðxÞ; uðx þ x0 ÞÞ; x; x0 2 Rn ; ðA:159Þ
defines the translated distribution f ðx  x0 Þ. In particular, for the Dirac
distribution dðx  x0 Þ we may write
ðdðx  x0 Þ; uðxÞÞ ¼ uðx0 Þ; x; x0 2 Rn ; ðA:1590 Þ
For the distributions subjected to a homothetic transformation with respect to
the independent variable, we shall use—by definition—the formula
ðf ðax; uðxÞÞ ¼ jajn ðf ðxÞ; uðx=aÞÞ; x 2 Rn : ðA:160Þ

For a ¼ 1 we get the property of symmetry and we may write


ðf ðxÞ; uðxÞÞ ¼ ðf ðxÞ; uðxÞÞ; x 2 Rn ; ðA:161Þ
in particular, we notice that dðxÞ ¼ dðxÞ, hence the Dirac distribution is even
with respect to the independent variable x 2 Rn .
766 Appendix

The equality of two distributions f ðxÞ and gðxÞ is defined by the relation
ðf ; uÞ ¼ ðg; uÞ; 8u 2 K; ðA:162Þ
hence, we may write
f ¼ g: ðA:163Þ
If the distributions f and g are generated by continuous functions f ðxÞ and gðxÞ,
then the equality (A.163) occurs in the usual sense, i.e. punctual, because—in this
case—the distributions f and g coincide everywhere with the functions f and g. If
the functions f and g are locally integrable and coincide almost everywhere, then
the distributions generated by them will be equal and the relation (A.163) holds.
A distribution f ðxÞ is equal to zero (f ¼ 0) if, for any fundamental function
uðxÞ, we have ðf ðxÞ; uðxÞÞ ¼ 0.
We define the Heaviside function (the unit function) on the real axis in the form
(Fig. A.10)

1; x 0;
hðxÞ ¼ ðA:164Þ
0; x \ 0;

the distribution generated by it will be called the Heaviside distribution.


If f ðxÞ is a function of variable x, defined on the real axis, then we will call its
positive part the function fþ , defined by the relation (Fig. A.11a)

f ðxÞ; x 0;
fþ ðxÞ ¼ f ðxÞhðxÞ ¼ ðA:165Þ
0; x \ 0:

Fig. A.10 Heaviside


function
θ(x)

O x

Fig. A.11 Positive part of a


f+(x) x+
function; the functions fþ ðxÞ
(a) and xþ (b)
1

O x O x
(a) (b)
A.3 Elements of the Theory of Distributions 767

Fig. A.12 Function 1ðxÞ


1(x)
1

O x

In particular, for the function f ðxÞ ¼ x, we can introduce the positive part
(Fig. A.11b)

x; x 0;
xþ ¼ xhðxÞ ¼ ðA:1650 Þ
0; x \ 0;

obviously, to these functions correspond certain distributions of function type. We


introduce also the distribution generated by the function 1ðxÞ ¼ 1; 8x 2 R
(Fig. A.12). The characteristic function corresponding to the interval ½a; a is
defined in the form

0; jxj [ a;
hðxÞ ¼ a [ 0; ðA:166Þ
1; jxj
a;

being thus led to a distribution of function type; this distribution may be expressed
also by means of the Heaviside distribution
hðxÞ ¼ hðx þ aÞ  hðx  aÞ ¼ hða  jxjÞ; x 2 R: ðA:1660 Þ
If wðxÞ is a function of class C1 , then we can write the equality
wðxÞdðxÞ  wð0ÞdðxÞ; x 2 R; ðA:167Þ
in particular, for wðxÞ ¼ xn , we obtain
xn dðxÞ ¼ 0; n ¼ 1; 2; . . . ðA:168Þ

so that the product of a distribution by an indefinite differentiable function may be


equal to zero, even if no one of its factors vanishes. Similarly
wðxÞdðx  aÞ ¼ wðaÞdðx  aÞ: ðA:1670 Þ
We call support of the distribution f (supp f ) the component of the union of open
sets on which this distribution vanishes; therefore, the support of a distribution is a
closed set. If the support of a distribution is contained in a set A, then we say that
the distribution is concentrated on the set A. Thus, we can say that the Dirac
distribution dðxÞ is concentrated at a point (the origin). Similarly, one can define
distributions concentrated on curves or surfaces, in general distributions
concentrated on a manifold of a space Rn .
We say that a sequence of functions fn ðx1 ; x2 ; . . .; xm Þ is a d representative
sequence if, in the sense of the topology of K 0 (K 0 is the topological dual of K,
containing the distributions defined on the fundamental space) we have (in
768 Appendix

Fig. A.13 Representative


dðxÞ sequence fn(x)
fn

f2
f1

O x

Fig. A.13 is given a d representative sequence for m ¼ 1)


lim fn ðx1 ; x2 ; . . .; xm Þ ¼ dðx1 ; x2 ; . . .; xm Þ; ðA:169Þ
n!1

obviously, this condition is equivalent to


lim ðfn ðx1 ; x2 ; . . .; xm Þ; uðx1 ; x2 ; . . .; xm ÞÞ ¼ uð0; 0; . . .; 0Þ: ðA:1690 Þ
n!1

A.3.1.2 Differentiation of Distributions

From the very beginning, we mention that the distributions admit derivatives of
any order, which is a great advantage with respect to usual functions.
Let f ðxÞ be a function of class C1 and uðxÞ a fundamental function belonging to
the fundamental space K; considering the corresponding distribution of function
type, we obtain the rule of differentiation in the form
ðf 0 ; uÞ ¼ ðf ; u0 Þ: ðA:170Þ
In particular, we have
h0 ðxÞ ¼ dðxÞ: ðA:171Þ
In case of a distribution of several variables f ðx1 ; x2 ; . . .; xn Þ we can write

o
f ðx1 ; x2 ; :::; xn Þ; uðx1 ; x2 ; :::; xn Þ
oxi

o ðA:172Þ
¼ f ðx1 ; x2 ; :::; xn Þ; uðx1 ; x2 ; :::; xn Þ ; i ¼ 1; 2; :::; n;
oxi
one obtains also the property

o2 f o2 f
¼ ; i; j ¼ 1; 2; . . .; n; ðA:173Þ
oxi oxj oxj oxi

which shows that, in case of distributions, the derivatives do not depend on the
order of differentiation.
A.3 Elements of the Theory of Distributions 769

Let f ðxÞ be a function of class C1 everywhere, excepting at the point x0 , where


the function has a discontinuity of the first species, the corresponding jump being
given by
s0 ¼ f ðx0 þ 0Þ  f ðx0  0Þ: ðA:174Þ
As well, we denote by f 0 ðxÞ the derivative of the distribution f ðxÞ (in the sense of
the theory of distributions) and by ~f 0 ðxÞ the distribution corresponding to the
derivative of the function which generated the distribution, in the usual sense,
wherever this derivative exists; we obtain the relation

f 0 ðxÞ ¼ ~f 0 ðxÞ þ s0 dðx  x0 Þ: ðA:175Þ


It is worth to note that if the function f ðxÞ is continuous at the point x0 , then the
jump s0 vanishes and the formula (A.175) becomes

f 0 ðxÞ ¼ ~f 0 ðxÞ; ðA:176Þ


i.e. the derivative in the sense of the theory of distributions coincides with the
derivative in the usual sense.
If the function f ðxÞ is of class C1 everywhere, excepting the points
xi ; i ¼ 1; 2; . . .; n, where it has discontinuities of the first species, and if we
denote by si the jump of the function at the point xi , then, by a similar procedure,
we obtain a more general formula, namely
X
n
f 0 ðxÞ ¼ ~f 0 ðxÞ þ si dðx  xi Þ: ðA:177Þ
i¼1

A last property, which is worth to be revealed, is the following: If the derivative


of a distribution is equal to zero, then the distribution is a constant.

A.3.2 Composition of Distributions

In general, the product of two distributions has no meaning; we have seen that the
products of a distribution by a function of class C1 has sense. That is why we will
define products of a special type (composition of distributions). We introduce thus
the direct (or tensor) product and the convolution product.

A.3.2.1 Direct Product of Two Distributions

Let be x  ðx1 ; x2 ; . . .; xn Þ a point of the n-dimensional Euclidean space X n and


y  ðy1 ; y2 ; . . .; ym Þ a point of the m-dimensional Euclidean space Y m , By direct
Cartesian product X n  Y m of the two Euclidean spaces we mean a new ðn þ mÞ-
dimensional Euclidean space, built up of the points ðx; yÞ  ðx1 ; x2 ; . . .; xn ;
y1 ; y2 ; . . .; ym Þ, where—obviously—we have put in evidence the co-ordinates of
a point of this space, in the order in which they have been written.
770 Appendix

The direct product f ðxÞ  gðyÞ of two distributions f ðxÞ and gðyÞ; defined on the
basic spaces Kx ðx 2 X n Þ and Ky ðy 2 Y m Þ; respectively, is given by the relation
ðf ðxÞ  gðyÞ; uðx; yÞÞ ¼ ðf ðxÞ; ðgðyÞ; uðx; yÞÞÞ; ðA:178Þ
where uðx; yÞ is a fundamental function defined on X n  Y m ; this product is a
distribution defined on the fundamental space Kx  Ky . In the case of usual
functions, this product coincides with their usual product.
The direct product is commutative
f ðxÞ  gðyÞ ¼ gðyÞ  f ðxÞ ðA:179Þ
and associative
½f ðxÞ  gðyÞ  hðzÞ ¼ f ðxÞ  ½gðyÞ  hðzÞ ¼ f ðxÞ  gðyÞ  hðzÞ: ðA:180Þ
The first of these properties allows to write the definition relation (A.178) also in
the form
ðf ðxÞ  gðyÞ; uðx; yÞÞ ¼ ðgðyÞ; ðf ðxÞ; uðx; yÞÞÞ: ðA:1780 Þ

The second property shows that the direct product may be defined for an arbitrary
finite number of distributions.
Let Dx and Dy be two differential operators with respect to the variables x and y,
respectively; we may write the relation
Dx Dy ½f ðxÞ  gðyÞ ¼ Dx f ðxÞ  Dy gðyÞ: ðA:181Þ

In particular, we get
o2 ohðxÞ ohðyÞ
½hðxÞ  hðyÞ ¼  ¼ dðxÞ  dðyÞ ¼ dðx; yÞ: ðA:182Þ
oxoy ox oy

A.3.2.2 Convolution Product of Two Distributions

Let f ðxÞ and gðxÞ be locally integrable functions of x; their convolution product is
the function defined by
Z1
f ðxÞ gðxÞ ¼ f ðnÞgðx  nÞdn; ðA:183Þ
1

obviously, the definition remains valid for x 2 Rn :


If the functions f ðxÞ and gðxÞ are continuous, then their convolution product is a
continuous function too. In order that the convolution product may exist, it is
necessary that the functions f ðxÞ and gðxÞ should satisfy certain conditions; thus, a
A.3 Elements of the Theory of Distributions 771

sufficient condition in this respect is that the support of the two functions f ðxÞ and
gðxÞ be compact.
If f ðxÞ and gðxÞ are two distributions on Rn , then their convolution product
f ðxÞ gðxÞ represents a new distribution on Rn , defined by the formula
ðf ðxÞ gðxÞ; uðxÞÞ ¼ ðf ðxÞ  gðyÞ; uðx þ yÞÞ
¼ ðf ðxÞ; ðgðyÞ; uðx þ yÞÞÞ ¼ ðgðyÞ; ðf ðxÞ; uðx þ yÞÞÞ; ðA:184Þ

this definition is reduced to the first one in the case of usual functions. We may
show that the convolution product has a meaning if one of the following conditions
is satisfied:
1. one of the distributions f ðxÞ; gðxÞ has a compact support;
2. the distributions f ðxÞ and gðxÞ have the support bounded on the same side;
thus, if f ðxÞ ¼ 0 for x\a and gðxÞ ¼ 0 for x\b, then the supports of the two
distributions are bounded on the same side.
We remark that the convolution product may be defined for an arbitrary finite
number of distributions. Under the conditions required for the existence of the
convolution product, one may prove the property of commutativity
f ðxÞ gðxÞ ¼ gðxÞ f ðxÞ ðA:185Þ
and the property of associativity
½f ðxÞ gðxÞ hðxÞ ¼ f ðxÞ ½gðxÞ hðxÞ ¼ f ðxÞ gðxÞ hðxÞ: ðA:186Þ

We notice that
dðxÞ f ðxÞ ¼ f ðxÞ dðxÞ ¼ f ðxÞ; ðA:187Þ
hence, Dirac’s distribution is a unit element for the convolution product.
If D is an arbitrary differential operator, then we may write
D½f ðxÞ gðxÞ ¼ Df ðxÞ gðxÞ ¼ f ðxÞ DgðxÞ: ðA:188Þ

A.3.3 Integral Transforms in Distributions

A strong tool for the integration of differential equations is the method of integral
transforms. We give, in the following, some general results concerning Fourier and
Laplace transforms.
772 Appendix

A.3.3.1 Fourier Transform of a Distribution

If f ðxÞ is a real or complex function of the real variable x 2 R, which satisfies


Dirichlet’s conditions (it is bounded, piecewise monotone and has at most a finite
number of points of discontinuity of the first species) and is absolutely integrable,
then the function
Z1 pffiffiffiffiffiffiffi
FðuÞ ¼ f ðxÞ eiux dx; i ¼ 1; ðA:189Þ
1

exists and is called the Fourier transform of the function f ðxÞ; we shall write

F½f ðxÞ ¼ FðuÞ ¼ ~f ðuÞ; ðA:1890 Þ

noting that the variable u is real. In general, the image function FðuÞ is complex,
although the function f ðxÞ may be a real function.
Assuming that the function FðuÞ is given, the equality (A.189) may be
considered as an integral equation with respect to the unknown function f ðxÞ under
the integral symbol; the solution of this integral equation is written in the form
Z1
1
f ðxÞ ¼ FðuÞ eiux du: ðA:190Þ
2p
1

The function f ðxÞ is called the inverse Fourier transform of the function FðuÞ; we
have

f ðxÞ ¼ F1 ½F½f ðxÞ ¼ F1 ½FðuÞ: ðA:1900 Þ


Let uðxÞ be a complex fundamental function of a real variable x; hence, uðxÞ 2
C1 and has a compact support (e.g., jxj
a). By the formula (A.189), the Fourier
transform of this fundamental function is
Z1
F½uðxÞ ¼ wðuÞ ¼ uðxÞ eiux dx: ðA:191Þ
1

The function wðuÞ may be definite for complex values s ¼ u þ iv too, namely
Z1 Z1
wðsÞ ¼ isx
uðxÞ e dx ¼ uðxÞ evx eiux dx: ðA:1910 Þ
1 1

The set of functions wðsÞ ¼ F½uðxÞ, where the support of the fundamental
functions is included in the segment ½a; a, forms the vector space ZðaÞ. We
denote by
Z ¼ [ ZðaÞ; K ¼ [ KðaÞ ðA:192Þ
a a
A.3 Elements of the Theory of Distributions 773

0
the new complex linear space; then, Z is the set of linear and continuous
functional defined on Z (ultradistributions).
If FðsÞ is a distribution defined on Z and f ðxÞ is a distribution defined on K, then
0
the functional FðsÞ 2 Z specified by the equality of Parseval type
ðFðsÞ; wðsÞÞ ¼ 2pðf ðxÞ; uðxÞÞ; ðA:193Þ
is called the Fourier transform of the distribution f ðxÞ and is denoted by

FðsÞ ¼ F½f ðxÞ ¼ ~f ðsÞ: ðA:194Þ


We can also write
ðF½f ðxÞ; F½uðxÞÞ ¼ 2pðf ðxÞ; uðxÞÞ: ðA:1930 Þ
Analogically, one may introduce the Fourier transform of a distribution of
several variables.
The classical properties of the Fourier transform are maintained in the form:

d d
P FðsÞ ¼ P F½f ðxÞ ¼ F½PðixÞf ðxÞ; P polynomial; ðA:195Þ
ds ds


d
F P f ðxÞ ¼ PðisÞF½f ðxÞ ¼ PðisÞFðsÞ; P polynomial; ðA:196Þ
dx

F1 ½F½f ðxÞ ¼ f ðxÞ; ðA:197Þ


F½F½f ðxÞ ¼ 2pf ðxÞ; ðA:1970 Þ
0
where by F1 is denoted the inverse operator, defined on Z .
We can prove the relation
F½f ðxÞ  gðyÞ ¼ F½f ðxÞ  F½gðyÞ ðA:198Þ
for the direct product of two distributions. As well, in connection with the
convolution product, one may show the relations
F½f ðxÞ gðxÞ ¼ F½f ðxÞF½gðxÞ; ðA:199Þ
F½f ðxÞgðxÞ ¼ F½f ðxÞ F½gðxÞ: ðA:1990 Þ
0
The first relation is valid if f ðxÞ 2 S and gðxÞ is a distribution with bounded
0
support; the second relation is valid if f ðxÞ 2 S and the function gðxÞ 2 C1 is such
0
that f ðxÞgðxÞ 2 S and the support of the Fourier transform is bounded.
We mention the Fourier transforms
F½dðxÞ ¼ 1; ðA:200Þ
F½dðx1 ; x2 ; . . .; xn Þ ¼ 1; ðA:2000 Þ

F½dðx  aÞ ¼ eiua ; ðA:201Þ


774 Appendix

1
F½hðxÞ ¼ pdðuÞ þ ; ðA:202Þ
u
1 0
F½xþ  ¼   ipd ðuÞ; ðA:203Þ
u2
F½1ðxÞ ¼ 2pdðuÞ; ðA:204Þ

where 1ðxÞ is the function which takes the value 1 for any x.

A.3.3.2 Laplace Transform of a Distribution

Let f ðxÞ be a complex function of a real variable, which satisfies the conditions:
1. f ðxÞ ¼ 0 for x\0,
2. f ðxÞ is piecewise differentiable,
3. jf ðxÞj
Meax , where M is a positive constant, while the non-negative constant a
represents the incremental ratio of the function.The function LðpÞ of the
complex variable p ¼ u þ iv, defined by the expression

Z1
LðpÞ ¼ f ðxÞepx dx; ðA:205Þ
0

is called the Laplace transform of the function f ðxÞ and is denoted by

L½f ðxÞ ¼ LðpÞ ¼ ~f ðpÞ: ðA:2050 Þ


The function f ðxÞ is called the original function and the function LðpÞ the image
function. To the Laplace transform thus defined there corresponds an inverse
Laplace transform, given by
Z
uþi1
1 1
L ½LðpÞ ¼ f ðxÞ ¼ LðpÞepx dp; u [ a: ðA:206Þ
2pi
ui1

If f ðxÞ is a distribution having its support on the half-line x 0 and is such that
the distribution f ðxÞepx is a temperate distribution, then

L½f ðxÞ ¼ ðf ðxÞ; epx Þ ¼ ~f ðpÞ ðA:207Þ

represents the Laplace transform of that distribution. It is obvious that the relation
(A.207) generalizes the relation (A.205)
One verifies easily the delay theorem
L½f ðx  aÞ ¼ epa L½f ðxÞ; ðA:208Þ

the theorem of similitude


A.3 Elements of the Theory of Distributions 775

1 p
L½f ðkxÞ ¼ L ; k [ 0; ðA:209Þ
k k
and the theorem of translation (the damping theorem)
L½f ðxÞeqx  ¼ Lðp  qÞ: ðA:210Þ
In the case of a derivative of a distribution one may write
L½f 0 ðxÞ ¼ pL½f ðxÞ: ðA:211Þ
One may give analogous results for the distributions of several variables.
For a convolution product it results
L½f ðxÞ gðxÞ ¼ L½f ðxÞL½gðxÞ: ðA:212Þ
We may write the Laplace transforms
L½dðxÞ ¼ 1: ðA:213Þ
L½dðx1 ; x2 ; . . .; xn Þ ¼ 1: ðA:2130 Þ

L½dðmÞ ðxÞ ¼ pm ; m ¼ 0; 1; 2; . . . ðA:214Þ

A.3.4 Applications to the Study of Differential Equations.


Fundamental Solutions

The theory of distributions is particularly useful in the study of ordinary or partial


differential equations, as well as in the case of various boundary value problems.
Firstly, we shall give some general results concerning the fundamental solutions
and then we shall deal with the problem of obtaining them for some particular
differential equations.

A.3.4.1 Ordinary Differential Equations

Let be the linear ordinary differential equation with constant coefficients

DyðxÞ ¼ yðnÞ ðxÞ þ a1 yðn1Þ ðxÞ þ    þ an yðxÞ ¼ f ðxÞ; ðA:215Þ


where f ðxÞ is a distribution. The distribution EðxÞ which satisfies the equation
DEðxÞ ¼ dðxÞ; ðA:216Þ
is called the fundamental solution of this equation and is of the form
EðxÞ ¼ YðxÞ þ Eþ ðxÞ; ðA:217Þ

where YðxÞ is the general solution of the homogeneous equation


776 Appendix

DYðxÞ ¼ 0; ðA:218Þ
while Eþ ðxÞ is a particular fundamental solution {corresponding to the non-
homogeneous equations (A.216)}. We shall give a simple method for determining
this solution; to this end, we determine first the solution YðxÞ of the homogeneous
equation, which verifies the initial conditions
00
Yð0Þ ¼ Y 0 ð0Þ ¼ Y ð0Þ ¼    ¼ Y ðn2Þ ð0Þ ¼ 0; Y ðn1Þ ð0Þ ¼ 1; ðA:219Þ
one can prove that a fundamental particular solution is, in this case, given by
Eþ ðxÞ ¼ hðxÞYðxÞ ðA:220Þ
The fundamental solution of a differential equation is useful for the determination
of the general solution; thus, the general solution of the Eq. (A.215) is given by
yðxÞ ¼ EðxÞ f ðxÞ: ðA:221Þ
Let now be again the Eq. (A.215) with x 0, f ðxÞ being a continuous function
with the support in ½0; 1Þ; in the case of initial conditions of Cauchy type

yðkÞ ð0Þ ¼ yk ; k ¼ 0; 1; 2; . . .; n  1; ðA:222Þ

the solution of the Eq. (A.215) is expressed in the form


X
n1
dk
yðxÞ ¼ Eþ ðxÞ f ðxÞhðxÞ þ hk Eþ ðxÞ; ðA:223Þ
k¼0
dxk

where the coefficients hk ; k ¼ 0; 1; 2; . . .; n  1 are given by


hk ¼ ynk1 þ a1 ynk2 þ    þ ank1 y0 : ðA:224Þ
The solution of the homogeneous equation
DyðxÞ ¼ 0 ðA:225Þ
for x 0, with the initial conditions of Cauchy type (A.222),
X
n1
dk
yðxÞ ¼ hk Eþ ðxÞ: ðA:226Þ
k¼0
dxk

In particular, in the case of the differential equation

yðnÞ ðxÞ ¼ f ðxÞ; ðA:227Þ


with initial conditions of the form (A.222), we obtain the fundamental particular
solution
A.3 Elements of the Theory of Distributions 777

xn1 1 n1
Eþ ðxÞ ¼ hðxÞ ¼ x ; x 2 R; ðA:228Þ
ðn  1Þ! ðn  1Þ! þ
the solution of the boundary value problem is given by
x2 xn1
yðxÞ ¼ y0 þ y1 x þ y2 þ    þ yn1
2 ðn  1Þ!
Zx
1
þ ðx  nÞn1 f ðnÞ dn; ðA:229Þ
ðn  1Þ!
0

where the latter integral, which represents the convolution product, is known as the
Cauchy formula; for n  2, we get
Eþ ðxÞ ¼ xþ : ðA:230Þ
The above ideas may be extended to systems of ordinary differential equations
with constant coefficients.

A.3.4.2 General Considerations on Partial Differential Equations

Problems similar to those in the preceding subsection may be put in the case of
partial differential equations. Let thus be

o o o o
P ; ; . . .; ; uðx1 ; x2 ; . . .; xm ; tÞ ¼ 0 ðA:231Þ
ox1 ox2 oxm ot

a homogeneous linear partial differential equation of nth order with respect to the
variable t, with constant coefficients. For instance, the Cauchy problem for this
equation consists in the determination of the function uðx1 ; x2 ; . . .; xm ; tÞ which
satisfies the Eq. (A.231) and the initial conditions
uðx1 ; x2 ; . . .; xm ; t0 Þ ¼ u0 ðx1 ; x2 ; . . .; xm Þ;
o
uðx1 ; x2 ; . . .; xm ; t0 Þ ¼ u1 ðx1 ; x2 ; . . .; xm Þ;
ot
. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . . ðA:232Þ

on1
uðx1 ; x2 ; . . .; xm ; t0 Þ ¼ un1 ðx1 ; x2 ; . . .; xm Þ:
otn1
To solve this problem, let us consider the function
uðx1 ; x2 ; . . .; xm ; tÞ ¼ uðx1 ; x2 ; . . .; xm ; tÞhðt  t0 Þ;
 ðA:233Þ
as well as the corresponding regular distribution; taking into account the formula
which links the derivative in the sense of the theory of distributions to the
derivative in the usual sense and using the initial conditions (A.232), it results
778 Appendix

o ~
o
uðx1 ; x2 ; . . .; xm ; tÞ ¼  uðx1 ; x2 ; . . .; xm ; tÞ
ot ot
þu0 ðx1 ; x2 ; . . .; xm Þdðt  t0 Þ;
o 2 ~
o2

u ðx 1 ; x 2 ; . . .; x m ; tÞ ¼ uðx1 ; x2 ; . . .; xm ; tÞ

ot2 ot2
þu1 ðx1 ; x2 ; . . .; xm Þdðt  t0 Þ þ u0 ðx1 ; x2 ; . . .; xm Þdðt _  t0 Þ;
ðA:234Þ
. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .

on ~
on

u ðx 1 ; x 2 ; . . .; xm ; tÞ ¼ uðx1 ; x2 ; . . .; xm ; tÞ

otn otn
þun1 ðx1 ; x2 ; . . .; xm Þdðt  t0 Þ þ un2 ðx1 ; x2 ; . . .; xm Þdðt _  t0 Þ
þ. . . þ u0 ðx1 ; x2 ; . . .; xm Þdðn1Þ ðt  t0 Þ:

Noting that the derivatives in the usual sense with respect to the temporal variable
of the function  uðx1 ; x2 ; . . .; xm ; tÞ are equal to the corresponding ones of the
function uðx1 ; x2 ; . . .; xm ; tÞ for t t0 , the Eq. (A.231) takes the form

o o o o
P ; ; . . .; ; uðx1 ; x2 ; . . .; xm ; tÞ ¼ f ðx1 ; x2 ; . . .; xm ; tÞ;
 ðA:2310 Þ
ox1 ox2 oxm ot
in distributions, where f ðx1 ; x2 ; . . .; xm ; tÞ is a given distribution, which contains the
initial conditions considered above.
Thus, we call fundamental solution of the Eq. (A.2310 ) the distribution
Eðx1 ; x2 ; . . .; xm ; tÞ which satisfies the equation

o o o o
P ; ; . . .; ; Eðx1 ; x2 ; . . .; xm ; tÞ ¼ dðx1 ; x2 ; . . .; xm ; tÞ: ðA:235Þ
ox1 ox2 oxm ot
The solution of the above Cauchy problem is given by (A.233) where
uðx1 ; x2 ; . . .; xm ; tÞ ¼ Eðx1 ; x2 ; . . .; xm ; tÞ f ðx1 ; x2 ; . . .; xm ; tÞ;
 ðA:236Þ
the convolution product corresponding to all m þ 1 variables.
It should be noted that some equations of mathematical physics cannot be
always deduced directly in the space of distributions, owing to the difficulties
encountered in modelling physical phenomena. In general, the equations which
describe such phenomena are obtained first by classical method. Next, an
extension is effected, where the unknown functions take zero values, so that they
be defined on the whole space; the derivatives, considered in the usual sense, are
expressed by relations which connect derivatives in the sense of the theory of
distributions to the derivatives in the usual sense of a distribution corresponding to
an almost everywhere continuous function, having a finite number of
discontinuities of the first species. In this way, the unknowns of the problem
will be regular distributions; then it will be assumed that these unknowns may be
arbitrary distributions. Another possibility, which is frequently used, is to suppose
A.3 Elements of the Theory of Distributions 779

from the very beginning, that the unknowns of the problem are arbitrary
distributions, assuming the same form in distributions for the differential equation
obtained by classical methods (obviously, these ones are no longer valid for the
whole space). However, there are not general methods for passing to differential
equations in distributions.

A.3.4.3 Equations of Elliptic Type. Fundamental Solutions

Let be Poisson’s equation


Duðx1 ; x2 ; x3 Þ ¼ f ðx1 ; x2 ; x3 Þ; ðA:237Þ

where f ðx1 ; x2 ; x3 Þ is a given distribution; the fundamental solution is of the form


1 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Eðx1 ; x2 ; x3 Þ ¼  ; R ¼ x21 þ x22 þ x23 : ðA:2370 Þ
4pR
Analogically, for the equation
DDuðx1 ; x2 ; x3 Þ ¼ f ðx1 ; x2 ; x3 Þ ðA:238Þ
we have
1
Eðx1 ; x2 ; x3 Þ ¼  R: ðA:2380 Þ
8p
In the case of the equation

Duðx1 ; x2 ; x3 Þ þ k2 uðx1 ; x2 ; x3 Þ ¼ f ðx1 ; x2 ; x3 Þ ðA:239Þ


we may use the fundamental solution
1
Eðx1 ; x2 ; x3 Þ ¼  cos kR; ðA:2390 Þ
4pR
we also notice that an integral of the homogeneous Helmholtz equation

Duðx1 ; x2 ; x3 Þ þ k2 uðx1 ; x2 ; x3 Þ ¼ 0 ðA:240Þ


is given by
1
uðx1 ; x2 ; x3 Þ ¼ sin kR: ðA:2400 Þ
R
pffiffiffiffiffiffiffi
Replacing k by ik; i ¼ 1, we find the fundamental solution
1
Eðx1 ; x2 ; x3 Þ ¼  cosh kR ðA:241Þ
4pR
for the equation
Duðx1 ; x2 ; x3 Þ  k2 uðx1 ; x2 ; x3 Þ ¼ f ðx1 ; x2 ; x3 Þ; ðA:2410 Þ
780 Appendix

similarly, we notice that


1
uðx1 ; x2 ; x3 Þ ¼  sinh kR ðA:242Þ
R
is an integral of the homogeneous equation

Duðx1 ; x2 ; x3 Þ  k2 uðx1 ; x2 ; x3 Þ ¼ 0: ðA:2420 Þ


In the case of Poisson’s equation in two variables
Duðx1 ; x2 Þ ¼ f ðx1 ; x2 Þ; ðA:243Þ
where f ðx1 ; x2 Þ is a given distribution, a fundamental solution is
1 1 qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Eðx1 ; x2 Þ ¼  ln ; r ¼ x21 þ x22 : ðA:2430 Þ
2p r
As well, we get the fundamental solution
1 2 1
Eðx1 ; x2 Þ ¼  r ln ðA:244Þ
8p r
for the equation
DDuðx1 ; x2 Þ ¼ f ðx1 ; x2 Þ: ðA:2440 Þ

A.3.4.4 Equations of Hyperbolic Type. Fundamental Solutions

We consider the waves equations


hi uðx1 ; x2 ; x3 ; tÞ ¼ fi ðx1 ; x2 ; x3 ; tÞ; i ¼ 1; 2; ðA:245Þ
where fi ðx1 ; x2 ; x3 ; tÞ are given distributions; a fundamental solution of these
equations is of the form

1 R
Ei ðx1 ; x2 ; x3 ; tÞ ¼  d t ; i ¼ 1; 2: ðA:2450 Þ
4pR ci

Analogically, for the equation


hi uðx1 ; x2 ; x3 ; tÞ þ 4p,ðtÞ  dðx1 ; x2 ; x3 Þ ¼ 0; i ¼ 1; 2; ðA:246Þ
where jðtÞ is a distribution, one obtains

1 R
uðx1 ; x2 ; x3 ; tÞ ¼ , t ; i ¼ 1; 2: ðA:2460 Þ
R ci
The solution of the equation
h1 h2 uðx1 ; x2 ; x3 ; tÞ þ 4p,ðtÞ  dðx1 ; x2 ; x3 Þ ¼ 0 ðA:247Þ
A.3 Elements of the Theory of Distributions 781

reads



c2 c2 R R
uðx1 ; x2 ; x3 ; tÞ ¼  2 1 2 2 t  t ,ðtÞ; ðA:2470 Þ
c1  c2 c1 þ c2 þ ðtÞ

where the convolution product concerns only the time variable.


The fundamental solution of the equation
h1 h2 uðx1 ; x2 ; x3 ; tÞ ¼ f ðx1 ; x2 ; x3 ; tÞ ðA:248Þ
is of the form



c2 c2 R R
Eðx1 ; x2 ; x3 ; tÞ ¼   21 2 2  t  t : ðA:2480 Þ
4p c1  c2 c1 þ c2 þ

In the case of only two space variables, we consider the equations


hi uðx1 ; x2 ; tÞ ¼ f ðx1 ; x2 ; tÞ; i ¼ 1; 2; ðA:249Þ
h1 h2 uðx1 ; x2 ; tÞ ¼ f ðx1 ; x2 ; tÞ; ðA:250Þ
where fi ðx1 ; x2 ; tÞ and f ðx1 ; x2 ; tÞ are given distributions. Introducing the
distribution defined by the function



r 1 r
f0 t; ¼ L K0 p
ci ci


1=2
r r2
¼h t t2  2 ; t [ 0; i ¼ 1; 2; ðA:251Þ
ci ci

where K0 is the modified Bessel function of order zero, we obtain the fundamental
solutions

1 r
Eðx1 ; x2 ; tÞ ¼  f0 t; ; i ¼ 1; 2; ðA:2490 Þ
2p ci

for the Eq. (A.249). As well, if we introduce the distribution defined by the
function



r 1 1 r
f2 t; ¼L K0 p
ci p2 ci

" sffiffiffiffiffiffiffiffiffiffiffiffiffiffi! sffiffiffiffiffiffiffiffiffiffiffiffiffiffi#
r r2 r2
¼h t t ln t þ t2  2  t2  2 ; t [ 0; i ¼ 1; 2;
ci ci ci
ðA:252Þ

then we get the fundamental solution


782 Appendix




c2 c2 r r
Eðx1 ; x2 ; tÞ ¼  21 2 2  f2 t;  f2 t; ; ðA:2500 Þ
2p c1  c2 c1 c2

corresponding to the Eq. (A.250).

A.3.4.5 Equations of Parabolic Type: Fundamental Solutions

Let be the caloric equation


\uðx1 ; x2 ; x3 ; tÞ ¼ f ðx1 ; x2 ; x3 ; tÞ; ðA:253Þ
where f ðx1 ; x2 ; x3 ; tÞ is a given distribution; a fundamental solution is of the form
1 2
Eðx1 ; x2 ; x3 ; tÞ ¼ pffiffiffiffiffiffiffi hðtÞeR =4at ; t 2 ð1; 1Þ; ðA:2530 Þ
8pat pat
or of the form
1 2
Eðx1 ; x2 ; x3 ; tÞ ¼ pffiffiffiffiffiffiffi eR =4at ; t 0: ðA:25300 Þ
8pat pat
In the case of two space variables, let be the equation
\uðx1 ; x2 ; tÞ ¼ f ðx1 ; x2 ; tÞ; ðA:254Þ
where f ðx1 ; x2 ; tÞ is a given distribution; a fundamental solution is of the form
1 2
Eðx1 ; x2 ; tÞ ¼ hðtÞer =4at ; t 2 ð1; 1Þ; ðA:2540 Þ
4pat
which can be expressed also in the form
1 r2 =4at
Eðx1 ; x2 ; tÞ ¼ e ; t 0: ðA:25400 Þ
4pat
Similarly, we can consider also other equations which occur in the study of the
mechanical phenomena.

A.4 Notations and Integrals

We give, in the following, some functions used in the study of the problems of the
elastic parallelepiped to simplify the representations of the results. As well, we
give final forms for some integrals which appear in the formulae thus obtained.
A.4 Notations and Integrals 783

A.4.1 Notations

We present the functions U; W and W0 which appear in the study of the problems
specified above and also put in evidence the differential relations which exist
between them.

A.4.1.1 The U Functions

Let be the functions

U1 ðhsÞ ¼ ð2 þ hsÞehs ; U2 ðhsÞ ¼ ð1 þ hsÞehs ; U3 ðhsÞ ¼ hsehs ; ðA:255Þ

U4 ðhsÞ ¼ ð1  hsÞehs ; U5 ðhsÞ ¼ ð2  hsÞehs ; U6 ðhsÞ ¼ 2ehs : ðA:2550 Þ


We remark that between these functions take place the differential relations
dU1 ðhsÞ dU2 ðhsÞ
¼ U2 ðhsÞ; ¼ U3 ðhsÞ;
dðhsÞ dðhsÞ
ðA:256Þ
dU3 ðhsÞ dU4 ðhsÞ
¼ U4 ðhsÞ; ¼ U5 ðhsÞ:
dðhsÞ dðhsÞ
We also introduce the notation
Upq ðhsÞ ¼ Up ðhsÞ  mUq ðhsÞ; p; q ¼ 1; 2; . . .; 6; ðA:257Þ

where 0
m
1=2 is Poisson’s ratio.

A.4.1.2 The W Functions

Let be the functions


hs
W1 ðhs; htÞ ¼ ½ð2 þ hs coth hsÞ sinh ht  ht cosh ht;
sinh hs
hs
W2 ðhs; htÞ ¼ ½ð1 þ hs coth hsÞ cosh ht  ht sinh ht;
sinh hs ðA:258Þ
hs
W3 ðhs; htÞ ¼ ðhs coth hs sinh ht  ht cosh htÞ;
sinh
hs
W4 ðhs; htÞ ¼ ½ð1  hs coth hsÞ cosh ht þ ht sinh ht;
sinh
hs hs
W5 ðhs; htÞ ¼ 2 cosh ht; W6 ðhs; htÞ ¼ 2 sinh ht: ðA:2580 Þ
sinh hs sinh hs
784 Appendix

The differential following relations take place


dW1 ðhs; htÞ dW2 ðhs; htÞ
¼ W2 ðhs; htÞ; ¼ W3 ðhs; htÞ;
dðhtÞ dðhtÞ
ðA:259Þ
dW3 ðhs; Þ
¼ W4 ðhs; htÞ;
dðhtÞ
dW5 ðhs; htÞ dW6 ðhs; htÞ
¼ W6 ðhs; htÞ; ¼ W5 ðhs; htÞ: ðA:2590 Þ
dðhtÞ dðhtÞ
One introduces the notations
Wpq ðhs; htÞ ¼ Wp ðhs; htÞ  mWq ðhs; htÞ; p; q ¼ 1; 2; . . .; 6; ðA:260Þ

too, m being Poisson’s ratio.


In particular, let be the function

hs
vðhsÞ ¼ W2 ðhs; hsÞ ¼ coth hs þ hs: ðA:261Þ
sinh2 hs

A.4.1.3 The W0 Functions

Analogically, let be the functions


hs
W01 ðhs; htÞ ¼ ½ð2 þ hs tanh hsÞ cosh ht  ht sinh ht;
cosh hs
hs
W02 ðhs; htÞ ¼ ½ð1 þ hs tanh hsÞ sinh ht  ht cosh ht;
cosh hs ðA:262Þ
hs
W03 ðhs; htÞ ¼  ðhs tanh hs cosh ht  ht sinh htÞ;
cosh hs
hs
W04 ðhs; htÞ ¼ ½ð1  hs tanh hsÞ sinh ht þ ht cosh ht;
cosh hs
hs hs
W05 ðhs; htÞ ¼ 2 sinh ht; W06 ðhs; htÞ ¼ 2 cosh : ðA:2620 Þ
cosh hs cosh hs
We remark the differential relations
dW01 ðhs; htÞ dW02 ðhs; htÞ
¼ W02 ðhs; htÞ; ¼ W03 ðhs; htÞ;
dðhtÞ dðhtÞ
ðA:263Þ
dW03 ðhs; htÞ
¼ W04 ðhs; htÞ;
dðhtÞ

dW05 ðhs; htÞ dW06 ðhs; htÞ


¼ W06 ðhs; htÞ; ¼ W05 ðhs; htÞ: ðA:2630 Þ
dðhtÞ dðhtÞ
A.4 Notations and Integrals 785

We also introduce the notations


W0pq ðhs; htÞ ¼ W0p ðhs; htÞ  mW0q ðhs; htÞ; p; q ¼ 1; 2; . . .; 6: ðA:264Þ

In particular, one introduces the function


hs
v0 ðhsÞ ¼ W02 ðhs; hsÞ ¼ tanh hs  hs: ðA:265Þ
cosh2 hs

A.4.2 Integrals
We shall give some results concerning integrals of rational functions and of
rational functions and radicals; as well, we give results for double Fourier
integrals.

A.4.2.1 Integrals of Rational Functions

Let be following integrals of rational functions, i.e.


Z1
dx p
¼ ; ðA:266Þ
p2 þx 2 2p
0

Z1
dx p
¼ ; ðA:2660 Þ
ðp2 þ x2 Þ2 4p3
0

Z1
x2 dx p
¼ ; ðA:267Þ
ð p2 þ x 2 Þ2 4p
0

Z1
xdx p
¼ ; ðA:2670 Þ
ðp2 þ x2 Þ2 2p2
0

Z1
dx p
¼ ; ðA:268Þ
ðp2 þ x2 Þðq2 þ x2 Þ 2pqðp þ qÞ
0

Z1
x2 dx p
¼ ; ðA:2680 Þ
ðp2 þ x2 Þðq2 þ x2 Þ 2pðp þ qÞ
0
786 Appendix

Z1
dx pð2p þ qÞ
¼ ; ðA:269Þ
ð p2 þ x2 Þ2 ðq2 þ x2 Þ 4p3 qðp þ qÞ2
0

Z1
x2 dx p
¼ ; ðA:2690 Þ
ð p2 þ x 2 Þ 2 ð q2 þ x2 Þ 4pðp þ qÞ2
0

Z1
x4 dx pðp þ 2qÞ
¼ ; ðA:26900 Þ
ð p2 þ x 2 Þ 2 ð q2 þ x2 Þ 2ðp þ qÞ2
0

Z1
x2 dx p
¼ ; ðA:270Þ
ðp2 þ x2 Þ2 ðq2 þ x2 Þ2 4pqðp þ qÞ3
0

Z1
x4 dx p
¼ : ðA:2700 Þ
ð p2 þ x2 Þ2 ðq2 þ x2 Þ2 4ðp þ qÞ3
0

A.4.2.2 Integrals of Rational Functions and Radicals

We give some integrals of rational functions and radicals, i.e.


Z1
dx 1
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ 2 ; ðA:271Þ
ð p2 þ x 2 Þ p2 þ x2 p
0

Z1
dx 2
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ 4 ; ðA:2710 Þ
ð p2 þ x2 Þ2 2
p þx 2 3p
0

Z1
dx 1 pþq
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ ln ; ðA:272Þ
ð p2 þ x2 Þ 2
q þx 2 2pq p  q
0

Z1
dx 1 1 pþq
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ 2 2 þ ln ðp 6¼ qÞ; ðA:2720 Þ
ð p2 þ x 2 Þ 2 q2 þ x2 2p ðp  q2 Þ 4p3 q p  q
0

Z1
x2 dx 1 1 pþq
2
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼  2  q2 Þ
þ ln ðp 6¼ qÞ; ðA:27200 Þ
2 2 2
ðp þ x Þ q þ x 2 2ðp 4pq pq
0
A.4 Notations and Integrals 787

Z1
dx
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ð p2 þ x 2 Þ ð q2 þ x 2 Þ r 2 þ x 2
0

1 1 pþr 1 qþr
¼ ln  ln ; p 6¼ r; q 6¼ r; ðA:273Þ
2rðp2  q2 Þ p p  r q q  r
Z1
dx 1
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼  2 2
ð p2 þ x 2 Þ 2 ð q2 þ x2 Þ 2
r þx 2 2p ðp  q 2 Þðp2  r 2 Þ
0
3p2  q2 pþr 1 qþr
 ln þ ln ; p 6¼ r; q 6¼ r; ðA:274Þ
4p3 ðp2  q2 Þ2 r p  r 2qrðp2  q2 Þ2 q  r
Z1
x2 dx 1
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼
ð p2 þ x2 Þ2 ðq2 þ x2 Þ r 2 þ x2 2ðp2  q2 Þðp2  r 2 Þ
0
p2 þ q2 pþr q qþr
þ ln  ln ; p 6¼ r; q 6¼ r; ðA:2740 Þ
4prðp2  q2 Þ 2 p  r 2rðp  q Þ
2 2 2 qr
Z1
x2 dx p2 þ q2  2r 2
p ffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼
ðp2 þ x2 Þ2 ðq2 þ x2 Þ2 r 2 þ x2 2ðp2  q2 Þðp2  r 2 Þðq2  r 2 Þ
0
2

1 3p þ q2 p þ r 3q2 þ p2 q þ r
 ln  ln ; p 6¼ r; q 6¼ r;
4rðp2  q2 Þ3 p pr q qr
ðA:275Þ
Z1
x4 dx p4  ðp2 þ q2 Þr 2 þ q4
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼
ð p2 þ x2 Þ2 ðq2 þ x2 Þ2 r 2 þ x2 2ðp2  q2 Þ2 ðp2  r 2 Þðq2  r 2 Þ
0

1 pþr
þ 3
pðp2 þ 3q2 Þ ln
2 2
4rðp  q Þ pr

qþr
qðq2 þ 3p2 Þ ln ; p 6¼ r; q 6¼ r; ðA:2750 Þ
qr
Z1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
dx 1 p2 þ q2 þ q
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ln pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ; p 6¼ 0; ðA:276Þ
2 2 2 2
ðp þ q þ x Þ p þ x 2 2
2q p þ q 2 p2 þ q2  q
0

Z1
x2 dx 1
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ 2 ; ðA:277Þ
ð p2 þ x 2 Þ ð q2 þ x 2 Þ q2 þ x 2 p  q2
0
788 Appendix

Z1
x2 dx 1
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ ; ðA:2770 Þ
ð p2 þ x2 Þ2 ðq2 þ x2 Þ 2
q þx 2 ð p  q2 Þ 2
2
0

A.4.2.3 Double Fourier Integrals

Using the notation


qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
c¼ a 2 þ b2 ; ðA:278Þ

as well as
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r ¼ x1 þ x2 þ x3 ; r3 ¼ x21 þ x22
2 2 2 ðA:279Þ

for the radii, one obtains, for some double Fourier integrals, the formulae
Z1 Z1
2 x3
ecx3 cos ax1 cos bx2 dadb ¼ ; ðA:280Þ
p r3
0 0

Z1 Z1 x 2
2 1 3
cecx3 cos ax1 cos bx2 dadb ¼  1  3 ; ðA:281Þ
p r3 r
0 0

Z1 Z1
2 1
c cos ax1 cos bx2 dadb ¼  ; ðA:2810 Þ
p r33
0 0

Z1 Z1
2 1 cx3 1
e cos ax1 cos bx2 dadb ¼ ; ðA:282Þ
p c r
0 0

Z1 Z1
2 1 1
cos ax1 cos bx2 dadb ¼ ; ðA:2820 Þ
p c r3
0 0

Z1 Z1
2 x3 x1
aecx3 sin ax1 cos bx2 dadb ¼ 3 ; ðA:283Þ
p r5
0 0

Z1 Z1
2 x2 x3
becx3 cos ax1 sin bx2 dadb ¼ 3 ; ðA:2830 Þ
p r5
0 0
A.4 Notations and Integrals 789

Z1 Z1
2 a cx3 x1
e sin ax1 cos bx2 dadb ¼ 3 ; ðA:284Þ
p c r
0 0

Z1 Z1
2 b cx3 x2
e cos ax1 sin bx2 dadb ¼ 3 ; ðA:2840 Þ
p c r
0 0

Z1 Z1
2 a x1
sin ax1 cos bx2 dadb ¼ 3 ; ðA:285Þ
p c r3
0 0

Z1 Z1
2 b x2
cos ax1 sin bx2 dadb ¼ 3 ; ðA:2850 Þ
p c r3
0 0

Z1 Z1 x 2
2 cx3 x1 3
ace sin ax1 cos bx2 dadb ¼ 3 5 1  5 ; ðA:286Þ
p r r
0 0

Z1 Z1 x 2
2 cx3 x2 3
bce cos ax1 sin bx2 dadb ¼ 3 5 1  5 ; ðA:2860 Þ
p r r
0 0

Z1 Z1
2 x1
ac sin ax1 cos bx2 dadb ¼ 3 ; ðA:287Þ
p rx53
0 0

Z1 Z1
2 x2
bc cos ax1 sin bx2 dadb ¼ 3 ; ðA:2870 Þ
p rx53
0 0

Z1 Z1 x 2
2 2 cx3 x3 1
ae cos ax1 cos bx2 dadb ¼ 3 5 1  5 ; ðA:288Þ
p r r
0 0

Z1 Z1 x 2
2 x3 2
b2 ecx3 cos ax1 cos bx2 dadb ¼ 3 1  5 ; ðA:2880 Þ
p r5 r
0 0

Z1 Z1
2 x1 x2 x3
abecx3 sin ax1 sin bx2 dadb ¼ 15 ; ðA:28800 Þ
p r7
0 0
790 Appendix

Z1 Z1
2 a cx3 x1
e sin ax1 cos bx2 dadb ¼ ; ðA:289Þ
p c2 rðr þ x3 Þ
0 0

Z1 Z1
2 b cx3 x2
e cos ax1 sin bx2 dadb ¼ ; ðA:2890 Þ
p c2 rðr þ x3 Þ
0 0

Z1 Z1
2 a x1
2
sin ax1 cos bx2 dadb ¼ 2 ; ðA:290Þ
p c rx 3
0 0

Z1 Z1
2 b x2
cos ax1 sin bx2 dadb ¼ 2 ; ðA:2900 Þ
p c2 rx 3
0 0

Z1 Z1
2
2 a ðr 2  x21 Þðr þ x3 Þ  rx21
ecx3 cos ax1 cos bx2 dadb ¼ ; ðA:291Þ
p c r 3 ðr þ x3 Þ2
0 0

Z1 Z1
2
2 b ðr 2  x22 Þðr þ x3 Þ  rx22
ecx3 cos ax1 cos bx2 dadb ¼ ; ðA:2910 Þ
p c r 3 ðr þ x3 Þ2
0 0

Z1 Z1
2 ab cx3 2ðr þ x3 Þx1 x2
e sin ax1 sin bx2 dadb ¼ ; ðA:292Þ
p c2 r 3 ðr þ x3 Þ2
0 0

Z1 Z1
2 ab cx3 3x1 x2
e sin ax1 sin bx2 dadb ¼ 5 ; ðA:2920 Þ
p c r
0 0

Z1 Z1
2
2 a x2  x2
cos ax1 cos bx2 dadb ¼  1 4 2 ; ðA:293Þ
p c r3
0 0

Z1 Z1
2
2 b x21  x22
cos ax1 cos bx2 dadb ¼ ; ðA:2930 Þ
p c r34
0 0

Z1 Z1
2 ab 2x1 x2
sin ax1 sin bx2 dadb ¼ 4 ; ðA:294Þ
p c2 r3
0 0
A.4 Notations and Integrals 791

Z1 Z1
2 ab 3x1 x2
sin ax1 sin bx2 dadb ¼ 5 ; ðA:2940 Þ
p c r3
0 0

Z1 Z1 x 2
2 a2 cx3 1 1
e cos ax1 cos bx2 dadb ¼ 3 1  3 ; ðA:295Þ
p c r r
0 0

Z1 Z1 x 2
2 b2 cx3 1 2
e cos ax1 cos bx2 dadb ¼ 3 1  3 ; ðA:2950 Þ
p c r r
0 0

Z1 Z1
2 a2 x2  x2
cos ax1 cos bx2 dadb ¼  1 5 2 ; ðA:296Þ
p c r3
0 0

Z1 Z1
2 b2 x2  x2
cos ax1 cos bx2 dadb ¼ 1 5 2 ; ðA:2960 Þ
p c r3
0 0

Z1 Z1 x 2
2 2 cx3 x3 3
c e cos ax1 cos bx2 dadb ¼ 3 5 3  5 : ðA:297Þ
p r r
0 0

By means of these integrals, one also obtains:


Z1 Z1
2 x33
ð1 þ cx3 Þecx3 cos ax1 cos bx2 dadb ¼ 3 ; ðA:298Þ
p r5
0 0

Z1 Z1 x 2
2 x3 3
ð1  cx3 Þecx3 cos ax1 cos bx2 dadb ¼ 2  3 ; ðA:2980 Þ
p r3 r
0 0

Z1 Z1
2
cð1 þ cx3 Þecx3 cos ax1 cos bx2 dadb
p
0 0
x 2 x 4
1 3 3
¼ 3 1þ6 15 ; ðA:299Þ
r r r
Z1 Z1
2
cð1  cx3 Þecx3 cos ax1 cos bx2 dadb
p
0 0
x 2 x 4
1 3 3
¼ 1  12 þ15 ; ðA:2990 Þ
r3 r r
792 Appendix

Z1 Z1
2
2 a
ð1  cx3 Þecx3 cos ax1 cos bx2 dadb
p c
0 0
" #
1 x 2 ðr 2  x2 Þðr þ x Þ  rx2
1 2 3 2
¼  3 3x3  ; ðA:300Þ
r r ðr þ x3 Þ2
Z1 Z1
2
2 b
ð1  cx3 Þecx3 cos ax1 cos bx2 dadb
p c
0 0
" #
1 x 2 ðr 2  x2 Þðr þ x Þ  rx2
2 3
¼  3 3x3  1 1
; ðA:3000 Þ
r r ðr þ x3 Þ2
Z1 Z1
2 a2
ð1  cx3 Þecx3 cos ax1 cos bx2 dadb
p c
0 0
x 2
1 2 x23 x21
¼ 3 23 15 4 ; ðA:301Þ
r r r
Z1 Z1
2 b2
ð1  cx3 Þecx3 cos ax1 cos bx2 dadb
p c
0 0
x 2
1 1 x2 x2
¼ 3 23 15 2 4 3 : ðA:3010 Þ
r r r
Subject Index

A C
Anisotropy, 138, 624 Computation methods, 266
axis of symmetry, 627 direct, 280
dynamics problems, 634 fundamental solutions, 288
general case, 138 Fourier representations, 272
elastic constants, 619 in elastodynamics, 295
orthotropy, 142, 622 in elastostaties, 292
plane of symmetry, 141, 625 particular integrals, 266
transverse isotropy, 619 polynomials, 269
volume forces, 632 wave equations, 276
Antiplane problems, 604 point matching, 302
state of strain, 608 successive approximations, 301
state of stress, 567 variational, 281
Bubnov-Galerkin, 287
least squares, 284
B Ritz, 286
Basic problems, 192 Concentrated loads, 357
first, 192 Cosserat type solution, 383
limit conditions, 192 centre of dilatation, 386
boundary, 192 directed, 385
initial, 197 force, 384
mixed, 194 moments, 385
second, 193 rotational, 386
Body, 8 elastic solutions, 377
bar, 8 dynamic, 380
block, 10 static, 376
connection, 262 force, 358
multiply, 263 system of, 360
simple, 262 moment, 361
distorsion, 264 dipole type, 367
incompressible, 150 directed, 361
plate, 9 linear, 367

P. P. Teodorescu, Treatise on Classical Elasticity, Mathematical and Analytical 793


Techniques with Applications to Engineering, DOI: 10.1007/978-94-007-2616-1,
Ó Springer Science+Business Media Dordrecht 2013
794 Subject Index

C (cont.) incompressible state, 150


plane, 368 infinitesimal, 43
spatial, 370 deviator, 49
rotational, 363 spherical, 49
tensor properties, 372 kinematics, 68
Co-ordinates, 35 rigid body motion, 54
material, 36 local rotation, 54
space, 37 rotation, 57
Cosserat type bodies, 307 thermodynamics of, 126
Cesàro type formulae, 319 of form, 149
continuity, 316 of volume, 149
dynamic case, 336 principles, 127
formulations, 336
representations, 337
equations, 328 E
of equilibrium, 328 Elastic domains, 393
of motion, 328 eights-space, 427
kinematics, 321 normal, local load, 430
rotations, 312 particular cases, 455
constrained, 312, 320, 331, 335 periodic normal load, 428
free, 312 tangential, local load, 446
states, 322 half-space, 399
of couple-stress, 323 concentrated moments, 422
of stress, 322 fundamental solution for displacements
tensors, 325 and stresses, 412, 418
static case, 333 local load, Boussinesq
formulations, 333 problem, 404, 409
representations, 333 periodical load, 400
theorems, 345 layer, 496
reciprocity, 346 local load, 501
uniqueness, 346 periodic load, 496
Crystallography, 638 parallelepiped, 482
classes of crystals, 648 quarter-space, 462
fundamental theory, 643 local normal load, 462
general properties, 639 local tangential load, 467
geometrical properties, 639 particular cases, 473
operations, 644 space, 395
symmetry, 640 concentrated loads, 397
Curvilinear co-ordinates, 753 solutions in displacements
cylindrical, 756, 762 and stresses, 395
differential operators, 757 solutions, volume forces, 395
orthogonal, 753 Elastodynamic formulations, 219
spherical, 756, 762 in displacements, 219
Lamé’s equations, 219
in stresses, 227
D Beltrami–Michell equations, 227
Deformation, 33 Elastodynamic problems, 512
Cèsaro’s formulae, 59 axisymmetrical, 512
continuity, 59 potential functions, 521
displacement, 33, 56 forced vibrations, 531
gradient, 44 elastic half-space, 531
elementary states, 60 free vibrations, 528
finite, 33 elastic half-space, 540
continuity of mass, 101 Rayleigh waves, 540
Subject Index 795

progressive waves, 522, 524 M


acoustic tensor, 523 Mechanics of deformable solids, 2
Elastodynamic representations, 222 Models, 3
in displacements, 221 elastic, 116
Lamé-Clebsch, 226 Castigliano’s formulae,
Somigliana–Iacovache, 222 potential, 132
Sternberg–Eubanks, 224 characteristic curve, 119
in stresses, 233 constants, Hooke’s law, 117, 138, 143,
Beltrami–Finzi type, 233 145
Maxwell type, 235 constitutive law, non-linear, 115, 116
Morera type, 235 curvilinear co-ordinates,
Teodorescu, 236, 238 Hooke’s law, 152
Elastostatic formulation, 199 temperature, Hooke’s law, 153
in displacements, 199 Gibbs’s conditions, potential, 137
Lamé’s equations, 199 Green’s formulae, potential, 132
in stresses, 205
Beltrami–Michell equations, 205
Elastostatic representations, 199 N
in displacements, 202 Non-homogeneity, 151, 652
Galerkins, 202 approximate methods, 654
Papkovich–Neuber, 204 elastic half-plane, 665
in stresses, 208 particular integrals, 662
Beltrami-Finzi, 208 problems, 652
Maxwell, 209 three-dimensional, 652
Morera, 210 two-dimensional, 656
Schaefer, 211 Notations, 782
Teodorescu, 216

P
F Particular states, 547
Force, 73 of strain, 592
lost, 78 a zero linear strain, 592
superficial, 73 two zero angular strains, 602
volume, 76 of stress, 548
a zero normal stress, 548
two zero tangential
H stresses, 577
History, 15 straight cylinder, 587
deformable solids, 14 Plane problems, 582
elastic solids, 18 state of strain, 597
dynamic, 605
static, 604
I state of stress, 573, 582, 604
Inelastic bodies, 155 Principles, 248
hardening material, 162 Clapeyron, 247
plastic models, 156 d’Alembert-Lagrange, 249
theories, 158 Hamilton, 251
viscoelastic models, 184 minimum internal work, 247
Burgers, 182 minimum potential energy, 252
creep, 170 reciprocity, 254
generalized; Kelvin, Maxwell, 177, 178 Betti, 254
grouping; Kelvin–Voigt, Maxwell, Maxwell, 255
relaxation, 171, 174, 175, 180 Saint–Venant, 260
Integrals, 785 virtual variations, 251
796 Subject Index

S classification, 764
Strain, 39 composition, 769
angular, 40 differential equations, 775
curvilinear co-ordinates, 65 elliptic, 779
cylindrical, 66 hyperbolic, 780
spherical, 67 ordinary, 775
initial, 155 parabolic, 782
linear, 39 partial, 777
octahedral, 51 differentiation, 768
principal, 48 integral transforms, 771
tensor, 45 Fourier, 772
volume, 42, 56 Laplace, 774
Strength of materials, 2 product, 769
Stress, 73 convolution, 770
Cauchy’s theorem, 79 direct, 769
curvilinear co-ordinates, 108 Thermoelasticity, 671
cylindrical, 110 applications, 685
spherical, 111 axial symmetry, 688
deviator, 92 elementary, 685
elementary states, 98 plane problems, 693
ellipsoid, 90 coupled problems, 684
equilibrium, 95 general equations, 672
finite, 108 in displacements, 674, 678
initial, 154 in stresses, 677, 679
Mohr’s circles, 90 heat conduction, 671
motion, 97 Maysel’s formulae, 683
octahedral, 92 reciprocity principle, 682
principal, 85 Thick plate, 505
symmetry, 82
tensor, 84
Cauchy, 105 V
Piola–Kirchhoff, 103 Viscoelasticity, 699
variation around a point, 78 applications, 720
vector, 73 complex modulus, 707
normal, 75 one-dimensional case, 710
tangential, 75 three-dimensional case, 712
constitutive laws, 700
creep, 700
T formulation, 715
Tensor calculus, 729 in displacements, 715
operators, 735 principle of correspondence, 717
algebraic, 735 quasi-static problems, 719
analytic, 740 relaxation, 700
integral formulae, 748
orthogonal affine, 729
differential operators of first W
and second order, 741, 746 Work, 124, 243
Theorems, 257 elementary, 124
existence, 257 external, 244
uniqueness, 257 formulae of Castigliano, 247
Kirchhoff, 257 formulae of Green, 246
Neumann, 265 of deformation, 245
Theory of distributions, 763
Author Index

A Berekhovskikh, L. M., 531


Achenbach, J. D., 311 Bernoulli, Daniel, 16
Adkins, J. E., 30 Bernoulli, Jacques, 16
Adomeit, G., 309 Bernoulli, Jacob, 3, 16, 28
Aero, E. L., 309, 311 Bernoulli, Jean, 16
Airy, George-Bidell, 5, 18, 19, Berry, D. S., 394
195, 575 Betti, Enrico, 19, 254, 295
Alain, M., 698 Bezukhov, N. I., 29
Alblas, J. B., 310 Bishop, R. E. D., 634
Alembert, Jean le Rond d’, 16, 78, Bleich, H. H., 531
102, 249 Blokh, V. I., 195
Alfrey, T., 717 Bodaszewski, S., 308
Almansi, Emilio, 38, 506, 547, 587, 751 Boggio, Tomaso, 199
André, P. A., 309 Boley, Bruno-A., 671
Anthony, K., 311 Boltzmann, James Joseph, 123, 186,
Appell, Paul-Émile, 23 308, 699
Archimedes, 15 Bondarenko, Boris Anisimovich, 198
Archytas of Tarentum, 15 Bonneau, L., 506
Borchardt, C. W., 64
Boussinesq, Joseph Valentin, 20, 399, 421
B Bragg, W. H., 639
Baı̆da Ē. N., 482 Bragg, W. L., 639
Barański, W., 310 Bravais, 640
Basheleı̆shvili, M. C., 257 Brdička, 29
Beju, Iulian, 531 Bressan, A., 310
Bekhterev, P., 618, 619 Bresse, Jacques-Antoine-Charles, 21
Belesß, Aurel A., 23 Brewster, David, 18
Belluzi, Odone, 30 Broberg, K. B., 531
Beltrami, Eugenio, 19, 22, 195, 198, 199, 207, Bubnov, Ivan Grigorievich, 21, 287
208, 238, 679, 680 Bulugin, A. N., 309
Belyaev, Nikolai Mikhailovich, 21 Burchuladzhe, 257

P. P. Teodorescu, Treatise on Classical Elasticity, Mathematical and Analytical 797


Techniques with Applications to Engineering, DOI: 10.1007/978-94-007-2616-1,
Ó Springer Science+Business Media Dordrecht 2013
798 Author Index

Burger, 182 E
Burzynski, W., 308 Eason, G., 529
Butty, H., 29 Engesser, Friedrich, 19
Ericksen, J. L., 309
Eringen, Cemal A., 309, 310
C Essmann, U., 311
Cagniard, L., 531 Eubanks, R. A., 195, 198, 224, 225, 344, 345
Carathéodory, 130 Euler, Leonhard, 16, 18, 35, 38, 101, 102, 124,
Carlslaw, H. C., 671 283, 284, 646
Carlslaw, M. S., 188, 671, 698 Ewing, W. M., 529
Carlson, D. E., 348
Castigliano, Alberto, 19, 132, 135, 246,
247, 287 F
Cauchy, Augustin-Louis, 17, 38, 46, 51, 79, Fedorov, E. S., 525, 637
82, 104, 120, 198, 308 Filipescu, Gheorghe Emanoil, 23
Cerruti, V., 399 Filon, Louis-Napoleon George, 22
Cesàro, Ernesto, 59, 64 Filonenko-Borodich, Mikhail Mitrofanovich,
Chadwick, P. D., 198 22, 479
Chao, C. C., 529 Finzi, Bruno, 195, 205, 208, 209, 233, 235
Chentsov, N. G., 18, 616 Flamant, Alfred Aimé, 20, 21
Chladni, 17 Flint, 647
Clapeyron, Benoît-Paul Émile, 17, 197, Flitman, L. M., 531
248, 250 Flügge, S., 30
Clausis, 129 Fontana, 16
Clebsch, R. F. A., 226, 676 Fourier, François-Marie-Charles, 18, 269, 272,
Clintock, F. A. Mc, 309 278, 672
Cosserat, Eugène-Maurice-Pierre, 22, 25, 310, Föppl, August, 19, 481
346 Föppl, Ludwig, 19, 20, 481
Cosserat, François, 22, 25, 310, 348 Franciosi, V., 30
Coulomb, Charles-Augustin de, 16 Fredholm, Erik Ivar, 18
Cove, Y. Le, 310 Fresnel, 526
Crăciun, Marius E., 613 Fridman, M. M., 633
Cristea, M., 618 Friedrichs, K. O., 259
Cristescu, Dan Nicolae, 613 Fung, V. C., 30

D G
Davidescu Moisil, Ana, 545 Gadolin, Aksel Wilhelmovich, 20, 639
Davidoglu, Anton, 23 Galerkin, Boris Grigorievich, 21, 194, 198,
Davies, R. M., 529 202, 203, 287, 333
Derevianko, N. I., 311 Galileo, Galilei, 15
Dhaliwal, R. S., 390 Galletto, Dionigi, 312
Dinnik, Aleksandr Nicolaevich, 21 Gauss, Karl Friedrich, 17, 64
Dirac, Paul Adrien Maurice, 771 Gazis, D. C., 311
Dixon, R. C., 310 Gegelya, T. G., 257
Djurić, S., 309, 311 Germain, Paul, 30
Dougall, 503 Germain, Sophie, 17
Drăgan, Mircea, 24 Gibbs, Josiah Dixon Willard, 137
Drucker, 160 Golitsyn, Boris Borisovich, 21
Duhamel, Jean-Marie-Constant, 18 Golovin, Harlampii Sergeevich, 20
Duhem, Paul, 197 Goodier, J. N., 675, 685
Author Index 799

Gordon, V. A., 310 Iesßan, Dorin, 310


Gorskiı̆, B. V., 310 Ignaczak, J., 228
Graff, K. F., 311 Ille, Vasile, 390, 612
Graffi, Dario, 255, 347 Ilyushin, A. A., 167, 717
Grammel, R., 29 Ionescu-Cazimir, Viorica, 311
Grashof, Franz, 16 Ionescu, Dan, 198
Green, A. E., 310, 627 Ionescu, Ion, 23
Green, George, 17–19, 37, 122, 124, 132, 134,
198, 246
Grioli, Gaetano, 309, 310 J
Grodski, G. D., 21, 194 Jaeger, J. C., 671
Gurtin, Morton E., 195, 198, 248, 257, Jaramillo, T. J., 308
259, 700 Jardetzky, W. S., 531
Guz’, A. N., 353 Jaunzemis, W., 240
Günther, W., 309, 311 Joel, N., 308
Gwither, R. F., 113 John, Fritz, 259
Joule, 127
Jukovski, M., 310
H Juravskiı̆, Dimitrii Ivanovich, 21
Hadamard, Jacques-Solomon, 22, 526
Haimovici, Mendel, 506
Hamburger, Leon, 671 K
Hamel, G., 38, 308 Kaliski, S., 309
Hamilton, William Rowan, 251, 680 Kárman, Theodor von, 22
Hangan, Mihail, 23 Kecs, Wilhelm, 30, 240, 304, 390, 391, 424,
Hartranft, R. J., 350 425, 545, 728
Haüy, R. J., 638 Kelly, P. D., 310, 350, 377
Hayes, M., 531 Kepler, Johannes, 638
Hearmon, R. F. S., 668 Kessel, Sigfried, 376, 311
Heaviside, 766 Khan, S. M., 390
Hehl, F., 309 Kirchhoff, Gustav Robert, 3, 18, 28, 30, 103,
Hellinger, E., 308 105, 257, 305
Helmholtz, Hermann Ludwig Ferdinand von, Kirichenko, A. M., 311
20, 131 Kirpichev, Viktor Lvovich, 21
Hencky, Heinrich, 22, 120, 167 Klein, Felix, 22, 308
Hermann, G., 311 Knopoff, L., 531
Hermite, R. L., 350 Knops, R. J., 260
Hertz, Heinrich Rudolf, 20, 399 Koiter, W. T., 310
Hessel, 639 Kolodner, 527
Heun, K., 308 Kolosov, G. V., 22
Hijab, W. A., 427 Kolski, H., 357, 727, 643
Hilbert, David, 22 Kondo, Kazuo, 26, 311
Hodge, P. G., Jr, 30 Kononenko, E. S., 481
Hoffman, O., 311 Korn, Arthur, 18
Hooke, Robert, 16, 115, 117, 138, 638 Kowalewski, Sonya, 198
Hoppman, W. H., 310 Krishnan, R. S., 308
Howink, R., 30 Kronecker, Leopold, 733
Huber, Maksymilian Tytus, 22, 23 Kr}
oner, Eckehard, 195, 309
Huygens, Christian, 638 Krylov, Aleksei Nicolaevich, 21
Kubenko, V. D., 309
Kubilin, Ivan Petrovich, 21
I Kunin, I. A., 310
Iacovache, Maria, 198, 222, 223, 519, Kupffer, Adolf Iakovlevich, 17
520, 620 Kupradzhe, V. D., 257
800 Author Index

Kuvshinskiı̆, E. V., 309, 311 Mishonov, M., 482


Misßicu, Mircea, 309, 506
Mohr, Christian Otto, 16, 19, 90, 91
L Moisil, Ana, 620
Lagrange, Joseph-Louis, 16, 17, 249, 286, 680 Moisil, Grigore C, 24, 191, 194, 198
Lai, Pham The, 399 Morera, Giacinto, 19, 195, 198, 205, 210, 212
Lamb, Horace, 20, 399, 531 Muki, R., 311
Lamé, Gabriel, 17, 64, 71, 90, 144, 197, 199, Muskhelishvili, Nikolai Ivanovich, 22
219, 226, 336, 342, 344, 481, 678 Müller, W., 31
Landau, Lev Davidovich, 23 Myller, Alexandru, 24
Lang, H. A., 531, 623
Langhaar, H., 242
Laplace, Pierre Simon de, 21, 774 N
Larmor, J., 108 Nadai, A., 4, 167
Laue, Max von, 639 Naghdi, P. M., 309, 310
Lauricella, Giuseppe, 22 Nagy, B. Sz, 771
Laval, J., 310 Navier, Louis-Marie-Henri, 17, 18, 21, 197
Lean, L. Mc, 32 Nemish, Iu N., 352
Lee, E. H., 717 Nernst, 130
Leı̆benzon, Leonid Samuilovich, 25 Netrebko, V. P., 482
Lejeune-Dirichlet, Peter Gustav, 16 Neuber, Heinz, 21, 194, 195, 204, 311, 334
Lekhnitski, S. G., 616, 617, 619, 633 Neumann, Franz Ernst, 17, 18, 153
Lévy, Maurice, 16, 20, 165 Newton, Isaac sir, 7
Lewy, H., 259 Nicolescu, Miron, 673
Lifshitz, Evgenii M, 23 Noether, Emmy, 308
Lifson, H., 531 Noll, Walter, 309
Lomakin, V. A., 309 Novozhilov, V. V., 31
Lomonosov, Mikhail Vasilievich, 16, 17 Nowacki, Witold, 310, 311, 345, 671
Love, August-Eduard-Hough, 20, 28, 300, 399,
531, 547
Lurje (Lur’e), A. I., 401, 497, 504, 505 O
Oravas, G. Æ., 32
Orlando. L., 505
M Ornstein, W., 241
Mandel, J., 399, 717 Oshima, N., 308
Marcolongo, Roberto, 22 Ostrogradskiı̆, Mikhail Vasilievich, 17,
Mariotte, Edmé, 16 284, 749
Marguerre, Karl, 195, 625, 627
Mateescu, Cristea, 23
Maxwell, James Clerk, 13, 16, 19, 20, 175, P
179, 184, 195, 198, 205, 209, 210, Pal’mov, V. A., 309
212, 255, 708 Papkovich, Piotr Fedorovich, 21, 194, 195,
Maysel, 683 204, 334
Melan, E., 671 Parkus, Heinz, 671
Menabrea, Ludovico Frederico, 19 Payne, L. E., 260
Meshkov, A. I., 481 Pearson, C. E., 31
Mesnager, Augustin-Charles-Marie, 21 Pearson, K., 20
Michell, John Henri, 19, 20, 206, 209, 238, Pekeris, C. C., 531
547, 680 Peretti, G., 195
Mikhlin, S. G., 633 Picard Émile, 23
Miklowitz, J., 531 Pietras, F., 311
Miller, W. H., 638 Pinney, E., 531
Mindlin, R. D., 310, 311, 383, 399 Piola, 101, 104, 106
Mises, Richard von, 22, 159, 165 Planck, Max, 31
Author Index 801

Plochocki, Z., 310 Seeger, A., 311


Podil’chuk, Ju. N., 311 Sherwood, J. W. C., 531
Poincaré, Henri,, 22 Shul’ga, N. A., 309, 311
Poisson, Siméon Denis, 17, 18, 147, 197, 284, Sneddon, Ian, N, 394, 505
308, 330, 779 Sokolnikoff, I. S., 623
Poncelet, Jean-Victor, 16 Sokolovski, Marek, 311
Prager, W., 163 Somigliana, Carlo, 19, 197, 198, 222, 223,
Prandtl, Ludwig, 4, 22, 165 292, 519, 520
Predeleanu, Mircea, 198, 301, 517, 519, Sommerfeld, Arnold, 23
521, 522 Soós, Eugen, 198, 309, 531, 615
Press, F., 531 Southwell, R. V., 31
Profiri, Nicolae, 23 Stefaniak, J., 353, 391
Stensen, N., 638
Sternberg, Eli, 194, 197, 198, 224, 225, 259,
R 300, 311, 700
Rabinovich, A. L., 259, 616, 617 Stevenson, A. C., 308
Rabotnov, Yu. N., 700, 717 Stippes, M., 527, 741
Radu, Elie., 23 Stoeckly, R. E., 308
Rajagopal, E. S., 308 Stoianović, Rastko, 309, 311
Rankine, William-John Macquorn, 16 Stokes, George Gabriel, 18, 21, 197, 200, 748
Read, W. T., 717 Strutt, John William (lord Rayleigh), 19
Reissner, Eric, 308, 309, 623 Sudria, J., 71
Reissner, H., 616 Suhubi, E. S., 309
Reuss, 165 Sylvester, 140
Ribière, Charles Henri, 22 S
ßandru, Nicolae, 310, 333, 337,
Rieci, Gregorio, 734 340, 383
Rieder, Georg, 196
Riesz, F., 702
Ritz, Walter, 20, 286, 287 T
Rivlin, R. S., 310, 531 Tait, Peter-Guthrie, 19
Rymarz, G., 311 Tanimoto, Benosuke, 399
Rogula, D., 310 Tedone, O., 197, 300, 671
Röntgen, W. K., 639 Teodorescu, Constantin, C., 23
Teodorescu, Nicolae, 24
Teodorescu, Petre, P., 198, 205, 216, 232, 235,
S 238, 301, 309, 310, 531
Sackmann, J., 531 Teodosiu, Cristian, 309
Saint-Venant, Adhémar Jean Claude Barré de, Terezawa, K., 399
4, 16, 18–20, 37, 156, 165, 260 Thomson, William (lord Kelvin), 16, 18, 64,
Saligny, Anghel, 23 173, 177–179, 184, 197, 292,
Sanielevici, Simion, 24 377, 708
Santer, F., 531 Tiersten, H. F., 383
Savin, G. N., 310, 311, 633 Tiffen, R., 308
Savov, L. N., 309 Timoshenko, Stepan, Prokofievich, 22
Schade, K. D., 310 Timpe, Aloys, 22
Schaefer, Hermann, 195, 198, 199, 211, 211, Todhunter, Isaac, 20
212, 309, 310 Toien, H. S., 717
Shahman, F. O. F., 309 Tolokonikov, L. A., 310
Schechter, M., 304 Toupin, R. A., 260, 309, 310
Schleicher, F., 399 Träuble, H., 311
Schwarz, 740 Trefftz, Erich, Immannuel, 18, 287
Schwedt, K. R., 309 Tresca, Henri-Édouard, 16, 160
Sedov, Leonid, I., 240 Trowbridge, E. A., 198
Seebeck, Ludwig, Friederich, 18 Truesdell, Clifford Ambrose, 13, 309, 310
802 Author Index

Tsien, H. S., 717, 728 Westergaard, Harald-Malcom, 22


Wheeler, L. T., 255, 259, 300
Wieghardt, K., 5
V Wilmański, K., 310
Vinci, Leonardo da, 15 Winkler, Emil, 19
Vinokurov, L. A., 311 Woinowski-Krieger, S., 506
Vitruvius, 15 Wooster, W. A., 308
Vlasov, Vasilii Zakharovich, 22 Wo_zniak, Cz, 310
Vodička, V., 394 Wulff, G. V., 639
Voigt, Woldemar, 16, 18, 20, 173, 708 Wyrwiński, J., 311
Volterra, Enrico, 22, 265
Volterra, Vito, 22, 265, 717
Vujośević, L., 309, 311 Y
Yamamoto, Y., 309
Yih-Hsing, Pao, 311
W Young, Thomas, 18
Wainright, W. L., 310
Wallis, R. F., 311
Wan, F. Y. M., 309 Z
Wang, C. C., 32, 71 Zaremba, S., 259
Weng, Chi Teh, 32 Zener, 180
Weber, C., 242 Zerna, W., 30
Weiner, J. H., 671 Ziegler, Fr., 163
Weitsman, Y., 311 _
Zukowski, M., 355
Wertheim, Wilhelm, 17
Weselowski, Zdenek, 310

You might also like