Sie sind auf Seite 1von 0

Draft

DRAFT

Lecture Notes
Introduction to

CONTINUUM MECHANICS
and Elements of

Elasticity/Structural Mechanics

c VICTOR

E. SAOUMA

Dept. of Civil Environmental and Architectural Engineering


University of Colorado, Boulder, CO 80309-0428

Draft
02

Victor Saouma

Introduction to Continuum Mechanics

Draft

03

PREFACE

Une des questions fondamentales que lingnieur des Matriaux se pose est de conna le comportee
e
itre
ment dun materiel sous leet de contraintes et la cause de sa rupture. En dnitive, cest prcisment la
e
e e
rponse a c/mat es deux questions qui vont guider le dveloppement de nouveaux matriaux, et dterminer
e
`
e
e
e
leur survie sous direntes conditions physiques et environnementales.
e
Lingnieur en Matriaux devra donc possder une connaissance fondamentale de la Mcanique sur le
e
e
e
e
plan qualitatif, et tre capable deectuer des simulations numriques (le plus souvent avec les Elments
e
e
e
Finis) et den extraire les rsultats quantitatifs pour un probl`me bien pos.
e
e
e
Selon lhumble opinion de lauteur, ces nobles buts sont idalement atteints en trois tapes. Pour
e
e
commencer, ll`ve devra tre confront aux principes de base de la Mcanique des Milieux Continus.
ee
e
e
e
Une prsentation dtaille des contraintes, dformations, et principes fondamentaux est essentiel. Par
e
e
e
e
la suite une briefe introduction a lElasticit (ainsi qu` la thorie des poutres) convaincra ll`ve quun
`
e
a
e
ee
probl`me gnral bien pos peut avoir une solution analytique. Par contre, ceci nest vrai (` quelques
e
e e
e
a
exceptions prts) que pour des cas avec de nombreuses hypoth`ses qui simplient le probl`me (lasticit
e
e
e
e
e
linaire, petites dformations, contraintes/dformations planes, ou axisymmetrie). Ainsi, la troisi`me
e
e
e
e
et derni`re tape consiste en une briefe introduction a la Mcanique des Solides, et plus prcisment
e e
`
e
e e
au Calcul Variationel. A travers la mthode des Puissances Virtuelles, et celle de Rayleigh-Ritz, ll`ve
e
ee
sera enn prt ` un autre cours dlments nis. Enn, un sujet dintrt particulier aux tudiants en
e a
ee
e e
e
Matriaux a t ajout, ` savoir la Rsistance Thorique des Matriaux cristallins. Ce sujet est capital
e
ee
e a
e
e
e
pour une bonne comprhension de la rupture et servira de lien a un ventuel cours sur la Mcanique de
e
`
e
e
la Rupture.
Ce polycopi a t enti`rement prpar par lauteur durant son anne sabbatique a lEcole Polye ee
e
e e
e
`
technique Fdrale de Lausanne, Dpartement des Matriaux. Le cours tait donn aux tudiants en
e e
e
e
e
e
e
deuxi`me anne en Franais.
e
e
c
Ce polycopi a t crit avec les objectifs suivants. Avant tout il doit tre complet et rigoureux. A
e ee e
e
tout moment, ll`ve doit tre ` mme de retrouver toutes les tapes suivies dans la drivation dune
ee
e
a e
e
e
quation. Ensuite, en allant a travers toutes les drivations, ll`ve sera a mme de bien conna les
e
`
e
ee
` e
itre
limitations et hypoth`ses derri`re chaque model. Enn, la rigueur scientique adopte, pourra servir
e
e
e
dexemple a la solution dautres probl`mes scientiques que ltudiant pourrait tre emmen ` rsoudre
`
e
e
e
ea e
dans le futur. Ce dernier point est souvent nglig.
e e
Le polycopi est subdivis de faon tr`s hirarchique. Chaque concept est dvelopp dans un parae
e
c
e
e
e
e
graphe spar. Ceci devrait faciliter non seulement la comprhension, mais aussi le dialogue entres levs
e e
e
e e
eux-mmes ainsi quavec le Professeur.
e
Quand il a t jug ncessaire, un bref rappel mathmatique est introduit. De nombreux exemples
ee
e e
e
sont prsents, et enn des exercices solutionns avec Mathematica sont prsents dans lannexe.
e
e
e
e
e
Lauteur ne se fait point dillusions quand au complet et a lexactitude de tout le polycopi. Il a t
`
e
ee
enti`rement dvelopp durant une seule anne acadmique, et pourrait donc bncier dune rvision
e
e
e
e
e
e e
e
extensive. A ce titre, corrections et critiques seront les bienvenues.
Enn, lauteur voudrait remercier ses levs qui ont diligemment suivis son cours sur la Mcanique
e e
e
de Milieux Continus durant lanne acadmique 1997-1998, ainsi que le Professeur Huet qui a t son
e
e
ee
hte au Laboratoire des Matriaux de Construction de lEPFL durant son sjour a Lausanne.
o
e
e
`

Victor Saouma
Ecublens, Juin 1998

Victor Saouma

Introduction to Continuum Mechanics

Draft
04

PREFACE

One of the most fundamental question that a Material Scientist has to ask him/herself is how a
material behaves under stress, and when does it break. Ultimately, it its the answer to those two
questions which would steer the development of new materials, and determine their survival in various
environmental and physical conditions.
The Material Scientist should then have a thorough understanding of the fundamentals of Mechanics
on the qualitative level, and be able to perform numerical simulation (most often by Finite Element
Method) and extract quantitative information for a specic problem.
In the humble opinion of the author, this is best achieved in three stages. First, the student should
be exposed to the basic principles of Continuum Mechanics. Detailed coverage of Stress, Strain, General
Principles, and Constitutive Relations is essential. Then, a brief exposure to Elasticity (along with Beam
Theory) would convince the student that a well posed problem can indeed have an analytical solution.
However, this is only true for problems problems with numerous simplifying assumptions (such as linear
elasticity, small deformation, plane stress/strain or axisymmetry, and resultants of stresses). Hence, the
last stage consists in a brief exposure to solid mechanics, and more precisely to Variational Methods.
Through an exposure to the Principle of Virtual Work, and the Rayleigh-Ritz Method the student will
then be ready for Finite Elements. Finally, one topic of special interest to Material Science students
was added, and that is the Theoretical Strength of Solids. This is essential to properly understand the
failure of solids, and would later on lead to a Fracture Mechanics course.
These lecture notes were prepared by the author during his sabbatical year at the Swiss Federal
Institute of Technology (Lausanne) in the Material Science Department. The course was oered to
second year undergraduate students in French, whereas the lecture notes are in English. The notes were
developed with the following objectives in mind. First they must be complete and rigorous. At any time,
a student should be able to trace back the development of an equation. Furthermore, by going through
all the derivations, the student would understand the limitations and assumptions behind every model.
Finally, the rigor adopted in the coverage of the subject should serve as an example to the students of
the rigor expected from them in solving other scientic or engineering problems. This last aspect is often
forgotten.
The notes are broken down into a very hierarchical format. Each concept is broken down into a small
section (a byte). This should not only facilitate comprehension, but also dialogue among the students
or with the instructor.
Whenever necessary, Mathematical preliminaries are introduced to make sure that the student is
equipped with the appropriate tools. Illustrative problems are introduced whenever possible, and last
but not least problem set using Mathematica is given in the Appendix.
The author has no illusion as to the completeness or exactness of all these set of notes. They were
entirely developed during a single academic year, and hence could greatly benet from a thorough review.
As such, corrections, criticisms and comments are welcome.
Finally, the author would like to thank his students who bravely put up with him and Continuum
Mechanics in the AY 1997-1998, and Prof. Huet who was his host at the EPFL.

Victor E. Saouma
Ecublens, June 1998

Victor Saouma

Introduction to Continuum Mechanics

Draft
Contents
I

CONTINUUM MECHANICS

07

1 MATHEMATICAL PRELIMINARIES; Part I Vectors and Tensors


1.1 Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.1.1 Operations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.1.2 Coordinate Transformation . . . . . . . . . . . . . . . . . . . . . . .
1.1.2.1 General Tensors . . . . . . . . . . . . . . . . . . . . . . . .
1.1.2.1.1 Contravariant Transformation . . . . . . . . . . .
1.1.2.1.2 Covariant Transformation . . . . . . . . . . . . . .
1.1.2.2 Cartesian Coordinate System . . . . . . . . . . . . . . . . .
1.2 Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2.1 Indicial Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2.2 Tensor Operations . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2.2.1 Sum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2.2.2 Multiplication by a Scalar . . . . . . . . . . . . . . . . . . .
1.2.2.3 Contraction . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2.2.4 Products . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2.2.4.1 Outer Product . . . . . . . . . . . . . . . . . . . .
1.2.2.4.2 Inner Product . . . . . . . . . . . . . . . . . . . .
1.2.2.4.3 Scalar Product . . . . . . . . . . . . . . . . . . . .
1.2.2.4.4 Tensor Product . . . . . . . . . . . . . . . . . . .
1.2.2.5 Product of Two Second-Order Tensors . . . . . . . . . . . .
1.2.3 Dyads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2.4 Rotation of Axes . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2.5 Trace . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2.6 Inverse Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2.7 Principal Values and Directions of Symmetric Second Order Tensors
1.2.8 Powers of Second Order Tensors; Hamilton-Cayley Equations . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

11
11
12
14
14
15
16
16
18
18
110
110
110
110
111
111
111
111
111
113
113
113
114
114
114
115

2 KINETICS
2.1 Force, Traction and Stress Vectors . . . . . . . . . . . .
2.2 Traction on an Arbitrary Plane; Cauchys Stress Tensor
E 2-1 Stress Vectors . . . . . . . . . . . . . . . . . . . .
2.3 Symmetry of Stress Tensor . . . . . . . . . . . . . . . .
2.3.1 Cauchys Reciprocal Theorem . . . . . . . . . . .
2.4 Principal Stresses . . . . . . . . . . . . . . . . . . . . . .
2.4.1 Invariants . . . . . . . . . . . . . . . . . . . . . .
2.4.2 Spherical and Deviatoric Stress Tensors . . . . .
2.5 Stress Transformation . . . . . . . . . . . . . . . . . . .
E 2-2 Principal Stresses . . . . . . . . . . . . . . . . . .
E 2-3 Stress Transformation . . . . . . . . . . . . . . .
2.5.1 Plane Stress . . . . . . . . . . . . . . . . . . . . .
2.5.2 Mohrs Circle for Plane Stress Conditions . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

21
21
23
24
25
25
26
28
28
29
29
210
210
210

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

Draft
02

2.6

E 2-4 Mohrs Circle in Plane Stress . . . .


2.5.3 Mohrs Stress Representation Plane
Simplied Theories; Stress Resultants . . .
2.6.1 Arch . . . . . . . . . . . . . . . . . .
2.6.2 Plates . . . . . . . . . . . . . . . . .

CONTENTS
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

3 MATHEMATICAL PRELIMINARIES; Part II VECTOR


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2 Derivative WRT to a Scalar . . . . . . . . . . . . . . . . . . .
E 3-1 Tangent to a Curve . . . . . . . . . . . . . . . . . . .
3.3 Divergence . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.1 Vector . . . . . . . . . . . . . . . . . . . . . . . . . . .
E 3-2 Divergence . . . . . . . . . . . . . . . . . . . . . . . .
3.3.2 Second-Order Tensor . . . . . . . . . . . . . . . . . . .
3.4 Gradient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4.1 Scalar . . . . . . . . . . . . . . . . . . . . . . . . . . .
E 3-3 Gradient of a Scalar . . . . . . . . . . . . . . . . . . .
E 3-4 Stress Vector normal to the Tangent of a Cylinder . .
3.4.2 Vector . . . . . . . . . . . . . . . . . . . . . . . . . . .
E 3-5 Gradient of a Vector Field . . . . . . . . . . . . . . . .
3.4.3 Mathematica Solution . . . . . . . . . . . . . . . . . .
3.5 Curl . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
E 3-6 Curl of a vector . . . . . . . . . . . . . . . . . . . . . .
3.6 Some useful Relations . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

212
214
214
214
217

DIFFERENTIATION
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

31
31
31
33
34
34
36
36
36
36
38
38
39
310
311
311
311
313

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

41
41
41
42
42
44
44
47
48
48
48
49
410
410
411
411
411
412
412
414
414
415
415
415
416
416
417
417
418

4 KINEMATIC
4.1 Elementary Denition of Strain . . . . . . . . . . . . . . . . . . . . . . .
4.1.1 Small and Finite Strains in 1D . . . . . . . . . . . . . . . . . . .
4.1.2 Small Strains in 2D . . . . . . . . . . . . . . . . . . . . . . . . .
4.2 Strain Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2.1 Position and Displacement Vectors; (x, X) . . . . . . . . . . . . .
E 4-1 Displacement Vectors in Material and Spatial Forms . . . . . . .
4.2.1.1 Lagrangian and Eulerian Descriptions; x(X, t), X(x, t) .
E 4-2 Lagrangian and Eulerian Descriptions . . . . . . . . . . . . . . .
4.2.2 Gradients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2.2.1 Deformation; (x X , X x ) . . . . . . . . . . . . . . . .
4.2.2.1.1 Change of Area Due to Deformation . . . . .
4.2.2.1.2 Change of Volume Due to Deformation . . .
E 4-3 Change of Volume and Area . . . . . . . . . . . . . . . . . . . . .
4.2.2.2 Displacements; (u X , u x ) . . . . . . . . . . . . . . .
4.2.2.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . .
E 4-4 Material Deformation and Displacement Gradients . . . . . . . .
4.2.3 Deformation Tensors . . . . . . . . . . . . . . . . . . . . . . . . .
4.2.3.1 Cauchys Deformation Tensor; (dX)2 . . . . . . . . . .
4.2.3.2 Greens Deformation Tensor; (dx)2 . . . . . . . . . . . .
E 4-5 Greens Deformation Tensor . . . . . . . . . . . . . . . . . . . . .
4.2.4 Strains; (dx)2 (dX)2 . . . . . . . . . . . . . . . . . . . . . . . .
4.2.4.1 Finite Strain Tensors . . . . . . . . . . . . . . . . . . .
4.2.4.1.1 Lagrangian/Greens Strain Tensor . . . . . . .
E 4-6 Lagrangian Tensor . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2.4.1.2 Eulerian/Almansis Tensor . . . . . . . . . . .
4.2.4.2 Innitesimal Strain Tensors; Small Deformation Theory
4.2.4.2.1 Lagrangian Innitesimal Strain Tensor . . . .
4.2.4.2.2 Eulerian Innitesimal Strain Tensor . . . . . .
Victor Saouma

.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

Introduction to Continuum Mechanics

Draft
CONTENTS

03

4.2.4.3 Examples . . . . . . . . . . . . . . . . . . . . . . .
E 4-7 Lagrangian and Eulerian Linear Strain Tensors . . . . . . .
4.2.5 Physical Interpretation of the Strain Tensor . . . . . . . .
4.2.5.1 Small Strain . . . . . . . . . . . . . . . . . . . . .
4.2.5.2 Finite Strain; Stretch Ratio . . . . . . . . . . . . .
4.3 Strain Decomposition . . . . . . . . . . . . . . . . . . . . . . . . .
4.3.1 Linear Strain and Rotation Tensors . . . . . . . . . . . . .
4.3.1.1 Small Strains . . . . . . . . . . . . . . . . . . . . .
4.3.1.1.1 Lagrangian Formulation . . . . . . . . . .
4.3.1.1.2 Eulerian Formulation . . . . . . . . . . .
4.3.1.2 Examples . . . . . . . . . . . . . . . . . . . . . . .
E 4-8 Relative Displacement along a specied direction . . . . . .
E 4-9 Linear strain tensor, linear rotation tensor, rotation vector .
4.3.2 Finite Strain; Polar Decomposition . . . . . . . . . . . . . .
E 4-10 Polar Decomposition I . . . . . . . . . . . . . . . . . . . . .
E 4-11 Polar Decomposition II . . . . . . . . . . . . . . . . . . . .
E 4-12 Polar Decomposition III . . . . . . . . . . . . . . . . . . . .
4.4 Summary and Discussion . . . . . . . . . . . . . . . . . . . . . . .
4.5 Compatibility Equation . . . . . . . . . . . . . . . . . . . . . . . .
E 4-13 Strain Compatibility . . . . . . . . . . . . . . . . . . . . . .
4.6 Lagrangian Stresses; Piola Kircho Stress Tensors . . . . . . . . .
4.6.1 First . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.6.2 Second . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
E 4-14 Piola-Kircho Stress Tensors . . . . . . . . . . . . . . . . .
4.7 Hydrostatic and Deviatoric Strain . . . . . . . . . . . . . . . . . .
4.8 Principal Strains, Strain Invariants, Mohr Circle . . . . . . . . . .
E 4-15 Strain Invariants & Principal Strains . . . . . . . . . . . . .
E 4-16 Mohrs Circle . . . . . . . . . . . . . . . . . . . . . . . . . .
4.9 Initial or Thermal Strains . . . . . . . . . . . . . . . . . . . . . . .
4.10 Experimental Measurement of Strain . . . . . . . . . . . . . . . .
4.10.1 Wheatstone Bridge Circuits . . . . . . . . . . . . . . . . . .
4.10.2 Quarter Bridge Circuits . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

418
418
419
419
421
422
422
422
422
424
425
425
425
426
426
427
428
430
430
432
432
433
433
434
434
436
436
438
439
439
440
441

5 MATHEMATICAL PRELIMINARIES; Part III VECTOR INTEGRALS


5.1 Integral of a Vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2 Line Integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.3 Integration by Parts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.4 Gauss; Divergence Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.5 Stokes Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.6 Green; Gradient Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
E 5-1 Physical Interpretation of the Divergence Theorem . . . . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

51
51
51
52
52
52
52
52

6 FUNDAMENTAL LAWS of CONTINUUM MECHANICS


6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.1.1 Conservation Laws . . . . . . . . . . . . . . . . . . . . .
6.1.2 Fluxes . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.2 Conservation of Mass; Continuity Equation . . . . . . . . . . .
6.2.1 Spatial Form . . . . . . . . . . . . . . . . . . . . . . . .
6.2.2 Material Form . . . . . . . . . . . . . . . . . . . . . . .
6.3 Linear Momentum Principle; Equation of Motion . . . . . . . .
6.3.1 Momentum Principle . . . . . . . . . . . . . . . . . . . .
E 6-1 Equilibrium Equation . . . . . . . . . . . . . . . . . . .
6.3.2 Moment of Momentum Principle . . . . . . . . . . . . .
6.3.2.1 Symmetry of the Stress Tensor . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

61
61
61
62
63
63
64
64
64
65
66
66

Victor Saouma

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

Introduction to Continuum Mechanics

Draft
04

6.4

CONTENTS
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

7 CONSTITUTIVE EQUATIONS; Part I LINEAR


7.1 Thermodynamic Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.1.1 State Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.1.2 Gibbs Relation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.1.3 Thermal Equation of State . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.1.4 Thermodynamic Potentials . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.1.5 Elastic Potential or Strain Energy Function . . . . . . . . . . . . . . . . . . .
7.2 Experimental Observations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.2.1 Hookes Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.2.2 Bulk Modulus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.3 Stress-Strain Relations in Generalized Elasticity . . . . . . . . . . . . . . . . . . . . .
7.3.1 Anisotropic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.3.2 Monotropic Material . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.3.3 Orthotropic Material . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.3.4 Transversely Isotropic Material . . . . . . . . . . . . . . . . . . . . . . . . . .
7.3.5 Isotropic Material . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.3.5.1 Engineering Constants . . . . . . . . . . . . . . . . . . . . . . . . . .
7.3.5.1.1 Isotropic Case . . . . . . . . . . . . . . . . . . . . . . . . .
7.3.5.1.1.1
Youngs Modulus . . . . . . . . . . . . . . . . . . . .
7.3.5.1.1.2
Bulks Modulus; Volumetric and Deviatoric Strains .
7.3.5.1.1.3
Restriction Imposed on the Isotropic Elastic Moduli
7.3.5.1.2 Transversly Isotropic Case . . . . . . . . . . . . . . . . . .
7.3.5.2 Special 2D Cases . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.3.5.2.1 Plane Strain . . . . . . . . . . . . . . . . . . . . . . . . . .
7.3.5.2.2 Axisymmetry . . . . . . . . . . . . . . . . . . . . . . . . . .
7.3.5.2.3 Plane Stress . . . . . . . . . . . . . . . . . . . . . . . . . .
7.4 Linear Thermoelasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.5 Fourrier Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.6 Updated Balance of Equations and Unknowns . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

71
. 71
. 71
. 72
. 73
. 73
. 74
. 75
. 75
. 75
. 76
. 76
. 77
. 77
. 78
. 79
. 710
. 710
. 710
. 711
. 712
. 712
. 713
. 713
. 713
. 714
. 714
. 715
. 716

6.5

6.6
6.7

Conservation of Energy; First Principle of Thermodynamics


6.4.1 Spatial Gradient of the Velocity . . . . . . . . . . . .
6.4.2 First Principle . . . . . . . . . . . . . . . . . . . . .
Equation of State; Second Principle of Thermodynamics . .
6.5.1 Entropy . . . . . . . . . . . . . . . . . . . . . . . . .
6.5.1.1 Statistical Mechanics . . . . . . . . . . . .
6.5.1.2 Classical Thermodynamics . . . . . . . . .
6.5.2 Clausius-Duhem Inequality . . . . . . . . . . . . . .
Balance of Equations and Unknowns . . . . . . . . . . . . .
Elements of Heat Transfer . . . . . . . . . . . . . . . . . .
6.7.1 Simple 2D Derivation . . . . . . . . . . . . . . . . .
6.7.2 Generalized Derivation . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

8 INTERMEZZO

II

81

ELASTICITY/SOLID MECHANICS

9 BOUNDARY VALUE PROBLEMS in


9.1 Preliminary Considerations . . . . . .
9.2 Boundary Conditions . . . . . . . . . .
9.3 Boundary Value Problem Formulation
9.4 Compacted Forms . . . . . . . . . . .
9.4.1 Navier-Cauchy Equations . . .

Victor Saouma

67
67
67
69
69
610
610
611
611
612
613
614

83

ELASTICITY
. . . . . . . . . .
. . . . . . . . . .
. . . . . . . . . .
. . . . . . . . . .
. . . . . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

91
91
91
93
93
93

Introduction to Continuum Mechanics

Draft
CONTENTS

05

9.4.2 Beltrami-Mitchell Equations . . . . . . . . . . .


9.4.3 Ellipticity of Elasticity Problems . . . . . . . .
Strain Energy and Extenal Work . . . . . . . . . . . .
Uniqueness of the Elastostatic Stress and Strain Field
Saint Venants Principle . . . . . . . . . . . . . . . . .
Cylindrical Coordinates . . . . . . . . . . . . . . . . .
9.8.1 Strains . . . . . . . . . . . . . . . . . . . . . . .
9.8.2 Equilibrium . . . . . . . . . . . . . . . . . . . .
9.8.3 Stress-Strain Relations . . . . . . . . . . . . . .
9.8.3.1 Plane Strain . . . . . . . . . . . . . .
9.8.3.2 Plane Stress . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

10 SOME ELASTICITY PROBLEMS


10.1 Semi-Inverse Method . . . . . . . . . . . . . . . . . . .
10.1.1 Example: Torsion of a Circular Cylinder . . . .
10.2 Airy Stress Functions . . . . . . . . . . . . . . . . . .
10.2.1 Cartesian Coordinates; Plane Strain . . . . . .
10.2.1.1 Example: Cantilever Beam . . . . . .
10.2.2 Polar Coordinates . . . . . . . . . . . . . . . .
10.2.2.1 Plane Strain Formulation . . . . . . .
10.2.2.2 Axially Symmetric Case . . . . . . . .
10.2.2.3 Example: Thick-Walled Cylinder . . .
10.2.2.4 Example: Hollow Sphere . . . . . . .
10.2.2.5 Example: Stress Concentration due to

.
.
.
.
.
.
.
.
.
.
a

. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
Circular

9.5
9.6
9.7
9.8

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

95
95
95
95
96
96
97
98
99
910
910

. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
Hole

. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
in a Plate

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

101
. 101
. 101
. 103
. 103
. 105
. 107
. 107
. 107
. 108
. 1010
. 1010

11 THEORETICAL STRENGTH OF PERFECT CRYSTALS


11.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
11.2 Theoretical Strength . . . . . . . . . . . . . . . . . . . . . . . .
11.2.1 Ideal Strength in Terms of Physical Parameters . . . . .
11.2.2 Ideal Strength in Terms of Engineering Parameter . . .
11.3 Size Eect; Grith Theory . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

111
. 111
. 113
. 113
. 115
. 116

12 BEAM THEORY
12.1 Introduction . . . . . . . . . . . . . . . . . . . .
12.2 Statics . . . . . . . . . . . . . . . . . . . . . . .
12.2.1 Equilibrium . . . . . . . . . . . . . . . .
12.2.2 Reactions . . . . . . . . . . . . . . . . .
12.2.3 Equations of Conditions . . . . . . . . .
12.2.4 Static Determinacy . . . . . . . . . . . .
12.2.5 Geometric Instability . . . . . . . . . . .
12.2.6 Examples . . . . . . . . . . . . . . . . .
E 12-1 Simply Supported Beam . . . . . . . . .
12.3 Shear & Moment Diagrams . . . . . . . . . . .
12.3.1 Design Sign Conventions . . . . . . . . .
12.3.2 Load, Shear, Moment Relations . . . . .
12.3.3 Examples . . . . . . . . . . . . . . . . .
E 12-2 Simple Shear and Moment Diagram . .
12.4 Beam Theory . . . . . . . . . . . . . . . . . . .
12.4.1 Basic Kinematic Assumption; Curvature
12.4.2 Stress-Strain Relations . . . . . . . . . .
12.4.3 Internal Equilibrium; Section Properties
12.4.3.1 Fx = 0; Neutral Axis . . . .
12.4.3.2 M = 0; Moment of Inertia .
12.4.4 Beam Formula . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

121
. 121
. 121
. 121
. 123
. 123
. 124
. 124
. 125
. 125
. 126
. 126
. 126
. 128
. 128
. 1210
. 1210
. 1211
. 1211
. 1211
. 1212
. 1212

Victor Saouma

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

Introduction to Continuum Mechanics

Draft
06

CONTENTS

12.4.5 Limitations of the Beam Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1213


12.4.6 Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1213
E 12-3 Design Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1213
13 VARIATIONAL METHODS
13.1 Preliminary Denitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
13.1.1 Internal Strain Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . .
13.1.2 External Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
13.1.3 Virtual Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
13.1.3.1 Internal Virtual Work . . . . . . . . . . . . . . . . . . . . . . .
13.1.3.2 External Virtual Work W . . . . . . . . . . . . . . . . . . . .
13.1.4 Complementary Virtual Work . . . . . . . . . . . . . . . . . . . . . . . .
13.1.5 Potential Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
13.2 Principle of Virtual Work and Complementary Virtual Work . . . . . . . . . .
13.2.1 Principle of Virtual Work . . . . . . . . . . . . . . . . . . . . . . . . . .
E 13-1 Tapered Cantiliver Beam, Virtual Displacement . . . . . . . . . . . . . .
13.2.2 Principle of Complementary Virtual Work . . . . . . . . . . . . . . . . .
E 13-2 Tapered Cantilivered Beam; Virtual Force . . . . . . . . . . . . . . . . .
13.3 Potential Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
13.3.1 Derivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
13.3.2 Rayleigh-Ritz Method . . . . . . . . . . . . . . . . . . . . . . . . . . . .
E 13-3 Uniformly Loaded Simply Supported Beam; Polynomial Approximation
13.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
14 INELASTICITY (incomplete)

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

131
. 131
. 132
. 133
. 134
. 134
. 135
. 135
. 136
. 136
. 136
. 137
. 139
. 1310
. 1311
. 1311
. 1313
. 1314
. 1316
1

A SHEAR, MOMENT and DEFLECTION DIAGRAMS for BEAMS

A1

B SECTION PROPERTIES

B1

C MATHEMATICAL PRELIMINARIES;
C.1 Euler Equation . . . . . . . . . . . . . .
E C-1 Extension of a Bar . . . . . . . .
E C-2 Flexure of a Beam . . . . . . . .

Part IV VARIATIONAL METHODS


C1
. . . . . . . . . . . . . . . . . . . . . . . . . . . . C1
. . . . . . . . . . . . . . . . . . . . . . . . . . . . C4
. . . . . . . . . . . . . . . . . . . . . . . . . . . . C5

D MID TERM EXAM

D1

E MATHEMATICA ASSIGNMENT and SOLUTION

E1

Victor Saouma

Introduction to Continuum Mechanics

Draft
List of Figures
1.1
1.2
1.3
1.4
1.5
1.6
1.7

Direction Cosines (to be corrected) .


Vector Addition . . . . . . . . . . . .
Cross Product of Two Vectors . . . .
Cross Product of Two Vectors . . . .
Coordinate Transformation . . . . .
Arbitrary 3D Vector Transformation
Rotation of Orthonormal Coordinate

. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
System

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

12
12
13
14
15
17
18

2.1
2.2
2.3
2.4
2.5
2.6
2.7
2.8
2.9
2.10
2.11
2.12
2.13

Stress Components on an Innitesimal Element . . .


Stresses as Tensor Components . . . . . . . . . . . .
Cauchys Tetrahedron . . . . . . . . . . . . . . . . .
Cauchys Reciprocal Theorem . . . . . . . . . . . . .
Principal Stresses . . . . . . . . . . . . . . . . . . . .
Mohr Circle for Plane Stress . . . . . . . . . . . . . .
Plane Stress Mohrs Circle; Numerical Example . . .
Unit Sphere in Physical Body around O . . . . . . .
Mohr Circle for Stress in 3D . . . . . . . . . . . . . .
Dierential Shell Element, Stresses . . . . . . . . . .
Dierential Shell Element, Forces . . . . . . . . . . .
Dierential Shell Element, Vectors of Stress Couples
Stresses and Resulting Forces in a Plate . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

22
22
23
26
27
211
213
214
215
215
216
216
217

3.1
3.2
3.3
3.4
3.5
3.6
3.7
3.8
3.9
3.10
3.11
3.12

Examples of a Scalar and Vector Fields . . . . . . . . . .


Dierentiation of position vector p . . . . . . . . . . . . .
Curvature of a Curve . . . . . . . . . . . . . . . . . . . . .
Mathematica Solution for the Tangent to a Curve in 3D .
Vector Field Crossing a Solid Region . . . . . . . . . . . .
Flux Through Area dA . . . . . . . . . . . . . . . . . . . .
Innitesimal Element for the Evaluation of the Divergence
Mathematica Solution for the Divergence of a Vector . . .
Radial Stress vector in a Cylinder . . . . . . . . . . . . . .
Gradient of a Vector . . . . . . . . . . . . . . . . . . . . .
Mathematica Solution for the Gradients of a Scalar and of
Mathematica Solution for the Curl of a Vector . . . . . .

.
.
.
.
.
.
.
.
.
.
a
.

. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
Vector .
. . . . .

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

32
32
33
34
34
35
35
37
38
310
311
312

4.1
4.2
4.3
4.4
4.5
4.6
4.7
4.8

Elongation of an Axial Rod . . . . . . . . . . . . . . . . . .


Elementary Denition of Strains in 2D . . . . . . . . . . . .
Position and Displacement Vectors . . . . . . . . . . . . . .
Position and Displacement Vectors, b = 0 . . . . . . . . . .
Undeformed and Deformed Congurations of a Continuum
Physical Interpretation of the Strain Tensor . . . . . . . . .
Relative Displacement du of Q relative to P . . . . . . . . .
Mohr Circle for Strain . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

41
43
45
46
413
420
423
436

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

Draft
02

LIST OF FIGURES

4.9
4.10
4.11
4.12

Bonded Resistance Strain Gage . .


Strain Gage Rosette . . . . . . . .
Quarter Wheatstone Bridge Circuit
Wheatstone Bridge Congurations

5.1

Physical Interpretation of the Divergence Theorem . . . . . . . . . . . . . . . . . . . . . . 53

6.1
6.2
6.3
6.4
6.5

Flux Through Area dS . . . . . . . . . . . . . .


Equilibrium of Stresses, Cartesian Coordinates
Flux vector . . . . . . . . . . . . . . . . . . . .
Flux Through Sides of Dierential Element . .
*Flow through a surface . . . . . . . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

62
66
613
614
614

9.1
9.2
9.3
9.4
9.5
9.6
9.7

Boundary Conditions in Elasticity Problems


Boundary Conditions in Elasticity Problems
Fundamental Equations in Solid Mechanics
St-Venants Principle . . . . . . . . . . . . .
Cylindrical Coordinates . . . . . . . . . . .
Polar Strains . . . . . . . . . . . . . . . . .
Stresses in Polar Coordinates . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

92
93
94
96
97
97
99

10.1
10.2
10.3
10.4

Torsion of a Circular Bar . . . .


Pressurized Thick Tube . . . . .
Pressurized Hollow Sphere . . . .
Circular Hole in an Innite Plate

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

102
109
1010
1010

11.1
11.2
11.3
11.4
11.5

Elliptical Hole in an Innite Plate . . . . . . . . . .


Griths Experiments . . . . . . . . . . . . . . . . .
Uniformly Stressed Layer of Atoms Separated by a0
Energy and Force Binding Two Adjacent Atoms . .
Stress Strain Relation at the Atomic Level . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

111
112
113
114
115

12.1
12.2
12.3
12.4
12.5
12.6
12.7

Types of Supports . . . . . . . . . . . . . . . . . . . . . . . . .
Inclined Roller Support . . . . . . . . . . . . . . . . . . . . . .
Examples of Static Determinate and Indeterminate Structures .
Geometric Instability Caused by Concurrent Reactions . . . . .
Shear and Moment Sign Conventions for Design . . . . . . . . .
Free Body Diagram of an Innitesimal Beam Segment . . . . .
Deformation of a Beam under Pure Bending . . . . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

123
124
124
125
126
127
1210

13.1
13.2
13.3
13.4
13.5
13.6
13.7
13.8

*Strain Energy and Complementary Strain Energy . . . . . . . . . . . . . . . . . .


Tapered Cantilivered Beam Analysed by the Vitual Displacement Method . . . . .
Tapered Cantilevered Beam Analysed by the Virtual Force Method . . . . . . . . .
Single DOF Example for Potential Energy . . . . . . . . . . . . . . . . . . . . . . .
Graphical Representation of the Potential Energy . . . . . . . . . . . . . . . . . . .
Uniformly Loaded Simply Supported Beam Analyzed by the Rayleigh-Ritz Method
Summary of Variational Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Duality of Variational Principles . . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

132
138
1310
1312
1313
1315
1317
1318

14.1
14.2
14.3
14.4
14.5
14.6

test . .
mod1 .
v-kv .
vis .
vis .
comp .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

1
2
2
3
3
3

.
.
.
.
.
.

Victor Saouma

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

439
440
441
442

Introduction to Continuum Mechanics

Draft

LIST OF FIGURES

03

14.7 epp . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
14.8 ehs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
C.1 Variational and Dierential Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C2

Victor Saouma

Introduction to Continuum Mechanics

Draft

LIST OF FIGURES

Victor Saouma

Introduction to Continuum Mechanics

04

Draft

LIST OF FIGURES

Symbol

05

NOTATION

Denition
SCALARS
A
Area
c
Specic heat
e
Volumetric strain
E
Elastic Modulus
g
Specicif free enthalpy
h
Film coecient for convection heat transfer
h
Specic enthalpy
I
Moment of inertia
J
Jacobian
K
Bulk modulus
K
Kinetic Energy
L
Length
p
Pressure
Q
Rate of internal heat generation
r
Radiant heat constant per unit mass per unit time
s
Specic entropy
S
Entropy
t
Time
T
Absolute temperature
u
Specic internal energy
U
Energy
Complementary strain energy
U
W
Work
W
Potential of External Work

Potential energy

Coecient of thermal expansion

Shear modulus

Poissons ratio

mass density
ij
Shear strains
1
ij
Engineering shear strain
2

Lames coecient

Stretch ratio
G
Lames coecient

Lames coecient

Airy Stress Function

(Helmholtz) Free energy


First stress and strain invariants
I , IE
II , IIE Second stress and strain invariants
III , IIIE Third stress and strain invariants

Temperature

Dimension

SI Unit

L2

m2

N.D.
L1 M T 2
L2 T 2

Pa
JKg 1

L2 T 2
L4

JKg 1
m4

L1 M T 2
L2 M T 2
L
L1 M T 2
L2 M T 3
M T 3 L4
L2 T 2 1
M L2 T 2 1
T

L2 T 2
L2 M T 2
L2 M T 2
L2 M T 2
L2 M T 2
L2 M T 2
1
L1 M T 2
N.D.
M L3
N.D.
N.D.
L1 M T 2
N.D.
L1 M T 2
L1 M T 2

Pa
J
m
Pa
W
W m6
JKg 1 K 1
JK 1
s
K
JKg 1
J
J
J
J
J
T 1
Pa
Kgm3
Pa
Pa
Pa

L2 M T 2

TENSORS order 1
b
b
q
t
t
u

Body force per unit massLT 2


Base transformation
Heat ux per unit area
Traction vector, Stress vector
Specied tractions along t
Displacement vector

Victor Saouma

N Kg 1
M T 3
L1 M T 2
L1 M T 2
L

W m2
Pa
Pa
m

Introduction to Continuum Mechanics

Draft
06

u(x)
u
x
X
0
(i)

LIST OF FIGURES
L
L
L
L
L1 M T 2
L1 M T 2

Specied displacements along u


Displacement vector
Spatial coordinates
Material coordinates
Initial stress vector
Principal stresses

m
m
m
m
Pa
Pa

TENSORS order 2
B1
C
D
E
E
E
F
H
I
J
k
K
L
R
T0

T
U
V
W
0
k

, T
T

Cauchys deformation tensor


N.D.
Greens deformation tensor; metric tensor,
right Cauchy-Green deformation tensor
N.D.
Rate of deformation tensor; Stretching tensor
N.D.
Lagrangian (or Greens) nite strain tensor
N.D.
Eulerian (or Almansi) nite strain tensor
N.D.
Strain deviator
N.D.
Material deformation gradient
N.D.
Spatial deformation gradient
N.D.
Idendity matrix
N.D.
Material displacement gradient
N.D.
Thermal conductivity
LM T 3 1
Spatial displacement gradient
N.D.
Spatial gradient of the velocity
Orthogonal rotation tensor
First Piola-Kircho stress tensor, Lagrangian Stress Tensor L1 M T 2
Second Piola-Kircho stress tensor
L1 M T 2
Right stretch tensor
Left stretch tensor
Spin tensor, vorticity tensor. Linear lagrangian rotation tensor
Initial strain vector
Conductivity
Curvature
Cauchy stress tensor
L1 M T 2
Deviatoric stress tensor
L1 M T 2
Linear Eulerian rotation tensor
Linear Eulerian rotation vector

W m1 K 1
-

Pa
Pa

Pa
Pa

TENSORS order 4
D

L1 M T 2

Constitutive matrix

Pa

L2
L2
L2
L2
L2
L2
L2
L3

m2
m2
m2
m2
m2
m2
m2
m3

CONTOURS, SURFACES, VOLUMES


C
S

t
u
T
c
q
, V

Contour line
Surface of a body
Surface
Boundary along which
Boundary along which
Boundary along which
Boundary along which
Boundary along which
Volume of body

surface tractions, t are specied


displacements, u are specied
temperatures, T are specied
convection ux, qc are specied
ux, qn are specied

FUNCTIONS, OPERATORS
Victor Saouma

Introduction to Continuum Mechanics

Draft

LIST OF FIGURES

u
2

07

Neighbour function to u(x)


Variational operator
Linear dierential operator relating displacement to strains
Divergence, (gradient operator) on scalar T
x
y
z
u
Divergence, (gradient operator) on vector (div . u = ux + yy +
x
Laplacian Operator

Victor Saouma

uz
z

Introduction to Continuum Mechanics

Draft

LIST OF FIGURES

Victor Saouma

Introduction to Continuum Mechanics

08

Draft

Part I

CONTINUUM MECHANICS

Draft

Draft
Chapter 1

MATHEMATICAL
PRELIMINARIES; Part I Vectors
and Tensors
Physical laws should be independent of the position and orientation of the observer. For this reason,
physical laws are vector equations or tensor equations, since both vectors and tensors transform
from one coordinate system to another in such a way that if the law holds in one coordinate system, it
holds in any other coordinate system.

1.1

Vectors

A vector is a directed line segment which can denote a variety of quantities, such as position of point
with respect to another (position vector), a force, or a traction.

A vector may be dened with respect to a particular coordinate system by specifying the components
of the vector in that system. The choice of the coordinate system is arbitrary, but some are more suitable
than others (axes corresponding to the major direction of the object being analyzed).

The rectangular Cartesian coordinate system is the most often used one (others are the cylindrical, spherical or curvilinear systems). The rectangular system is often represented by three mutually
perpendicular axes Oxyz, with corresponding unit vector triad i, j, k (or e1 , e2 , e3 ) such that:

ij = k;

jk = i;

ki = j;

(1.1-a)

ii = jj = kk = 1

(1.1-b)

ij = jk = ki = 0

(1.1-c)

Such a set of base vectors constitutes an orthonormal basis.


5

An arbitrary vector v may be expressed by


v = vx i + vy j + vz k

(1.2)

where
vx
vy

=
=

vi = v cos
vj = v cos

(1.3-a)
(1.3-b)

vz

vk = v cos

(1.3-c)

are the projections of v onto the coordinate axes, Fig. 1.1.

Draft
12

MATHEMATICAL PRELIMINARIES; Part I Vectors and Tensors


Y

Figure 1.1: Direction Cosines (to be corrected)

The unit vector in the direction of v is given by


ev =

v
= cos i + cos j + cos k
v

(1.4)

Since v is arbitrary, it follows that any unit vector will have direction cosines of that vector as its
Cartesian components.
7

The length or more precisely the magnitude of the vector is denoted by

v =

2
2
2
v1 + v2 + v3 .

k
8 We will denote the contravariant components of a vector by superscripts v , and its covariant
components by subscripts vk (the signicance of those terms will be claried in Sect. 1.1.2.1.

1.1.1

Operations

Addition: of two vectors a + b is geometrically achieved by connecting the tail of the vector b with the
head of a, Fig. 1.2. Analytically the sum vector will have components a1 + b1 a2 + b2 a3 + b3 .

u+v

Figure 1.2: Vector Addition


Scalar multiplication: a will scale the vector into a new one with components

a1

a2

a3 .

Vector Multiplications of a and b comes in three varieties:

Victor Saouma

Introduction to Continuum Mechanics

Draft
1.1 Vectors

13

Dot Product (or scalar product) is a scalar quantity which relates not only to the lengths of the
vector, but also to the angle between them.
3

ab a

cos (a, b) =

ai bi

(1.5)

i=1

where cos (a, b) is the cosine of the angle between the vectors a and b. The dot product
measures the relative orientation between two vectors.
The dot product is both commutative
ab = ba

(1.6)

a(b + c) = (ab) + (ac)

(1.7)

and distributive
The dot product of a with a unit vector n gives the projection of a in the direction of n.
The dot product of base vectors gives rise to the denition of the Kronecker delta dened
as
ei ej = ij

(1.8)

where
ij =

1
0

if
if

i=j
i=j

(1.9)

Cross Product (or vector product) c of two vectors a and b is dened as the vector
c = ab = (a2 b3 a3 b2 )e1 + (a3 b1 a1 b3 )e2 + (a1 b2 a2 b1 )e3

(1.10)

which can be remembered from the determinant expansion of


ab =

e2
a2
b2

e1
a1
b1

e3
a3
b3

(1.11)

and is equal to the area of the parallelogram described by a and b, Fig. 1.3.
axb

A(a,b)=||a x b||
b

Figure 1.3: Cross Product of Two Vectors

A(a, b) = ab
Victor Saouma

(1.12)

Introduction to Continuum Mechanics

Draft
14

MATHEMATICAL PRELIMINARIES; Part I Vectors and Tensors

The cross product is not commutative, but satises the condition of skew symmetry
ab = ba

(1.13)

a(b + c) = (ab) + (ac)

(1.14)

The cross product is distributive

Triple Scalar Product: of three vectors a, b, and c is desgnated by (ab)c and it corresponds
to the (scalar) volume dened by the three vectors, Fig. 1.4.
n=a x b
||a x b||

c
c.n

Figure 1.4: Cross Product of Two Vectors


V (a, b, c)

=
=

(ab)c = a(bc) (1.15)


ax ay az
bx by bz
(1.16)
cx cy cz

The triple scalar product of base vectors represents a fundamental operation

1 if (i, j, k) are in cyclic order


0 if any of (i, j, k) are equal
(ei ej )ek = ijk

1 if (i, j, k) are in acyclic order

(1.17)

The scalars ijk is the permutation tensor. A cyclic permutation of 1,2,3 is 1 2 3 1,


an acyclic one would be 1 3 2 1. Using this notation, we can rewrite
c = ab ci = ijk aj bk

(1.18)

Vector Triple Product is a cross product of two vectors, one of which is itself a cross product.
a(bc) = (ac)b (ab)c = d

(1.19)

and the product vector d lies in the plane of b and c.

1.1.2

Coordinate Transformation

1.1.2.1

General Tensors

9 Let us consider two bases bj (x1 , x2 , x3 ) and bj (x1 , x2 x3 ), Fig. 1.5. Each unit vector in one basis must
be a linear combination of the vectors of the other basis

bj = ap bp and bk = bk bq
q
j
Victor Saouma

(1.20)

Introduction to Continuum Mechanics

Draft
1.1 Vectors

15

(summed on p and q respectively) where ap (subscript new, superscript old) and bk are the coecients
q
j
for the forward and backward changes respectively from b to b respectively. Explicitly
1 1 1

b1 b2 b3 e1
a1 a2 a3 e1
e1
e1
1
1
e2
e2
e2
e2
= b2 b2 b2
and
= a1 a2 a3
(1.21)
1
2
3
2
2
2

3
3
3
e3
b1 b2 b3
e3
e3
a1 a2 a3
e3
3
3
3

X2
X2
X1
-1

2
cos a1

X1

X3
X3
Figure 1.5: Coordinate Transformation

10

The transformation must have the determinant of its Jacobian


J=

x1
1
x2
x
1
x3
x
x1

x1
2
x2
x
2
x3
x
x2

x1
3
x2
x
3
x3
x
x3

=0

(1.22)

dierent from zero (the superscript is a label and not an exponent).


11 It is important to note that so far, the coordinate systems are completely general and may be Cartesian, curvilinear, spherical or cylindrical.

1.1.2.1.1

12

Contravariant Transformation

The vector representation in both systems must be the same


v = v q bq = v k bk = v k (bq bq ) (v q v k bq )bq = 0
k
k

(1.23)

since the base vectors bq are linearly independent, the coecients of bq must all be zero hence
v q = bq v k and inversely v p = ap v j
j
k

(1.24)

showing that the forward change from components v k to v q used the coecients bq of the backward
k
change from base bq to the original bk . This is why these components are called contravariant.
13 Generalizing, a Contravariant Tensor of order one (recognized by the use of the superscript)
transforms a set of quantities rk associated with point P in xk through a coordinate transformation into

Victor Saouma

Introduction to Continuum Mechanics

Draft
16

MATHEMATICAL PRELIMINARIES; Part I Vectors and Tensors

a new set rq associated with xq


rq =

xq k
r
xk

(1.25)

bq
k
14 By extension, the Contravariant tensors of order two requires the tensor components to obey
the following transformation law

rij =

1.1.2.1.2

xi xj rs
r
xr xs

(1.26)

Covariant Transformation

15 Similarly to Eq. 1.24, a covariant component transformation (recognized by subscript) will be


dened as

v j = ap vp and inversely vk = bk v q
q
j

(1.27)

We note that contrarily to the contravariant transformation, the covariant transformation uses the same
transformation coecients as the ones for the base vectors.
16

Finally transformation of tensors of order one and two is accomplished through


rq
rij

1.1.2.2

=
=

xk
rk
xq
r
x xs
rrs
xi xj

(1.28)
(1.29)

Cartesian Coordinate System

17 If we consider two dierent sets of cartesian orthonormal coordinate systems {e1 , e2 , e3 } and {e1 , e2 , e3 },
any vector v can be expressed in one system or the other

v = vj ej = v j ej

(1.30)

18 To determine the relationship between the two sets of components, we consider the dot product of v
with one (any) of the base vectors
ei v = v i = vj (ei ej )
(1.31)

(since v j (ej ei ) = v j ij = v i )
19

We can thus dene the nine scalar values


aj ei ej = cos(xi , xj )
i

(1.32)

which arise from the dot products of base vectors as the direction cosines. (Since we have an orthonormal system, those values are nothing else than the cosines of the angles between the nine pairing
of base vectors.)
20 Thus, one set of vector components can be expressed in terms of the other through a covariant
transformation similar to the one of Eq. 1.27.

Victor Saouma

Introduction to Continuum Mechanics

Draft
1.1 Vectors

17
vj
vk

ap vp
j

(1.33)

bk v q
q

(1.34)

we note that the free index in the rst and second equations appear on the upper and lower index
respectively.
21

Because of the orthogonality of the unit vector we have as as = pq and am an = mn .


p q
r r

As a further illustration of the above derivation, let us consider the transformation of a vector V from
(X, Y, Z) coordinate system to (x, y, z), Fig. 1.6:

22

Figure 1.6: Arbitrary 3D Vector Transformation

23

Eq. 1.33 would then result in


Vx = aX VX + aY VY + aZ VZ
x
x
x
X

ax
Vx
Vy
= aX
y

Vz
aX
z

or

aY
x
aY
y
aY
z

aZ VX
x
aZ
VY
y

VZ
aZ
z

(1.35)

(1.36)

and aj is the direction cosine of axis i with respect to axis j


i
aj = (ax X, aY , aZ ) direction cosines of x with respect to X, Y and Z
x
x
x
aj = (ay X, aY , aZ ) direction cosines of y with respect to X, Y and Z
y
y
y
aj = (az X, aY , aZ ) direction cosines of z with respect to X, Y and Z
z
z
z
24

Finally, for the 2D case and from Fig. 1.7, the transformation matrix is written as
T =

but since = + , and =


2
matrix becomes

a1
1
a1
2

cos
cos

cos
cos

(1.37)

, then cos = sin and cos = sin , thus the transformation


T =

Victor Saouma

a2
1
a2
2

cos
sin

sin
cos

(1.38)

Introduction to Continuum Mechanics

Draft
18

MATHEMATICAL PRELIMINARIES; Part I Vectors and Tensors


X

X2

X1

X1

Figure 1.7: Rotation of Orthonormal Coordinate System

1.2

Tensors

25 We now seek to generalize the concept of a vector by introducing the tensor (T), which essentially
exists to operate on vectors v to produce other vectors (or on tensors to produce other tensors!). We
designate this operation by Tv or simply Tv.

26

We hereby adopt the dyadic notation for tensors as linear vector operators
u
u

= Tv or ui = Tij vj
= vS where S = TT

(1.39-a)
(1.39-b)

In general the vectors may be represented by either covariant or contravariant components vj or v j .


Thus we can have dierent types of linear transformations
27

ui
ui

=
=

Tij v j ; ui
Ti.j vj ; ui

=
=

T ij vj
i
T.j v j

(1.40)

involving the covariant components Tij , the contravariant components T ij and the mixed comi
ponents T.j or Ti.j .
28 Whereas a tensor is essentially an operator on vectors (or other tensors), it is also a physical quantity,
independent of any particular coordinate system yet specied most conveniently by referring to an
appropriate system of coordinates.
29 Tensors frequently arise as physical entities whose components are the coecients of a linear relationship between vectors.
30 A tensor is classied by the rank or order. A Tensor of order zero is specied in any coordinate system
by one coordinate and is a scalar. A tensor of order one has three coordinate components in space, hence
it is a vector. In general 3-D space the number of components of a tensor is 3n where n is the order of
the tensor.

31

A force and a stress are tensors of order 1 and 2 respectively.

1.2.1

Indicial Notation

32 Whereas the Engineering notation may be the simplest and most intuitive one, it often leads to long
and repetitive equations. Alternatively, the tensor and the dyadic form will lead to shorter and more
compact forms.

Victor Saouma

Introduction to Continuum Mechanics

Draft
1.2 Tensors

19

33 While working on general relativity, Einstein got tired of writing the summation symbol with its range
n=3
of summation below and above (such as i=1 aij bi ) and noted that most of the time the upper range
(n) was equal to the dimension of space (3 for us, 4 for him), and that when the summation involved a
product of two terms, the summation was over a repeated index (i in our example). Hence, he decided
that there is no need to include the summation sign
if there was repeated indices (i), and thus any
repeated index is a dummy index and is summed over the range 1 to 3. An index that is not repeated
is called free index and assumed to take a value from 1 to 3.

34

Hence, this so called indicial notation is also referred to Einsteins notation.

35

The following rules dene indicial notation:


1. If there is one letter index, that index goes from i to n (range of the tensor). For instance:

a1
a2
ai = ai = a1 a2 a3 =
i = 1, 3
(1.41)

a3
assuming that n = 3.
2. A repeated index will take on all the values of its range, and the resulting tensors summed. For
instance:
(1.42)
a1i xi = a11 x1 + a12 x2 + a13 x3
3. Tensors order:
First order tensor (such as force) has only one free index:
ai = ai =

a1

a2

a3

(1.43)

other rst order tensors aij bj , Fikk , ijk uj vk


Second order tensor (such as stress or strain) will have two free indeces.

D11 D22 D13


Dij = D21 D22 D23
D31 D32 D33

(1.44)

other examples Aijip , ij uk vk .


A fourth order tensor (such as Elastic constants) will have four free indeces.
4. Derivatives of tensor with respect to xi is written as , i. For example:

xi

36

= ,i

vi
xi

= vi,i

vi
xj

= vi,j

Ti,j
xk

= Ti,j,k

(1.45)

Usefulness of the indicial notation is in presenting systems of equations in compact form. For instance:
xi = cij zj

(1.46)

this simple compacted equation, when expanded would yield:


x1
x2

=
=

c11 z1 + c12 z2 + c13 z3


c21 z1 + c22 z2 + c23 z3

x3

c31 z1 + c32 z2 + c33 z3

(1.47-a)

Similarly:
Aij = Bip Cjq Dpq
Victor Saouma

(1.48)

Introduction to Continuum Mechanics

Draft
110

MATHEMATICAL PRELIMINARIES; Part I Vectors and Tensors

A11
A12
A21
A22

37

= B11 C11 D11 + B11 C12 D12 + B12 C11 D21 + B12 C12 D22
= B11 C11 D11 + B11 C12 D12 + B12 C11 D21 + B12 C12 D22
= B21 C11 D11 + B21 C12 D12 + B22 C11 D21 + B22 C12 D22
= B21 C21 D11 + B21 C22 D12 + B22 C21 D21 + B22 C22 D22

(1.49-a)

Using indicial notation, we may rewrite the denition of the dot product

ab = ai bi

(1.50)

ab = pqr aq br ep

(1.51)

and of the cross product

we note that in the second equation, there is one free index p thus there are three equations, there are
two repeated (dummy) indices q and r, thus each equation has nine terms.

1.2.2

Tensor Operations

1.2.2.1

Sum

38

The sum of two (second order) tensors is simply dened as:

(1.52)

Sij = Tij + Uij

1.2.2.2
39

Multiplication by a Scalar

The multiplication of a (second order) tensor by a scalar is dened by:

(1.53)

Sij = Tij

1.2.2.3

Contraction

40 In a contraction, we make two of the indeces equal (or in a mixed tensor, we make a ubscript equal to
the superscript), thus producing a tensor of order two less than that to which it is applied. For example:

Tij
u i vj
Amr
..sn
Eij ak
Ampr
qs

Victor Saouma

Tii ;
u i vi ;
r
Amr = B.s ;
..sm
Eij ai = cj ;
mp
Ampr = Bq ;
qr

2
2
4
3
5

0
0
2
1
3

(1.54)

Introduction to Continuum Mechanics

Draft
1.2 Tensors
1.2.2.4
1.2.2.4.1

111

Products
Outer Product

41 The outer product of two tensors (not necessarily of the same type or order) is a set of tensor
components obtained simply by writing the components of the two tensors beside each other with no
repeated indices (that is by multiplying each component of one of the tensors by every component of
the other). For example

ai bj
.k
A Bj

=
=

Tij
C i.k .j

(1.55-a)
(1.55-b)

vi Tjk

Sijk

(1.55-c)

1.2.2.4.2

Inner Product

42 The inner product is obtained from an outer product by contraction involving one index from each
tensor. For example

ai bj
ai Ejk
Eij Fkm
.k
Ai Bi

1.2.2.4.3

43

ai bi
ai Eik = fk

(1.56-a)
(1.56-b)

Eij Fjm = Gim


.k
Ai Bi = Dk

(1.56-c)
(1.56-d)

Scalar Product

The scalar product of two tensors is dened as


(1.57)

T : U = Tij Uij
in any rectangular system.
44

The following inner-product axioms are satised:


T:U = U:T
T : (U + V) = T : U + T : V
(T : U) = (T) : U = T : (U)
T : T > 0 unless T = 0

1.2.2.4.4

(1.58-a)
(1.58-b)
(1.58-c)
(1.58-d)

Tensor Product

45 Since a tensor primary objective is to operate on vectors, the tensor product of two vectors provides
a fundamental building block of second-order tensors and will be examined next.

Victor Saouma

Introduction to Continuum Mechanics

Draft
112

MATHEMATICAL PRELIMINARIES; Part I Vectors and Tensors

The Tensor Product of two vectors u and v is a second order tensor u v which in turn operates
on an arbitrary vector w as follows:

46

[u v]w (vw)u

(1.59)

In other words when the tensor product u v operates on w (left hand side), the result (right hand
side) is a vector that points along the direction of u, and has length equal to (vw)||u||, or the original
length of u times the dot (scalar) product of v and w.
47 Of particular interest is the tensor product of the base vectors ei ej . With three base vectors, we
have a set of nine second order tensors which provide a suitable basis for expressing the components of a
tensor. Again, we started with base vectors which themselves provide a basis for expressing any vector,
and now the tensor product of base vectors in turn provides a formalism to express the components of
a tensor.
48 The second order tensor T can be expressed in terms of its components Tij relative to the base
tensors ei ej as follows:

Tij [ei ej ]

(1.60-a)

Tij [ei ej ] ek

(1.60-b)

i=1 j=1
3

Tek

=
i=1 j=1

[ei ej ] ek

(ej ek )ei = jk ei

(1.60-c)

Tek

Tik ei

(1.60-d)

i=1

Thus Tik is the ith component of Tek . We can thus dene the tensor component as follows
Tij = ei Tej

(1.61)

Now we can see how the second order tensor T operates on any vector v by examining the components
of the resulting vector Tv:

49

Tv =

Tij [ei ej ]

i=1 j=1

Tij vk [ei ej ]ek

vk ek

(1.62)

i=1 j=1 k=1

k=1

which when combined with Eq. 1.60-c yields


3

Tv =

Tij vj ei

(1.63)

i=1 j=1

which is clearly a vector. The ith component of the vector Tv being


3

(Tv)i =

Tij vj

(1.64)

i=1

50

The identity tensor I leaves the vector unchanged Iv = v and is equal to


I ei ei

Victor Saouma

(1.65)
Introduction to Continuum Mechanics

Draft
1.2 Tensors

113

51 A simple example of a tensor and its operation on vectors is the projection tensor P which generates
the projection of a vector v on the plane characterized by a normal n:

PInn

(1.66)

the action of P on v gives Pv = v (vn)n. To convince ourselves that the vector Pv lies on the plane,

its dot product with n must be zero, accordingly Pvn = vn (vn)(nn) = 0 .


1.2.2.5
52

Product of Two Second-Order Tensors

The product of two tensors is dened as


P = TU;

(1.67)

Pij = Tik Ukj

in any rectangular system.


53

The following axioms hold


(TU)R = T(UR)
T(R + U) = TR + tU

(1.68-a)
(1.68-b)

(R + U)T = RT + UT
(TU) = (T)U = T(U)

(1.68-c)
(1.68-d)

1T =

T1 = T

(1.68-e)

Note again that some authors omit the dot.


Finally, the operation is not commutative

1.2.3

Dyads

54 The indeterminate vector product of a and b dened by writing the two vectors in juxtaposition as
ab is called a dyad. A dyadic D corresponds to a tensor of order two and is a linear combination of
dyads:
(1.69)
D = a1 b1 + a2 b2 an bn

The conjugate dyadic of D is written as


Dc = b1 a1 + b2 a2 bn an

1.2.4
55

(1.70)

Rotation of Axes

The rule for changing second order tensor components under rotation of axes goes as follow:
ui

= aj uj
i
= aj Tjq vq
i
= aj Tjq aq v p
p
i

From Eq. 1.33


From Eq. 1.39-a
From Eq. 1.33

(1.71)

But we also have ui = T ip v p (again from Eq. 1.39-a) in the barred system, equating these two expressions
we obtain
T ip (aj aq Tjq )v p = 0
(1.72)
i p
hence
T ip
Tjq
Victor Saouma

aj aq Tjq in Matrix Form [T ] = [A]T [T ][A]


i p

(1.73)

aj aq T ip
i p

(1.74)

in Matrix Form [T ] = [A][T ][A]

Introduction to Continuum Mechanics

Draft
114

MATHEMATICAL PRELIMINARIES; Part I Vectors and Tensors

By extension, higher order tensors can be similarly transformed from one coordinate system to another.
56

If we consider the 2D case, From Eq. 1.38

cos sin 0
A = sin cos 0
0
0
1

Txx Txy 0
T = Txy Tyy 0
0
0
0

T xx T xy 0
T = AT T A = T xy T yy 0
0
0
0

2
2
cos Txx + sin Tyy + sin 2Txy
= 1 ( sin 2Txx + sin 2Tyy + 2 cos 2Txy
2
0

(1.75-a)

(1.75-b)

(1.75-c)
1
2 ( sin 2Txx + sin 2Tyy + 2 cos 2Txy
sin2 Txx + cos (cos Tyy 2 sin Txy

0
0
0
(1.75-d)

alternatively, using sin 2 = 2 sin cos and cos 2 = cos2 sin2 , this last equation can be rewritten
as

sin2
2 sin cos
cos2
Txx
T xx
(1.76)
Tyy
=
sin2
cos2
2 sin cos
T

yy

Txy
sin cos cos sin cos2 sin2
T xy

1.2.5

Trace

The trace of a second-order tensor, denoted tr T is a scalar invariant function of the tensor and is
dened as

57

tr T Tii

(1.77)

Thus it is equal to the sum of the diagonal elements in a matrix.

1.2.6
58

Inverse Tensor

An inverse tensor is simply dened as follows


T1 (Tv) = v and T(T1 v) = v

(1.78)

1
1
alternatively T1 T = TT1 = I, or Tik Tkj = ij and Tik Tkj = ij

1.2.7

Principal Values and Directions of Symmetric Second Order Tensors

59 Since the two fundamental tensors in continuum mechanics are of the second order and symmetric
(stress and strain), we examine some important properties of these tensors.
60 For every symmetric tensor Tij dened at some point in space, there is associated with each direction
(specied by unit normal nj ) at that point, a vector given by the inner product

vi = Tij nj
Victor Saouma

(1.79)

Introduction to Continuum Mechanics

Draft
1.2 Tensors

115

If the direction is one for which vi is parallel to ni , the inner product may be expressed as
Tij nj = ni

(1.80)

and the direction ni is called principal direction of Tij . Since ni = ij nj , this can be rewritten as
(Tij ij )nj = 0

(1.81)

which represents a system of three equations for the four unknowns ni and .
(T11 )n1 + T12 n2 + T13 n3
T21 n1 + (T22 )n2 + T23 n3

=
=

0
0

T31 n1 + T32 n2 + (T33 )n3

(1.82-a)

To have a non-trivial slution (ni = 0) the determinant of the coecients must be zero,
|Tij ij | = 0

61

(1.83)

Expansion of this determinant leads to the following characteristic equation


3 IT 2 + IIT IIIT = 0

(1.84)

the roots are called the principal values of Tij and


IT

IIT

IIIT

Tij = tr Tij
1
(Tii Tjj Tij Tij )
2
|Tij | = det Tij

(1.85)
(1.86)
(1.87)

are called the rst, second and third invariants respectively of Tij .
62

It is customary to order those roots as 1 > 2 > 3

63 For a symmetric tensor with real components, the principal values are also real. If those values are
distinct, the three principal directions are mutually orthogonal.

1.2.8
64

Powers of Second Order Tensors; Hamilton-Cayley Equations

When expressed in term of the principal axes, the tensor array can be written in matrix form as

(1)
0
0
0
(2)
T = 0
(1.88)
0
0
(3)

65 By direct matrix multiplication, the quare of the tensor Tij is given by the inner product Tik Tkj , the
cube as Tik Tkm Tmn . Therefore the nth power of Tij can be written as

0
0
n
(1)

0
n
Tn= 0
(1.89)

(2)
0
0
n
(3)

Victor Saouma

Introduction to Continuum Mechanics

Draft
116

MATHEMATICAL PRELIMINARIES; Part I Vectors and Tensors

Since each of the principal values satises Eq. 1.84 and because the diagonal matrix form of T given
above, then the tensor itself will satisfy Eq. 1.84.
T 3 IT T 2 + IIT T IIIT I = 0

(1.90)

where I is the identity matrix. This equation is called the Hamilton-Cayley equation.

Victor Saouma

Introduction to Continuum Mechanics

Draft
Chapter 2

KINETICS
Or How Forces are Transmitted

2.1
1

Force, Traction and Stress Vectors

There are two kinds of forces in continuum mechanics

body forces: act on the elements of volume or mass inside the body, e.g. gravity, electromagnetic
elds. dF = bdV ol.
surface forces: are contact forces acting on the free body at its bounding surface. Those will be dened
in terms of force per unit area.
The surface force per unit area acting on an element dS is called traction or more accurately stress
vector.
tdS = i
tx dS + j
ty dS + k
tz dS
(2.1)

Most authors limit the term traction to an actual bounding surface of a body, and use the term stress
vector for an imaginary interior surface (even though the state of stress is a tensor and not a vector).
The traction vectors on planes perpendicular to the coordinate axes are particularly useful. When the
vectors acting at a point on three such mutually perpendicular planes is given, the stress vector at
that point on any other arbitrarily inclined plane can be expressed in terms of the rst set of tractions.

4 A stress, Fig 2.1 is a second order cartesian tensor, ij where the 1st subscript (i) refers to the
direction of outward facing normal, and the second one (j) to the direction of component force.


11 12 13
t1
t2
= ij = 21 22 23 =
(2.2)

31 32 33
t3

5 In fact the nine rectangular components ij of turn out to be the three sets of three vector components
(11 , 12 , 13 ), (21 , 22 , 23 ), (31 , 32 , 33 ) which correspond to the three tractions t1 , t2 and t3 which
are acting on the x1 , x2 and x3 faces (It should be noted that those tractions are not necesarily normal
to the faces, and they can be decomposed into a normal and shear traction if need be). In other words,
stresses are nothing else than the components of tractions (stress vector), Fig. 2.2.

The state of stress at a point cannot be specied entirely by a single vector with three components; it
requires the second-order tensor with all nine components.

Draft
22

KINETICS

X3

33

32

31

23

13

21

X3

22

X2

12

X1

11
X2
X1

Figure 2.1: Stress Components on an Innitesimal Element

X3
X3

V3

33
t3

t2

13

21

X1

t1

32

23

31

11

V2
X2

22

V1
X1

X2

(Components of a vector are scalars)

12

Stresses as components of a traction vector

(Components of a tensor of order 2 are vectors)

Figure 2.2: Stresses as Tensor Components

Victor Saouma

Introduction to Continuum Mechanics

Draft

2.2 Traction on an Arbitrary Plane; Cauchys Stress Tensor

2.2

23

Traction on an Arbitrary Plane; Cauchys Stress Tensor

Let us now consider the problem of determining the traction acting on the surface of an oblique plane
(characterized by its normal n) in terms of the known tractions normal to the three principal axis, t1 , t2
and t3 . This will be done through the so-called Cauchys tetrahedron shown in Fig. 2.3.
7

X2

-t
1

-t

S3

*
tn S

h N

n
A

X1

-t 2 S2
*

X3
*

Figure 2.3: Cauchys Tetrahedron

The components of the unit vector n are the direction cosines of its direction:
n1 = cos( AON );

n2 = cos( BON );

n3 = cos( CON );

(2.3)

The altitude ON , of length h is a leg of the three right triangles AN O, BN O and CN O with hypothenuses OA, OB and OC. Hence
h = OAn1 = OBn2 = OCn3
9

(2.4)

The volume of the tetrahedron is one third the base times the altitude
V =

1
1
1
1
hS = OAS1 = OBS2 = OCS3
3
3
3
3

(2.5)

which when combined with the preceding equation yields


S1 = Sn1 ;

S2 = Sn2 ;

S3 = Sn3 ;

(2.6)

or Si = Sni .
10 In Fig. 2.3 are also shown the average values of the body force and of the surface tractions (thus the
asterix). The negative sign appears because t denotes the average traction on a surface whose outward
i
normal points in the negative xi direction. We seek to determine t .
n
11 We invoke the momentum principle of a collection of particles (more about it later on) which
is postulated to apply to our idealized continuous medium. This principle states that the vector sum

Victor Saouma

Introduction to Continuum Mechanics

Draft
24

KINETICS

of all external forces acting on the free body is equal to the rate of change of the total momentum1 .
vdm. By the mean-value theorem of the integral calculus, this is equal

The total momentum is


m

to v m where v is average value of the velocity. Since we are considering the momentum of a given

collection of particles, m does not change with time and m dv = V dv where is the average
dt
dt
density. Hence, the momentum principle yields
t S + b V t S1 t S2 t S3 = V
n
1
2
3

dv
dt

(2.7)

Substituting for V , Si from above, dividing throughout by S and rearanging we obtain


1
1
dv
t + h b = t n1 + t n2 + t n3 + h
n
1
2
3
3
3
dt

(2.8)

and now we let h 0 and obtain


tn = t1 n1 + t2 n2 + t3 n3 = ti ni

(2.9)

We observe that we dropped the asterix as the length of the vectors approached zero.
12 It is important to note that this result was obtained without any assumption of equilibrium and that
it applies as well in uid dynamics as in solid mechanics.
13 This equation is a vector equation, and the corresponding algebraic equations for the components of
tn are
tn1 = 11 n1 + 21 n2 + 31 n3
tn2 = 12 n1 + 22 n2 + 32 n3
tn3 = 13 n1 + 23 n2 + 33 n3
(2.10)
Indicial notation tni = ji nj
dyadic notation
tn = n = T n

14 We have thus established that the nine components ij are components of the second order tensor,
Cauchys stress tensor.
15 Note that this stress tensor is really dened in the deformed space (Eulerian), and this issue will be
revisited in Sect. 4.6.

Example 2-1: Stress Vectors


if the stress tensor at point P is given by

7 5
= 5 3
0
1


0
t1
1 =
t2

t3
2

(2.11)

We seek to determine the traction (or stress vector) t passing through P and parallel to the plane ABC
where A(4, 0, 0), B(0, 2, 0) and C(0, 0, 6). Solution:
The vector normal to the plane can be found by taking the cross products of vectors AB and AC:
N

= ABAC =
=

1 This

12e1 + 24e2 + 8e3

is really Newtons second law F = ma =

Victor Saouma

e2
2
0

e1
4
4

e3
0
6

(2.12-a)
(2.12-b)

m dv
dt

Introduction to Continuum Mechanics

Draft

2.3 Symmetry of Stress Tensor

25

The unit normal of N is given by


n=

3
6
2
e1 + e2 + e3
7
7
7

Hence the stress vector (traction) will be

3
7

and thus t = 9 e1 + 5 e2 +
7
7

2.3
16

6
7

2
7

7 5
5 3
0
1

0
1 =
2

(2.13)

9
7

5
7

10
7

(2.14)

10
7 e3

Symmetry of Stress Tensor

From Fig. 2.1 the resultant force exerted on the positive X1 face is
11 X2 X3

12 X2 X3

13 X2 X3

(2.15)

similarly the resultant forces acting on the positive X2 face are


21 X3 X1

22 X3 X1

23 X3 X1

(2.16)

We now consider moment equilibrium (M = Fd). The stress is homogeneous, and the normal
force on the opposite side is equal opposite and colinear. The moment (X2 /2)31 X1 X2 is likewise
balanced by the moment of an equal component in the opposite face. Finally similar argument holds for
32 .
17

18

The net moment about the X3 axis is thus


M = X1 (12 X2 X3 ) X2 (21 X3 X1 )

(2.17)

which must be zero, hence 12 = 21 .


19 We generalize and conclude that in the absence of distributed body forces, the stress matrix is symmetric,

(2.18)

ij = ji

20

A more rigorous proof of the symmetry of the stress tensor will be given in Sect. 6.3.2.1.

2.3.1

Cauchys Reciprocal Theorem

21 If we consider t1 as the traction vector on a plane with normal n1 , and t2 the stress vector at the
same point on a plane with normal n2 , then

t1 = n1 and t2 = n2

(2.19)

{t1 } = n1 [] and {t2 } = n2 []

(2.20)

or in matrix form as
If we postmultiply the rst equation by n2 and the second one by n1 , by virtue of the symmetry of []
we have
(2.21)
[n1 ]n2 = [n2 ]n1
Victor Saouma

Introduction to Continuum Mechanics

Draft
26

KINETICS

n
tn

11111111111111111111111
00000000000000000000000
11111111111111111111111
00000000000000000000000
11111111111111111111111
00000000000000000000000
11111111111111111111111
00000000000000000000000
11111111111111111111111
00000000000000000000000
11111111111111111111111
00000000000000000000000
11111111111111111111111
00000000000000000000000
11111111111111111111111
00000000000000000000000
11111111111111111111111
00000000000000000000000
11111111111111111111111
00000000000000000000000
11111111111111111111111
00000000000000000000000
111
000
11
00
11111111111111111111111
00000000000000000000000
111
000
11
00
11111111111111111111111
00000000000000000000000
11
00
11
00
11111111111111111111111
00000000000000000000000
11
00
11111111111111111111111
00000000000000000000000
11
00
11
00
11111111111111111111111
00000000000000000000000
11
00
1111
0000

11
00
1111
0000
11
00
1111
0000
11
00
1111
0000
11
00
11
00
11
00
11
00
11
00

t -n

-n
Figure 2.4: Cauchys Reciprocal Theorem
or
t1 n2 = t2 n1

22

(2.22)

In the special case of two opposite faces, this reduces to


tn = tn

(2.23)

23 We should note that this theorem is analogous to Newtons famous third law of motion To every
action there is an equal and opposite reaction.

2.4

Principal Stresses

24 Regardless of the state of stress (as long as the stress tensor is symmetric), at a given point, it is
always possible to choose a special set of axis through the point so that the shear stress components
vanish when the stress components are referred to this system of axis. these special axes are called
principal axes of the principal stresses.

To determine the principal directions at any point, we consider n to be a unit vector in one of the
unknown directions. It has components ni . Let represent the principal-stress component on the plane
whose normal is n (note both n and are yet unknown). Since we know that there is no shear stress
component on the plane perpendicular to n,
the stress vector on this plane must be parallel to n and
25

tn = n
26

(2.24)

From Eq. 2.10 and denoting the stress tensor by we get


n = n

(2.25)

nr rs = ns

(2.26)

in indicial notation this can be rewritten as

Victor Saouma

Introduction to Continuum Mechanics

Draft

2.4 Principal Stresses

27

tn = t
12
n2

12
11

11= t

n1

Initial (X1) Plane

tn
s

s =0

n2

=tn
n

t n2

n1

t n2
n1

tn1

Arbitrary Plane

Principal Plane

Figure 2.5: Principal Stresses


or
(rs rs )nr = 0

(2.27)

n ([] [I]) = 0

(2.28)

in matrix notation this corresponds to


where I corresponds to the identity matrix. We really have here a set of three homogeneous algebraic
equations for the direction cosines ni .
27

Since the direction cosines must also satisfy


n2 + n2 + n2 = 1
1
2
3

(2.29)

they can not all be zero. hence Eq.2.28 has solutions which are not zero if and only if the determinant
of the coecients is equal to zero, i.e
11
21
31

12
22
32

13
23
33

(2.30)

|rs rs | =

(2.31)

| I| =

(2.32)

28 For a given set of the nine stress components, the preceding equation constitutes a cubic equation for
the three unknown magnitudes of .
29 Cauchy was rst to show that since the matrix is symmetric and has real elements, the roots are all
real numbers.
30 The three lambdas correspond to the three principal stresses (1) > (2) > (3) .
When any one
of them is substituted for in the three equations in Eq. 2.28 those equations reduce to only two
independent linear equations, which must be solved together with the quadratic Eq. 2.29 to determine
the direction cosines ni of the normal ni to the plane on which i acts.
r
31

The three directions form a right-handed system and


n3 = n1 n2

Victor Saouma

(2.33)

Introduction to Continuum Mechanics

Draft
28

32

KINETICS

In 2D, it can be shown that the principal stresses are given by:
1,2 =

2.4.1

x + y

x y
2

2
2
+ xy

(2.34)

Invariants

33 The principal stresses are physical quantities, whose values do not depend on the coordinate system
in which the components of the stress were initially given. They are therefore invariants of the stress
state.

34

When the determinant in the characteristic Eq. 2.32 is expanded, the cubic equation takes the form
3 I 2 II III = 0

(2.35)

where the symbols I , II and III denote the following scalar expressions in the stress components:
I
II

=
=
=
=

III

35

11 + 22 + 33 = ii = tr
2
2
2
(11 22 + 22 33 + 33 11 ) + 23 + 31 + 12
1
1
1 2
(ij ij ii jj ) = ij ij I
2
2
2
1
2
( : I )
2
1
det = eijk epqr ip jq kr
6

(2.38)
(2.39)
(2.40)

In terms of the principal stresses, those invariants can be simplied into


I
II

2.4.2

=
=

(1) + (2) + (3)


((1) (2) + (2) (3) + (3) (1) )

(2.41)
(2.42)

III

36

(2.36)
(2.37)

(1) (2) (3)

(2.43)

Spherical and Deviatoric Stress Tensors

If we let denote the mean normal stress p


= p =

1
1
1
(11 + 22 + 33 ) = ii = tr
3
3
3

(2.44)

then the stress tensor can be written as the sum of two tensors:
Hydrostatic stress in which each normal stress is equal to p and the shear stresses are zero. The
hydrostatic stress produces volume change without change in shape in an isotropic medium.

p 0
0
p 0
(2.45)
hyd = pI = 0
0
0
p
Deviatoric Stress: which causes the change in shape.

11 12
22
dev = 21
31
32
Victor Saouma

13

23
33

(2.46)

Introduction to Continuum Mechanics

Draft

2.5 Stress Transformation

2.5
37

29

Stress Transformation

From Eq. 1.73 and 1.74, the stress transformation for the second order stress tensor is given by
ip

aj aq jq in Matrix Form [] = [A]T [][A]


i p

(2.47)

jq
38

=
=

aj aq ip in Matrix Form [] = [A][][A]T


i p

(2.48)

For the 2D plane stress case we rewrite Eq. 1.76


sin2
cos2
xx
2
yy
=
sin
cos2

xy
sin cos cos sin

2 sin cos
xx
yy
2 sin cos

xy
cos2 sin2

(2.49)

Example 2-2: Principal Stresses


The stress tensor is given at a point by

3
= 1
1

1
0
2

1
2
0

determine the principal stress values and the corresponding directions.


Solution:
From Eq.2.32 we have
3
1
1
1
0
2
=0
1
2
0

(2.50)

(2.51)

Or upon expansion (and simplication) ( + 2)( 4)( 1) = 0, thus the roots are (1) = 4, (2) = 1
and (3) = 2. We also note that those are the three eigenvalues of the stress tensor.
If we let x1 axis be the one corresponding to the direction of (3) and n3 be the direction cosines of
i
this axis, then from Eq. 2.28 we have

(3 + 2)n3 + n3 + n3 = 0
1
2
3
1
1
n3 + 2n3 + 2n3 = 0 n3 = 0;
n3 =
(2.52)
n3 = ;
1
2
3
1
2
3

2
2
n3 + 2n3 + 2n3 = 0
1
2
3
Similarly If we let x2 axis be the one corresponding to the direction of (2) and n2 be the direction
i
cosines of this axis,

2n2 + n2 + n2 = 0
1
2
3
1
1
1
n2 n2 + 2n2 = 0 n2 = ;
n2 = ;
n2 =
(2.53)
1
2
3
1
2
3
2
3
3
3
n1 + 2n2 n2 = 0
2
3
Finally, if we let x3 axis be the one corresponding to the direction of (1) and n1 be the direction cosines
i
of this axis,

n1 + n1 + n1 = 0
1
2
3
2
1
1
n1 4n1 + 2n1 = 0 n1 = ;
n1 = ;
n1 =
(2.54)
1
2
3
1
2
3
1
6
6
6
1
1
n1 + 2n2 4n3 = 0
Finally, we can convince ourselves that the two stress tensors have the same invariants I , II and III .

Victor Saouma

Introduction to Continuum Mechanics

Draft
210

KINETICS

Example 2-3: Stress Transformation


Show that the transformation tensor of direction cosines previously determined transforms the original stress tensor into the diagonal principal axes stress tensor.
Solution:
From Eq. 2.47

0
1

3
2
6

2
= 0
0

2.5.1
39

2
1
3
1
6

0 0
1 0
0 4

1
2
3
1
3 1

1
1
6

1
0
2

0
1
1
2 2
1
2
2

3
1
3
1
3

2
6
1
6

1
6

(2.55-a)

(2.55-b)

Plane Stress

Plane stress conditions prevail when 3i = 0, and thus we have a biaxial stress eld.

40 Plane stress condition prevail in (relatively) thin plates, i.e when one of the dimensions is much smaller
than the other two.

2.5.2

Mohrs Circle for Plane Stress Conditions

41 The Mohr circle will provide a graphical mean to contain the transformed state of stress ( xx , yy , xy )
at an arbitrary plane (inclined by ) in terms of the original one (xx , yy , xy ).
42

Substituting
cos2
cos 2

=
=

1+cos 2
2
2

cos sin
2

sin2
sin 2

= 1cos 2
2
= 2 sin cos

(2.56)

into Eq. 2.49 and after some algebraic manipulation we obtain


xx
xy

1
1
(xx + yy ) + (xx yy ) cos 2 + xy sin 2
2
2
1
= xy cos 2 (xx yy ) sin 2
2
=

(2.57-a)
(2.57-b)

Points (xx , xy ), (xx , 0), (yy , 0) and [(xx +yy )/2, 0] are plotted in the stress representation of Fig.
2.6. Then we observe that
1
(xx yy ) = R cos 2
(2.58-a)
2
(2.58-b)
xy = R sin 2

43

where
R
tan 2
Victor Saouma

=
=

1
2
(xx yy )2 + xy
4
2xy
xx yy

(2.59-a)
(2.59-b)

Introduction to Continuum Mechanics

Draft

2.5 Stress Transformation

211

yy

yx

y
yy

yx

xy

xx

xy

xx

xy

xx

xx

yx

yx

xy

yy

yy

(b)

(a)

n
xy

xx

X( xx xy )
,

xy

xx

x
O

yy

xy

X( xx xy )
,

2 2
2

xx

D
yx

yx
1 ( + )
xx yy
2

yy

1 ( + )
1
2
2

1 ( - )
xx yy
2

1( - )
1
2
2

(c)

(d)

Figure 2.6: Mohr Circle for Plane Stress

Victor Saouma

Introduction to Continuum Mechanics

Draft
212

KINETICS

then after substitution and simpliation, Eq. 2.57-a and 2.57-b would result in
xx

xy

1
(xx + yy ) + R cos(2 2)
2
R sin(2 2)

(2.60)
(2.61)

We observe that the form of these equations, indicates that xx and xy are on a circle centered at
1
2 (xx + yy ) and of radius R. Furthermore, since xx , yy , R and are denite numbers for a given
state of stress, the previous equations provide a graphical solution for the evaluation of the rotated
stress xx and xy for various angles .
44

By eliminating the trigonometric terms, the Cartesian equation of the circle is given by
1
[ xx (xx + yy )]2 + 2 = R2
xy
2

45

(2.62)

Finally, the graphical solution for the state of stresses at an inclined plane is summarized as follows
1. Plot the points (xx , 0), (yy , 0), C : [ 1 (xx + yy ), 0], and X : (xx , xy ).
2
2. Draw the line CX, this will be the reference line corresponding to a plane in the physical body
whose normal is the positive x direction.
3. Draw a circle with center C and radius R = CX.
4. To determine the point that represents any plane in the physical body with normal making a
counterclockwise angle with the x direction, lay o angle 2 clockwise from CX. The terminal
side CX of this angle intersects the circle in point X whose coordinates are ( xx , xy ).
5. To determine yy , consider the plane whose normal makes an angle + 1 with the positive x axis
2
in the physical plane. The corresponding angle on the circle is 2 + measured clockwise from
the reference line CX. This locates point D which is at the opposite end of the diameter through
X. The coordinates of D are ( yy , xy )

Example 2-4: Mohrs Circle in Plane Stress


An element in plane stress is subjected to stresses xx = 15, yy = 5 and xy = 4. Using the
Mohrs circle determine: a) the stresses acting on an element rotated through an angle = +40o
(counterclockwise); b) the principal stresses; and c) the maximum shear stresses. Show all results on
sketches of properly oriented elements.
Solution:
With reference to Fig. 2.7:
1. The center of the circle is located at
1
1
(xx + yy ) = (15 + 5) = 10.
2
2

(2.63)

2. The radius and the angle 2 are given by


R
tan 2

Victor Saouma

=
=

1
(15 5)2 + 42 = 6.403
4
2(4)
= 0.8 2 = 38.66o ;
15 5

(2.64-a)
= 19.33o

(2.64-b)

Introduction to Continuum Mechanics

Draft

2.5 Stress Transformation

213
n
=25.7

6.4

X(15,4) =0

15

=109.3

38.66

=19.3

15

15

80

41.34

4
4

=40

=90

=64.3

10

10.00

3.6
5.19

14.81
o

40

16.4
o

19.3

6.40

4.23

25.7
10.00

Figure 2.7: Plane Stress Mohrs Circle; Numerical Example


3. The stresses acting on a plane at = +40o are given by the point making an angle of 80o
(clockwise) with respect to point X(15, 4) or 80o + 38.66o = 41.34o with respect to the axis.
4. Thus, by inspection the stresses on the x face are
xx
xy

10 + 6.403 cos 41.34o = 14.81

(2.65-a)

6.403 sin 41.34 = 4.23

(2.65-b)

5. Similarly, the stresses at the face y are given by


yy
xy

10 + 6.403 cos(180o 41.34o ) = 5.19

(2.66-a)

6.403 sin(180 41.34 ) = 4.23

(2.66-b)

6. The principal stresses are simply given by


(1)

10 + 6.4 = 16.4

(2.67-a)

(2)

10 6.4 = 3.6

(2.67-b)

(1) acts on a plane dened by the angle of +19.3o clockwise from the x axis, and (2) acts at an
o
o
angle of 38.66 2+180 = 109.3o with respect to the x axis.
7. The maximum and minimum shear stresses are equal to the radius of the circle, i.e 6.4 at an angle
of
90o 38.66o
= 25.70
(2.68)
2

Victor Saouma

Introduction to Continuum Mechanics

Draft
214

2.5.3

KINETICS

Mohrs Stress Representation Plane

46 There can be an innite number of planes passing through a point O, each characterized by their own
normal vector along ON , Fig. 2.8. To each plane will correspond a set of n and n .

Y II
B
E

J
C

Z III
Figure 2.8: Unit Sphere in Physical Body around O
47 It can be shown that all possible sets of n and n which can act on the point O are within the shaded
area of Fig. 2.9.

2.6

Simplied Theories; Stress Resultants

48 For many applications of continuum mechanics the problem of determining the three-dimensional
stress distribution is too dicult to solve. However, in many (civil/mechanical)applications, one or more
dimensions is/are small compared to the others and possess certain symmetries of geometrical shape and
load distribution.
49 In those cases, we may apply engineering theories for shells, plates or beams. In those problems,
instead of solving for the stress components throughout the body, we solve for certain stress resultants
(normal, shear forces, and Moments and torsions) resulting from an integration over the body. We
consider separately two of those three cases.
50 Alternatively, if a continuum solution is desired, and engineering theories prove to be either too
restrictive or inapplicable, we can use numerical techniques (such as the Finite Element Method) to
solve the problem.

2.6.1

Arch

51 Fig. 2.10 illustrates the stresses acting on a dierential element of a shell structure. The resulting
forces in turn are shown in Fig. 2.11 and for simplication those acting per unit length of the middle
surface are shown in Fig. 2.12. The net resultant forces are given by:

Victor Saouma

Introduction to Continuum Mechanics

Draft

2.6 Simplied Theories; Stress Resultants

215

1 ( - )

2

1 ( - )

2

III

O
C

CII

1 ( - )

2

II

C III

1 ( + )

2
1 ( + )

2
1

Figure 2.9: Mohr Circle for Stress in 3D

Figure 2.10: Dierential Shell Element, Stresses

Victor Saouma

Introduction to Continuum Mechanics

Draft
216

KINETICS

Figure 2.11: Dierential Shell Element, Forces

Figure 2.12: Dierential Shell Element, Vectors of Stress Couples

Victor Saouma

Introduction to Continuum Mechanics

Draft

2.6 Simplied Theories; Stress Resultants


Membrane Force

+h
2

z
dz
1
h
r
2

Bending Moments

+h
2

M =
h
2

z 1

z
dz
r

Transverse Shear Forces


+h
2

2.6.2
52

z
dz
1
r
h
2

217

xx

Nyy

Nxy

N
yx

Mxx

Myy

Mxy

M
yx

+h
2

=
h
2
+h
2

=
h
2
+h
2

=
h
2
+h
2

=
h
2
+h
2

yy 1

z
rx

dz

xy 1

z
ry

dz

xy 1

z
rx

dz

z
yy z 1
rx
h
2

dz

xx z 1

+h
2

h
2

+h
2

=
h
2

Qy

dz

dz

h
2
+h
2

Qx

z
ry

z
ry

xx 1

+h
2
h
2
+h
2
h
2

xy z 1

xy z 1

z
ry

z
rx

xz 1

z
ry
z
rx

dz
dz

dz

yz 1

(2.69)

dz

Plates

Considering an arbitrary plate, the stresses and resulting forces are shown in Fig. 2.13, and resultants

Figure 2.13: Stresses and Resulting Forces in a Plate

Victor Saouma

Introduction to Continuum Mechanics

Draft
218

KINETICS

per unit width are given by

t
2

Membrane Force

N =
t
2

t
2

Bending Moments

M =
t
2

t
2

Transverse Shear Forces

=
t
2

dz

zdz

dz

Nxx =

=
N
yy

xy =

xx =

Myy =

M
xy =

Vx =

V
y

t
2
t
2

xx dz

t
2

t
2

yy dz

t
2

t
2

xy dz

t
2

t
2

xx zdz
(2.70-a)

t
2

t
2

yy zdz

t
2

t
2

t
2

t
2

xy zdz

xz dz

t
2

=
t
2

yz dz

53 Note that in plate theory, we ignore the eect of the membrane forces, those in turn will be accounted
for in shells.

Victor Saouma

Introduction to Continuum Mechanics

Draft
Chapter 3

MATHEMATICAL
PRELIMINARIES; Part II
VECTOR DIFFERENTIATION
3.1

Introduction

A eld is a function dened over a continuous region. This includes, Scalar Field g(x), Vector
Field v(x), Fig. 3.1 or Tensor Field T(x).

We rst introduce the dierential vector operator Nabla denoted by

i+
j+
k
x
y
z

(3.1)

We also note that there are as many ways to dierentiate a vector eld as there are ways of multiplying
vectors, the analogy being given by Table 3.1.

Multiplication
uv
dot
uv
cross
u v tensor

Dierentiation
v
divergence
v curl
v
gradient

Tensor Order
?
6

Table 3.1: Similarities Between Multiplication and Dierentiation Operators

3.2
4

Derivative WRT to a Scalar

The derivative of a vector p(u) with respect to a scalar u, Fig. 3.2 is dened by
dp
p(u + u) p(u)
lim
du u0
u

(3.2)

If p(u) is a position vector p(u) = x(u)i + y(u)j + z(u)k, then


dx
dy
dz
dp
=
i+
j+
k
du
du
du
du

(3.3)

Draft
32

MATHEMATICAL PRELIMINARIES; Part II VECTOR DIFFERENTIATION


mfields.nb

Scalar and Vector Fields


ContourPlot@Exp@Hx ^ 2 + y ^ 2LD, 8x, 2, 2<, 8y, 2, 2<, ContourShading > FalseD

-1

-2
-2

-1

ContourGraphics
Plot3D@Exp@Hx ^ 2 + y ^ 2LD, 8x, 2, 2<, 8y, 2, 2<, FaceGrids > AllD

1
0.75
0.5
0.25
0
-2
2

2
1
0
-1
-1

0
1
2 -2
SurfaceGraphics

Figure 3.1: Examples of a Scalar and Vector Fields


p=p (u+ u)- p(u)
C

)
u
(u+
p

p (u)

Figure 3.2: Dierentiation of position vector p


Victor Saouma

Introduction to Continuum Mechanics

Draft

3.2 Derivative WRT to a Scalar

33

is a vector along the tangent to the curve.


6

If u is the time t, then

dp
dt

is the velocity

In dierential geometry, if we consider a curve C dened by the function p(u) then dp is a vector
du
tangent ot C, and if u is the curvilinear coordinate s measured from any point along the curve, then dp
ds
is a unit tangent vector to C T, Fig. 3.3. and we have the following relations
7

Figure 3.3: Curvature of a Curve


dp
= T
ds
dT
= N
ds
B = TN

we also note that p dp = 0 if


ds

dp
ds

(3.4)
(3.5)
(3.6)

curvature
(3.7)
1
Radius of Curvature (3.8)
=

= 0.

Example 3-1: Tangent to a Curve


Determine the unit vector tangent to the curve: x = t2 + 1, y = 4t 3, z = 2t2 6t for t = 2.
Solution:

dp
dt
dp
dt

=
=

T =
=

d
(t2 + 1)i + (4t 3)j + (2t2 6t)k = 2ti + 4j + (4t 6)k
dt
(2t)2 + (4)2 + (4t 6)2

(3.9-a)
(3.9-b)

2ti + 4j + (4t 6)k


(2t)2 + (4)2 + (4t 6)2
2
1
2
4i + 4j + 2k
= i + j + k for t = 2
2 + (4)2 + (2)2
3
3
3
(4)

(3.9-c)
(3.9-d)

Mathematica solution is shown in Fig. 3.4

Victor Saouma

Introduction to Continuum Mechanics

Draft
34

MATHEMATICAL PRELIMINARIES; Part II VECTOR DIFFERENTIATION


mpar3d.nb

Parametric Plot in 3D
ParametricPlot3D@8t ^ 2 + 1, 4 t 3, 2 t ^ 2 6 t<, 8t, 0, 4<D

10
5
0

0
5
10
15
Graphics3D

Figure 3.4: Mathematica Solution for the Tangent to a Curve in 3D

3.3
3.3.1

Divergence
Vector

The divergence of a vector eld of a body B with boundary , Fig. 3.5 is dened by considering
that each point of the surface has a normal n, and that the body is surrounded by a vector eld v(x).
The volume of the body is v(B).
8

v(x)

Figure 3.5: Vector Field Crossing a Solid Region


9

The divergence of the vector eld is thus dened as


div v(x) lim

v(B)0

1
v(B)

vndA

(3.10)

where v.n is often referred as the ux and represents the total volume of uid that passes through
dA in unit time, Fig. 3.6 This volume is then equal to the base of the cylinder dA times the height of
Victor Saouma

Introduction to Continuum Mechanics

Draft

3.3 Divergence

35

n
dA
v
v.n

Figure 3.6: Flux Through Area dA


the cylinder vn. We note that the streamlines which are tangent to the boundary do not let any uid
out, while those normal to it let it out most eciently.
10

The divergence thus measure the rate of change of a vector eld.

11 The denition is clearly independent of the shape of the solid region, however we can gain an insight
into the divergence by considering a rectangular parallelepiped with sides x1 , x2 , and x3 , and with
normal vectors pointing in the directions of the coordinate axies, Fig. 3.7. If we also consider the corner

x3
-e

e3
-e

x3
e2

e1

x2
-e

x2

x1

x1
Figure 3.7: Innitesimal Element for the Evaluation of the Divergence
closest to the origin as located at x, then the contribution (from Eq. 3.10) of the two surfaces with
normal vectors e1 and e1 is
1
x1 ,x2 ,x3 0 x1 x2 x3
lim

[v(x + x1 e1 )e1 + v(x)(e1 )]dx2 dx3

(3.11)

x2 x3

or
lim

x1 ,x2 ,x3 0

1
x2 x3

v(x + x1 e1 ) v(x)
e1 dx2 dx3
x1
x2 x3

=
=

lim

x1 0

v
e1 (3.12-a)
x1

v
e1
x1

(3.12-b)

hence, we can generalize


div v(x) =

Victor Saouma

v(x)
ei
xi

(3.13)

Introduction to Continuum Mechanics

Draft
36

12

MATHEMATICAL PRELIMINARIES; Part II VECTOR DIFFERENTIATION

or alternatively
div v

=
=

e1 +
e2 +
e3 )(v1 e1 + v2 e2 + v3 e3 ) (3.14)
x1
x2
x3
v1
v2
v3
vi
+
+
=
= i vi = vi,i
(3.15)
x1
x2
x3
xi
v = (

13

The divergence of a vector is a scalar.

14

We note that the Laplacian Operator is dened as


2

(3.16)

F = F,ii

Example 3-2: Divergence


Determine the divergence of the vector A = x2 zi 2y 3 z 2 j + xy 2 zk at point (1, 1, 1).
Solution:
v

i+
j+
k (x2 zi 2y 3 z 2 j + xy 2 zk)
x
y
z
2y 3 z 2
xy 2 z
x2 z
+
+
=
x
y
z
2 2
2
= 2xz 6y z + xy
= 2(1)(1) 6(1)2 (1)2 + (1)(1)2 = 3 at (1, 1, 1)
=

(3.17-a)
(3.17-b)
(3.17-c)
(3.17-d)

Mathematica solution is shown in Fig. 3.8

3.3.2
15

Second-Order Tensor

By analogy to Eq. 3.10, the divergence of a second-order tensor eld T is


T = div T(x) lim

v(B)0

1
v(B)

TndA

(3.18)

which is the vector eld


T =

3.4
3.4.1

Tpq
eq
xp

(3.19)

Gradient
Scalar

16 The gradient of a scalar eld g(x) is a vector eld


g(x) such that for any unit vector v, the
directional derivative dg/ds in the direction of v is given by

dg
=
ds
Victor Saouma

gv

(3.20)
Introduction to Continuum Mechanics

Draft
3.4 Gradient

37

mdiver.nb

Divergence of a Vector
<< CalculusVectorAnalysis
V = 8x ^ 2 z, 2 y ^ 3 z ^ 2, x y ^ 2 z<;
Div@V, Cartesian@x, y, zDD
-6 z2 y2 + x y2 + 2 x z
<< GraphicsPlotField3D
PlotVectorField3D@8x ^ 2 z, 2 y ^ 3 z ^ 2, x y ^ 2 z<, 8x, 10, 10<, 8y, 10, 10<, 8z, 10, 10<,
Axes > Automatic, AxesLabel > 8"X", "Y", "Z"<D

10
5

0
-5
-10
10
10

5
Z

0
-5
-10
-10
-5
0
X

5
10

Graphics3D
Div@Curl@V, Cartesian@x, y, zDD, Cartesian@x, y, zDD
0

Figure 3.8: Mathematica Solution for the Divergence of a Vector

where v = dp We note that the denition made no reference to any coordinate system. The gradient is
ds
thus a vector invariant.
17

To nd the components in any rectangular Cartesian coordinate system we use


v

dg
ds

dxi
dp
=
ei
ds
ds
g dxi
xi ds

(3.21-a)
(3.21-b)

which can be substituted and will yield


g=

g
ei
xi

(3.22)

or

=

i+
j+
k
x
y
z

i+
j+
k
x
y
z

(3.23-a)
(3.23-b)

and note that it denes a vector eld.


18 The physical signicance of the gradient of a scalar eld is that it points in the direction in which the
eld is changing most rapidly (for a three dimensional surface, the gradient is pointing along the normal
to the plane tangent to the surface). The length of the vector || g(x)|| is perpendicular to the contour
lines.

Victor Saouma

Introduction to Continuum Mechanics

Draft
38

19

MATHEMATICAL PRELIMINARIES; Part II VECTOR DIFFERENTIATION

g(x)n gives the rate of change of the scalar eld in the direction of n.

Example 3-3: Gradient of a Scalar


Determine the gradient of = x2 yz + 4xz 2 at point (1, 2, 1) along the direction 2i j 2k.
Solution:
=
(x2 yz + 4xz 2 ) = (2xyz + 4z 2 )i + (x2 zj + (x2 y + 8xz)k
= 8i j 10k at (1, 2, 1)
1
2
2
2i j 2k
= i j k
n =
2 + (1)2 + (2)2
3
3
3
(2)

(8i j 10k)

1
2
2
i j k
3
3
3

16 1 20
37
+ +
=
3
3
3
3

(3.24-a)
(3.24-b)
(3.24-c)
(3.24-d)

Since this last value is positive, increases along that direction.

Example 3-4: Stress Vector normal to the Tangent of a Cylinder


The stress tensor throughout a continuum is given with respect to Cartesian axes as

3x1 x2 5x2
0
2
0
2x2
= 5x2
2
3
0
2x3
0

(3.25)

Determine the stress vector (or traction) at the point P (2, 1, 3) of the plane that is tangent to the
cylindrical surface x2 + x2 = 4 at P , Fig. 3.9.
2
3
x3
n

x2
P

3
1

x1

Figure 3.9: Radial Stress vector in a Cylinder

Victor Saouma

Introduction to Continuum Mechanics

Draft
3.4 Gradient

39

Solution:
At point P , the stress tensor is given by

6
5
0
= 5
0 2 3

2 3
0

(3.26)

The unit normal to the surface at P is given from


(x2 + x2 4) = 2x2 22 + 2x3 e3
2
3

(3.27)

(x2 + x2 4) = 222 + 2 3e3


2
3

(3.28)

At point P ,

1
3
e3
n = e1 +
2
2
Thus the traction vector will be determined from

6
5
0
0 5/2

3
1/2
0
=
= 5
2 3

3/2
3
0
0 2 3
and thus the unit normal at P is

or tn = 5 e1 + 3e2 +
2

3.4.2

(3.29)

(3.30)

3e3

Vector

We can also dene the gradient of a vector eld. If we consider a solid domain B with boundary ,
Fig. 3.5, then the gradient of the vector eld v(x) is a second order tensor dened by

20

x v(x)

lim

v(B)0

1
v(B)

v ndA

(3.31)

and with a construction similar to the one used for the divergence, it can be shown that
x v(x)

vi (x)
[ei ej ]
xj

(3.32)

where summation is implied for both i and j.


21 The components of
x v are simply the various partial derivatives of the component functions with
respect to the coordinates:

x v]

[v

Victor Saouma

x]

vx
x
vx
y
vx
z
vx
x
vy
x
vz
x

x
vy
y
vy
z
vx
y
vy
y
vz
y

vz
x
vz
y
vz
z
vx
z
vy
z
vz
z

(3.33)

(3.34)

Introduction to Continuum Mechanics

Draft
310

MATHEMATICAL PRELIMINARIES; Part II VECTOR DIFFERENTIATION

that is [ v]ij gives the rate of change of the ith component of v with respect to the jth coordinate axis.
22 Note the diference between v
other.

23

and

x v.

In matrix representation, one is the transpose of the

The gradient of a vector is a tensor of order 2.

24 We can interpret the gradient of a vector geometrically, Fig. 3.10. If we consider two points a and b
that are near to each other (i.e s is very small), and let the unit vector m points in the direction from a
to b. The value of the vector eld at a is v(x) and the value of the vector eld at b is v(x + sm). Since
the vector eld changes with position in the domain, those two vectors are dierent both in length and
orientation. If we now transport a copy of v(x) and place it at b, then we compare the dierences between
those two vectors. The vector connecting the heads of v(x) and v(x + sm) is v(x + sm) v(x), the
change in vector. Thus, if we divide this change by s, then we get the rate of change as we move in
the specied direction. Finally, taking the limit as s goes to zero, we obtain

lim

s0

v(x + sm) v(x)


Dv(x)m
s

(3.35)

v(x+ s m ) -v(x)
v(x+ s m )
x3

v(x)
a sm b
x2

x1
Figure 3.10: Gradient of a Vector
The quantity Dv(x)m is called the directional derivative because it gives the rate of change of
the vector eld as we move in the direction m.

Example 3-5: Gradient of a Vector Field


Determine the gradient of the following vector eld v(x) = x1 x2 x3 (x1 e1 + x2 e2 + x3 e3 ).
Solution:

x v(x)

2x1 x2 x3 [e1 e1 ] + x2 x3 [e1 e2 ] + x2 x2 [e1 e3 ]


1
1
+x2 x3 [e2 e1 ] + 2x1 x2 x3 [e2 e2 ] + x1 x2 [e2 e3 ]
2
2

e1 ] +
e2 ] + 2x1 x2 x3 [e3 e3 ]

2
x1 /x2 x1 /x3
2
x2 /x3
= x1 x2 x3 x2 /x1
x3 /x1 x3 /x2
2
+x2 x2 [e3
3

Victor Saouma

(3.36-a)

x1 x2 [e3
3

(3.36-b)

Introduction to Continuum Mechanics

Draft
3.5 Curl

3.4.3
25

311

Mathematica Solution

Mathematica solution of the two preceding examples is shown in Fig. 3.11.


mgrad.nb

2
1

PlotVectorField3D@vecfield, 8x1, -10, 10<, 8x2, -10, 10<, 8x3, -10, 10<, Axes -> Automatic,
AxesLabel -> 8"x1", "x2", "x3"<D

Gradient

x2

10

Scalar
-10

f = x ^ 2 y z + 4 x z ^ 2;

10
Gradf = Grad@f, Cartesian@x, y, zDD
8 x3+ 2 x y z, x2 z, y x2 + 8 z x<
4 z2

<< GraphicsPlotField3D
PlotGradientField3D@f, 8x, 0, 2<, 8y, -3, -1<, 8z, -2, 0<D
-10

-10
0
x1

10

Graphics3D
MatrixForm@Grad@vecfield, Cartesian@x1, x2, x3DDD
i2 x1 x2 x3 x12 x3
x12 x2 y
j
z
j
z
j
j
z
j x22 x3 2 x1 x2 x3 x1 x22 z
z
j
z
j
z
j
z
j
z
j
z
j
z
2
2
x1 x3
2 x1 x2 x3 {
k x2 x3

Graphics3D
x = 1; y = -2; z = -1;
vect = 82, -1, -2< Sqrt@4 + 1 + 4D
2
1
2
9 ,- ,- =
3
3
3
Gradf . vect
37
3

Gradient of a Vector
vecfield = x1 x2 x3 8x1, x2, x3<

Figure 3.11: Mathematica Solution for the Gradients of a Scalar and of a Vector
8 2 x2 x3, x1 x22 x3, x1 x2 x32 <
x1

3.5

Curl

When the vector operator


operates in a manner analogous to vector multiplication, the result is a
vector, curl v called the curl of the vector eld v (sometimes called the rotation).

26

curl v

v =

e1

x1
v1

=
=

Victor Saouma

v2
v3

x2
x3
eijk j vk

e2

x2
v2

e1 +

e3

x3

(3.37)

v3

v3
v1

x3
x1

e2 +

v1
v2

x1
x2

e3

(3.38)
(3.39)

Introduction to Continuum Mechanics

Draft
312

MATHEMATICAL PRELIMINARIES; Part II VECTOR DIFFERENTIATION

Example 3-6: Curl of a vector


Determine the curl of the following vector A = xz 3 i 2x2 yzj + 2yz 4 k at (1, 1, 1).
Solution:

i+
j+
k (xz 3 i 2x2 yzj + 2yz 4 k)
x
y
z

A =

i
=

x
3

y
2

2x yz

xz

(3.40-b)

2yz 4

2x2 yz
2yz 4

y
z

(3.40-a)

i+

2yz 4
xz 3

z
x

xz 3
2x2 yz

x
y

k (3.40-c)

(2z 4 + 2x2 y)i + 3xz 2 j 4xyzk

j+

3j + 4k at (1, 1, 1)

(3.40-d)
(3.40-e)

Mathematica solution is shown in Fig. 3.12.


mcurl.nb

Curl
<< CalculusVectorAnalysis
A = 8x z ^ 3, 2 x ^ 2 y z, 2 y z ^ 4<;
CurlOfA = Curl@A, Cartesian@x, y, zDD
82 z4 + 2 x2 y, 3 x z2 , -4 x y z<
<< GraphicsPlotField3D
PlotVectorField3D@CurlOfA, 8x, 0, 2<, 8y, 2, 0<, 8z, 0, 2<, Axes > Automatic, AxesLabel > 8"x", "y", "z"<D

0
y
-0.5
-1
-1.5
-2
2
2
1.5
z
1
0.5
0
0
0.5
1
1.5
x

Graphics3D
Div@CurlOfA, Cartesian@x, y, zDD
0
x = 1; y = 1; z = 1;
CurlOfA
80, 3, 4<

Figure 3.12: Mathematica Solution for the Curl of a Vector

Victor Saouma

Introduction to Continuum Mechanics

Draft

3.6 Some useful Relations

3.6
27

313

Some useful Relations

Some useful relations


d(AB)
d(AB)

= AdB + dAB

(3.41-a)

= AdB + dAB

(3.41-b)

( + )

(A + B)

A +

v
(A)

(3.41-c)
B

(3.41-d)

= v
= ( )A + ( A)

(3.41-e)
(3.41-f)

(AB) = B( A) A( B)
(AB) = (B )A + (A )B + B( A) + A( B)
2 2 2
2

+ 2 + 2 Laplacian Operator
( )
x2
y
z
( A) = 0
( )

Victor Saouma

= 0

(3.41-g)
(3.41-h)
(3.41-i)
(3.41-j)
(3.41-k)

Introduction to Continuum Mechanics

Draft
314

MATHEMATICAL PRELIMINARIES; Part II VECTOR DIFFERENTIATION

Victor Saouma

Introduction to Continuum Mechanics

Draft
Chapter 4

KINEMATIC
Or on How Bodies Deform

4.1

Elementary Denition of Strain

We begin our detailed coverage of strain by a simplied and elementary set of denitions for the 1D
and 2D cases. Following this a mathematically rigorous derivation of the various expressions for strain
will follow.

4.1.1

Small and Finite Strains in 1D

2 We begin by considering an elementary case, an axial rod with initial lenght l0 , and subjected to a
deformation l into a nal deformed length of l, Fig. 4.1.

l0

l
Figure 4.1: Elongation of an Axial Rod
3 We seek to quantify the deformation of the rod and even though we only have 2 variables (l0 and l),
there are dierent possibilities to introduce the notion of strain. We rst dene the stretch of the rod
as
l
(4.1)

l0

This stretch is one in the undeformed case, and greater than one when the rod is elongated.
4

Using l0 , l and we next introduce four possible denitions of the strain in 1D:

Draft
42

KINEMATIC

Engineering Strain

Natural Strain

Lagrangian Strain

Eulerian Strain E

l l0
=1
l0
1
l l0
=1

2
1 l 2 l0

=
2
2
l0
2
1 l 2 l0
=

2
l2

(4.2)
(4.3)
1 2
( 1)
2
1
1
1 2
2

(4.4)
(4.5)

we note the strong analogy between the Lagrangian and the engineering strain on the one hand, and the
Eulerian and the natural strain on the other.
The choice of which strain denition to use is related to the stress-strain relation (or constitutive law)
that we will later adopt.

4.1.2
6

Small Strains in 2D

The elementary denition of strains in 2D is illustrated by Fig. 4.2 and are given by
xx

yy

xy

xy

ux
X
uy
Y
uy

ux
= 2 + 1 =
+
2
Y
X
uy
1
1 ux
xy
+
2
2 Y
X

(4.6-a)
(4.6-b)
(4.6-c)
(4.6-d)

In the limit as both X and Y approach zero, then


xx =

ux
;
X

yy =

uy
;
Y

xy =

1
1
xy =
2
2

uy
ux
+
Y
X

(4.7)

We note that in the expression of the shear strain, we used tan which is applicable as long as is
small compared to one radian.
We have used capital letters to represent the coordinates in the initial state, and lower case letters
for the nal or current position coordinates (x = X + ux ). This corresponds to the Lagrangian strain
representation.
7

4.2

Strain Tensor

Following the simplied (and restrictive) introduction to strain, we now turn our attention to a rigorous
presentation of this important deformation tensor.

The approach we will take in this section is as follows:


1. Dene Material (xed, Xi ) and Spatial (moving, xi ) coordinate systems.
2. Introduce the notion of a position and of a displacement vector, U, u, (with respect to either
coordinate system).
3. Introduce Lagrangian and Eulerian descriptions.
4. Introduce the notion of a material deformation gradient and spatial deformation gradient

Victor Saouma

Introduction to Continuum Mechanics

Draft

Introduction to Continuum Mechanics

Victor Saouma

1 u y

43

ux

Pure Shear Without Rotation

4.2 Strain Tensor

Draft
44

KINEMATIC

5. Introduce the notion of a material displacement gradient and spatial displacement gradient.
6. Dene Cauchys and Greens deformation tensors (in terms of (dX)2 and (dx)2 respectively.
7. Introduce the notion of strain tensor in terms of (dx)2 (dX)2 as a measure of deformation
in terms of either spatial coordinates or in terms of displacements.

4.2.1

Position and Displacement Vectors; (x, X)

We consider in Fig. 4.3 the undeformed conguration of a material continuum at time t = 0 together
with the deformed conguration at coordinates for each conguration.

10

11

In the initial conguration P0 has the position vector


X = X 1 I1 + X2 I2 + X3 I3

(4.8)

which is here expressed in terms of the material coordinates (X1 , X2 , X3 ).


12 In the deformed conguration, the particle P0 has now moved to the new position P and has the
following position vector
(4.9)
x = x1 i1 + x2 i2 + x3 i3

which is expressed in terms of the spatial coordinates.


Note certain similarity with Fig. 4.1, and Eq. 4.2-4.5 where the strains are dened in terms of l and
l0 rather than the displacement l.

13

14 The relative orientation of the material axes (OX1 X2 X3 ) and the spatial axes (ox1 x2 x3 ) is specied
through the direction cosines aX .
x
15 The displacement vector u connecting P0 and P is the displacement vector which can be expressed
in both the material or spatial coordinates

U = UK IK
u = u k ik

(4.10-a)
(4.10-b)

again Uk and uk are interrelated through the direction cosines ik = aK IK . Substituting above we
k
obtain
(4.11)
u = uk (aK IK ) = UK IK = U UK = aK uk
k
k
16 The vector b relates the origin o with respect to O. From geometry X+u = b+x, thus u = b+xX
or if the origins are the same (superimposed axis), Fig. 4.4:

u = xX
uk = xk aK XK
k

(4.12-a)
(4.12-b)

kK

uk

xk Xk

(4.12-c)

Example 4-1: Displacement Vectors in Material and Spatial Forms

Victor Saouma

Introduction to Continuum Mechanics

Draft

45

Spatial

X1

I3
O

Material

I2

X3

t=0

P0

x3

X2

x1

i3
o

i1

t=t

i2

x2

4.2 Strain Tensor

Figure 4.3: Position and Displacement Vectors

Victor Saouma

Introduction to Continuum Mechanics

Draft
46

KINEMATIC
t=t
X

,x

t=0
U
P0
I 3 , 3i

Oo
I

I 2 , 2i

X2

,i

,x
2

Material/Spatial

X1

u
x

b=0

,x
1

Figure 4.4: Position and Displacement Vectors, b = 0


With respect to superposed material axis Xi and spatial axes xi , the displacement eld of a continuum
body is given by: x1 = X1 , x2 = X2 + AX3 , and x3 = AX2 + X3 where A is constant.
1. Determine the displacement vector components, u, in both the material and spatial form.
2. Determine the displacements of the edges of a cube with edges along the coordinate axes of length
dXi = dX, and sketch the displaced conguration for A = 1/2.
3. Determine the displaced location of material particles which originally comprises the plane circular
2
2
surface X1 = 0, X2 + X3 = 1/(1 A2 ) if A = 1/2.
Solution:
1. From Eq. 4.12-c the displacement eld can be written in material coordinates as
u1
u2
u3

= x1 X1 = 0
= x2 X2 = AX3
= x3 X3 = AX2

2. The position vector can be written in matrix form as


1 0 0 X1
x1
x2
X2
= 0 1 A

x3
X3
0 A 1
or upon inversion

1 A2
X1
1
X2
0
=

1 A2
X3
0

0
0
x1
x2
1
A

x3
A
1

(4.13-a)
(4.13-b)
(4.13-c)

(4.14)

(4.15)

that is X1 = x1 , X2 = (x2 Ax3 )/(1 A2 ), and X3 = (x3 Ax2 )/(1 A2 ).


3. The displacement eld can now be written in spatial coordinates as
u1
u2

u3
Victor Saouma

x1 X1 = 0
A(x3 Ax2 )
x2 X2 =
1 A2
A(x2 Ax3 )
x3 X3 =
a A2

(4.16-a)
(4.16-b)
(4.16-c)

Introduction to Continuum Mechanics

Draft

4.2 Strain Tensor

47

4. The displacements for the edge of the cube are determined as follows:
(a) L1 : Edge X1 = X1 , X2 = X3 = 0, u1 = u2 = u3 = 0
(b) L2 :X1 = X3 = 0, X2 = X2 , u1 = u2 = 0, u3 = AX2 .
(c) L3 : Edge X1 = X2 = 0, X3 = X3 , u1 = u3 = 0, u2 = AX3 , thus points along this edge are
displaced in the X2 direction proportionally to their distance from the origin.
5. For the circular surface, and by direct substitution of X2 = (x2 Ax3 )/(1 A2 ), and X3 =
2
2
(x3 Ax2 )/(1 A2 ) in X2 + X3 = 1/(1 A2 ), the circular surface becomes the elliptical surface
(1 + A2 )x2 4Ax2 x3 + (1 + A2 )x2 = (1 A2 ) or for A = 1/2, 5x2 8x2 x3 + 5x2 = 3 . When
2
3
2
3

expressed in its principal axes, Xi (at /4), it has the equation x2 + 9x2 = 3
2
3

X3
*
X
2

dX

1/ 3

L
2/ 3

dX

dX
2

X +X =4/3

4.2.1.1

Lagrangian and Eulerian Descriptions; x(X, t), X(x, t)

17 When the continuum undergoes deformation (or ow), the particles in the continuum move along
various paths which can be expressed in either the material coordinates or in the spatial coordinates
system giving rise to two dierent formulations:

Lagrangian Formulation: gives the present location xi of the particle that occupied the point (X1 X2 X3 )
at time t = 0, and is a mapping of the initial conguration into the current one.

xi = xi (X1 , X2 , X3 , t) or x = x(X, t)

(4.17)

Eulerian Formulation: provides a tracing of its original position of the particle that now occupies the
location (x1 , x2 , x3 ) at time t, and is a mapping of the current conguration into the initial one.
Xi = Xi (x1 , x2 , x3 , t) or X = X(x, t)

(4.18)

and the independent variables are the coordinates xi and t.


18

(X, t) and (x, t) are the Lagrangian and Eulerian variables respectivly.

19

If X(x, t) is linear, then the deformation is said to be homogeneous and plane sections remain plane.

Victor Saouma

Introduction to Continuum Mechanics

Draft
48

KINEMATIC

20 For both formulation to constitute a one-to-one mapping, with continuous partial derivatives, they
must be the unique inverses of one another. A necessary and unique condition for the inverse functions
to exist is that the determinant of the Jacobian should not vanish

|J| =

xi
=0
Xi

(4.19)

For example, the Lagrangian description given by


x1 = X1 + X2 (et 1); x2 = X1 (et 1) + X2 ; x3 = X3

(4.20)

has the inverse Eulerian description given by


X1 =

x1 + x2 (et 1)
x1 (et 1) x2
; X2 =
; X3 = x3
t et
1e
1 et et

(4.21)

Example 4-2: Lagrangian and Eulerian Descriptions


The Lagrangian description of a deformation is given by x1 = X1 +X3 (e2 1), x2 = X2 +X3 (e2 e2 ),
and x3 = e2 X3 where e is a constant. Show that the jacobian does not vanish and determine the Eulerian
equations describing the motion.
Solution:
The Jacobian is given by
1 0
0 1
0 0

(e2 1)
(e2 e2 )
e2

= e2 = 0

(4.22)

Inverting the equation

1 0
0 1
0 0

1
(e2 1)
1
(e2 e2 ) = 0
e2
0

4.2.2

Deformation; (x

(e2 1)
X1
X2
(e4 1)

X3
e2

=
=
=

x1 + (e2 1)x3
x2 + (e4 1)x3
e2 x3

(4.23)

Gradients

4.2.2.1

0
1
0

X, X

x)

21 Partial dierentiation of Eq.


4.17 with respect to Xj produces the tensor xi /Xj which is the
material deformation gradient. In symbolic notation xi /Xj is represented by the dyadic

Fx
The matrix form of F is

x1
x2
F=

x3

Victor Saouma

x
x
x
xi
e1 +
e2 +
e3 =
X1
X2
X3
Xj

X1

X2

X3

x1
X1
x2
X1
x3
X1

x1
X2
x2
X2
x3
X2

x1
X3
x2
X3
x3
X3

(4.24)

xi
Xj

(4.25)

Introduction to Continuum Mechanics

Draft

4.2 Strain Tensor

22

49

Similarly, dierentiation of Eq. 4.18 with respect to xj produces the spatial deformation gradient
H=X

The matrix form of H is

X1
X2
H=

X3
23

X
X
X
Xi
e1 +
e2 +
e3 =
x1
x2
x3
xj

x1

x2

x3

X1
x1
X2
x1
X3
x1

X1
x2
X2
x2
X3
x2

X1
x3
X2
x3
X3
x3

(4.26)

Xi

=
xj

(4.27)

The material and spatial deformation tensors are interrelated through the chain rule
Xi xj
xi Xj
=
= ik
Xj xk
xj Xk

(4.28)

H = F1

(4.29)

and thus F 1 = H or

24 The deformation gradient characterizes the rate of change of deformation with respect to coordinates.
It reects the stretching and rotation of the domain in the innitesimal neighborhood at point x (or X).
25 The deformation gradient is often called a two point tensor because the basis ei Ej has one leg
in the spatial (deformed), and the other in the material (undeformed) conguration.

F is a tensor of order two which when operating on a unit tangent vector in the undeformed conguration will produce a tangent vector in the deformed conguration. Similarly H is a tensor of order two
which when operating on a unit tangent vector in the deformed conguration will produce a tangent
vector in the undeformed conguration.

26

4.2.2.1.1 Change of Area Due to Deformation 27 In order to facilitate the derivation of the
Piola-Kircho stress tensor later on, we need to derive an expression for the change in area due to
deformation.
If we consider two material element dX(1) = dX1 e1 and dX(2) = dX2 e2 emanating from X, the
rectangular area formed by them at the reference time t0 is
28

dA0 = dX(1) dX(2) = dX1 dX2 e3 = dA0 e3


29

(4.30)

At time t, dX(1) deforms into dx(1) = FdX(1) and dX(2) into dx(2) = FdX(2) , and the new area is
dA

= FdX(1) FdX(2) = dX1 dX2 Fe1 Fe2 = dA0 Fe1 Fe2


= dAn

(4.31-a)
(4.31-b)

where the orientation of the deformed area is normal to Fe1 and Fe2 which is denoted by the unit
vector n. Thus,
(4.32)
Fe1 dAn = Fe2 dAn = 0
and recalling that abc is equal to the determinant whose rows are components of a, b, and c,
Fe3 dA = dA0 (Fe3 Fe1 Fe2 )

(4.33)

det(F)

Victor Saouma

Introduction to Continuum Mechanics

Draft
410

KINEMATIC

or
e3 FT n =

dA0
det(F)
dA

(4.34)

and FT n is in the direction of e3 so that


FT n =

dA0
det Fe3 dAn = dA0 det(F)(F1 )T e3
dA

(4.35)

which implies that the deformed area has a normal in the direction of (F1 )T e3 . A generalization of the
preceding equation would yield
dAn = dA0 det(F)(F1 )T n0

4.2.2.1.2 Change of Volume Due to Deformation


it has the following volume in material coordinate system:

30

(4.36)

If we consider an innitesimal element

d0 = (dX1 e1 dX2 e2 )dX3 e3 = dX1 dX2 dX3

(4.37)

d = (dx1 e1 dx2 e2 )dx3 e3

(4.38)

in spatial cordiantes:
If we dene
Fi =

xi
ei
Xj

(4.39)

then the deformed volume will be


d = (F1 dX1 F2 dX2 )F3 dX3 = (F1 F2 F3 )dX1 dX2 dX3

(4.40)

d = det Fd0

(4.41)

or

and J is called the Jacobian and is the determinant of the deformation gradient F

J=

x1
X1
x2
X1
x3
X1

x1
X2
x2
X2
x3
X2

x1
X3
x2
X3
x3
X3

(4.42)

and thus the Jacobian is a measure of deformation.


31

We observe that if a material is incompressible than det F = 1.

Example 4-3: Change of Volume and Area


For the following deformation: x1 = 1 X1 , x2 = 3 X3 , and x3 = 2 X2 , nd the deformed volume
for a unit cube and the deformed area of the unit square in the X1 X2 plane.
Solution:

[F]
det F
Victor Saouma

1
0
0

1 2 3

0
0
2

0
3
0

(4.43-a)
(4.43-b)
Introduction to Continuum Mechanics

Draft

4.2 Strain Tensor

411
V

1 2 3

(4.43-c)

A0
n0

=
=

1
e3

(4.43-d)
(4.43-e)

An

(1)(det F)(F1 )T
1
0
1
0
0
= 1 2 3
1
0 2
=

0
0 0
1
3
0
1 2
=


0
1
0

An = 1 2 e2

4.2.2.2

(4.43-f)
(4.43-g)
(4.43-h)

Displacements; (u

X, u

x)

32 We now turn our attention to the displacement vector ui as given by Eq. 4.12-c. Partial dierentiation
of Eq. 4.12-c with respect to Xj produces the material displacement gradient

xi
ui
=
ij or J u
Xj
Xj
The matrix form of J is

u1
u2
J =

u3

X1

X2

X3

u1
X1
u2
X1
u3
X1

=FI

u1
X2
u2
X2
u3
X2

u1
X3
u2
X3
u3
X3

(4.44)

ui
Xj

(4.45)

33 Similarly, dierentiation of Eq. 4.12-c with respect to xj produces the spatial displacement gradient

Xi
ui
= ij
or K u
xj
xj
The matrix form of K is

u1
u2
K=

u3
4.2.2.3

x1

x2

x3

u1
x1
u2
x1
u3
x1

=IH

u1
x2
u2
x2
u3
x2

u1
x3
u2
x3
u3
x3

(4.46)

ui

=
xj

(4.47)

Examples

Example 4-4: Material Deformation and Displacement Gradients


2
2
2
A displacement eld is given by u = X1 X3 e1 + X1 X2 e2 + X2 X3 e3 , determine the material deformation gradient F and the material displacement gradient J, and verify that J = F I.
Solution:

1. Since x = u + X, the displacement eld is given by


2
2
2
x = X1 (1 + X3 ) e1 + X2 (1 + X1 ) e2 + X3 (1 + X2 ) e3
x1

Victor Saouma

x2

(4.48)

x3

Introduction to Continuum Mechanics

Draft
412

KINEMATIC

2. Thus
F

x1
X1
x2
X1
x3
X1

x
x
x
xi
e1 +
e2 +
e3 =
X1
X2
X3
Xj

x1
X2
x2
X2
x3
X2

2
1 + X3
2X1 X2
0

0
2
1 + X1
2X2 X3

3. The material deformation gradient is:


ui
=J=u
Xj

x1
X3
x2
X3
x3
X3

(4.49-b)

2X1 X3

0
2
1 + X2

uX1
X1
uX2
X1
uX3
X1
2
X3

(4.49-a)

uX1
X2
uX2
X2
uX3
X2

2X1 X2
0

(4.49-c)

uX1
X3
uX2
X3
uX3
X3

0
2
X1
2X2 X3

2X1 X3

0
2
X2

(4.50-a)

(4.50-b)

We observe that the two second order tensors are related by J = F I.

4.2.3

Deformation Tensors

34 The deofrmation gradients, previously presented, can not be used to determine strains as embedded
in them is rigid body motion.

x
Having derived expressions for Xij and Xji we now seek to determine dx2 and dX 2 where dX and dx
x
correspond to the distance between points P and Q in the undeformed and deformed cases respectively.
35

36 We consider next the initial (undeformed) and nal (deformed) conguration of a continuum in which
the material OX1 , X2 , X3 and spatial coordinates ox1 x2 x3 are superimposed. Neighboring particles P0
and Q0 in the initial congurations moved to P and Q respectively in the nal one, Fig. 4.5.

4.2.3.1

Cauchys Deformation Tensor; (dX)2

The Cauchy deformation tensor, introduced by Cauchy in 1827, B1 (alternatively denoted as c)


gives the initial square length (dX)2 of an element dx in the deformed conguration.
37

38

This tensor is the inverse of the tensor B which will not be introduced until Sect. 4.3.2.

39

The square of the dierential element connecting Po and Q0 is


(dX)2 = dXdX = dXi dXi

(4.51)

however from Eq. 4.18 the distance dierential dXi is


dXi =

Xi
dxj or dX = Hdx
xj

(4.52)

thus the squared length (dX)2 in Eq. 4.51 may be rewritten as

Victor Saouma

Xk Xk
1
dxi dxj = Bij dxi dxj
xi xj

(4.53-a)

(dX)2

dxB1 dx

(4.53-b)
Introduction to Continuum Mechanics

Draft

413

P
0

dX

u +du

X 1, x 1

t=0

X 3 , x3

X+
d

t=t
Q

dx

X 2, x 2

4.2 Strain Tensor

Figure 4.5: Undeformed and Deformed Congurations of a Continuum

Victor Saouma

Introduction to Continuum Mechanics

Draft
414

KINEMATIC

in which the second order tensor


1
Bij =

Xk Xk
or B1 =
xi xj

x XX

(4.54)

Hc H

is Cauchys deformation tensor. It relates (dX)2 to (dx)2 .


4.2.3.2

Greens Deformation Tensor; (dx)2

The Green deformation tensor, introduced by Green in 1841, C (alternatively denoted as B1 ),


referred to in the undeformed conguration, gives the new square length (dx)2 of the element dX is
deformed.
40

41 The square of the dierential element connecting Po and Q0 is now evaluated in terms of the spatial
coordinates
(4.55)
(dx)2 = dxdx = dxi dxi

however from Eq. 4.17 the distance dierential dxi is


xi
dXj or dx = FdX
Xj

dxi =

(4.56)

thus the squared length (dx)2 in Eq. 4.55 may be rewritten as


(dx)2

=
=

xk xk
dXi dXj = Cij dXi dXj
Xi Xj
dXCdX

(4.57-a)
(4.57-b)

in which the second order tensor


Cij =

xk xk
or C =
Xi Xj

X xx

(4.58)

Fc F

is Greens deformation tensor also known as metric tensor, or deformation tensor or right
Cauchy-Green deformation tensor. It relates (dx)2 to (dX)2 .
42

Inspection of Eq. 4.54 and Eq. 4.58 yields


C1 = B1 or B1 = (F1 )T F1

(4.59)

Example 4-5: Greens Deformation Tensor


A continuum body undergoes the deformation x1 = X1 , x2 = X2 + AX3 , and x3 = X3 + AX2 where
A is a constant. Determine the deformation tensor C.
Solution:
From Eq. 4.58 C = Fc F where F was dened in Eq. 4.24 as
F

Victor Saouma

xi
Xj

1 0
= 0 1
0 A
=

(4.60-a)

0
A
(4.60-b)
1
Introduction to Continuum Mechanics

Draft

4.2 Strain Tensor

415

and thus
C =
=

4.2.4

Fc F

T
1 0 0
1
0 1 A 0
0 A 1
0

0 0
1
1 A = 0
0
A 1

(4.61-a)

0
1 + A2
2A

0
2A
1 + A2

(4.61-b)

Strains; (dx)2 (dX)2

With (dx)2 and (dX)2 dened we can now nally introduce the concept of strain through (dx)2
(dX)2 .
43

4.2.4.1

Finite Strain Tensors

44 We start with the most general case of nite strains where no constraints are imposed on the deformation (small).

4.2.4.1.1

Lagrangian/Greens Strain Tensor

The dierence (dx)2 (dX)2 for two neighboring particles in a continuum is used as the measure
of deformation. Using Eqs. 4.57-a and 4.51 this dierence is expressed as
45

(dx)2 (dX)2

=
=

xk xk
ij dXi dXj = 2Eij dXi dXj
Xi Xj
dX(Fc F I)dX = 2dXEdX

(4.62-a)
(4.62-b)

in which the second order tensor


Eij =

1
2

xk xk
ij
Xi Xj

or E =

1
(
2

X xx

I)

(4.63)

Fc F=C

is called the Lagrangian (or Greens) nite strain tensor which was introduced by Green in 1841
and St-Venant in 1844.
46

The Lagrangian stress tensor is one half the dierence between the Green deformation tensor and I.

47 Note similarity with Eq. 4.4 where the Lagrangian strain (in 1D) was dened as the dierence between
the square of the deformed length and the square of the original length divided by twice the square of
2
l2 l0
). Eq. 4.62-a can be rewritten as
the original length (E 1
2
l2
0

(dx)2 (dX)2 = 2Eij dXi dXj Eij =

(dx)2 (dX)2
2dXi dXj

(4.64)

which gives a clearer physical meaning to the Lagrangian Tensor.


48 To express the Lagrangian tensor in terms of the displacements, we substitute Eq.
4.44 in the
preceding equation, and after some simple algebraic manipulations, the Lagrangian nite strain tensor
can be rewritten as

Eij =
Victor Saouma

1
2

uj
uk uk
ui
+
+
Xj
Xi
Xi Xj

or E =

1
(u
2

Xu +

X uu

X)

(4.65)

J+Jc
Jc J
Introduction to Continuum Mechanics

Draft
416

KINEMATIC

or:
E11

E12

u1
1
+
X1
2

u1
X1

1 u1
u2
+
2 X2
X1

+
+

u2
X1

u3
X1

(4.66-a)

1 u1 u1
u2 u2
u3 u3
+
+
2 X1 X2
X1 X2
X1 X2

(4.66-b)
(4.66-c)

Example 4-6: Lagrangian Tensor


Determine the Lagrangian nite strain tensor E for the deformation of example 4.2.3.2.
Solution:

1
0
0
2A
C = 0 1 + A2
0
2A
1 + A2
1
(C I)
E =
2

0 0
0
1
0 A2 2A
=
2
0 2A A2

(4.67-a)
(4.67-b)
(4.67-c)

Note that the matrix is symmetric.

4.2.4.1.2

Eulerian/Almansis Tensor

Alternatively, the dierence (dx)2 (dX)2 for the two neighboring particles in the continuum can be
expressed in terms of Eqs. 4.55 and 4.53-b this same dierence is now equal to
49

(dx)2 (dX)2

Xk Xk

dxi dxj = 2Eij dxi dxj


xi xj
= dx(I Hc H)dx = 2dxE dx
=

ij

(4.68-a)
(4.68-b)

in which the second order tensor

Eij =

1
2

ij

Xk Xk
xi xj

or E =

1
(I
2

x XX
Hc

x)

(4.69)

H=B1

is called the Eulerian (or Almansi) nite strain tensor.


50

The Eulerian strain tensor is one half the dierence between I and the Cauchy deformation tensor.

51 Note similarity with Eq. 4.5 where the Eulerian strain (in 1D) was dened as the dierence between
the square of the deformed length and the square of the original length divided by twice the square of
2
l2 l0
). Eq. 4.68-a can be rewritten as
the deformed length (E 1
2
l2

(dx)2 (dX)2 = 2Eij dxi dxj Eij =

Victor Saouma

(dx)2 (dX)2
2dxi dxj

(4.70)

Introduction to Continuum Mechanics

Draft

4.2 Strain Tensor

417

which gives a clearer physical meaning to the Eulerian Tensor.


52

For innitesimal strain it was introduced by Cauchy in 1827, and for nite strain by Almansi in 1911.

53 To express the Eulerian tensor in terms of the displacements, we substitute 4.46 in the preceding
equation, and after some simple algebraic manipulations, the Eulerian nite strain tensor can be rewritten
as

Eij =

54

1
2

uj
uk uk
ui
+

xj
xi
xi xj

or E =

1
(u
2

xu

x uu

x)

(4.71)

Kc K

K+Kc

Expanding

E11

E12

4.2.4.2

u1
1

x1
2

u1
x1

1 u1
u2
+
2 x2
x1

u2
x1

u3
x1

(4.72-a)

1 u1 u1
u2 u2
u3 u3
+
+
2 x1 x2
x1 x2
x1 x2

(4.72-b)
(4.72-c)

Innitesimal Strain Tensors; Small Deformation Theory

55 The small deformation theory of continuum mechanics has as basic condition the requirement that
the displacement gradients be small compared to unity. The fundamental measure of deformation is the
dierence (dx)2 (dX)2 , which may be expressed in terms of the displacement gradients by inserting Eq.
4.65 and 4.71 into 4.62-b and 4.68-b respectively. If the displacement gradients are small, the nite strain
tensors in Eq. 4.62-b and 4.68-b reduce to innitesimal strain tensors and the resulting equations
represent small deformations.

For instance, if we were to evaluate + 2 , for = 103 and 101 , then we would obtain 0.001001
0.001 and 0.11 respectively. In the rst case 2 is negligible compared to , in the other it is not.
56

4.2.4.2.1

Lagrangian Innitesimal Strain Tensor

u
In Eq. 4.65 if the displacement gradient components Xi are each small compared to unity, then the
j
third term are negligible and may be dropped. The resulting tensor is the Lagrangian innitesimal
strain tensor denoted by
57

Eij =

1
2

uj
ui
+
Xj
Xi

or E =

1
(u
2

X u)

(4.73)

J+Jc

or:
E11

E12

u1
X1
1 u1
u2
+
2 X2
X1

(4.74-a)
(4.74-b)
(4.74-c)

Note the similarity with Eq. 4.7.

Victor Saouma

Introduction to Continuum Mechanics

Draft
418

4.2.4.2.2

KINEMATIC

Eulerian Innitesimal Strain Tensor

ui
Similarly, inn Eq. 4.71 if the displacement gradient components xj are each small compared to
unity, then the third term are negligible and may be dropped. The resulting tensor is the Eulerian
innitesimal strain tensor denoted by
58

Eij =

59

1
2

uj
ui
+
xj
xi

or E =

1
(u
2

x u)

(4.75)

K+Kc

Expanding

E11

E12

4.2.4.3

u1
x1
1 u1
u2
+
2 x2
x1

(4.76-a)
(4.76-b)
(4.76-c)

Examples

Example 4-7: Lagrangian and Eulerian Linear Strain Tensors


A displacement eld is given by x1 = X1 +AX2 , x2 = X2 +AX3 , x3 = X3 +AX1 where A is constant.
Calculate the Lagrangian and the Eulerian linear strain tensors, and compare them for the case where
A is very small.
Solution:
The displacements are obtained from Eq. 4.12-c uk = xk Xk or
u1

x1 X1 = X1 + AX2 X1 = AX2

(4.77-a)

u2
u3

=
=

x2 X2 = X2 + AX3 X2 = AX3
x3 X3 = X3 + AX1 X3 = AX1

(4.77-b)
(4.77-c)

then from Eq. 4.44

Ju
From Eq. 4.73:

0
= 0
A

0
2E = (J + Jc ) = 0
A

0 A A
= A 0 A
A A 0
To determine the Eulerian tensor,
displacement eld given above:

1 A 0 X1
x1
x2
X2
= 0 1 A

x3
X3
A 0 1
Victor Saouma

A
0
0

0
A
0


A 0
0
0 A + A
0 0
0

(4.78)

0
0
A

A
0
0

(4.79-a)

(4.79-b)

we need the displacement u in terms of x, thus inverting the

1
A A2
x1
X1
1 2
X2
A
1
A
x2

1 + A3
X3
1
x3
A A2
Introduction to Continuum

(4.80)

Mechanics

Draft

4.2 Strain Tensor

419

thus from Eq. 4.12-c uk = xk Xk we obtain


u1

u2

u3

1
A(A2 x1 + x2 Ax3 )
(x1 Ax2 + A2 x3 ) =
1 + A3
1 + A3
1
A(Ax1 + A2 x2 + x3 )
x2 X2 = x2
(A2 x1 + x2 Ax3 ) =
3
1+A
1 + A3
1
A(x1 Ax2 + A2 x3 )
x3 X3 = x3
(Ax1 + A2 x2 + x3 ) =
1 + A3
1 + A3
x1 X1 = x1

From Eq. 4.46


Ku

1
A
A
A
1
A A2
=
1 + A3
1
A A2

(4.81-a)
(4.81-b)
(4.81-c)

(4.82)

Finally, from Eq. 4.71


2E

=
=

K + Kc

A
1 + A3

A
1 + A3

A2
A
1
A
A
1
A
1 +
A A2
1
A2 A
1 + A3
1
A A2
A
1
A2

2A2 1 A 1 A
1 A 2A2 1 A
1 A 1 A 2A2

(4.83-a)
(4.83-b)

(4.83-c)

as A is very small, A2 and higher power may be neglected with the results, then E E.

4.2.5

Physical Interpretation of the Strain Tensor

4.2.5.1

Small Strain

60 We nally show that the linear lagrangian tensor in small deformation Eij is nothing else than the
strain as was dened earlier in Eq.4.7.

61

We rewrite Eq. 4.62-b as


(dx)2 (dX)2
(dx)2 (dX)2

=
or
=

(dx dX)(dx + dX) = 2Eij dXi dXj

(4.84-a)

(dx dX)(dx + dX) = dX2EdX

(4.84-b)

but since dx dX under current assumption of small deformation, then the previous equation can be
rewritten as
du
dXi dXj
dx dX
= Eij
= Eij i j = E
dX
dX dX

(4.85)

62 We recognize that the left hand side is nothing else than the change in length per unit original length,
and is called the normal strain for the line element having direction cosines dXi .
dX
63 With reference to Fig. 4.6 we consider two cases: normal and shear strain.

Normal Strain: When Eq. 4.85 is applied to the dierential element P0 Q0 which lies along the X2
axis, the result will be the normal strain because since dX1 = dX3 = 0 and dX2 = 1. Therefore,
dX
dX
dX
Eq. 4.85 becomes (with ui = xi Xi ):
u2
dx dX
= E22 =
dX
X2
Victor Saouma

(4.86)

Introduction to Continuum Mechanics

Draft
420

KINEMATIC
X3

x3

Q0

X2

dX

Normal

e3

n3
n2

X1

e2

X3

e1

M0
dX

Shear

P0

Q0
dX

X1

Figure 4.6: Physical Interpretation of the Strain Tensor


Likewise for the other 2 directions. Hence the diagonal terms of the linear strain tensor represent
normal strains in the coordinate system.
Shear Strain: For the diagonal terms Eij we consider the two line elements originally located along
the X2 and the X3 axes before deformation. After deformation, the original right angle between
u
the lines becomes the angle . From Eq. 4.101 (dui = Xi
dXj ) a rst order approximation
j
gives the unit vector at P in the direction of Q, and M as:
n2

n3

P0

u1
u3
e1 + e2 +
e3
X2
X2
u1
u2
e1 +
e2 + e3
X3
X3

(4.87-a)
(4.87-b)

and from the denition of the dot product:


cos = n2 n3 =

u1 u1
u2
u3
+
+
X2 X3
X3
X2

(4.88)

or neglecting the higher order term


cos =

u2
u3
+
= 2E23
X3
X2

(4.89)

64 Finally taking the change in right angle between the elements as 23 = /2 , and recalling
that for small strain theory 23 is very small it follows that

23 sin 23 = sin(/2 ) = cos = 2E23 .


Victor Saouma

(4.90)

Introduction to Continuum Mechanics

Draft

4.2 Strain Tensor

421

Therefore the o diagonal terms of the linear strain tensor represent one half of the angle change
between two line elements originally at right angles to one another. These components are called
the shear strains.
64 The Engineering shear strain is dened as one half the tensorial shear strain, and the resulting
tensor is written as

1
1
11
2 12
2 13
1
1

(4.91)
Eij = 2 12 22
2 23
1
1
33
2 13
2 23
65 We note that a similar development paralleling the one just presented can be made for the linear
Eulerian strain tensor (where the straight lines and right angle will be in the deformed state).

4.2.5.2

Finite Strain; Stretch Ratio

66 The simplest and most useful measure of the extensional strain of an innitesimal element is the
stretch or stretch ratio as dx which may be dened at point P0 in the undeformed conguration or
dX
at P in the deformed one (Refer to the original denition given by Eq, 4.1).

Hence, from Eq. 4.57-a, and Eq. 4.63 the squared stretch at P0 for the line element along the unit
vector m = dX is given by
dX
67

2
m

dx
dX

= Cij
P0

dXi dXj
or 2 = mCm
m
dX dX

(4.92)

Thus for an element originally along X2 , Fig. 4.6, m = e2 and therefore dX1 /dX = dX3 /dX = 0 and
dX2 /dX = 1, thus Eq. 4.92 (with Eq. ??) yields
22 = C22 = 1 + 2E22
e

(4.93)

and similar results can be obtained for 21 and 23 .


e
e
Similarly from Eq. 4.53-b, the reciprocal of the squared stretch for the line element at P along the
unit vector n = dx is given by
dx

68

2
n

dX
dx

2
P

1
= Bij

1
dxi dxj
or 2 = nB1 n
dx dx
n

(4.94)

Again for an element originally along X2 , Fig. 4.6, we obtain


1

= 1 2E22
2 2
e

(4.95)

69 we note that in general e = e since the element originally along the X2 axis will not be along the
2
2
x2 after deformation. Furthermore Eq. 4.92 and 4.94 show that in the matrices of rectangular cartesian
components the diagonal elements of both C and B1 must be positive, while the elements of E must
be greater than 1 and those of E must be greater than + 1 .
2
2
70

The unit extension of the element is


dx
dx dX
=
1 = m 1
dX
dX

(4.96)

and for the element P0 Q0 along the X2 axis, the unit extension is
dx dX
= E(2) = e2 1 =
dX
Victor Saouma

1 + 2E22 1

(4.97)

Introduction to Continuum Mechanics

Draft
422

KINEMATIC

for small deformation theory E22 << 1, and


1
dx dX
= E(2) = (1 + 2E22 ) 2 1
dX

1
1 + 2E22 1
2

E22

(4.98)

which is identical to Eq. 4.86.


For the two dierential line elements of Fig. 4.6, the change in angle 23 =
both e2 and e3 by
2E23
2E23

sin 23 =
=
e2 e3
1 + 2E22 1 + 2E33
71

is given in terms of
(4.99)

Again, when deformations are small, this equation reduces to Eq. 4.90.

4.3

Strain Decomposition

72 In this section we rst seek to express the relative displacement vector as the sum of the linear
(Lagrangian or Eulerian) strain tensor and the linear (Lagrangian or Eulerian) rotation tensor. This is
restricted to small strains.
73 For nite strains, the former additive decomposition is no longer valid, instead we shall consider the
strain tensor as a product of a rotation tensor and a stretch tensor.

4.3.1

Linear Strain and Rotation Tensors

74 Strain components are quantitative measures of certain type of relative displacement between neighboring parts of the material. A solid material will resist such relative displacement giving rise to internal
stresses.
75 Not all kinds of relative motion give rise to strain (and stresses). If a body moves as a rigid body,
the rotational part of its motion produces relative displacement. Thus the general problem is to express
the strain in terms of the displacements by separating o that part of the displacement distribution
which does not contribute to the strain.

4.3.1.1

Small Strains

From Fig. 4.7 the displacements of two neighboring particles are represented by the vectors uP0 and
u and the vector
(4.100)
dui = uQ0 uP0 or du = uQ0 uP0
i
i
76

Q0

is called the relative displacement vector of the particle originally at Q0 with respect to the one
originally at P0 .
4.3.1.1.1

77

Lagrangian Formulation

Neglecting higher order terms, and through a Taylor expansion


dui =

ui
Xj

dXj or du = (u
P0

X )P 0

dX

(4.101)

78 We also dene a unit relative displacement vector dui /dX where dX is the magnitude of the
dierential distance dXi , or dXi = i dX, then

ui dXj
ui
du
dui
=
=
=u
j or
dX
Xj dX
Xj
dX
Victor Saouma

= J

(4.102)

Introduction to Continuum Mechanics

Draft

4.3 Strain Decomposition

423

Q
Q

du

dx

Q0

dX
P0

P0
Figure 4.7: Relative Displacement du of Q relative to P
u
The material displacement gradient Xi can be decomposed uniquely into a symmetric and an antij
symetric part, we rewrite the previous equation as

79

dui

uj
ui
+
Xj
Xi

1
2

Eij

or

1
= (u
2

X u) +

1
(u
2

or

E=

u1
X1
1
2
1
2

u1
X2
u1
X3

(4.103-a)

Wij

du

dXj

uj
ui

Xj
Xi

+
+

X u) dX

(4.103-b)

W
u1
X2 +
u2
X2
u2
X3 +

1
2
u2
X1
u3
X1

1
2

u2
X1

u1
X3 +
u2
X3 +
u3
X3

1
2
1
2

u3
X2

u3
X1
u3
X2

X u)

u1
X3
u2
X3

(4.104)

We thus introduce the linear lagrangian rotation tensor


Wij =
in matrix form:

1
2

uj
ui

Xj
Xi

W = 1
2
1
2

1
2

0
u1
X2
u1
X3

or W =

u2
X1
u3
X1

u1
X2

1
(u
2

u2
X1

u2
X3

1
2
1
2

0
1
2

u3
X2

u3
X1
u3
X2

(4.105)

(4.106)

80 In a displacement for which Eij is zero in the vicinity of a point P0 , the relative displacement at that
point will be an innitesimal rigid body rotation. It can be shown that this rotation is given by the

Victor Saouma

Introduction to Continuum Mechanics

Draft
424

KINEMATIC

linear Lagrangian rotation vector


1
2

wi =

ijk Wkj

1
2

or w =

X u

(4.107)

or
w = W23 e1 W31 e2 W12 e3
4.3.1.1.2

81

Eulerian Formulation

The derivation in an Eulerian formulation parallels the one for Lagrangian formulation. Hence,
dui =

82

ui
dxj or du = Kdx
xj

ui
ui dxj
du
=
=u
j or
xj dx
xj
dx

uj
ui
+
xj
xi

1
2

u1
x1
1
2
1
2

u1
x2
u1
x3

(4.111-a)

ij

x u) +

1
(u
2

E=

dxj

1
= (u
2

(4.110)

results in

du

or

= K

uj
ui

xj
xi

Eij

or

ui
xj

The decomposition of the Eulerian displacement gradient

dui

84

(4.109)

The unit relative displacement vector will be


dui =

83

(4.108)

+
+

x u) dx

(4.111-b)

u1
x2 +
u2
x2
u2
x3 +

1
2
u2
x1
u3
x1

1
2

u2
x1

u1
x3 +
u2
x3 +
u3
x3

1
2
1
2

u3
x2

u3
x1
u3
x2

(4.112)

We thus introduced the linear Eulerian rotation tensor


wij =

in matrix form:

1
2

uj
ui

xj
xi

1
W = 2

1
2

u2
x1
u3
x1

u1
x2

1
2

0
u1
x2
u1
x3

or =

1
(u
2

u2
x1

u2
x3

1
2
1
2

0
1
2

u3
x2

u1
x3
u2
x3

(4.113)

x u)

u3
x1
u3
x2

(4.114)

and the linear Eulerian rotation vector will be


i =

Victor Saouma

1
2

ijk kj

or =

1
2

x u

(4.115)

Introduction to Continuum Mechanics

Draft

4.3 Strain Decomposition


4.3.1.2

425

Examples

Example 4-8: Relative Displacement along a specied direction


2
2
2
A displacement eld is specied by u = X1 X2 e1 + (X2 X3 )e2 + X2 X3 e3 . Determine the relative
displacement vector du in the direction of the X2 axis at P (1, 2, 1). Determine the relative displacements uQi uP for Q1 (1, 1, 1), Q2 (1, 3/2, 1), Q3 (1, 7/4, 1) and Q4 (1, 15/8, 1) and compute their
directions with the direction of du.
Solution:
From Eq. 4.44, J = u X or

2
2X1 X2
X1
0
ui
0
1
2X3
=
(4.116)
Xj
2
0
2X2 X3
X2

thus from Eq. 4.101 du = (u

X )P

dX in the direction

4 1 0
{du} = 0 1 2

0 4 4

of X2 or

0 1
1
1
=

0
4

(4.117)

By direct calculation from u we have


uP
uQ1

2e1 + e2 4e3

(4.118-a)

e1 e3

(4.118-b)

thus
uQ1 uP

uQ2 uP

uQ3 uP

uQ4 uP

e1 e2 + 3e3
1
(e1 e2 + 3.5e3 )
2
1
(e1 e2 + 3.75e3 )
4
1
(e1 e2 + 3.875e3 )
8

(4.119-a)
(4.119-b)
(4.119-c)
(4.119-d)

and it is clear that as Qi approaches P , the direction of the relative displacements of the two particles
approaches the limiting direction of du.

Example 4-9: Linear strain tensor, linear rotation tensor, rotation vector
Under the restriction of small deformation theory E = E , a displacement eld is given by u =
(x1 x3 )2 e1 + (x2 + x3 )2 e2 x1 x2 e3 . Determine the linear strain tensor, the linear rotation tensor and
the rotation vector at point P (0, 2, 1).
Solution:
the matrix form of the displacement gradient is

0
2(x1 x3 )
2(x1 x3 )
ui
0
2(x2 + x3 ) 2(x2 + x3 )
(4.120-a)
] =
[
xj
x1
0
x2

2 0 2
ui
= 0 2 2
(4.120-b)
xj P
2 0 0
Victor Saouma
Introduction to Continuum Mechanics

Draft
426

KINEMATIC

Decomposing this matrix into symmetric and antisymmetric components give:


2 0 2
0 0 0
[Eij ] + [wij ] = 0 2 1 + 0 0 1
0 1 0
2 1 0

(4.121)

and from Eq. Eq. 4.108


w = W23 e1 W31 e2 W12 e3 = 1e1

4.3.2

(4.122)

Finite Strain; Polar Decomposition

u
When the displacement gradients are nite, then we no longer can decompose Xi (Eq. 4.101) or
j
(Eq. 4.109) into a unique sum of symmetric and skew parts (pure strain and pure rotation).
85

ui
xj

86 Thus in this case, rather than having an additive decomposition, we will have a multiplicative
decomposition.
87 we call this a polar decomposition and it should decompose the deformation gradient in the product
of two tensors, one of which represents a rigid-body rotation, while the other is a symmetric positivedenite tensor.

88

We apply this decomposition to the deformation gradient F:


Fij

xi
= Rik Ukj = Vik Rkj or F = RU = VR
Xj

(4.123)

where R is the orthogonal rotation tensor, and U and V are positive symmetric tensors known as
the right stretch tensor and the left stretch tensor respectively.
89

The interpretation of the above equation is obtained by inserting the above equation into dxi =

xi
Xj dXj

dxi = Rik Ukj dXj = Vik Rkj dXj or dx = RUdX = VRdX

(4.124)

and we observe that in the rst form the deformation consists of a sequential stretching (by U) and
rotation (R) to be followed by a rigid body displacement to x. In the second case, the orders are
reversed, we have rst a rigid body translation to x, followed by a rotation (R) and nally a stretching
(by V).
90

To determine the stretch tensor from the deformation gradient


FT F = (RU)T (RU) = UT RT RU = UT U

(4.125)

Recalling that R is an orthonormal matrix, and thus RT = R1 then we can compute the various
tensors from

U =
FT F (4.126)
R = FU1
V = FRT
91

(4.127)
(4.128)

It can be shown that


U = C1/2 and V = B1/2

(4.129)

Example 4-10: Polar Decomposition I


Victor Saouma

Introduction to Continuum Mechanics

Draft

4.3 Strain Decomposition

427

Given x1 = X1 , x2 = 3X3 , x3 = 2X2 , nd the deformation gradient F, the right stretch tensor U,
the rotation tensor R, and the left stretch tensor V.
Solution:
From Eq. 4.25

x1
X1
x2
X1
x3
X1

F=
From Eq. 4.126

1
U2 = FT F = 0
0

x1
X2
x2
X2
x3
X2

0
0
3

1
U= 0
0
From Eq. 4.127

R = FU1

1
= 0
0

0
0
2

1

= 0
0

1
0
2 0
0
0

thus

x1
X3
x2
X3
x3
X3

0
2
0

1
0
3 0
0
0

0 0
0 3
2 0


1
0
3 = 0
0
0

0
0
2

(4.130)

0
4
0

0
0
9

(4.131)

0
0
3

(4.132)


0
1
0 = 0
1
0
3

0
1
2

0 0
0 1
1 0

(4.133)

Finally, from Eq. 4.128

1
V = FRT = 0
0

0 0
1
0 3 0
0
2 0

0
0
1


1
0
1 = 0
0
0

0
3
0

0
0
2

(4.134)

Example 4-11: Polar Decomposition II


For the following deformation: x1 = 1 X1 , x2 = 3 X3 , and x3 = 2 X2 , nd the rotation tensor.
Solution:

[F]
2

1
= 0
0

0
0
2

0
3
0

(4.135)

[F]T [F]

0
0
0
1 0
1
0
2 0
0 3 =
= 0
0 3 0
0 2
0

0
1 0
[U] = 0 2 0
0
0 3
1

0
1 0
0
1
1
0 3 0 2
[R] = [F][U]1 = 0
0 2
0
0
0

[U]

2
1
0
0

0
2
2
0

(4.136)

0
0
2
3

(4.137)

(4.138)

0
1
0 = 0
1
0
3

0
0
1

0
1
0

(4.139)

Thus we note that R corresponds to a 90o rotation about the e1 axis.


Victor Saouma

Introduction to Continuum Mechanics

Draft
428

KINEMATIC

Example 4-12: Polar Decomposition III

Victor Saouma

Introduction to Continuum Mechanics

Draft

4.3 Strain Decomposition

429

mpolar.nb
2

In[4]:=

8v1, v2, v3< = N@Eigenvectors@CSTD, 4D

Determine U and U -1 with respect to the ei basis


0 0 1. y
i
z
j
j
Polar Decomposition Using Mathematica
z
j -2.414 1. 0 z
z
j
Out[4]=

z
j
z
j
z
j
In[10]:=
. Ueigen the principal values of
Given x1 =XU_e = ,N@vnormalizedObtain C, b). vnormalized, 3DC and the corresponding directions, c) the
1 0.4142 1.
3
k +2X2 x2 =X2 ,0x{ =X3 , a)
-1 with respect to the principal directions, d) Obtain the matrix U and U -1 with respect to the e bas
matrix U and U
i
obtain the matrix R with respect to the ei basis.
i 0.707 0.707 0. y
z
j LinearAlgebraOrthogonalization
z
j
In[5]:=
<<
j 0.707 2.12 0. z
z
j
Out[10]=
z
j
z
j
z
j
0.
0. 1. {
k
In[6]:=
vnormalized = GramSchmidt@8v3, v2, v1<D
Determine the F matrix
In[11]:=
In[1]:=
Out[6]=
Out[11]=
Out[1]=
In[7]:=

U_einverse = N@Inverse@%D, 3D
F = 881, 2, 0<, 80, 1, 0<, 80, 0, 1<<
0.92388 0 y
i 0.382683
j
z
j
z
j 0.92388 -0.382683 0 z
j 2.12 -0.707 0.
z
j
z
j
i
y
j
z
j
z 1. z
j1 2 0 0
z
0
j
k -0.707 y 0.707 0. z
{
j
z
i
z
j
z
z
j
z
j0 1 0z
z
j
z
0. z
0. 1. {
j
k
j
z
j
z
k0 0 1{
CSTeigen = Chop@N@vnormalized . CST . vnormalized, 4DD

i 5.828

0y

j
z
j
Determine R 0 0.1716 0 z to the ei basis
with respect
z
j
z
j
Out[7]=
z
j
z
j
Solve for C
z
j
In[12]:=
In[2]:=

0
0 1. {
k
R = N@F . %, 3D
CST = Transpose@FD . F
i 0.707 0.707 0. y
j
z
j
z

j 1 2 0 y 0.707 0. z
j
z
Out[12]=
i -0.707 z
j
z
z
j
Determine U with respect
z
z
j2 5 0z
j
Out[2]=
0. z
0. 1. {
z
j
k
In[8]:=

to the principal directions

z
j
z
j
k0 0 1{
Ueigen = N@Sqrt@CSTeigenD, 4D

0 0y
i 2.414
z
j
z
j
j
z
j
0 0.4142 0 z
z
j
z
j Eigenvalues and
Determine
z
j
0
0 1. {
k
Out[8]=

Eigenvectors

In[3]:=

N@Eigenvalues@CSTDD

In[9]:=

Ueigenminus1 = Inverse@UeigenD

Out[3]=

Out[9]=

Victor Saouma

81., 0.171573, 5.82843<


0. 0. y
i 0.414214
j
z
j
z
j
j
z
0. 2.41421 0. z
j
z
j
z
j
z
0.
0. 1. {
k

Introduction to Continuum Mechanics

Draft
430

4.4
92

KINEMATIC

Summary and Discussion

From the above, we deduce the following observations:


u
ui
1. If both the displacement gradients and the displacements themselves are small, then Xi xj and
j

thus the Eulerian and the Lagrangian innitesimal strain tensors may be taken as equal Eij = Eij .

2. If the displacement gradients are small, but the displacements are large, we should use the Eulerian
innitesimal representation.
3. If the displacements gradients are large, but the displacements are small, use the Lagrangian nite
strain representation.
4. If both the displacement gradients and the displacements are large, use the Eulerian nite strain
representation.

4.5

Compatibility Equation

If ij = 1 (ui,j + uj,i ) then we have six dierential equations (in 3D the strain tensor has a total
2
of 9 terms, but due to symmetry, there are 6 independent ones) for determining (upon integration)
three unknowns displacements ui . Hence the system is overdetermined, and there must be some linear
relations between the strains.
93

94 It can be shown (through appropriate successive dierentiation of the strain expression) that the
compatibility relation for strain reduces to:

2 jj
2 jk
2 ij
2 ik
+

= 0. or
xj xj
xi xk
xi xj
xj xk

x L

=0

(4.140)

There are 81 equations in all, but only six are distinct


2 11
2 22
+
x2
x2
2
1
2
2 33
22
+
x2
x2
3
2
2
2 11
33
+
x2
x2
1
3
23
31
12

+
+
x1
x1
x2
x3
31
12
23

+
x2 x1
x2
x3

31
12
23
+

x3 x1
x2
x3

=
=
=
=
=
=

2 12
x1 x2
2 23
2
x2 x3
2 31
2
x3 x1
2 11
x2 x3
2 22
x3 x1
2 33
x1 x2
2

(4.141-a)
(4.141-b)
(4.141-c)
(4.141-d)
(4.141-e)
(4.141-f)

In 2D, this results in (by setting i = 2, j = 1 and l = 2):


2 22
2 12
2 11
+
=
x2
x2
x1 x2
2
1
Victor Saouma

(4.142)

Introduction to Continuum Mechanics

Draft

431

P
x

u +d u
X

Oo

I2 ,i

X2

I 1 , 1i

b=0

Material/Spatial

X1

,x
2

,x
1

LAGRANGIAN
Material
Position Vector

x = x(X, t)

Deformation

F=x

Displacement

ui
Xj

Deformation

=
J=u

xi
Xj

ij or
=FI

xi
Xj
X

dX 2 = dxB1 dx
Cauchy
1
k
k
Bij = Xi Xj or
x x
B1 = x XX x = Hc H

Lagrangian
dx2 dX 2 = dX2EdX
Finite Strain

Eij = 1
2
E= 1(
2
Eij = 1
2
E = 1 (u
2

Small
Deformation
Victor Saouma
Small
deformation

EULERIAN
Spatial
X = X(x, t)
GRADIENTS
H = X x Xji
x
H = F1
ui
Xi
xj = ij xj or
Ku x =IH
TENSOR
dx2 = dXCdX
Green
x x
Cij = Xk Xk or
i
j
C = X xx X = Fc F
C1 = B1
STRAINS
Eulerian/Almansi
dx2 dX 2 = dx2E dx

ij or
X xx X I)
xk xk
Xi Xj

Fc F
uj
ui
Xj + Xi
X

X 1, x 1

,i

X 3 , x3

P0

U
I

t=0

t=0

dX

,x

t=t
X

X+
d

t=t
Q

dx

X 2, x 2

4.5 Compatibility Equation

u u
+ Xk Xk or
i
j
Xu +
X uu X )

k
k
Eij = 1 ij Xi Xj or
2
x x
E = 1 (I x XX x )
2

Hc H
u
ui

Eij = 1 xj + xj uk uk or
2
xi xj
i
E = 1 (u x + x u x uu x )
2

K+Kc Kc K
u
ui

Eij = 1 xj + xj
2
i
+ Jc )
E = 1 (u x + x u) = 1 (K + Kc )
2
2
ROTATION TENSORS
Introduction touj
Continuum Mechanics
u
u
uj
u
u
1 ui
1 ui
dxj
[ 1 Xi + Xji + 1 Xi Xji ]dXj
2
2
2 xj + xi + 2 xj xi
j
j
1
1
1
1
[ (u X + X u) + (u X X u)]dX [ (u x + x u) + (u x x u)]dx
2
2
2
2

Eij = 1
2
E = 1 (u
2

ui
Xj

+
X+

J+Jc +Jc J
uj
Xi
1
X u) = 2 (J

Draft
432

KINEMATIC

(recall that 212 = 12 .)


95

When he compatibility equation is written in term of the stresses, it yields:


22 2
2 22
2 11
2 21
2 11

= 2 (1 + )
x2
x2
x2
x2
x1 x2
2
2
1
1

(4.143)

Example 4-13: Strain Compatibility


For the following strain eld

X1
2
2
2(X1 +X2 )

X 2X2 2
+X

0
0

X1
2
2
2(X1 +X2 )

0
0

(4.144)

does there exist a single-valued continuous displacement eld?


Solution:

2 E11
2
X2

E11
X2
E12
2
X1
E22
2
X1
2 E22
+
2
X1

=
=

2
2
2
2
(X1 + X2 ) X2 (2X2 )
X2 X 1
=
2 + X 2 )2
2 + X 2 )2
(X1
(X1
2
2
2
2
2
2
(X1 + X2 ) X1 (2X1 )
X2 X 1
=
2
2
2
2
(X1 + X2 )2
(X1 + X2 )2

(4.145-a)
(4.145-b)
(4.145-c)

2 E12
X1 X2

(4.145-d)

Actually, it can be easily veried that the unique displacement eld is given by
u1 = arctan

X2
;
X1

u2 = 0;

u3 = 0

(4.146)

to which we could add the rigid body displacement eld (if any).

4.6

Lagrangian Stresses; Piola Kircho Stress Tensors

In Sect. 2.2 the discussion of stress applied to the deformed conguration dA (using spatial coordiantes
x), that is the one where equilibrium must hold. The deformed conguration being the natural one in
which to characterize stress. Hence we had

96

df

= tdA

t = Tn

(4.147-a)
(4.147-b)

(note the use of T instead of ). Hence the Cauchy stress tensor was really dened in the Eulerian
space.
97 However, there are certain advantages in referring all quantities back to the undeformed conguration
(Lagrangian) of the body because often that conguration has geometric features and symmetries that
are lost through the deformation.
98 Hence, if we were to dene the strain in material coordinates (in terms of X), we need also to express
the stress as a function of the material point X in material coordinates.

Victor Saouma

Introduction to Continuum Mechanics

Draft

4.6 Lagrangian Stresses; Piola Kircho Stress Tensors

4.6.1

433

First

99 The rst Piola-Kircho stress tensor T0 is dened in the undeformed geometry in such a way that it
results in the same total force as the traction in the deformed conguration (where Cauchys stress
tensor was dened). Thus, we dene
(4.148)
df t0 dA0

where t0 is a pseudo-stress vector in that being based on the undeformed area, it does not describe
the actual intensity of the force, however it has the same direction as Cauchys stress vector t.
100 The rst Piola-Kircho stress tensor (also known as Lagrangian Stress Tensor) is thus the linear
transformation T0 such that
(4.149)
t0 = T0 n0

and for which


df = t0 dA0 = tdA t0 =

dA
t
ddA0

(4.150)

using Eq. 4.147-b and 4.149 the preceding equation becomes


dA
dA
Tn = T
n
dA0
dA0

T0 n0 =
and using Eq. 4.36 dAn = dA0 (det F) F1

(4.151)

n0 we obtain

T0 n0 = T(det F) F1

n0

(4.152)

the above equation is true for all n0 , therefore


T0

(det F)T F1
1
1
T0 FT or Tij =
(T0 )im Fjm
(det F)
(det F)

(4.153)
(4.154)

The rst Piola-Kircho stress tensor is not symmetric in general, and is not energitically correct.
That is multiplying this stress tensor with the Green-Lagrange tensor will not be equal to the product
of the Cauchy stress tensor multiplied by the deformation strain tensor.

101

102 To determine the corresponding stress vector, we solve for T0 rst, then for dA0 and n0 from
1
dA0 n0 = det F FT n (assuming unit area dA), and nally t0 = T0 n0 .

4.6.2

Second

The second Piola-Kircho stress tensor, T is formulated dierently. Instead of the actual force df
related to the force df in the same way that a material vector dX at X is
on dA, it gives the force df
related by the deformation to the corresponding spatial vector dx at x. Thus, if we let
103

d
f
df

= 0
tdA
and
= Fd
f

(4.155-a)
(4.155-b)

where d is the pseudo dierential force which transforms, under the deformation gradient F, the
f
(actual) dierential force df at the deformed position (note similarity with dx = FdX). Thus, the
pseudo vector t is in general in a diernt direction than that of the Cauchy stress vector t.
104

The second Piola-Kircho stress tensor is a linear transformation T such that


= Tn0
t

Victor Saouma

(4.156)
Introduction to Continuum Mechanics

Draft
434

KINEMATIC

thus the preceding equations can be combined to yield

df = FTn0 dA0

(4.157)

df = t0 dA0 = T0 n0 ddA0

(4.158)

we also have from Eq. 4.148 and 4.149

and comparing the last two equations we note that

T = F1 T0

(4.159)

which gives the relationship between the rst Piola-Kircho stress tensor T0 and the second Piola
Kircho stress tensor T.
Finally the relation between the second Piola-Kircho stress tensor and the Cauchy stress tensor can
be obtained from the preceding equation and Eq. 4.153

105

T = (det F) F1 T F1

(4.160)

and we note that this second Piola-Kircho stress tensor is always symmetric (if the Cauchy stress tensor
is symmetric). It can also be shown that it is energitically correct.
106

To determine the corresponding stress vector, we solve for T rst, then for dA0 and n0 from dA0 n0 =

= Tn0 .
(assuming unit area dA), and nally t

1
T
det F F n

Example 4-14: Piola-Kircho Stress Tensors

4.7

Hydrostatic and Deviatoric Strain

85 The lagrangian and Eulerian linear strain tensors can each be split into spherical and deviator
tensor as was the case for the stresses. Hence, if we dene

1
1
e = tr E
3
3

(4.161)

then the components of the strain deviator E are given by


1
1
Eij = Eij eij or E = E e1
3
3

(4.162)

We note that E measures the change in shape of an element, while the spherical or hydrostatic strain
1
3 e1 represents the volume change.

Victor Saouma

Introduction to Continuum Mechanics

Draft

4.7 Hydrostatic and Deviatoric Strain

435

mpiola.nb
2

mpiola.nb
3

MatrixForm@Transpose@FD . n detFD

PiolaKirchoff Stress Tensors


Second PiolaKirchoff Stress Tensor

i0y
j z
The deformed4configuration of a body is described by x1 =X 1 2, x2 =X2 /2, x3 =4X3 ; If the Cauchy stress tensor is
j z
j z
j z
100 0 0 y = Inverse@FD . Tfirst
Tsecond
i j z
j k0{ z
j
z
j 0 0 0 z MPa; What are the corresponding first and second PiolaKirchoff stress tensors, and calculate the
z
j
given byj
z
j
z
z
j
k 0 0 0{
25
980, 0, 0 on n0 we obtain
Thus n0 =e2 and using t0<,0the e3 plane in the 80, 0, 0<=
respective stress tensors =T 90, , 0=, deformed state.
4

F tensor

t01st = MatrixForm@Tfirst . 80, 1, 0<D

MatrixForm@%D
i 0 y
z
j
j 0 z
j
z
j 0 z0 0
j
i 25 z
CST = 880, y 0<, 80, 0, 0<, 80, 0, 100<<
0,
j
z
{
k
j
z
25
j
z
j
z
j 0 0 z
4
j
z
j
z
k 0 0 0<, 80,
880, 0, 0 {
We note that this vector is in the0, 0<, 80, 0, 100<<
same direction as the Cauchy stress vector, its magnitude is one fourth of that of the
Cauchy stress vector, because the undeformed area is 4 times that of the deformed area

F = 881 2, 0,
Cuchy stress vector 0<, 80, 0, 1 2<, 80, 4, 0<<
PseudoStress vector associated with the Second PiolaKirchoff stress
tensor
Can be obtained from t=CST n
1
1
99 , 0, 0=, 90, 0, - =, 80, 4, 0<=
2
2

tcauchy == MatrixForm@Tsecond 0, 1<D 0<D


t0second MatrixForm@CST . 80, . 80, 1,
Finverse = Inverse@FD
y
i 0 y
i 0 z
z
j
z
j 25 z
z
j 0 z
z
j z
z
j 4 z
j
j 100z
j
z
{
k
1
0 { 0<, 90, 0, =, 80, -2, 0<=
k
982, 0,
4
We see that this pseudo stress vector is in a different direction from that of the Cauchy stress vector (and we note that
the tensor F transforms e2 into e3 ).

PseudoStress vector associated with the First PiolaKirchoff stress tensor


First PiolaKirchoff Stress Tensor
FT n
For a unit area in the deformed state in the e3 direction, its undeformed area dA0 n0 is given by dA0 n0 =
det F
TfirstDet@FD CST . Transpose@FinverseD
detF = = Det@FD
880, 0, 0<, 80, 0, 0<, 80, 25, 0<<
1

MatrixForm@%D
n = 80, 0, 1<
i0 0 0y
j
j 0 0, 1< z
z
j
80, 0 0 z
z
j
z
j
k 0 25 0 {

Victor Saouma

Introduction to Continuum Mechanics

Draft
436

KINEMATIC

III

II

Figure 4.8: Mohr Circle for Strain

4.8

Principal Strains, Strain Invariants, Mohr Circle

86 Determination of the principal strains (E(3) < E(2) < E(1) , strain invariants and the Mohr circle for
strain parallel the one for stresses (Sect. 2.4) and will not be repeated here.

3 IE 2 IIE IIIE = 0

(4.163)

where the symbols IE , IIE and IIIE denote the following scalar expressions in the strain components:
IE
IIE

=
=
=
=

IIIE

87

E11 + E22 + E33 = Eii = tr E


2
2
2
(E11 E22 + E22 E33 + E33 E11 ) + E23 + E31 + E12
1
1
1 2
(Eij Eij Eii Ejj ) = Eij Eij IE
2
2
2
1
2
(E : E IE )
2
1
detE = eijk epqr Eip Ejq Ekr
6

(4.166)
(4.167)
(4.168)

In terms of the principal strains, those invariants can be simplied into


IE

E(1) + E(2) + E(3)

(4.169)

IIE
IIIE
88

(4.164)
(4.165)

=
=

(E(1) E(2) + E(2) E(3) + E(3) E(1) )


E(1) E(2) E(3)

(4.170)
(4.171)

The Mohr circle uses the Engineering shear strain denition of Eq. 4.91, Fig. 4.8
Example 4-15: Strain Invariants & Principal Strains

Victor Saouma

Introduction to Continuum Mechanics

Draft

4.8 Principal Strains, Strain Invariants, Mohr Circle

437

Determine the planes of principal strains for the following strain tensor

1
3 0

3 0 0
0
0 1

(4.172)

Solution:
The strain invariants are given by
IE
IIE
IIIE

= Eii = 2
1
(Eij Eij Eii Ejj ) = 1 + 3 = +2
=
2
= |Eij | = 3

(4.173-a)
(4.173-b)
(4.173-c)

The principal strains by

3
0
0

0
0
1

1 + 13
(1 )
2

1 + 13
= 2.3
(1) =
2
(2) = 1

1 13
= 1.3
(3) =
2

Eij ij

=
E(1)

E(2)

E(3)

The eigenvectors for E(1) =

1+2 13

3
0

1+2 13
0

(4.174-a)

13

(4.174-b)

(4.174-c)
(4.174-d)
(4.174-e)

1+ 13
2

give the principal directions n(1) :

(1)
(1)
(1)
1 1+2 13 n1 + 3n2

0
n1

(1)
(1)
(1)
3n1 1+2 13 n2
=
0
n2
(1)

(1)
n3
1 1+2 13

1 1+ 13 n
3


0
0
=

(4.175)

which gives
(1)
n1
(1)

n3

1 + 13 (1)
n2
2 3

= 0

n(1) n(1)

n(1)

(4.176-a)

0.8

0.6

1 1

3
0

3
1
0

(1)

n2

= 1 n1 = 0.8;
2

For the second eigenvector (2) = 1:

(4.176-b)

1 + 2 13 + 13
+1
12

(2)
n1
0

(2)

0
n
2
(2)
11
n3

(4.176-c)
(4.176-d)

(2)

3n
(2) 2 (2)
=
3n1 n2


0


0
0
=

(4.177)

which gives (with the requirement that n(2) n(2) = 1)


n(2) =
Victor Saouma

(4.178)

Introduction to Continuum Mechanics

Draft
438

KINEMATIC

Finally, the third eigenvector can be obrained by the same manner, but more easily from
e1
n(3) = n(1) n(2) = det 0.8
0
Therefore

e2
0.6
0

(1)
0.8
n

(2)
aj =
= 0
n
i
(3)
0.6
n

and this results can be checked via

0.8 0.6 0
1

0
1 3
[a][E][a]T = 0
0.6 0.8 0
0

0.8
3 0
0 0 0.6
0
0 1

e3
0
1

= 0.6e1 0.8e2

(4.179)

0.6 0
0
1
0.8 0

(4.180)


0.6
2.3
0.8 = 0
0
0

0
0
1

0
1
0

0
0
1.3

(4.181)

Example 4-16: Mohrs Circle


Construct the Mohrs circle for the following

0
0
0

plane strain case:

0
0
3
5

3 3

(4.182)

Solution:

1
2

60o
1

We note that since E(1) = 0 is a principal value for plane strain, ttwo of the circles are drawn as
shown.

Victor Saouma

Introduction to Continuum Mechanics

Draft

4.9 Initial or Thermal Strains

4.9
89

439

Initial or Thermal Strains

Initial (or thermal strain) in 2D:


T
0

ij =

0
T

= (1 + )

Plane Stress

T
0

0
T

(4.183)

Plane Strain

note there is no shear strains caused by thermal expansion.

4.10

Experimental Measurement of Strain

90 Typically, the transducer to measure strains in a material is the strain gage. The most common type
of strain gage used today for stress analysis is the bonded resistance strain gage shown in Figure 4.9.

Figure 4.9: Bonded Resistance Strain Gage


91 These gages use a grid of ne wire or a metal foil grid encapsulated in a thin resin backing. The gage
is glued to the carefully prepared test specimen by a thin layer of epoxy. The epoxy acts as the carrier
matrix to transfer the strain in the specimen to the strain gage. As the gage changes in length, the tiny
wires either contract or elongate depending upon a tensile or compressive state of stress in the specimen.
The cross sectional area will increase for compression and decrease in tension. Because the wire has an
1
electrical resistance that is proportional to the inverse of the cross sectional area, R A , a measure of
the change in resistance can be converted to arrive at the strain in the material.
92 Bonded resistance strain gages are produced in a variety of sizes, patterns, and resistance. One type
of gage that allows for the complete state of strain at a point in a plane to be determined is a strain
gage rosette. It contains three gages aligned radially from a common point at dierent angles from each
other, as shown in Figure 4.10. The strain transformation equations to convert from the three strains a
t any angle to the strain at a point in a plane are:

a
b
c

=
=
=

cos2 a +
2

cos b +
cos2 c +
x
x

sin2 a + xy sin a cos a


2

sin b + xy sin b cos b


sin2 c + xy sin c cos c
y
y

(4.184)
(4.185)
(4.186)

93 When the measured strains


a , b , and c , are measured at their corresponding angles from the
reference axis and substituted into the above equations the state of strain at a point may be solved,
namely, x , y , and xy . In addition the principal strains may then be computed by Mohrs circle or the
principal strain equations.
94 Due to the wide variety of styles of gages, many factors must be considered in choosing the right
gage for a particular application. Operating temperature, state of strain, and stability of installation all
inuence gage selection. Bonded resistance strain gages are well suited for making accurate and practical
strain measurements because of their high sensitivity to strains, low cost, and simple operation.

Victor Saouma

Introduction to Continuum Mechanics

Draft
440

KINEMATIC

Figure 4.10: Strain Gage Rosette


95 The measure of the change in electrical resistance when the strain gage is strained is known as the
gage factor. The gage factor is dened as the fractional change in resistance divided by the fractional
R
R
change in length along the axis of the gage. GF = L Common gage factors are in the range of 1.5-2
L
for most resistive strain gages.
96 Common strain gages utilize a grid pattern as opposed to a straight length of wire in order to reduce
the gage length. This grid pattern causes the gage to be sensitive to deformations transverse to the
gage length. Therefore, corrections for transverse strains should be computed and applied to the strain
data. Some gages come with the tranverse correction calculated into the gage factor. The transverse
sensitivity factor, Kt , is dened as the transverse gage factor divided by the longitudinal gage factor.
GFtransverse
Kt = GFlongitudinal These sensitivity values are expressed as a percentage and vary from zero to ten
percent.
97 A nal consideration for maintaining accurate strain measurement is temperature compensation. The
resistance of the gage and the gage factor will change due to the variation of resistivity and strain sensitivity with temperature. Strain gages are produced with dierent temperature expansion coecients.
In order to avoid this problem, the expansion coecient of the strain gage should match that of the
specimen. If no large temperature change is expected this may be neglected.
98 The change in resistance of bonded resistance strain gages for most strain measurements is very small.
From a simple calculation, for a strain of 1 ( = 106 ) with a 120 gage and a gage factor of 2, the
change in resistance produced by the gage is R = 1 106 120 2 = 240 106 . Furthermore,
it is the fractional change in resistance that is important and the number to be measured will be in
the order of a couple of ohms. For large strains a simple multi-meter may suce, but in order to
acquire sensitive measurements in the range a Wheatstone bridge circuit is necessary to amplify this
resistance. The Wheatstone bridge is described next.

4.10.1

Wheatstone Bridge Circuits

99 Due to their outstanding sensitivity, Wheatstone bridge circuits are very advantageous for the measurement of resistance, inductance, and capacitance. Wheatstone bridges are widely used for strain
measurements. A Wheatstone bridge is shown in Figure 4.11. It consists of 4 resistors arranged in a
diamond orientation. An input DC voltage, or excitation voltage, is applied between the top and bottom
of the diamond and the output voltage is measured across the middle. When the output voltage is zero,
the bridge is said to be balanced. One or more of the legs of the bridge may be a resistive transducer,
such as a strain gage. The other legs of the bridge are simply completion resistors with resistance equal
to that of the strain gage(s). As the resistance of one of the legs changes, by a change in strain from
a resistive strain gage for example, the previously balanced bridge is now unbalanced. This unbalance
causes a voltage to appear across the middle of the bridge. This induced voltage may be measured with
a voltmeter or the resistor in the opposite leg may be adjusted to re-balance the bridge. In either case
the change in resistance that caused the induced voltage may be measured and converted to obtain the

Victor Saouma

Introduction to Continuum Mechanics

Draft

4.10 Experimental Measurement of Strain

441

engineering units of strain.

Figure 4.11: Quarter Wheatstone Bridge Circuit

4.10.2

Quarter Bridge Circuits

If a strain gage is oriented in one leg of the circuit and the other legs contain xed resistors as shown
Rgage
in Figure 4.11, the circuit is known as a quarter bridge circuit. The circuit is balanced when R1 = R3 .
R2

100

When the circuit is unbalanced Vout = Vin ( R1R1 2


+R

Rgage
Rgage +R3 ).

Wheatstone bridges may also be formed with two or four legs of the bridge being composed of
resistive transducers and are called a half bridge and full bridge respectively. Depending upon the type
of application and desired results, the equations for these circuits will vary as shown in Figure 4.12. Here
E0 is the output voltage in mVolts, E is the excitation voltage in Volts, is strain and is Poissons
ratio.
101

In order to illustrate how to compute a calibration factor for a particular experiment, suppose a
single active gage in uniaxial compression is used. This will correspond to the upper Wheatstone bridge
conguration of Figure 4.12. The formula then is

102

Victor Saouma

Introduction to Continuum Mechanics

Draft
442

KINEMATIC

Figure 4.12: Wheatstone Bridge Congurations

Victor Saouma

Introduction to Continuum Mechanics

Draft

4.10 Experimental Measurement of Strain

443

F (103 )
E0
=
E
4 + 2F (106 )

(4.187)

The extra term in the denominator 2F (106 ) is a correction factor for non-linearity. Because
this term is quite small compared to the other term in the denominator it will be ignored. For most
measurements a gain is necessary to increase the output voltage from the Wheatstone bridge. The
gain relation for the output voltage may be written as V = GE0 (103 ), where V is now in Volts. so
Equation 4.187 becomes
103

V
EG(103 )
V

=
=

F (103 )
4
4
F EG

(4.188)

Here, Equation 4.188 is the calibration factor in units of strain per volt. For common values where
4
F = 2.07, G = 1000, E = 5, the calibration factor is simply (2.07)(1000)(5) or 386.47 microstrain per volt.

104

Victor Saouma

Introduction to Continuum Mechanics

Draft

KINEMATIC

Victor Saouma

Introduction to Continuum Mechanics

444

Draft
Chapter 5

MATHEMATICAL
PRELIMINARIES; Part III
VECTOR INTEGRALS
5.1
1

Integral of a Vector

The integral of a vector R(u) = R1 (u)e1 + R2 (u)e2 + R3 (u)e3 is dened as


R(u)du = e1

if a vector S(u) exists such that R(u) =

R1 (u)du + e2
d
du

R3 (u)du

(5.1)

(S(u)), then
d
(S(u)) du = S(u) + c
du

R(u)du =

5.2

R2 (u)du + e3

(5.2)

Line Integral

2 Given r(u) = x(u)e1 + y(u)e2 + z(u)e3 where r(u) is a position vector dening a curve C connecting
point P1 to P2 where u = u1 and u = u2 respectively, anf given A(x, y, z) = A1 e1 + A2 e2 + A3 e3 being
a vectorial function dened and continuous along C, then the integral of the tangential component of A
along C from P1 to P2 is given by

P2

Adr =

Adr =
C

P1

A1 dx + A2 dy + A3 dz

(5.3)

If A were a force, then this integral would represent the corresponding work.
3

If the contour is closed, then we dene the contour integral as


Adr =
C

It can be shown that if A =


P2

Adr

A1 dx + A2 dy + A3 dz

(5.4)

then
is independent of the path C connecting P1 to P2

(5.5-a)

along a closed contour line

(5.5-b)

P1

Adr = 0
C

Draft
52

5.3
5

MATHEMATICAL PRELIMINARIES; Part III VECTOR INTEGRALS

Integration by Parts

The integration by part formula is


b
a

5.4

u(x)v (x)dx = u(x)v(x)|a

v(x)u (x)dx

(5.6)

Gauss; Divergence Theorem

The divergence theorem (also known as Ostrogradskis Theorem) comes repeatedly in solid mechanics
and can be stated as follows:

vd =

v.nd or

vi,i d =

vi ni d

(5.7)

That is the integral of the outer normal component of a vector over a closed surface (which is the
volume ux) is equal to the integral of the divergence of the vector over the volume bounded by the
closed surface.
7

For 2D-1D transformations, we have


qdA =
A

(5.8)

This theorem is sometime refered to as Greens theorem in space.

5.5
9

qT nds
s

Stokes Theorem

Stokes theorem states that


( A)ndS =

Adr =
C

( A)dS

(5.9)

where S is an open surface with two faces conned by C

5.6
10

Green; Gradient Theorem

Greens theorem in plane is a special case of Stokes theorem.

(Rdx + Sdy) =

Example 5-1:

Victor Saouma

R
S

x
y

dxdy

(5.10)

Physical Interpretation of the Divergence Theorem

Introduction to Continuum Mechanics

Draft

5.6 Green; Gradient Theorem

53
Z

E
Z

C
V

P(X,Y,Z) H

X
B

a)
S

dV=dxdydz

n
Vt

dS

dS

b)
c)

Figure 5.1: Physical Interpretation of the Divergence Theorem


Provide a physical interpretation of the Divergence Theorem.
Solution:
A uid has a velocity eld v(x, y, z) and we rst seek to determine the net inow per unit time per
unit volume in a parallelepiped centered at P (x, y, z) with dimensions x, y, z, Fig. 5.1-a.
vx |x,y,z

vx

vx

xx/2,y,z

vx

vx

x+x/2,y,z

(5.11-a)

1 vx
x AFED
2 x
1 vx
x GHCB
vx +
2 x

(5.11-b)
(5.11-c)

The net inow per unit time across the x planes is


Vx

=
=

vx +

1 vx
1 vx
x yz vx
x yz
2 x
2 x

vx
xyz
x

(5.12-a)
(5.12-b)

Similarly
Vy

Vz

vy
xyz
y
vz
xyz
z

(5.13-a)
(5.13-b)

Hence, the total increase per unit volume and unit time will be given by
vx
x

vy
y

vz
z

xyz

xyz
= div v =

(5.14)

Furthermore, if we consider the total of uid crossing dS during t, Fig. 5.1-b, it will be given by
(vt)ndS = vndSt or the volume of uid crossing dS per unit time is vndS.
Victor Saouma

Introduction to Continuum Mechanics

Draft
54

MATHEMATICAL PRELIMINARIES; Part III VECTOR INTEGRALS

Thus for an arbitrary volume, Fig. 5.1-c, the total amount of uid crossing a closed surface S per
vdV (Eq. 5.14), thus

vndS. But this is equal to

unit time is
S

vdV

vndS =
S

(5.15)

which is the divergence theorem.

Victor Saouma

Introduction to Continuum Mechanics

Draft
Chapter 6

FUNDAMENTAL LAWS of
CONTINUUM MECHANICS
6.1

Introduction

We have thus far studied the stress tensors (Cauchy, Piola Kircho), and several other tensors which
describe strain at a point. In general, those tensors will vary from point to point and represent a tensor
eld.

We have also obtained only one dierential equation, that was the compatibility equation.

In this chapter, we will derive additional dierential equations governing the way stress and deformation
vary at a point and with time. They will apply to any continuous medium, and yet we will not have
enough equations to determine unknown tensor eld. For that we need to wait for the next chapter
where constitututive laws relating stress and strain will be introduced. Only with constitutive equations
and boundary and initial conditions would we be able to obtain a well dened mathematical problem to
solve for the stress and deformation distribution or the displacement or velocity elds.

In this chapter we shall derive dierential equations expressing locally the conservation of mass, momentum and energy. These dierential equations of balance will be derived from integral forms of the
equation of balance expressing the fundamental postulates of continuum mechanics.

6.1.1

Conservation Laws

Conservation laws constitute a fundamental component of classical physics. A conservation law establishes a balance of a scalar or tensorial quantity in voulme V bounded by a surface S. In its most
general form, such a law may be expressed as

d
dt

AdV +

dS

Rate of variation

Exchange by Diffusion

AdV

(6.1)

V
Source

where A is the volumetric density of the quantity of interest (mass, linear momentum, energy, ...) a,
A is the rate of volumetric density of what is provided from the outside, and is the rate of surface
density of what is lost through the surface S of V and will be a function of the normal to the surface n.
Hence, we read the previous equation as: The input quantity (provided by the right hand side) is equal
to what is lost across the boundary, and to modify A which is the quantity of interest. The dimensions
of various quantities are given by
6

dim(a)

dim(AL3 )

(6.2-a)

Draft
62

FUNDAMENTAL LAWS of CONTINUUM MECHANICS


dim()
dim(A)

=
=

dim(AL2 t1 )
3 1

dim(AL

(6.2-b)

(6.2-c)

Hence this chapter will apply the previous conservation law to mass, momentum, and energy. the
resulting dierential equations will provide additional interesting relation with regard to the imcompressibiltiy of solids (important in classical hydrodynamics and plasticity theories), equilibrium and
symmetry of the stress tensor, and the rst law of thermodynamics.

The enunciation of the preceding three conservation laws plus the second law of thermodynamics,
constitute what is commonly known as the fundamental laws of continuum mechanics.

6.1.2

Fluxes

Prior to the enunciation of the rst conservation law, we need to dene the concept of ux across a
bounding surface.

10 The ux across a surface can be graphically dened through the consideration of an imaginary surface
xed in space with continuous medium owing through it. If we assign a positive side to the surface,
and take n in the positive sense, then the volume of material owing through the innitesimal surface
area dS in time dt is equal to the volume of the cylinder with base dS and slant height vdt parallel to
the velocity vector v, Fig. 6.1 (If vn is negative, then the ow is in the negative direction). Hence, we

vdt

vn dt

n
dS

Figure 6.1: Flux Through Area dS


dene the volume ux as
Volume Flux =

vndS =
S

vj nj dS

(6.3)

where the last form is for rectangular cartesian components.


11

We can generalize this denition and dene the following uxes per unit area through dS:

Victor Saouma

Introduction to Continuum Mechanics

Draft

6.2 Conservation of Mass; Continuity Equation

Mass Flux

63

vndS =

vj nj dS

Momentum Flux

v(vn)dS =

=
S

Kinetic Energy Flux


Heat ux

qndS =
S

Electric ux

6.2.1

(6.5)

1
vi vi vj nj dS
S2

(6.6)

qj nj dS

(6.7)

Jj nj dS

(6.8)

JndS =
S

6.2

vk vj nj dS
S

1 2
v (vn)dS =
S2

(6.4)

Conservation of Mass; Continuity Equation


Spatial Form

12 If we consider an arbitrary volume V , xed in space, and bounded by a surface S. If a continuous


medium of density lls the volume at time t, then the total mass in V is

(x, t)dV

M=

(6.9)

where (x, t) is a continuous function called the mass density. We note that this spatial form in terms
of x is most common in uid mechanics.
13

The rate of increase of the total mass in the volume is


M
=
t

dV
t

(6.10)

14 The Law of conservation of mass requires that the mass of a specic portion of the continuum
remains constant. Hence, if no mass is created or destroyed inside V , then the preceding equation must
eqaul the inow of mass (of ux) through the surface. The outow is equal to vn, thus the inow
will be equal to vn.

(vn )dS =
S

must be equal to

M
t .

vndS =

(v)dV

+
t

(v) dV = 0

(6.11)

Thus
V

(6.12)

since the integral must hold for any arbitrary choice of dV , then we obtain

+
t
15

(v) or

(vi )
+
=0
t
xi

(6.13)

The chain rule will in turn give


vi
(vi )

=
+ vi
xi
xi
xi

(6.14)

16 It can be shown that the rate of change of the density in the neighborhood of a particle instantaneously
at x by

d
=
+ v =
+ vi
(6.15)
dt
t
t
xi
where the rst term gives the local rate of change of the density in the neighborhood of the place of x,
while the second term gives the convective rate of change of the density in the neighborhood of a

Victor Saouma

Introduction to Continuum Mechanics

Draft
64

FUNDAMENTAL LAWS of CONTINUUM MECHANICS

particle as it moves to a place having a dierent density. The rst term vanishes in a steady ow, while
the second term vanishes in a uniform ow.
17

Upon substitution in the last three equations, we obtain the continuity equation
vi
d
d
+
+ v = 0
= 0 or
dt
xi
dt

(6.16)

The vector form is independent of any choice of coordinates. This equation shows that the divergence of
the velocity vector eld equals (1/)(d/dt) and measures the rate of ow of material away from the
particle and is equal to the unit rate of decrease of density in the neighborhood of the particle.
18 If the material is incompressible, so that the density in the neighborhood of each material particle
remains constant as it moves, then the continuity equation takes the simpler form

vi
= 0 or
xi

v = 0

(6.17)

this is the condition of incompressibility

6.2.2

Material Form

19 If material coordinates X are used, the conservation of mass, and using Eq. 4.41 (dV = |J|dV0 ),
implies

(X, t0 )dV0 =
V0

(x, t)dV =

(x, t)|J|dV0

(6.18)

V0

or
[0 |J|]dV0 = 0

(6.19)

V0

and for an arbitrary volume dV0 , the integrand must vanish. If we also suppose that the initial density
0 is everywhere positive in V0 (no empty spaces), and at time t = t0 , J = 1, then we can write
J = 0

(6.20)

d
(J) = 0
dt

(6.21)

or

which is the continuity equation due to Euler, or the Lagrangian dierential form of the continuity equation.
20 We note that this is the same equation as Eq. 6.16 which was expressed in spatial form. Those two
equations can be derived one from the other.

21

The more commonly used form if the continuity equation is Eq. 6.16.

6.3
6.3.1

Linear Momentum Principle; Equation of Motion


Momentum Principle

The momentum principle states that the time rate of change of the total momentum of a given set of
particles equals the vector sum of all external forces acting on the particles of the set, provided Newtons
Third Law applies. The continuum form of this principle is a basic postulate of continuum mechanics.

22

tdS +
S

Victor Saouma

bdV =
V

d
dt

vdV

(6.22)

Introduction to Continuum Mechanics

Draft

6.3 Linear Momentum Principle; Equation of Motion

65

Then we substitute ti = Tij nj and apply the divergence theorm to obtain

Tij
+ bi dV
xj
V
dvi
Tij
dV
+ bi
xj
dt

dvi
dV
dt

(6.23-a)
(6.23-b)

or for an arbitrary volume


dvi
Tij
or
+ bi =
xj
dt

T + b =

dv
dt

(6.24)

which is Cauchys (rst) equation of motion, or the linear momentum principle, or more simply
equilibrium equation.
23

When expanded in 3D, this equation yields:


T11
T12
T13
+
+
+ b1
x1
x2
x3
T22
T23
T21
+
+
+ b2
x1
x2
x3
T32
T33
T31
+
+
+ b3
x1
x2
x3

= 0
= 0

(6.25-a)

= 0

24 We note that these equations could also have been derived from the free body diagram shown in Fig.
6.2 with the assumption of equilibrium (via Newtons second law) considering an innitesimal element
of dimensions dx1 dx2 dx3 . Writing the summation of forces, will yield

(6.26)

Tij,j + bi = 0
where is the density, bi is the body force (including inertia).

Example 6-1: Equilibrium Equation


In the absence of body forces, does the following stress distribution
2

2x1 x2
0
x2 + (x2 x2 )
x
1

x2 + (x2 x2 )
0
2x1 x2
1
2
1
2
2
0
0
(x1 + x2 )

(6.27)

where is a constant, satisfy equilibrium?


Solution:

T1j
xj
T2j
xj
T3j
xj

=
=
=

T11
T12
T13
+
+
= 2x1 2x1 = 0
x1
x2
x3

T21
T22
T23
+
+
= 2x2 + 2x2 = 0
x1
x2
x3

T31
T32
T33
+
+
=0
x1
x2
x3

(6.28-a)
(6.28-b)
(6.28-c)

Therefore, equilibrium is satised.

Victor Saouma

Introduction to Continuum Mechanics

Draft
66

FUNDAMENTAL LAWS of CONTINUUM MECHANICS

yy
+

dy

yy d y
y

+
yx

yx y
d
y
xx
+

xx
xy

xx d x
x

xy x
d
x

+
xy
yx
yy
dx

Figure 6.2: Equilibrium of Stresses, Cartesian Coordinates

6.3.2

Moment of Momentum Principle

25 The moment of momentum principle states that the time rate of change of the total moment of
momentum of a given set of particles equals the vector sum of the moments of all external forces acting
on the particles of the set.

Thus, in the absence of distributed couples (this theory of Cosserat will not be covered in this
course) we postulate the same principle for a continuum as

26

(rt)dS +

(rb)dV =

6.3.2.1

d
dt

(6.29)

(rv)dV
V

Symmetry of the Stress Tensor

27 We observe that the preceding equation does not furnish any new dierential equation of motion. If
we substitute tn = Tn and the symmetry of the tensor is assumed, then the linear momentum principle
(Eq. 6.24) is satised.
28 Alternatively, we may start by using Eq. 1.18 (ci = ijk aj bk ) to express the cross product in indicial
form and substitute above:

(rmn xm tn )dS +
S

d
dt

(rmn xm bn )dV =
V

(rmn xm vn )dV

(6.30)

we then substitute tn = Tjn nj , and apply Gauss theorem to obtain


rmn
V

xm Tjn
+ xm bn dV =
xj

rmn
V

d
(xm vn )dV
dt

(6.31)

but since dxm /dt = vm , this becomes


rmn xm
V

Victor Saouma

Tjn
+ bn
xj

+ mj Tjn dV =

rmn vm vn + xm
V

dvn
dt

dV

(6.32)

Introduction to Continuum Mechanics

Draft

6.4 Conservation of Energy; First Principle of Thermodynamics

67

but rmn vm vn = 0 since vm vn is symmetric in the indeces mn while rmn is antisymmetric, and the last
term on the right cancels with the rst term on the left, and nally with mj Tjn = Tmn we are left with
rmn Tmn dV = 0

(6.33)

rmn Tmn = 0

(6.34)

or for an arbitrary volume V ,

at each point, and this yields


for r
for r
for r

= 1
= 2
= 3

T23 T32
T31 T13
T12 T21

=
=
=

0
0
0

(6.35)

establishing the symmetry of the stress matrix without any assumption of equilibrium or of uniformity
of stress distribution as was done in Sect. 2.3.
29

The symmetry of the stress matrix is Cauchys second law of motion (1827).

6.4

Conservation of Energy; First Principle of Thermodynamics

30 The rst principle of thermodynamics relates the work done on a (closed) system and the heat transfer
into the system to the change in energy of the system. We shall assume that the only energy transfers
to the system are by mechanical work done on the system by surface traction and body forces, by heat
transfer through the boundary.

6.4.1

Spatial Gradient of the Velocity

31 We dene L as the spatial gradient of the velocity and in turn this gradient can be decomposed
into a symmetric rate of deformation tensor D (or stretching tensor) and a skew-symmeteric
tensor W called the spin tensor or vorticity tensor1 .

Lij

vi,j or L = v

D+W
1
(v x +
=
2

(6.36)

(6.37)
1
(v
x v) and W =
2

x v)

(6.38)

this term will be used in the derivation of the rst principle.

6.4.2

First Principle

If mechanical quantities only are considered, the principle of conservation of energy for the
continuum may be derived directly from the equation of motion given by Eq. 6.24. This is accomplished
by taking the integral over the volume V of the scalar product between Eq. 6.24 and the velocity vi .

32

vi Tji,j dV +
1 Note

similarity with Eq. 4.111-b.

Victor Saouma

bi vi dV =
V

vi
V

dvi
dV
dt

(6.39)

Introduction to Continuum Mechanics

Draft
68

FUNDAMENTAL LAWS of CONTINUUM MECHANICS

If we consider the right hand side


vi
V

d
dvi
dV =
dt
dt

d
1
vi vi dV =
2
dt

dK
1 2
v dV =
2
dt

(6.40)

which represents the time rate of change of the kinetic energy K in the continuum.
33 Also we have vi Tji,j = (vi Tji ),j vi,j Tji and from Eq. 6.37 we have vi,j = Lij + Wij . It can be shown

that since Wij is skew-symmetric, and T is symmetric, that Tij Wij = 0, and thus Tij Lij = Tij Dij . TD
is called the stress power.

34

If we consider thermal processes, the rate of increase of total heat into the continuum is given by
Q=

qi ni dS +

rdV

(6.41)

Q has the dimension of power, that is M L2 T 3 , and the SI unit is the Watt (W). q is the heat ux
per unit area by conduction, its dimension is M T 3 and the corresponding SI unit is W m2 . Finally,
r is the radiant heat constant per unit mass, its dimension is M T 3 L4 and the corresponding SI
unit is W m6 .
35

36

We thus have

dK
+
dt

Dij Tij dV =
V

(vi Tji ),j dV +


V

vi bi dV + Q

(6.42)

We next convert the rst integral on the right hand side to a surface integral by the divergence theorem
vdV = S v.ndS) and since ti = Tij nj we obtain
V
dK
+
dt

Dij Tij dV

vi ti dS +
S

dU
dK
+
dt
dt

vi bi dV + Q (6.43)
V

dW
+Q
dt

(6.44)

this equation relates the time rate of change of total mechanical energy of the continuum on the left side
to the rate of work done by the surface and body forces on the right hand side.
37 If both mechanical and non mechanical energies are to be considered, the rst principle states that the
time rate of change of the kinetic plus the internal energy is equal to the sum of the rate of work plus all
other energies supplied to, or removed from the continuum per unit time (heat, chemical, electromagnetic,
etc.).
38 For a thermomechanical continuum, it is customary to express the time rate of change of internal
energy by the integral expression
d
dU
=
udV
(6.45)
dt
dt V
where u is the internal energy per unit mass or specic internal energy. We note that U appears
only as a dierential in the rst principle, hence if we really need to evaluate this quantity, we need to
have a reference value for which U will be null. The dimension of U is one of energy dim U = M L2 T 2 ,
and the SI unit is the Joule, similarly dim u = L2 T 2 with the SI unit of Joule/Kg.

39

In terms of energy integrals, the rst principle can be rewritten as


Exchange

Rate of increae
d
dt

d
1
vi vi dV +
2
dt
dK
dt

udV =
V
dU
dt

Source

ti vi dS +
S
dW
dt

rdV

vi bi dV +
V

Exchange

Source
V

qi ni dS

(6.46)

S
Q

we apply Gauss theorem to convert the surface integral, collect terms and use the fact that dV is arbitrary
to obtain
Victor Saouma

Introduction to Continuum Mechanics

Draft

6.5 Equation of State; Second Principle of Thermodynamics

du
dt

69

T:D + r

(6.47)

Tij Dij + r

qj
xj

(6.48)

or

du

dt

This equation expresses the rate of change of internal energy as the sum of the stress power plus
the heat added to the continuum.

40

41 In ideal elasticity, heat transfer is considered insignicant, and all of the input work is assumed
converted into internal energy in the form of recoverable stored elastic strain energy, which can be
recovered as work when the body is unloaded.
42 In general, however, the major part of the input work into a deforming material is not recoverably
stored, but dissipated by the deformation process causing an increase in the bodys temperature and
eventually being conducted away as heat.

6.5

Equation of State; Second Principle of Thermodynamics

The complete characterization of a thermodynamic system is said to describe the state of a system
(here a continuum). This description is specied, in general, by several thermodynamic and kinematic
state variables. A change in time of those state variables constitutes a thermodynamic process.
Usually state variables are not all independent, and functional relationships exist among them through
equations of state. Any state variable which may be expressed as a single valued function of a set of
other state variables is known as a state function.
43

44 The rst principle of thermodynamics can be regarded as an expression of the interconvertibility of


heat and work, maintaining an energy balance. It places no restriction on the direction of the process.
In classical mechanics, kinetic and potential energy can be easily transformed from one to the other in
the absence of friction or other dissipative mechanism.

The rst principle leaves unanswered the question of the extent to which conversion process is reversible or irreversible. If thermal processes are involved (friction) dissipative processes are irreversible
processes, and it will be up to the second principle of thermodynamics to put limits on the direction of
such processes.

45

6.5.1

Entropy

46 The basic criterion for irreversibility is given by the second principle of thermodynamics through
the statement on the limitation of entropy production. This law postulates the existence of two
distinct state functions: the absolute temperature and S the entropy with the following properties:

1. is a positive quantity.
2. Entropy is an extensive property, i.e. the total entropy is in a system is the sum of the entropies
of its parts.
47

Thus we can write


ds = ds(e) + ds(i)

where ds

(e)

(6.49)
(i)

is the increase due to interaction with the exterior, and ds

is the internal increase, and

ds(e)

Victor Saouma

>

0 irreversible process

(6.50-a)

ds(i)

0 reversible process

(6.50-b)

Introduction to Continuum Mechanics

Draft
610

48

FUNDAMENTAL LAWS of CONTINUUM MECHANICS

Entropy expresses a variation of energy associated with a variation in the temperature.

6.5.1.1

Statistical Mechanics

49 In statistical mechanics, entropy is related to the probability of the occurrence of that state among
all the possible states that could occur. It is found that changes of states are more likely to occur in the
direction of greater disorder when a system is left to itself. Thus increased entropy means increased
disorder.
50 Hence Boltzmans principle postulates that entropy of a state is proportional to the logarithm of its
probability, and for a gas this would give

3
S = kN [ln V + ln] + C
2

(6.51)

where S is the total entropy, V is volume, is absolute temperature, k is Boltzmans constant, and C
is a constant and N is the number of molecules.
6.5.1.2
51

Classical Thermodynamics

In a reversible process (more about that later), the change in specic entropy s is given by
ds =

52

dq

(6.52)
rev

If we consider an ideal gas governed by


pv = R

(6.53)

where R is the gas constant, and assuming that the specic energy u is only a function of temperature
, then the rst principle takes the form
du = dq pdv

(6.54)

du = dq = cv d

(6.55)

and for constant volume this gives


wher cv is the specic heat at constant volume. The assumption that u = u() implies that cv is a
function of only and that
(6.56)
du = cv ()d
53

Hence we rewrite the rst principle as


dq = cv ()d + R

dv
v

(6.57)

or division by yields
s s0 =

p,v

dq
=
p0 ,v0

cv ()
0

v
d
+ R ln

v0

(6.58)

which gives the change in entropy for any reversible process in an ideal gas. In this case, entropy is a
state function which returns to its initial value whenever the temperature returns to its initial value that
is p and v return to their initial values.

Victor Saouma

Introduction to Continuum Mechanics

Draft

6.6 Balance of Equations and Unknowns

6.5.2
54

611

Clausius-Duhem Inequality

We restate the denition of entropy as heat divided by temperature, and write the second principle
d
dt

r
dV

Rate of Entropy Increase

dS
dt

q
ndS +
S

Sources

Exchange

Q
+ ;

(6.59)

Internal production

(6.60)

= 0 for reversible processes, and > 0 in irreversible ones. The dimension of S =

sdV is one
v

of energy divided by temperature or L2 M T 2 1 , and the SI unit for entropy is Joule/Kelvin.


55 The second principle postulates that the time rate of change of total entropy S in a continuum
occupying a volume V is always greater or equal than the sum of the entropy inux through the continuum
surface plus the entropy produced internally by body sources.
56 The previous inequality holds for any arbitrary volume, thus after transformation of the surface
integral into a volume integral, we obtain the following local version of the Clausius-Duhem inequality
which must holds at every point

ds
dt

Rate of Entropy Increase

57

Sources

(6.61)

Exchange

We next seek to express the Clausius-Duhem inequality in terms of the stress tensor,

1
q
=

thus

q q

1
ds

dt

1
1
=

q +

1
q
2

1
r
q +
2

(6.62)

(6.63)

but since is always positive,


ds
1
q + r + q
dt

where q + r is the heat input into V and appeared in the rst principle Eq. 6.47

du
= T:D + r
dt

(6.64)

(6.65)

1
q 0

(6.66)

hence, substituting, we obtain


T:D

6.6

ds
du

dt
dt

Balance of Equations and Unknowns

58 In the preceding sections several equations and unknowns were introduced. Let us count them. for
both the coupled and uncoupled cases.

Victor Saouma

Introduction to Continuum Mechanics

Draft
612

FUNDAMENTAL LAWS of CONTINUUM MECHANICS

d
vi
dt + xi = 0
Tij
dvi
xj + bi = dt
du = Tij Dij + r
dt

Coupled
1
3

Uncoupled
1
3

1
5

Continuity Equation
Equation of motion

xj Energy equation
j
Total number of equations

59 Assuming that the body forces bi and distributed heat sources r are prescribed, then we have the
following unknowns:

Density

Velocity (or displacement) vi (ui )


Stress components
Tij
Heat ux components
qi
Specic internal energy
u
Entropy density
s
Absolute temperature

Total number of unknowns


and in addition the Clausius-Duhem inequality
hold.

ds
dt

Coupled
1
3
6
3
1
1
1
16

1
div

Uncoupled
1
3
6
10

which governs entropy production must

We thus need an additional 16 5 = 11 additional equations to make the system determinate. These
will be later on supplied by:

60

6
3
2
11

constitutive equations
temperature heat conduction
thermodynamic equations of state
Total number of additional equations

61 The next chapter will thus discuss constitutive relations, and a subsequent one will separately discuss
thermodynamic equations of state.

62

We note that for the uncoupled case


1. The energy equation is essentially the integral of the equation of motion.
2. The 6 missing equations will be entirely supplied by the constitutive equations.
3. The temperature eld is regarded as known, or at most, the heat-conduction problem must be
solved separately and independently from the mechanical problem.

6.7

Elements of Heat Transfer

63 One of the relations which we will need is the one which relates temperature to heat ux.
This
constitutive realtion will be discussed in the next chapter under Fourriers law.
64 However to place the reader in the right frame of reference to understand Fourriers law, this section
will provide some elementary concepts of heat transfer.

65

There are three fundamental modes of heat transfer:

Conduction: takes place when a temperature gradient exists within a material and is governed by
Fouriers Law, Fig. 6.3 on q :

Victor Saouma

Introduction to Continuum Mechanics

Draft

6.7 Elements of Heat Transfer

613

Figure 6.3: Flux vector

qx

qy

T
x
T
ky
y
kx

(6.67)
(6.68)

where T = T (x, y) is the temperature eld in the medium, qx and qy are the componenets of the
heat ux (W/m2 or Btu/h-ft2 ), k is the thermal conductivity (W/m.o C or Btu/h-ft-o F) and T ,
x
T
are the temperature gradients along the x and y respectively. Note that heat ows from hot
y
to cool zones, hence the negative sign.
Convection: heat transfer takes place when a material is exposed to a moving uid which is at dierent
temperature. It is governed by the Newtons Law of Cooling
q = h(T T ) on c

(6.69)

where q is the convective heat ux, h is the convection heat transfer coecient or lm coecient
(W/m2 .o C or Btu/h-ft2 .o F). It depends on various factors, such as whether convection is natural or
forced, laminar or turbulent ow, type of uid, and geometry of the body; T and T are the surface
and uid temperature, respectively. This mode is considered as part of the boundary condition.
Radiation: is the energy transferred between two separated bodies at dierent temperatures by means
of electromagnetic waves. The fundamental law is the Stefan-Boltmans Law of Thermal Radiation
for black bodies in which the ux is proportional to the fourth power of the absolute temperature.,
which causes the problem to be nonlinear. This mode will not be covered.

6.7.1
66

Simple 2D Derivation

If we consider a unit thickness, 2D dierential body of dimensions dx by dy, Fig. 6.4 then
1. Rate of heat generation/sink is
I2 = Qdxdy

(6.70)

2. Heat ux across the boundary of the element is shown in Fig. ?? (note similarity with equilibrium
equation)
I1 =

qx +

Victor Saouma

qx
dx qx dx dy +
x

qy +

qy
qy
qx
dy qy dy dx =
dxdy +
dydx
y
x
y

(6.71)

Introduction to Continuum Mechanics

Draft
614

FUNDAMENTAL LAWS of CONTINUUM MECHANICS

6y +
q

qx

qy
y dy

qx +

qx
x dx

dy
?

6


qy

dx

Figure 6.4: Flux Through Sides of Dierential Element


Figure 6.5: *Flow through a surface
3. Change in stored energy is
d
.dxdy
(6.72)
dt
where we dene the specic heat c as the amount of heat required to raise a unit mass by one
degree.
I3 = c

67 From the rst law of thermodaynamics, energy produced I2 plus the net energy across the boundary
I1 must be equal to the energy absorbed I3 , thus

I1 + I2 I3
qy
d
qx
dxdy +
dydx + Qdxdy c dxdy
x
y
dt
I2

I1

6.7.2
68

(6.73-a)

(6.73-b)

I3

Generalized Derivation

The amount of ow per unit time into an element of volume and surface is
I1 =

q(n)d =

D .nd

(6.74)

where n is the unit exterior normal to , Fig. 6.5


69

Using the divergence theorem


vnd =

div vd

(6.75)

div (D )d

(6.76)

Eq. 6.74 transforms into


I1 =

Victor Saouma

Introduction to Continuum Mechanics

Draft

6.7 Elements of Heat Transfer

615

70 Furthermore, if the instantaneous volumetric rate of heat generation or removal at a point x, y, z


inside is Q(x, y, z, t), then the total amount of heat/ow produced per unit time is

I2 =

Q(x, y, z, t)d

(6.77)

Finally, we dene the specic heat of a solid c as the amount of heat required to raise a unit mass
by one degree. Thus if is a temperature change which occurs in a mass m over a time t, then the
corresponding amount of heat that was added must have been cm, or

71

I3 =

cd

(6.78)

where is the density, Note that another expression of I3 is t(I1 + I2 ).


72 The balance equation, or conservation law states that the energy produced I2 plus the net energy
across the boundary I1 must be equal to the energy absorbed I3 , thus

I1 + I2 I3

d
div (D ) + Q c
t

(6.79-a)

(6.79-b)

but since t and are both arbitrary, then

=0
t

(6.80)

(6.81)

qy

qx
+
+ Q = c
x
y
t

(6.82)

div (D ) + Q c
or

div (D ) + Q = c
This equation can be rewritten as

1. Note the similarity between this last equation, and the equation of equilibrium
xy
xx
+
+ bx
x
y
xy
yy
+
+ by
y
x

=
=

2 ux
t2
2 uy
m 2
t
m

(6.83-a)
(6.83-b)

2. For steady state problems, the previous equation does not depend on t, and for 2D problems, it
reduces to

kx
+
ky
+Q=0
(6.84)
x
x
y
y
3. For steady state isotropic problems,
2 2 2 Q
=0
+ 2 + 2 +
x2
y
z
k

(6.85)

which is Poissons equation in 3D.

Victor Saouma

Introduction to Continuum Mechanics

Draft
616

FUNDAMENTAL LAWS of CONTINUUM MECHANICS

4. If the heat input Q = 0, then the previous equation reduces to


2 2 2
+ 2 + 2 =0
x2
y
z

(6.86)

which is an Elliptic (or Laplace) equation. Solutions of Laplace equations are termed harmonic
functions (right hand side is zero) which is why Eq. 6.84 is refered to as the quasi-harmonic
equation.
5. If the function depends only on x and t, then we obtain
c

=
t
x

kx

+Q

(6.87)

which is a parabolic (or Heat) equation.

Victor Saouma

Introduction to Continuum Mechanics

Draft
Chapter 7

CONSTITUTIVE EQUATIONS;
Part I LINEAR
ceiinosssttuu
Hooke, 1676
Ut tensio sic vis
Hooke, 1678

7.1
7.1.1

Thermodynamic Approach
State Variables

The method of local state postulates that the thermodynamic state of a continuum at a given
point and instant is completely dened by several state variables (also known as thermodynamic
or independent variables). A change in time of those state variables constitutes a thermodynamic
process. Usually state variables are not all independent, and functional relationships exist among them
through equations of state. Any state variable which may be expressed as a single valued function of
a set of other state variables is known as a state function.

The time derivatives of these variables are not involved in the denition of the state, this postulate
implies that any evolution can be considered as a succession of equilibrium states (therefore ultra rapid
phenomena are excluded).

3 The thermodynamic state is specied by n + 1 variables 1 , 2 , , n and s where i are the


thermodynamic substate variables and s the specic entropy. The former have mechanical (or electromagnetic) dimensions, but are otherwise left arbitrary in the general formulation. In ideal elasticity
we have nine substate variables the components of the strain or deformation tensors.

The basic assumption of thermodynamics is that in addition to the n substate variables, just one
additional dimensionally independent scalar paramer suces to determine the specic internal energy
u. This assumes that there exists a caloric equation of state

u = u(s, , X)

In general the internal energy u can not be experimentally measured but rather its derivative.

(7.1)

Draft
72

CONSTITUTIVE EQUATIONS; Part I LINEAR

6 For instance we can dene the thermodynamic temperature and the thermodynamic tension j as

u
s

u
j

j = 1, 2, , n

(7.2)

s,i(i=j)

where the subscript outside the parenthesis indicates that the variables are held constant.
7 By extension Ai = i would be the thermodynamic force and its dimension depends on the
one of i .

7.1.2
8

Gibbs Relation

From the chain rule we can express


u
s

du
=
dt
9

ds
dp
+ p
dt
dt

(7.3)

substituting into Clausius-Duhem inequality of Eq. 6.66


ds
du

dt
dt

T:D
we obtain
T:D +

ds

dt

u
s

1
q 0

+ Ap

(7.4)

1
dp
q 0
dt

(7.5)

but the second principle must be satised for all possible evolution and in particular the one for which
d
D = 0, dtp = 0 and = 0 for any value of ds thus the coecient of ds is zero or
dt
dt
=

u
s

(7.6)

thus
T:D + Ap

1
dp
q 0
dt

(7.7)

and Eq. 7.3 can be rewritten as


ds
du
dp
= + p
dt
dt
dt
and if we adopt the dierential notation, we obtain Gibbs relation

(7.8)

(7.9)

du = ds + p dp

10

For uid, the Gibbs relation takes the form


du = ds pdv; and

u
s

;
v

u
v

(7.10)
s

where p is the thermodynamic pressure; and the thermodynamic tension conjugate to the specic volume
v is p, just as is conjugate to s.

Victor Saouma

Introduction to Continuum Mechanics

Draft

7.1 Thermodynamic Approach

7.1.3

73

Thermal Equation of State

11 From the caloric equation of state, Eq. 7.1, and the the denitions of Eq. 7.2 it follows that the
temperature and the thermodynamic tensions are functions of the thermodynamic state:

= (s, );

j = j (s, )

(7.11)

we assume the rst one to be invertible


s = s(, )

(7.12)

and substitute this into Eq. 7.1 to obtain an alternative form of the caloric equation of state with
corresponding thermal equations of state (obtained by simple substitution).
u = u(, , bX)
i = i (, , X)
i = i (, , X)

(7.13)
(7.14)
(7.15)

12 The thermal equations of state resemble stress-strain relations, but some caution is necessary in
interpreting the tesnisons as stresses and the j as strains.

7.1.4

Thermodynamic Potentials

13 Based on the assumed existence of a caloric equation of state, four thermodynamic potentials are
introduced, Table 7.1. Those potentials are derived through the Legendre-Fenchel transformation

Potential
Internal energy
Helmholtz free energy
Enthalpy
Free enthalpy

h
g

Relation to u
u
= u s
h = u j j
g = u s j j

Independent Variables
s, j
, j
s, j
, j

Table 7.1: Thermodynamic Potentials


on the basis of selected state variables best suited for a given problem.
14 By means of the preceding equations, any one of the potentials can be expressed in terms of any of
the four choices of state variables listed in Table 7.1.

15

In any actual or hypothetical change obeying the equations of state, we have


du

= ds + j dj

(7.16-a)

d = sd + j dj
dh = ds j dj
dg

(7.16-b)
(7.16-c)

= sd j dj

(7.16-d)

and from these dierentials we obtain the following partial derivative expressions
=
s=

u
;
s

h
;
s
g
;
=

=

Victor Saouma

j =
j =
j =
j =

u
j

s,i(i=j)

(7.17-a)

h
j

s,i(i=j)

g
j

(7.17-b)

(7.17-c)
(7.17-d)

Introduction to Continuum Mechanics

Draft
74

CONSTITUTIVE EQUATIONS; Part I LINEAR

where the free energy is the portion of the internal energy available for doing work at constant
temperature, the enthalpy h (as dened here) is the portion of the internal energy that can be released
as heat when the thermodynamic tensions are held constant.

7.1.5

Elastic Potential or Strain Energy Function

16 Green dened an elastic material as one for which a strain-energy function exists. Such a material
is called Green-elastic or hyperelastic if there exists an elastic potential function W or strain
energy function, a scalar function of one of the strain or deformation tensors, whose derivative with
respect to a strain component determines the corresponding stress component.

17

For the fully recoverable case of isothermal deformation with reversible heat conduction we have

TIJ = 0

EIJ

(7.18)

hence W = 0 is an elastic potential function for this case, while W = 0 u is the potential for adiabatic
isentropic case (s = constant).
18 Hyperelasticity ignores thermal eects and assumes that the elastic potential function always exists,
it is a function of the strains alone and is purely mechanical

W (E)

TIJ =
EIJ

(7.19)

and W (E) is the strain energy per unit undeformed volume. If the displacement gradients are
small compared to unity, then we obtain
Tij =

W
Eij

(7.20)

which is written in terms of Cauchy stress Tij and small strain Eij .
19 We assume that the elastic potential is represented by a power series expansion in the small-strain
components.
1
1
(7.21)
W = c0 + cij Eij + cijkm Eij Ekm + cijkmnp Eij Ekm Enp +
2
3
where c0 is a constant and cij , cijkm , cijkmnp denote tensorial properties required to maintain the invariant
property of W . Physically, the second term represents the energy due to residual stresses, the third one
refers to the strain energy which corresponds to linear elastic deformation, and the fourth one indicates
nonlinear behavior.

Neglecting terms higher than the second degree in the series expansion, then W is quadratic in terms
of the strains

20

c0 + c1 E11 + c2 E22 + c3 E33 + 2c4 E23 + 2c5 E31 + 2c6 E12


2
+ 1 c1111 E11 + c1122 E11 E22 + c1133 E11 E33 + 2c1123 E11 E23 + 2c1131 E11 E31 + 2c1112 E11 E12
2
2
+ 1 c2222 E22 + c2233 E22 E33 + 2c2223 E22 E23 + 2c2231 E22 E31 + 2c2212 E22 E12
2
1
2
+ 2 c3333 E33 + 2c3323 E33 E23 + 2c3331 E33 E31 + 2c3312 E33 E12
2
+2c2323 E23 + 4c2331 E23 E31 + 4c2312 E23 E12
2
+2c3131 E31 + 4c3112 E31 E12
2
+2c1212 E12
(7.22)
we require that W vanish in the unstrained state, thus c0 = 0.
W

21

We next apply Eq. 7.20 to the quadratic expression of W and obtain for instance
T12 =

W
= 2c6 + c1112 E11 + c2212 E22 + c3312 E33 + c1212 E12 + c1223 E23 + c1231 E31
E12

Victor Saouma

(7.23)

Introduction to Continuum Mechanics

Draft

7.2 Experimental Observations

75

if the stress must also be zero in the unstrained state, then c6 = 0, and similarly all the coecients in
the rst row of the quadratic expansion of W . Thus the elastic potential function is a homogeneous
quadratic function of the strains and we obtain Hookes law

7.2

Experimental Observations

We shall discuss two experiments which will yield the elastic Youngs modulus, and then the bulk
modulus. In the former, the simplicity of the experiment is surrounded by the intriguing character of
Hooke, and in the later, the bulk modulus is mathematically related to the Green deformation tensor
C, the deformation gradient F and the Lagrangian strain tensor E.

22

7.2.1

Hookes Law

23 Hookes Law is determined on the basis of a very simple experiment in which a uniaxial force is
applied on a specimen which has one dimension much greater than the other two (such as a rod). The
elongation is measured, and then the stress is plotted in terms of the strain (elongation/length). The
slope of the line is called Youngs modulus.
24 Hooke anticipated some of the most important discoveries and inventions of his time but failed to carry
many of them through to completion. He formulated the theory of planetary motion as a problem in
mechanics, and grasped, but did not develop mathematically, the fundamental theory on which Newton
formulated the law of gravitation.
His most important contribution was published in 1678 in the paper De Potentia Restitutiva. It
contained results of his experiments with elastic bodies, and was the rst paper in which the elastic
properties of material was discussed.

Take a wire string of 20, or 30, or 40 ft long, and fasten the upper part thereof to a nail,
and to the other end fasten a Scale to receive the weights: Then with a pair of compasses take
the distance of the bottom of the scale from the ground or oor underneath, and set down the
said distance, then put inweights into the said scale and measure the several stretchings of
the said string, and set them down. Then compare the several stretchings of the said string,
and you will nd that they will always bear the same proportions one to the other that the
weights do that made them.
This became Hookes Law
(7.24)

= E

25 Because he was concerned about patent rights to his invention, he did not publish his law when rst
discovered it in 1660. Instead he published it in the form of an anagram ceiinosssttuu in 1676 and
the solution was given in 1678. Ut tensio sic vis (at the time the two symbols u and v were employed
interchangeably to denote either the vowel u or the consonant v), i.e. extension varies directly with force.

7.2.2

Bulk Modulus

26 If, instead of subjecting a material to a uniaxial state of stress, we now subject it to a hydrostatic
pressure p and measure the change in volume V .

27

From the summary of Table 4.1 we know that:


V
det F

Victor Saouma

(det F)V0

=
det C =

(7.25-a)

det[I + 2E]

(7.25-b)

Introduction to Continuum Mechanics

Draft
76

CONSTITUTIVE EQUATIONS; Part I LINEAR

therefore,
V + V
= det[I + 2E]
V
we can expand the determinant of the tensor det[I + 2E] to nd

(7.26)

det[I + 2E] = 1 + 2IE + 4IIE + 8IIIE

(7.27)

IIE
IIIE since the rst term is linear in E, the second is quadratic, and
but for small strains, IE
the third is cubic. Therefore, we can approximate det[I+2E] 1+2IE , hence we dene the volumetric
dilatation as
V
e IE = tr E
V

(7.28)

this quantity is readily measurable in an experiment.

7.3
7.3.1
28

Stress-Strain Relations in Generalized Elasticity


Anisotropic

From Eq. 7.22 and 7.23 we obtain the stress-strain relation for homogeneous anisotropic material

c1111 c1112 c1133 c1112 c1123 c1131


E11
T11

T22

c2222 c2233 c2212 c2223 c2231


E22

T33
c3333 c3312 c3323 c3331
E33

=
c1212 c1223 c1231 2E12 (12 )
T12
(7.29)

T23

SYM.
c2323 c2331 2E23 (23 )

T31
c3131
2E31 (31 )
cijkm

Tij

Ekm

which is Hookes law for small strain in linear elasticity.


29

We also observe that for symmetric cij we retrieve Clapeyron formula


W =

1
Tij Eij
2

(7.30)

30 In general the elastic moduli cij relating the cartesian components of stress and strain depend on the
orientation of the coordinate system with respect to the body. If the form of elastic potential function
W and the values cij are independent of the orientation, the material is said to be isotropic, if not it
is anisotropic.

31

cijkm is a fourth order tensor resulting with




c1,2,1,1
c1,1,1,1 c1,1,1,2 c1,1,1,3
c1,1,2,1 c1,1,2,2 c1,1,2,3 c1,2,2,1

c1,1,3,1 c1,1,3,2 c1,1,3,3 c1,2,3,1

c2,1,1,1 c2,1,1,2 c2,1,1,3


c2,2,1,1

c2,1,2,1 c2,1,2,2 c2,1,2,3 c2,2,2,1

c2,1,3,1 c2,1,3,2 c2,1,3,3 c2,2,3,1

c3,2,1,1
c3,1,1,1 c3,1,1,2 c3,1,1,3

c3,1,2,1 c3,1,2,2 c3,1,2,3 c3,2,2,1


c3,1,3,1 c3,1,3,2 c3,1,3,3
c3,2,3,1

34 = 81 terms.
c1,2,1,2
c1,2,2,2
c1,2,3,2
c2,2,1,2
c2,2,2,2
c2,2,3,2
c3,2,1,2
c3,2,2,2
c3,2,3,2

c1,2,1,3
c1,2,2,3
c1,2,3,3
c2,2,1,3
c2,2,2,3
c2,2,3,3
c3,2,1,3
c3,2,2,3
c3,2,3,3

c1,3,1,1
c1,3,2,1
c1,3,3,1
c2,3,1,1
c2,3,2,1
c2,3,3,1
c3,3,1,1
c3,3,2,1
c3,3,3,1

c1,3,1,2
c1,3,2,2
c1,3,3,2
c2,3,1,2
c2,3,2,2
c2,3,3,2
c3,3,1,2
c3,3,2,2
c3,3,3,2

c1,3,1,3
c1,3,2,3
c1,3,3,3
c2,3,1,3
c2,3,2,3
c2,3,3,3
c3,3,1,3
c3,3,2,3
c3,3,3,3


(7.31)

Victor Saouma

Introduction to Continuum Mechanics

Draft

7.3 Stress-Strain Relations in Generalized Elasticity

77

But the matrix must be symmetric thanks to Cauchys second law of motion (i.e symmetry of both the
stress and the strain), and thus for anisotropic material we will have a symmetric 6 by 6 matrix with
(6)(6+1)
= 21 independent coecients.
2
32 By means of coordinate transformation we can relate the material properties in one coordinate system
(old) xi , to a new one xi , thus from Eq. 1.27 (v j = ap vp ) we can rewrite
j

W =

1
1
1
crstu Ers Etu = crstu ar as at au E ij E km = cijkm E ij E km
i j k m
2
2
2

(7.32)

thus we deduce
cijkm = ar as at au crstu
i j k m

(7.33)

that is the fourth order tensor of material constants in old coordinates may be transformed into a new
coordinate system through an eighth-order tensor ar as at au
i j k m

7.3.2

Monotropic Material

A plane of elastic symmetry exists at a point where the elastic constants have the same values
for every pair of coordinate systems which are the reected images of one another with respect to the
plane. The axes of such coordinate systems are referred to as equivalent elastic directions.

33

34

If we assume x1 = x1 , x2 = x2 and x3 = x3 , then the transformation xi = aj xj is dened through


i

1 0 0
aj = 0 1 0
(7.34)
i
0 0 1

where the negative sign reects the symmetry of the mirror image with respect to the x3 plane.
We next substitute in Eq.7.33, and as an example we consider c1123 = ar as at au crstu = a1 a1 a2 a3 c1123 =
1 1 2 3
1 1 2 3
(1)(1)(1)(1)c1123 = c1123 , obviously, this is not possible, and the only way the relation can remanin
valid is if c1123 = 0. We note that all terms in cijkl with the index 3 occurring an odd number of times
will be equal to zero. Upon substitution, we obtain

c1111 c1122 c1133 c1112


0
0

0
0
c2222 c2233 c2212

0
0
c3333 c3312

(7.35)
cijkm =
0
0
c1212

SYM.
c2323 c2331
c3131
35

we now have 13 nonzero coecients.

7.3.3

Orthotropic Material

36 If the material possesses three mutually perpendicular planes of elastic symmetry, (that is symmetric
with respect to two planes x2 and x3 ), then the transformation xi = aj xj is dened through
i

1 0
0
aj = 0 1 0
(7.36)
i
0 0 1

Victor Saouma

Introduction to Continuum Mechanics

Draft
78

CONSTITUTIVE EQUATIONS; Part I LINEAR

where the negative sign reects the symmetry of


substitution in Eq.7.33 we now would have

c1111 c1122

c2222

cijkm =

SYM.

the mirror image with respect to the x3 plane. Upon


c1133
c2233
c3333

0
0
0

0
0
0
0

c1212

0
0
0
0
0

c2323

(7.37)

c3131
We note that in here all terms of cijkl with the indices 3 and 2 occuring an odd number of times are
again set to zero.
37

Wood is usually considered an orthotropic material and will have 9 nonzero coecients.

7.3.4

Transversely Isotropic Material

38 A material is transversely isotropic if there is a preferential direction normal to all but one of the
three axes. If this axis is x3 , then rotation about it will require that

cos
sin 0
(7.38)
aj = sin cos 0
i
0
0
1

substituting Eq. 7.33 into Eq. 7.41, using the above transformation matrix, we obtain
(cos4 )c1111 + (cos2 sin2 )(2c1122 + 4c1212 ) + (sin4 )c2222

c1111

c1122

(cos sin )c1111 + (cos )c1122 4(cos sin )c1212 + (sin )c2211
(7.39-b)
(7.39-c)
+(sin2 cos2 )c2222
2
2
= (cos )c1133 + (sin )c2233
(7.39-d)
4
2
2
4
= (sin )c1111 + (cos sin )(2c1122 + 4c1212 ) + (cos )c2222
(7.39-e)
2
2
2
2
2
2
4
= (cos sin )c1111 2(cos sin )c1122 2(cos sin )c1212 + (cos )c1212 (7.39-f)

c1133
c2222
c1212

(7.39-a)

+(sin2 cos2 )c2222 + sin4 c1212

(7.39-g)

.
.
.
But in order to respect our initial assumption about symmetry, these results require that
c1111
c1133

cijkm

c1111

(7.40-a)
(7.40-b)

=
=

c3131
1
(c1111 c1122 )
2

(7.40-c)

c1212

c2222
c2233

c2323

yielding

=
=

c1122
c2222

SYM.

c1133
c2233
c3333

0
0
0
1
(c1111 c1122 )
2

(7.40-d)
0
0
0
0
c2323

0
0
0
0
0

(7.41)

c3131
we now have 5 nonzero coecients.
39 It should be noted that very few natural or man-made materials are truly orthotropic (certain crystals
as topaz are), but a number are transversely isotropic (laminates, shist, quartz, roller compacted concrete,
etc...).

Victor Saouma

Introduction to Continuum Mechanics

Draft

7.3 Stress-Strain Relations in Generalized Elasticity

7.3.5

79

Isotropic Material

40 An isotropic material is symmetric with respect to every plane and every axis, that is the elastic
properties are identical in all directions.
41 To mathematically characterize an isotropic material, we require coordinate transformation with
rotation about x2 and x1 axes in addition to all previous coordinate transformations. This process will
enforce symmetry about all planes and all axes.

42

The rotation about the x2 axis is obtained through

cos 0 sin
j

1
0
ai = 0
sin 0 cos

(7.42)

we follow a similar procedure to the case of transversely isotropic material to obtain


c1111
c3131

43

=
=

c3333
1
(c1111 c1133 )
2

next we perform a rotation about the x1 axis

1
0
aj = 0 cos
i
0 sin

(7.43-a)
(7.43-b)

0
sin
cos

(7.44)

it follows that
c1122
c3131

c2323
which will nally give

cijkm

c1111

c1133
1
(c3333 c1133 )
2
1
(c2222 c2233 )
2
c1122
c2222

c1133
c2233
c3333

SYM.

0 0
0 0
0 0
a 0
b

(7.45-a)
(7.45-b)
(7.45-c)

0
0
0
0
0
c

(7.46)

with a = 1 (c1111 c1122 ), b = 1 (c2222 c2233 ), and c = 1 (c3333 c1133 ).


2
2
2
44 If we denote c1122 = c1133 = c2233 = and c1212 = c2323 = c3131 = then from the previous relations
we determine that c1111 = c2222 = c3333 = + 2, or

+ 2

0 0 0

+ 2

0 0 0

+ 2 0 0 0
(7.47)
cijkm =

0 0

SYM.
0

Victor Saouma

ij km + (ik jm + im kj )

(7.48)

Introduction to Continuum Mechanics

Draft
710

CONSTITUTIVE EQUATIONS; Part I LINEAR

and we are thus left with only two independent non zero coecients and which are called Lames
constants.
45

Substituting the last equation into Eq. 7.29,


Tij = [ij km + (ik jm + im kj )]Ekm

(7.49)

Or in terms of and , Hookes Law for an isotropic body is written as


Tij = ij Ekk + 2Eij
1

Tij
ij Tkk
Eij =
2
3 + 2

or
or

T = IE + 2E

1
IT +
T
E=
2(3 + 2)
2

(7.50)
(7.51)

It should be emphasized that Eq. 7.47 is written in terms of the Engineering strains (Eq. 7.29)
that is ij = 2Eij for i = j. On the other hand the preceding equations are written in terms of the
tensorial strains Eij
46

7.3.5.1

Engineering Constants

47 The stress-strain relations were expressed in terms of Lames parameters which can not be readily
measured experimentally. As such, in the following sections we will reformulate those relations in terms
of engineering constants (Youngs and the bulks modulus). This will be done for both the isotropic
and transversely isotropic cases.

7.3.5.1.1

Isotropic Case

7.3.5.1.1.1

Youngs Modulus

In order to avoid certain confusion between the strain E and the elastic constant E, we adopt the
usual engineering notation Tij ij and Eij ij

48

49

If we consider a simple uniaxial state of stress in the x1 direction, then from Eq. 7.51
+

(3 + 2)

= 33 =
2(3 + 2)
0 = 12 = 23 = 13
11

22

(7.52-a)

(7.52-b)
(7.52-c)

Yet we have the elementary relations in terms engineering constants E Youngs modulus and
Poissons ratio

(7.53-a)
11 =
E
22
33
=
=
(7.53-b)
11
11

50

then it follows that


1
E

Victor Saouma

=
=

; =
(3 + 2)
2( + )
E
E
; = G =
(1 + )(1 2)
2(1 + )

(7.54)
(7.55)

Introduction to Continuum Mechanics

Draft

7.3 Stress-Strain Relations in Generalized Elasticity

51

711

Similarly in the case of pure shear in the x1 x3 and x2 x3 planes, we have


21 = 12

212

all other ij = 0

(7.56-a)
(7.56-b)

and the is equal to the shear modulus G.


52

Hookes law for isotropic material in terms of engineering constants becomes


ij
ij

53

54

=
=

When the strain equation is expanded in

1
xx

yy

1
zz
=
xy (2xy ) E 0

yz (2yz )

zx (2zx )
0

3D cartesian coordinates it would yield:


0
0
0
xx

1
0
0
0 yy

zz
1
0
0
0

0
0 1+
0
0 xy

0
0
0
1+
0 yz

zx
0
0
0
0
1+

If we invert this equation, we obtain

xx

1
yy (1+)(12)


zz

=
xy

yz

zx
7.3.5.1.1.2

55

E
E

ij +
ij kk
+
I
or =
1+
1 2
1+
1 2

1+

1+
ij ij kk or =
I
E
E
E
E

1
G 0
0

0
1
0

(7.57)
(7.58)

(7.59)

xx

yy

zz

xy (2xy )
0

0 yz (2yz )

1
zx (2zx )

(7.60)

Bulks Modulus; Volumetric and Deviatoric Strains

We can express the trace of the stress I in terms of the volumetric strain I From Eq. 7.50
ii = ii kk + 2ii = (3 + 2)ii 3Kii

(7.61)

2
K =+
3

(7.62)

or

56 We can provide a complement to the volumetric part of the constitutive equations by substracting
the trace of the stress from the stress tensor, hence we dene the deviatoric stress and strains as as

1
(tr )I (7.63)
3
1
(tr )I (7.64)
3

and the corresponding constitutive relation will be

KeI + 2
1
p
I+

3K
2

(7.65)
(7.66)

where p 1 tr () is the pressure, and = pI is the stress deviator.


3
Victor Saouma

Introduction to Continuum Mechanics

Draft
712

7.3.5.1.1.3

57

CONSTITUTIVE EQUATIONS; Part I LINEAR


Restriction Imposed on the Isotropic Elastic Moduli

We can rewrite Eq. 7.20 as


dW = Tij dEij

(7.67)

but since dW is a scalar invariant (energy), it can be expressed in terms of volumetric (hydrostatic) and
deviatoric components as
(7.68)
dW = pde + ij dEij
substituting p = Ke and ij = 2GEij , and integrating, we obtain the following expression for the
isotropic strain energy
1
(7.69)
W = Ke2 + GEij Eij
2
and since positive work is required to cause any deformation W > 0 thus
2
+ GK
3
G

>

(7.70-a)

>

(7.70-b)

1
2

(7.71)

ruling out K = G = 0, we are left with


E > 0; 1 < <

58

The isotropic strain energy function can be alternatively expressed as


W =

59

From Table 7.2, we observe that =

+ 2
3
(3+2)
+

2(+)

1
2

1 2
e + GEij Eij
2

implies G =

E
3,

and

(7.72)
1
K

= 0 or elastic incompressibility.

E,

E,

K,

E
(1+)(12)
E
2(1+)
E
3(12)

2
12

(E2)
3E

2(1+)
3(12)

3K
1+
3K(12)
2(1+)

E
3(3E)

2(1 + )

E
2

E
1

K
3K(1 2)

Table 7.2: Conversion of Constants for an Isotropic Elastic Material


60

The elastic properties of selected materials is shown in Table 7.3.

7.3.5.1.2

61

Transversly Isotropic Case

For transversely isotropic, we can express the stress-strain relation in tems of

Victor Saouma

xx
yy
zz
xy
yz
xz

=
=
=
=
=
=

a11 xx + a12 yy + a13 zz


a12 xx + a11 yy + a13 zz
a13 (xx + yy ) + a33 zz
(7.73)
2(a11 a12 )xy
a44 xy
a44 xz
Introduction to Continuum Mechanics

Draft

7.3 Stress-Strain Relations in Generalized Elasticity


Material
A316 Stainless Steel
A5 Aluminum
Bronze
Plexiglass
Rubber
Concrete
Granite

713

E (MPa)
196,000
68,000
61,000
2,900
2
60,000
60,000

0.3
0.33
0.34
0.4
0.5
0.2
0.27

Table 7.3: Elastic Properties of Selected Materials at 200c


and
a11 =

1
;
E

a12 =

;
E

a13 =

;
E

a33 =

1
;
E

a44 =

(7.74)

where E is the Youngs modulus in the plane of isotropy and E the one in the plane normal to it.
corresponds to the transverse contraction in the plane of isotropy when tension is applied in the plane;
corresponding to the transverse contraction in the plane of isotropy when tension is applied normal
to the plane; corresponding to the shear moduli for the plane of isotropy and any plane normal to it,
and is shear moduli for the plane of isotropy.
7.3.5.2
62

Special 2D Cases

Often times one can make simplifying assumptions to reduce a 3D problem into a 2D one.

7.3.5.2.1

Plane Strain

For problems involving a long body in the z direction with no variation in load or geometry, then
zz = yz = xz = xz = yz = 0. Thus, replacing into Eq. 5.2 we obtain

(1 )

0
xx

xx

(1 )
0
yy

(7.75)
=

0 yy
zz (1 + )(1 2)

xy

12
xy
0
0
2

63

7.3.5.2.2

64

Axisymmetry

In solids of revolution, we can use a polar coordinate sytem and


rr

zz

rz

Victor Saouma

u
r
u
r
w
z
u w
+
z
r

(7.76-a)
(7.76-b)
(7.76-c)
(7.76-d)

Introduction to Continuum Mechanics

Draft
714

65

CONSTITUTIVE EQUATIONS; Part I LINEAR

The constitutive relation is again analogous to 3D/plane strain

rr

E
zz

1
=
(1 + )(1 2)

rz
0
0
0

7.3.5.2.3

0
0
0
0
12
2

rr

zz

rz

(7.77)

Plane Stress

If the longitudinal dimension in z direction is much smaller than in the x and y directions, then
yz = xz = zz = xz = yz = 0 throughout the thickness. Again, substituting into Eq. 5.2 we obtain:

1
0
xx
xx
1
1
0
yy
yy
=
(7.78-a)

1 2
xy
xy
0 0 1
2
1
(xx + yy )
(7.78-b)
zz =
1

66

7.4

Linear Thermoelasticity

67 If thermal eects are accounted for, the components of the linear strain tensor Eij may be considered
as the sum of
(T )
()
(7.79)
Eij = Eij + Eij

(T )

where Eij
eld.

()

is the contribution from the stress eld, and Eij

the contribution from the temperature

68 When a body is subjected to a temperature change 0 with respect to the reference state
temperature, the strain componenet of an elementary volume of an unconstrained isotropic body are
given by
()
(7.80)
Eij = ( 0 )ij

where is the linear coecient of thermal expansion.


69

Inserting the preceding two equation into Hookes law (Eq. 7.51) yields

Eij =

1
2

Tij

ij Tkk
3 + 2

+ ( 0 )ij

(7.81)

which is known as Duhamel-Neumann relations.


70

If we invert this equation, we obtain the thermoelastic constitutive equation:


Tij = ij Ekk + 2Eij (3 + 2)ij ( 0 )

(7.82)

71 Alternatively, if we were to consider the derivation of the Green-elastic hyperelastic equations, (Sect.
7.1.5), we required the constants c1 to c6 in Eq. 7.22 to be zero in order that the stress vanish in the
unstrained state. If we accounted for the temperature change 0 with respect to the reference state

Victor Saouma

Introduction to Continuum Mechanics

Draft

7.5 Fourrier Law

715

temperature, we would have ck = k ( 0 ) for k = 1 to 6 and would have to add like terms to Eq.
7.22, leading to
(7.83)
Tij = ij ( 0 ) + cijrs Ers

for linear theory, we suppose that ij is independent from the strain and cijrs independent of temperature
change with respect to the natural state. Finally, for isotropic cases we obtain
Tij = Ekk ij + 2Eij ij ( 0 )ij
which is identical to Eq. 7.82 with =

E
12 .

Hence

Tij =

72

(7.84)

E
1 2

(7.85)

In terms of deviatoric stresses and strains we have


Tij = 2Eij and Eij =

Tij
2

(7.86)

and in terms of volumetric stress/strain:


p = Ke + ( 0 ) and e =

7.5

p
+ 3( 0 )
K

(7.87)

Fourrier Law

73 Consider a solid through which there is a ow q of heat (or some other quantity such as mass, chemical,
etc...)

74

The rate of transfer per unit area is q

75 The direction of ow is in the direction of maximum potential (temperature in this case, but could
be, piezometric head, or ion concentration) decreases (Fourrier, Darcy, Fick...).


qx
x

qy
q=
= D
= D
(7.88)
y

qz

D is a three by three (symmetric) constitutive/conductivity matrix


The conductivity can be either
Isotropic

1
D = k 0
0
Anisotropic

kxx
D = kyx
kzx
Orthotropic

kxx
D= 0
0

0
1
0

0
0
1

(7.89)

kxy
kyy
kzy

kxz
kyz
kzz

(7.90)

0
kyy
0

0
0
kzz

(7.91)

Note that for ow through porous media, Darcys equation is only valid for laminar ow.
Victor Saouma

Introduction to Continuum Mechanics

Draft
716

7.6

CONSTITUTIVE EQUATIONS; Part I LINEAR

Updated Balance of Equations and Unknowns

76 In light of the new equations introduced in this chapter, it would be appropriate to revisit our balance
of equations and unknowns.

d
vi
dt + xi = 0
Tij
dvi
xj + bi = dt
du = Tij Dij + r
dt

Coupled
1
3

Continuity Equation
Equation of motion

qj
xj

Energy equation
T = IE + 2E
Hookes Law
q = D
Heat Equation (Fourrier)
= (s, ); j = j (s, ) Equations of state
Total number of equations

Uncoupled
1
3

1
6
3
2
16

10

and we repeat our list of unknowns


Density

Velocity (or displacement) vi (ui )


Stress components
Tij
Heat ux components
qi
Specic internal energy
u
Entropy density
s
Absolute temperature

Total number of unknowns


and in addition the Clausius-Duhem inequality
must hold.

ds
dt

Coupled
1
3
6
3
1
1
1
16
1
div

Uncoupled
1
3
6
10
which governs entropy production

77 Hence we now have as many equations as unknowns and are (almost) ready to pose and solve problems
in continuum mechanics.

Victor Saouma

Introduction to Continuum Mechanics

Draft
Chapter 8

INTERMEZZO
In light of the lengthy and rigorous derivation of the fundamental equations of Continuum Mechanics in
the preceding chapter, the reader may be at a loss as to what are the most important ones to remember.
Hence, since the complexity of some of the derivation may have eclipsed the nal results, this handout
seeks to summarize the most fundamental relations which you should always remember.
X3
X3

V3

33
t3

t2

13

21

11

23

31

t1

32

V2
X2

22

V1
X1

X2

(Components of a vector are scalars)

12

X 1 Stresses as components of a traction vector


(Components of a tensor of order 2 are vectors)

Stress Vector/Tensor
Strain Tensor

Engineering Strain
Equilibrium
Boundary Conditions
Energy Potential
Hookes Law

ti = Tij nj

1 ui
uj
uk uk

Eij =
+

2 xj
xi
xi xj

1
1
11
2 12
2 13
1
1

= 2 12 22
2 23
1
1
13 2 23 33
2

23 sin 23 = sin(/2 ) = cos = 2E23


dvi
Tij
+ bi =
xj
dt
= u + t
W
Tij =
Eij
Tij = ij Ekk + 2Eij

(8.1-a)
(8.1-b)

(8.1-c)
(8.1-d)
(8.1-e)
(8.1-f)
(8.1-g)
(8.1-h)

Draft
82

INTERMEZZO

xx

yy

zz
xy

yz

xz
Plane Stress
Plane Strain

Victor Saouma

zz = 0;
zz = 0;

1
E

0
0
0

zz = 0
zz = 0

E
1
E

0
0
0

0
0
0

0
0
0

1
G

1
E

0
0

0
0
0
0
1
G

0
0
0
0
0
1
G

xx
yy
zz
xy
yz
xz

(8.1-i)

(8.1-j)
(8.1-k)

Introduction to Continuum Mechanics

Draft

Part II

ELASTICITY/SOLID
MECHANICS

Draft

Draft
Chapter 9

BOUNDARY VALUE PROBLEMS


in ELASTICITY
9.1
1

Preliminary Considerations

All problems in elasticity require three basic components:

3 Equations of Motion (Equilibrium): i.e. Equations relating the applied tractions and body forces
to the stresses (3)
2 ui
Tij
+ bi = 2
(9.1)
Xj
t
6 Stress-Strain relations: (Hookes Law)
T = IE + 2E

(9.2)

6 Geometric (kinematic) equations: i.e. Equations of geometry of deformation relating displacement to strain (6)
1
(9.3)
E = (u x + x u)
2
2 Those 15 equations are written in terms of 15 unknowns: 3 displacement ui , 6 stress components Tij ,
and 6 strain components Eij .

In addition to these equations which describe what is happening inside the body, we must describe
what is happening on the surface or boundary of the body. These extra conditions are called boundary
conditions.

9.2
4

Boundary Conditions

In describing the boundary conditions (B.C.), we must note that:


1. Either we know the displacement but not the traction, or we know the traction and not the
corresponding displacement. We can never know both a priori.
2. Not all boundary conditions specications are acceptable. For example we can not apply tractions
to the entire surface of the body. Unless those tractions are specially prescribed, they may not
necessarily satisfy equilibrium.

Draft
92

BOUNDARY VALUE PROBLEMS in ELASTICITY

Properly specied boundary conditions result in well-posed boundary value problems, while improperly specied boundary conditions will result in ill-posed boundary value problem. Only the former
can be solved.

Thus we have two types of boundary conditions in terms of known quantitites, Fig. 9.1:

u
Figure 9.1: Boundary Conditions in Elasticity Problems
Displacement boundary conditions along u with the three components of ui prescribed on the
boundary. The displacement is decomposed into its cartesian (or curvilinear) components, i.e.
u x , uy
Traction boundary conditions along t with the three traction components ti = nj Tij prescribed
at a boundary where the unit normal is n. The traction is decomposed into its normal and shear(s)
components, i.e tn , ts .
Mixed boundary conditions where displacement boundary conditions are prescribed on a part of
the bounding surface, while traction boundary conditions are prescribed on the remainder.
We note that at some points, traction may be specied in one direction, and displacement at another.
Displacement and tractions can never be specied at the same point in the same direction.
Various terms have been associated with those boundary conditions in the litterature, those are suumarized in Table 9.1.

u, u
Dirichlet
Field Variable
Essential
Forced
Geometric

t, t
Neuman
Derivative(s) of Field Variable
Non-essential
Natural
Static

Table 9.1: Boundary Conditions in Elasticity


Often time we take advantage of symmetry not only to simplify the problem, but also to properly
dene the appropriate boundary conditions, Fig. 9.2.

Victor Saouma

Introduction to Continuum Mechanics

Draft

9.3 Boundary Value Problem Formulation

93

C
AB
BC
CD
DE
EA

E
A

u
ux uy
?
0
?
?
?
?
0
?
?
?

tn ts
?
0
0
0

0
?
0
0 0

Note: Unknown tractions=Reactions

Figure 9.2: Boundary Conditions in Elasticity Problems

9.3
9

Boundary Value Problem Formulation

Hence, the boundary value formulation is suumarized by


Tij
+ bi
Xj

2 ui
in
t2

(9.4)

1
(u x + x u)
(9.5)
2
T = IE + 2E in (9.6)
(9.7)
u = u in u
(9.8)
t = t in t

and is illustrated by Fig. 9.3. This is now a well posed problem.

9.4

Compacted Forms

10 Solving a boundary value problem with 15 unknowns through 15 equations is a formidable task.
Hence, there are numerous methods to reformulate the problem in terms of fewer unknows.

9.4.1

Navier-Cauchy Equations

11 One such approach is to substitute the displacement-strain relation into Hookes law (resulting in
stresses in terms of the gradient of the displacement), and the resulting equation into the equation of
motion to obtain three second-order partial dierential equations for the three displacement components
known as Naviers Equation

( + )

2 uk
2 ui
+
+ bi
Xi Xk
Xk Xk

( + ) ( u) +

u + b

2 ui
t2

(9.9)

2u
t2

(9.10)

or
=

(9.11)
Victor Saouma

Introduction to Continuum Mechanics

Draft
94

BOUNDARY VALUE PROBLEMS in ELASTICITY

Essential B.C.
ui :

?
Body Forces

Displacements

bi

ui

Equilibrium
Tij
xj

Kinematics
E =

+ bi = dvi
dt

1
(u
2

?
Stresses
Tij

x u)

?
-

Constitutive Rel.
T = IE + 2E

Strain
Eij

Natural B.C.
ti :

Figure 9.3: Fundamental Equations in Solid Mechanics

Victor Saouma

Introduction to Continuum Mechanics

Draft

9.5 Strain Energy and Extenal Work

9.4.2

95

Beltrami-Mitchell Equations

12 Whereas Navier-Cauchy equation was expressed in terms of the gradient of the displacement, we can
follow a similar approach and write a single equation in term of the gradient of the tractions.

Tij +

1
Tpp,ij
1+

1
Tpp,ij
Tij,pp +
1+

9.4.3

9.5

ij
1

ij bp,p (bi,j + bj,i )


1

or
=

(b) (bi,j + bj,i )

(9.12)
(9.13)

Ellipticity of Elasticity Problems

Strain Energy and Extenal Work

13 For the isotropic Hookes law, we saw that there always exist a strain energy function W which is
positive-denite, homogeneous quadratic function of the strains such that, Eq. 7.20

W
Eij

(9.14)

1
Tij Eij
2

(9.15)

Tij =
hence it follows that
W =
14

The external work done by a body in equilibrium under body forces bi and surface traction ti is
bi ui d +

equal to

becomes

ti ui d. Substituting ti = Tij nj and applying Gauss theorem, the second term

Tij nj ui d =

(Tij ui ),j d =

(Tij,j ui + Tij ui,j )d

(9.16)

but Tij ui,j = Tij (Eij + ij ) = Tij Eij and from equilibrium Tij,j = bi , thus
bi ui d +

ti ui d =

(Tij Eij bi ui )d

bi ui d +

(9.17)

or
bi ui d +

Tij Eij
d
2

ti ui d = 2

External Work

(9.18)

Internal Strain Energy

that is For an elastic system, the total strain energy is one half the work done by the external forces
acting through their displacements ui .

9.6

Uniqueness of the Elastostatic Stress and Strain Field

Because the equations of linear elasticity are linear equations, the principles of superposition may be
(1)
(1)
used to obtain additional solutions from those established. Hence, given two sets of solution Tij , ui ,
15

(2)

(2)

(2)

(1)

(2)

and Tij , ui , then Tij = Tij Tij , and ui = ui


solution.
16

(1)

ui

Hence for this dierence solution, Eq. 9.18 would yield


(2)

side is zero because ti = ti


Victor Saouma

(1)

ti

(2)

= 0 on u , and ui = ui

(2)

with bi = bi

(1)

bi

u d but the left hand

ti ui d = 2

(1)
ui

= 0 must also be a

u d = 0.

= 0 on t , thus

Introduction to Continuum Mechanics

Draft
96

BOUNDARY VALUE PROBLEMS in ELASTICITY

But u is positive-denite and continuous, thus the integral can vanish if and only if u = 0 everywhere,
and this is only possible if Eij = 0 everywhere so that
17

(2)

(1)

(2)

Eij = Eij Tij = T ij (1)

(9.19)

hence, there can not be two dierent stress and strain elds corresponding to the same externally imposed
body forces and boundary conditions1 and satisfying the linearized elastostatic Eqs 9.1, 9.14 and 9.3.

9.7

Saint Venants Principle

18 This famous principle of Saint Venant was enunciated in 1855 and is of great importance in applied
elasticity where it is often invoked to justify certain simplied solutions to complex problem.

In elastostatics, if the boundary tractions on a part 1 of the boundary are replaced by a


statically equivalent traction distribution, the eects on the stress distribution in the body
are negligible at points whose distance from 1 is large compared to the maximum distance
between points of 1 .
For instance the analysis of the problem in Fig. 9.4 can be greatly simplied if the tractions on 1
are replaced by a concentrated statically equivalent force.
19

dx
t

F=tdx

Figure 9.4: St-Venants Principle

9.8

Cylindrical Coordinates

20 So far all equations have been written in either vector, indicial, or engineering notation. The last two
were so far restricted to an othonormal cartesian coordinate system.
21 We now rewrite some of the fundamental relations in cylindrical coordinate system, Fig. 9.5, as
this would enable us to analytically solve some simple problems of great practical usefulness (torsion,
pressurized cylinders, ...). This is most often achieved by reducing the dimensionality of the problem
from 3 to 2 or even to 1.

1 This

theorem is attributed to Kircho (1858).

Victor Saouma

Introduction to Continuum Mechanics

Draft

9.8 Cylindrical Coordinates

97

Figure 9.5: Cylindrical Coordinates

9.8.1
22

Strains

With reference to Fig. 9.6, we consider the displacement of point P to P . the displacements can be

y
uy

P*

ur

ux

Figure 9.6: Polar Strains


expressed in cartesian coordinates as ux , uy , or in polar coordinates as ur , u . Hence,
ux
uy

=
=

ur cos u sin
ur sin + u cos

(9.20-a)
(9.20-b)

substituting into the strain denition for xx (for small displacements) we obtain
xx
ux

ux
r

x
r
x
xx
Victor Saouma

ux
ux r
ux
=
+
x
x
r x
ur
u
=
cos ur sin
sin u cos

ur
u
=
cos
sin
r
r
sin
=
r

(9.21-a)

(9.21-c)
(9.21-d)

cos

(9.21-b)

(9.21-e)

ur
u
cos + ur sin +
sin + u cos

sin
r

Introduction to Continuum Mechanics

Draft
98

BOUNDARY VALUE PROBLEMS in ELASTICITY


u
ur
cos
sin cos
r
r

(9.21-f)

Noting that as 0, xx rr , sin 0, and cos 1, we obtain


rr = xx |0 =
23

ur
r

(9.22)

Similarly, if /2, xx , sin 1, and cos 0. Hence,


1
ur
u +
(9.23)
r
r
as a function of ur , u and and noting that xy r as 0, we obtain
= xx |/2 =

nally, we may express xy

r =
24

u
1 ur
1
u
xy |0 =

+
2
r
r
r

(9.24)

In summary, and with the addition of the z components (not explicitely derived), we obtain
rr

rz

9.8.2

zz

ur
r
ur
1 u
+
r
r
uz
z
u
ut heta
1 1 ur
+

2 r
r
r
1 uz
1 u
+
2 z
r
ur
1 uz
+
2 r
z

(9.25)
(9.26)
(9.27)
(9.28)
(9.29)
(9.30)

Equilibrium

25 Whereas the equilibrium equation as given In Eq.


6.24 was obtained from the linear momentum
principle (without any reference to the notion of equilibrium of forces), its derivation (as mentioned)
could have been obtained by equilibrium of forces considerations. This is the approach which we will
follow for the polar coordinate system with respect to Fig. 9.7.
26 Summation of forces parallel to the radial direction through the center of the element with unit
thickness in the z direction yields:

Trr
dr (r + dr)d Trr (rd)
r
d
T
+ T dr sin
T +

2
d
Tr
d Tr dr cos
+ fr rdrd
+ Tr +

2
Trr +

(9.31-a)

(9.31-b)

we approximate sin(d/2) by d/2 and cos(d/2) by unity, divide through by rdrd,


Trr
1
Trr +
r
r

1+

dr
r

T
T d 1 Tr

+
+ fr = 0
r
dr
r

(9.32)

27 Similarly we can take the summation of forces in the direction. In both cases if we were to drop the
dr/r and d/r in the limit, we obtain

Victor Saouma

Introduction to Continuum Mechanics

Draft

9.8 Cylindrical Coordinates

99
T
T + d

Trr
d

fr
T
r d r
r

Tr
r d r
r

r +

rr +

T
r d

T
r

r +

r
r+dr

Figure 9.7: Stresses in Polar Coordinates


1 Tr
1
Trr
+
+ (Trr T ) + fr
r
r
r
1 T
1
Tr
+
+ (Tr Tr ) + f
r
r
r

(9.33)

(9.34)

28 It is often necessary to express cartesian stresses in terms of polar stresses and vice versa. This can
be done through the following relationships

cos
sin

sin
cos

Trr
Tr

Tr
T

cos
sin

sin
cos

Txy
Tyy

Txx
Tyy

=
=

Trr cos2 + T sin2 Tr sin 2


Trr sin2 + T cos2 + Tr sin 2

(9.36-a)
(9.36-b)

Txy

Txx
Txy

(Trr T ) sin cos + Tr (cos2 sin2 )

(9.36-c)

(9.35)

yielding

(recalling that sin2 = 1/2 sin 2, and cos2 = 1/2(1 + cos 2)).

9.8.3

Stress-Strain Relations

29 In orthogonal curvilinear coordinates, the physical components of a tensor at a point are merely the
Cartesian components in a local coordinate system at the point with its axes tangent to the coordinate
curves. Hence,

Trr
T

=
=

e + 2rr
e + 2

(9.37)
(9.38)

Tr
Tzz

=
=

2r
(Trr + T )

(9.39)
(9.40)

with e = rr + . alternatively,

Victor Saouma

Introduction to Continuum Mechanics

Draft
910

BOUNDARY VALUE PROBLEMS in ELASTICITY

Err
E

Erz

30

Er

9.8.3.1

1
(1 2 )Trr (1 + )T
E
1
(1 2 )T (1 + )Trr
E
1+
Tr
E
Ez = Ezz = 0

(9.41)
(9.42)
(9.43)
(9.44)

Plane Strain

For Plane strain problems, from Eq. 7.75:

(1 )
rr

zz (1 + )(1 2)

r
0

(1 )

0
0
0
12
2

rr

(9.45)

and zz = rz = z = rz = z = 0.
31

Inverting,

9.8.3.2
32

1 2
rr
1 (1 + )

E
r
0

(1 + )
1 2

0
rr

zz
0

r
2(1 +

(9.46)

Plane Stress

For plane stress problems, from Eq. 7.78-a

1
0
rr
rr
E
1
0

1 2
r
r
0 0 1
2
1
(rr + )
zz =
1

(9.47-a)
(9.47-b)

and rz = z = zz = rz = z = 0
33

Inverting

rr

Victor Saouma

1
1

E
0

1
0

0
rr

2(1 + )
r

(9.48-a)

Introduction to Continuum Mechanics

Draft
Chapter 10

SOME ELASTICITY PROBLEMS


Practical solutions of two-dimensional boundary-value problem in simply connected regions can be
accomplished by numerous techniques. Those include: a) Finite-dierence approximation of the differential equation, b) Complex function method of Muskhelisvili (most useful in problems with stress
concentration), c) Variational methods (which will be covered in subsequent chapters), d) Semi-inverse
methods, and e) Airy stress functions.

Only the last two methods will be discussed in this chapter.

10.1

Semi-Inverse Method

Often a solution to an elasticity problem may be obtained without seeking simulateneous solutions to
the equations of motion, Hookes Law and boundary conditions. One may attempt to seek solutions
by making certain assumptions or guesses about the components of strain stress or displacement while
leaving enough freedom in these assumptions so that the equations of elasticity be satised.

If the assumptions allow us to satisfy the elasticity equations, then by the uniqueness theorem, we
have succeeded in obtaining the solution to the problem.

This method was employed by Saint-Venant in his treatment of the torsion problem, hence it is often
referred to as the Saint-Venant semi-inverse method.

10.1.1

Example: Torsion of a Circular Cylinder

Let us consider the elastic deformation of a cylindrical bar with circular cross section of radius a and
length L twisted by equal and opposite end moments M1 , Fig. 10.1.

From symmetry, it is reasonable to assume that the motion of each cross-sectional plane is a rigid body
rotation about the x1 axis. Hence, for a small rotation angle , the displacement eld will be given by:

u = (e1 )r = (e1 )(x1 e1 + x2 e2 + x3 e3 ) = (x2 e3 x3 e2 )

(10.1)

or
u1 = 0;

u2 = x3 ;

u3 = x2

(10.2)

where = (x1 ).
8

The corresponding strains are given by


E11

E12

E22 = E33 = 0
1

x3
2 x1

(10.3-a)
(10.3-b)

Draft
102

SOME ELASTICITY PROBLEMS

X2

MT

a
n

X3

X1
MT

L
Figure 10.1: Torsion of a Circular Bar
1

x2
2 x1

E13

(10.3-c)

The non zero stress components are obtained from Hookes law

T12

T13

x1

x2
x1
x3

(10.4-a)
(10.4-b)

10 We need to check that this state of stress satises equilibrium Tij /xj = 0. The rst one j = 1 is
identically satised, whereas the other two yield

d2
dx2
1
d2
x2 2
dx1

x3

(10.5-a)

(10.5-b)

thus,
d
= constant
(10.6)
dx1
Physically, this means that equilibrium is only satised if the increment in angular rotation (twist per
unit length) is a constant.
11 We next determine the corresponding surface tractions. On the lateral surface we have a unit normal
1
vector n = a (x2 e2 + x3 e3 ), therefore the surface traction on the lateral surface is given by

0 T12 T13 0
x T
1
1 2 12
x2
T21
0
0
0
{t} = [T]{n} =
(10.7)
=

a
a
0
0
x3
T31
0

12

Substituting,

(x2 x3 + x2 x3 )e1 = 0
(10.8)
a
which is in agreement with the fact that the bar is twisted by end moments only, the lateral surface is
traction free.
t=

Victor Saouma

Introduction to Continuum Mechanics

Draft

10.2 Airy Stress Functions

13

103

On the face x1 = L, we have a unit normal n = e1 and a surface traction


t = Te1 = T21 e2 + T31 e3

(10.9)

this distribution of surface traction on the end face gives rise to the following resultants
R1

T11 dA = 0

(10.10-a)

R2

T21 dA =

x3 dA = 0

(10.10-b)

R3

T31 dA =

x2 dA = 0

(10.10-c)

M1

(x2 T31 x3 T21 )dA =

M2

(x2 + x2 )dA = J
2
3

M3 = 0

(10.10-d)
(10.10-e)

We note that (x2 + x3 )2 dA is the polar moment of inertia of the cross section and is equal to
2
3
J = a4 /2, and we also note that x2 dA = x3 dA = 0 because the area is symmetric with respect to
the axes.
14

From the last equation we note that


=

M
J

(10.11)

which implies that the shear modulus can be determined froma simple torsion experiment.
15

Finally, in terms of the twisting couple M , the stress tensor becomes

0
MJx3 MJx2
[T] = MJx3
0
0
M x2
0
0
J

10.2

Airy Stress Functions

10.2.1

(10.12)

Cartesian Coordinates; Plane Strain

16 If the deformation of a cylindrical body is such that there is no axial components of the displacement
and that the other components do not depend on the axial coordinate, then the body is said to be in a
state of plane strain. If e3 is the direction corresponding to the cylindrical axis, then we have

u1 = u1 (x1 , x2 ),

u2 = u2 (x1 , x2 ),

u3 = 0

(10.13)

and the strain components corresponding to those displacements are


E11

E22

E12

E13

u1
x1
u2
x2
1 u1
u2
+
2 x2
x1
E23 = E33 = 0

(10.14-a)
(10.14-b)
(10.14-c)
(10.14-d)

and the non-zero stress components are T11 , T12 , T22 , T33 where
T33 = (T11 + T22 )
Victor Saouma

(10.15)

Introduction to Continuum Mechanics

Draft
104

17

SOME ELASTICITY PROBLEMS

Considering a static stress eld with no body forces, the equilibrium equations reduce to:
T12
T11
+
x1
x2
T22
T12
+
x1
x2
T33
=0
x1

(10.16-a)

(10.16-b)
(10.16-c)

we note that since T33 = T33 (x1 , x2 ), the last equation is always satised.
18 Hence, it can be easily veried that for any arbitrary scalar variable , if we compute the stress
components from

T11

T22

T12

2
x2
2
2
x2
1
2

x1 x2

(10.17)
(10.18)
(10.19)

then the rst two equations of equilibrium are automatically satised. This function is called Airy
stress function.
19 However, if stress components determined this way are statically admissible (i.e.
they satisfy
equilibrium), they are not necessarily kinematically admissible (i.e. satisfy compatibility equations).
20 To ensure compatibility of the strain components, we obtain the strains components in terms of
from Hookes law, Eq. 5.1 and Eq. 10.15.

E11
E22

E12

21

1
1
2
2
(1 2 )T11 (1 + )T22 =
(1 2 ) 2 (1 + ) 2
E
E
x2
x1
2
1
1

2
(1 2 )T22 (1 + )T11 =
(1 2 ) 2 (1 + ) 2
E
E
x1
x2
2

1
1
(1 + )T12 = (1 + )
E
E
x1 x2

(10.20-a)
(10.20-b)
(10.20-c)

For plane strain problems, the only compatibility equation, 4.140, that is not automatically satised

is

2 E22
2 E12
2 E11
2 + x2 = 2 x x
x2
1
2
1

(10.21)

thus we obtain the following equation governing the scalar function


(1 )

4
4
4
4 + 2 x2 x2 + x4
x1
1
2
1

=0

(10.22)

=0

(10.23)

or
4
4
4
4 + 2 x2 x2 + x4 = 0 or
x1
1
2
1

Hence, any function which satises the preceding equation will satisfy both equilibrium and kinematic
and is thus an acceptable elasticity solution.

Victor Saouma

Introduction to Continuum Mechanics

Draft

10.2 Airy Stress Functions

105

22 We can also obtain from the Hookes law, the compatibility equation 10.21, and the equilibrium
equations the following

2
2
+
x2
x2
1
2

(T11 + T22 ) = 0 or

(T11 + T22 ) = 0

(10.24)

23 Any polynomial of degree three or less in x and y satises the biharmonic equation (Eq. 10.23). A
systematic way of selecting coecients begins with

Cmn xm y n

(10.25)

m=0 n=0

24

The stresses will be given by

Txx

n(n 1)Cmn xm y n2

(10.26-a)

m(m 1)Cmn xm1 y n

(10.26-b)

m=0 n=2

Tyy

m=2 n=0

Txy

mnCmn xm1 y n1

(10.26-c)

m=1 n=1

25

Substituting into Eq. 10.23 and regrouping we obtain

[(m+2)(m+1)m(m1)Cm+2,n2 +2m(m1)n(n1)Cmn +(n+2)(n+1)n(n1)Cm2,n+2 ]xm2 y n2 = 0


m=2 n=2

(10.27)
but since the equation must be identically satised for all x and y, the term in bracket must be equal to
zero.
(m+2)(m+1)m(m1)Cm+2,n2 +2m(m1)n(n1)Cmn +(n+2)(n+1)n(n1)Cm2,n+2 = 0 (10.28)
Hence, the recursion relation establishes relationships among groups of three alternate coecients which
can be selected from

0
0
C02 C03 C04 C05 C06

0
C11 C12 C13 C14 C15

C20 C21 C22 C23 C24

(10.29)

C30 C31 C32 C33

C40 C41 C42

C50 C

51

C60

For example if we consider m = n = 2, then


(4)(3)(2)(1)C40 + (2)(2)(1)(2)(1)C22 + (4)(3)(2)(1)C04 = 0

(10.30)

or 3C40 + C22 + 3C04 = 0

Victor Saouma

Introduction to Continuum Mechanics

Draft
106

10.2.1.1
26

SOME ELASTICITY PROBLEMS

Example: Cantilever Beam

We consider the homogeneous fourth-degree polynomial


4 = C40 x4 + C31 x3 y + C22 x2 y 2 + C13 xy 3 + C04 y 4

(10.31)

with 3C40 + C22 + 3C04 = 0,


27

The stresses are obtained from Eq. 10.26-a-10.26-c


Txx
Tyy
Txy

=
=
=

2C22 x2 + 6C13 xy + 12C04 y 2


2

12C40 x + 6C31 xy + 2C22 y


3C31 x2 4C22 xy 3C13 y 2

(10.32-a)
(10.32-b)
(10.32-c)

These can be used for the end-loaded cantilever beam with width b along the z axis, depth 2a and
length L.
28

If all coecients except C13 are taken to be zero, then


Txx

6C13 xy

(10.33-a)

Tyy
Txy

=
=

0
3C13 y 2

(10.33-b)
(10.33-c)

29 This will give a parabolic shear traction on the loaded end (correct), but also a uniform shear traction
Txy = 3C13 a2 on top and bottom. These can be removed by superposing uniform shear stress Txy =
+3C13 a2 corresponding to 2 = 3C13 a2 xy. Thus

Txy = 3C13 (a2 y 2 )

(10.34)

note that C20 = C02 = 0, and C11 = 3C13 a2 .


30

The constant C13 is determined by requiring that


a

P =b
a

Txy dy = 3bC13

hence
C13 =

a
a

(a2 y 2 )dy

P
4a3 b

(10.35)

(10.36)

and the solution is

Txx

Txy

Tyy

3P
P
xy 3 xy 3
4ab
4a b
3P
3 xy
2a b
3P
3 (a2 y 2 )
4a b
0

(10.37-a)
(10.37-b)
(10.37-c)
(10.37-d)

We observe that the second moment of area for the rectangular cross section is I = b(2a)3 /12 = 2a3 b/3,
hence this solution agrees with the elementary beam theory solution
31

Txx

Txy

Tyy
Victor Saouma

P
3P
xy 3 xy 3
4ab
4a b
y
M
P
xy = M =
I
I
S
P 2
2
(a y )
2I
0

C11 xy + C13 xy 3 =

(10.38-a)
(10.38-b)
(10.38-c)
(10.38-d)

Introduction to Continuum Mechanics

Draft

10.2 Airy Stress Functions

107

10.2.2

Polar Coordinates

10.2.2.1

Plane Strain Formulation

32

In polar coordinates, the strain components in plane strain are, Eq. 9.46
Err
E

Er

Erz

1
(1 2 )Trr (1 + )T
E
1
(1 2 )T (1 + )Trr
E
1+
Tr
E
Ez = Ezz = 0

(10.39-a)
(10.39-b)
(10.39-c)
(10.39-d)

and the equations of equilibrium are


1 Tr
T
1 Trr
+

r r
r
r
1 T
1 Tr
+
r2 r
r

33

(10.40-a)

(10.40-b)

Again, it can be easily veried that the equations of equilibrium are identically satised if
Trr

Tr

1 2
1
+ 2 2
r r
r
2
r2
1

r r

(10.41)
(10.42)
(10.43)

34 In order to satisfy the compatibility conditions, the cartesian stress components must also satisfy Eq.
10.24. To derive the equivalent expression in cylindrical coordinates, we note that T11 + T22 is the rst
scalar invariant of the stress tensor, therefore

T11 + T22 = Trr + T =


35

(10.44)

We also note that in cylindrical coordinates, the Laplacian operator takes the following form
2

36

1 2 2
1
+ 2 2 +
r r
r
r2

1 2
2
1
+ 2 2
+
r2
r r r

(10.45)

Thus, the function must satisfy the biharmonic equation


1 2
1
2
+ 2 2
+
2
r
r r r

Victor Saouma

1 2
1
2
+ 2 2
+
2
r
r r r

= 0 or

=0

(10.46)

Introduction to Continuum Mechanics

Draft
108

10.2.2.2
37

SOME ELASTICITY PROBLEMS

Axially Symmetric Case

If is a function of r only, we have


Trr =

and

38

1 d
;
r dr

T =

d2
;
dr2

Tr = 0

(10.47)

1 d2
1 d
d 4 2 d3
=0
+
2 2 + 3
dr4
r dr3
r dr
r dr

(10.48)

The general solution to this problem; using Mathematica:

DSolve[phi[r]+2 phi[r]/r-phi[r]/r^2+phi[r]/r^3==0,phi[r],r]
= A ln r + Br2 ln r + Cr2 + D
39

(10.49)

The corresponding stress eld is


Trr
T
Tr

A
+ B(1 + 2 ln r) + 2C
r2
A
= 2 + B(3 + 2 ln r) + 2C
r
= 0
=

(10.50)
(10.51)
(10.52)

and the strain components are (from Sect. 9.8.1)

Err

Er

40

1 (1 + )A
ur
=
+ (1 3 4 2 )B + 2(1 2 2 )B ln r + 2(1 2 2 )C
r
E
r2
ur
1
(1 + )A
1 u
+
=

+ (3 4 2 )B + 2(1 2 2 )B ln r + 2(1 2 2 )C
r
r
E
r2
0

Finally, the displacement components can be obtained by integrating the above equations
ur

10.2.2.3

(1 + )A
1

(1 + )Br + 2(1 2 2 )r ln rB + 2(1 2 2 )rC


E
r
4rB
(1 2 )
E

(10.56)
(10.57)

Example: Thick-Walled Cylinder

41 If we consider a circular cylinder with internal and external radii a and b respectively, subjected to
internal and external pressures pi and po respectively, Fig. 10.2, then the boundary conditions for the
plane strain problem are

Trr
Trr

Victor Saouma

=
=

pi at r = a
po at r = b

(10.58-a)
(10.58-b)

Introduction to Continuum Mechanics

(10.53)
(10.54)
(10.55)

Draft

10.2 Airy Stress Functions

109

Saint Venant

po

a
p
i

Figure 10.2: Pressurized Thick Tube

42

These Boundary conditions can be easily shown to be satised by the following stress eld
Trr

Tr

A
+ 2C
r2
A
2 + 2C
r
0

(10.59-a)
(10.59-b)
(10.59-c)

These equations are taken from Eq. 10.50, 10.51 and 10.52 with B = 0 and therefore represent a
possible state of stress for the plane strain problem.
We note that if we take B = 0, then u = 4rB (1 2 ) and this is not acceptable because if we were
E
to start at = 0 and trace a curve around the origin and return to the same point, than = 2 and the
displacement would then be dierent.
43

44

Applying the boundary condition we nd that


Trr

Tr

(b2 /r2 ) 1
1 (a2 /r2 )
p0
2 /a2 ) 1
(b
1 (a2 /b2 )
(b2 /r2 ) + 1
1 + (a2 /r2 )
p0
pi 2 2
(b /a ) 1
1 (a2 /b2 )
0
pi

(10.60)
(10.61)
(10.62)

45 We note that if only the internal pressure pi is acting, then Trr is always a compressive stress, and
T is always positive.
46

If the cylinder is thick, then the strains are given by Eq. 10.53, 10.54 and 10.55. For a very thin

Victor Saouma

Introduction to Continuum Mechanics

Draft
1010

SOME ELASTICITY PROBLEMS

cylinder in the axial direction, then the strains will be given by


Err

Ezz

Er

1
du
= (Trr T )
dr
E
1
u
= (T Trr )
r
E

dw
= (Trr + T )
dz
E
(1 + )
Tr
E

(10.63-a)
(10.63-b)
(10.63-c)
(10.63-d)

It should be noted that applying Saint-Venants principle the above solution is only valid away from
the ends of the cylinder.

47

10.2.2.4

Example: Hollow Sphere

48 We consider next a hollow sphere with internal and xternal radii ai and ao respectively, and subjected
to internal and external pressures of pi and po , Fig. 10.3.

ai
ao

Figure 10.3: Pressurized Hollow Sphere


With respect to the spherical ccordinates (r, , ), it is clear due to the spherical symmetry of the
geometry and the loading that each particle of the elastic sphere will expereince only a radial displacement
whose magnitude depends on r only, that is

49

ur = ur (r),
10.2.2.5

u = u = 0

(10.64)

Example: Stress Concentration due to a Circular Hole in a Plate

50 Analysing the innite plate under uniform tension with a circular hole of diameter a, and subjected
to a uniform stress 0 , Fig. 10.4.
51 The peculiarity of this problem is that the far-eld boundary conditions are better expressed in
cartesian coordinates, whereas the ones around the hole should be written in polar coordinate system.
52 First we select a stress function which satises the biharmonic Equation (Eq. 10.23), and the far-eld
boundary conditions. From St Venant principle, away from the hole, the boundary conditions are given
by:
Tyy = Txy = 0
(10.65)
Txx = 0 ;
2

Recalling (Eq. 10.19) that Txx = , this would would suggest a stress function of the form =
y 2
0 y 2 . Alternatively, the presence of the circular hole would suggest a polar representation of . Thus,
substituting y = r sin would result in = 0 r2 sin2 .
Victor Saouma

Introduction to Continuum Mechanics

Draft

10.2 Airy Stress Functions

1011

y
r

rr

rr
b

rr
b

II

Figure 10.4: Circular Hole in an Innite Plate


53

Since sin2 = 1 (1 cos 2), we could simplify the stress function into
2
= f (r) cos 2

(10.66)

Substituting this function into the biharmonic equation (Eq. 10.46) yields
1 2
1 2
1
2
2 1
+ 2 2
+ 2 2
+
+
r2
r r r
r2
r r
r
2
4
4f
1 d
1 df
d
d2 f
2
2
+
+
2
2
dr
r dr r
dr
r dr
r

54

= 0

(10.67-a)

= 0

(10.67-b)

The general solution of this ordinary linear fourth order dierential equation is
1
+D
r2

(10.68)

1
+ D cos 2
r2

(10.69)

f (r) = Ar2 + Br4 + C


thus the stress function becomes
=

Ar2 + Br4 + C

Using Eq. 10.41-10.43, the stresses are given by


Trr
T
Tr

1 2
1
6C
4D
+ 2 2 = 2A + 4 + 2 cos 2
r r
r
r
r
2

6C
=
= 2A + 12Br2 + 4 cos 2
r2
r
1
6C
2D
= 2A + 6Br2 4 2 sin 2
=
r r
r
r

(10.70-a)

(10.70-b)
(10.70-c)

55 Next we seek to solve for the four constants of integration by applying the boundary conditions. We
will identify two sets of boundary conditions:

1. Outer boundaries: around an innitely large circle of radius b inside a plate subjected to uniform
stress 0 , the stresses in polar coordinates are obtained from Eq. 9.35
Trr
Tr

Tr
T

cos
sin

sin
cos

0
0

0
0

cos
sin

sin
cos

(10.71)

yielding (recalling that sin2 = 1/2 sin 2, and cos2 = 1/2(1 + cos 2)).
(Trr )r=b
(Tr )r=b

(T )r=b
Victor Saouma

0 cos2 =

1
0 (1 + cos 2)
2

1
0 sin 2
2
0
(1 cos 2)
2

(10.72-a)
(10.72-b)
(10.72-c)

Introduction to Continuum Mechanics

Draft
1012

SOME ELASTICITY PROBLEMS

For reasons which will become apparent later, it is more convenient to decompose the state of
stress given by Eq. 10.72-a and 10.72-b, into state I and II:
(Trr )I
r=b
(Tr )I
r=b

(Trr )II
r=b

(Tr )II
r=b

1
0
2
0
1
0 cos 2
2
1
0 sin 2
2

(10.73-a)
(10.73-b)
(10.73-c)
(10.73-d)

Where state I corresponds to a thick cylinder with external pressure applied on r = b and of
magnitude 0 /2. This problem has already been previously solved. Hence, only the last two
equations will provide us with boundary conditions.
2. Around the hole: the stresses should be equal to zero:
(Trr )r=a
(Tr )r=a

=
=

0
0

(10.74-a)
(10.74-b)

56 Upon substitution in Eq. 10.70-a the four boundary conditions (Eq. 10.73-c, 10.73-d, 10.74-a, and
10.74-b) become

4D
b2
2D
2
b
4D
+ 2
a
2D
2
a

6C
b4
6C
2A + 6Bb2 4
b
6C
2A + 4
a
6C
2A + 6Ba2 4
a
2A +

57

Solving for the four unknowns, and taking


A=

0
;
4

a
b

B = 0;

1
0
2
1
0
2

(10.75-a)
(10.75-b)

(10.75-c)

(10.75-d)

= 0 (i.e. an innite plate), we obtain:


C=

a4
0 ;
4

D=

a2
0
2

(10.76)

58 To this solution, we must superimpose the one of a thick cylinder subjected to a uniform radial
traction 0 /2 on the outer surface, and with b much greater than a. These stresses were derived in Eqs.
10.60 and 10.61 yielding for this problem (carefull about the sign)

Trr

0
2
0
2

a2
r2
a2
1+ 2
r
1

(10.77-a)
(10.77-b)

Thus, upon substitution into Eq. 10.70-a, we obtain


Trr
T

Tr
Victor Saouma

a2
a4
4a2 1
0 cos 2
+ 1+3 4 2
r2
r
r
2
a2
3a4 1
1+ 2 1+ 4
0 cos 2
r
r
2
3a4
2a2 1
0 sin 2
1 4 + 2
r
r
2
0
2
0
2

(10.78-a)
(10.78-b)
(10.78-c)

Introduction to Continuum Mechanics

Draft

10.2 Airy Stress Functions

1013

59 We observe that as r , both Trr and Tr are equal to the values given in Eq. 10.72-a and 10.72-b
respectively.

60

Alternatively, at the edge of the hole when r = a we obtain Trr = Tr = 0 and


T )r=a = 0 (1 2 cos 2)

which for =

and

Victor Saouma

3
2

(10.79)

gives a stress concentration factor (SCF) of 3. For = 0 and = , T = 0 .

Introduction to Continuum Mechanics

Draft

SOME ELASTICITY PROBLEMS

Victor Saouma

Introduction to Continuum Mechanics

1014

Draft
Chapter 11

THEORETICAL STRENGTH OF
PERFECT CRYSTALS
This chapter (taken from the authors lecture notes in Fracture Mechanics) is of primary
interest to students in Material Science.

11.1

Introduction

In Eq. ?? we showed that around a circular hole in an innite plate under uniform traction, we do
have a stress concentration factor of 3.

Following a similar approach (though with curvilinear coordinates), it can be shown that if we have
an elliptical hole, Fig. ??, we would have

( )=0, = 0 1 + 2
=0

a
b

(11.1)

We observe that for a = b, we recover the stress concentration factor of 3 of a circular hole, and that for
a degenerated ellipse, i.e a crack there is an innite stress. Alternatively, the stress can be expressed in
x2

= o

2b

2a

Figure 11.1: Elliptical Hole in an Innite Plate

Draft
112

THEORETICAL STRENGTH OF PERFECT CRYSTALS

Strength (P/A)

Theoretical Strength

Diameter

Figure 11.2: Griths Experiments


terms of , the radius of curvature of the ellipse,
( )=0, = 0 1 + 2
=0

(11.2)

From this equation, we note that the stress concentration factor is inversely proportional to the radius
of curvature of an opening.
This equation, derived by Inglis, shows that if a = b we recover the factor of 3, and the stress
concentration factor increase as the ratio a/b increases. In the limit, as b = 0 we would have a crack
resulting in an innite stress concentration factor, or a stress singularity.

Around 1920, Grith was exploring the theoretical strength of solids by performing a series of experiments on glass rods of various diameters.

He observed that the tensile strength ( t ) of glass decreased with an increase in diameter, and that
1
for a diameter 10,000 in., t = 500, 000 psi; furthermore, by extrapolation to zero diameter he
obtained a theoretical maximum strength of approximately 1,600,000 psi, and on the other hand for very
large diameters the asymptotic values was around 25,000 psi.
5

Area A1
Failure Load P1
t
Failure Strength (P/A) 1

<
<
>

A2
P2
t
2

<
<
>

A3
P3
t
3

<
>
>

A4
P4
t
4

(11.3)

Furthermore, as the diameter was further reduced, the failure strength asymptotically approached a
limit which will be shown later to be the theoretical strength of glass, Fig. 11.2.
Clearly, one would have expected the failure strength to be constant, yet it was not. So Grith was
confronted with two questions:

1. What is this apparent theoretical strength, can it be derived?


2. Why is there a size eect for the actual strength?
The answers to those two questions are essential to establish a link between Mechanics and Materials.
In the next sections we will show that the theoretical strength is related to the force needed to break
a bond linking adjacent atoms, and that the size eect is caused by the size of imperfections inside a
solid.

Victor Saouma

Introduction to Continuum Mechanics

Draft

11.2 Theoretical Strength

113

Figure 11.3: Uniformly Stressed Layer of Atoms Separated by a0

11.2

Theoretical Strength

We start, [?] by exploring the energy of interaction between two adjacent atoms at equilibrium separated by a distance a0 , Fig. 11.3. The total energy which must be supplied to separate atom C from C
is
(11.4)
U0 = 2
8

where is the surface energy1 , and the factor of 2 is due to the fact that upon separation, we have
two distinct surfaces.

11.2.1

Ideal Strength in Terms of Physical Parameters

We shall rst derive an expression for the ideal strength in terms of physical parameters, and in the
next section the strength will be expressed in terms of engineering ones.

Solution I: Force being the derivative of energy, we have F = dU , thus F = 0 at a = a0 , Fig. 11.4, and
da
is maximum at the inection point of the U0 a curve. Hence, the slope of the force displacement
curve is the stiness of the atomic spring and should be related to E. If we let x = a a0 , then
x
F
the strain would be equal to = a0 . Furthermore, if we dene the stress as = a2 , then the
0
curve will be as shown in Fig. 11.5.
From this diagram, it would appear that the sine curve would be an adequate approximation to
this relationship. Hence,
x
theor
(11.5)
= max sin 2

theor
and the maximum stress max would occur at x = . The energy required to separate two atoms
4
is thus given by the area under the sine curve, and from Eq. 11.4, we would have

2 = U0

=
=

1 From

x
dx

0
theor
2x
2
max [ cos (
)] |0
2

theor
max sin 2

(11.6)
(11.7)

watching raindrops and bubbles it is obvious that liquid water has surface tension. When the surface of a liquid
is extended (soap bubble, insect walking on liquid) work is done against this tension, and energy is stored in the new
surface. When insects walk on water it sinks until the surface energy just balances the decrease in its potential energy. For
solids, the chemical bonds are stronger than for liquids, hence the surface energy is stronger. The reason why we do not
notice it is that solids are too rigid to be distorted by it. Surface energy is expressed in J/m2 and the surface energies
of water, most solids, and diamonds are approximately .077, 1.0, and 5.14 respectively.

Victor Saouma

Introduction to Continuum Mechanics

Draft
114

THEORETICAL STRENGTH OF PERFECT CRYSTALS

Energy

Interatomic
Distance

Repulsion

Attraction

Force

Interatomic
Distance

Figure 11.4: Energy and Force Binding Two Adjacent Atoms


1

2
theor

) + cos(0)]
[ cos (
2 max
2
2
theor
max

=
=

(11.8)
(11.9)

Also for very small displacements (small x) sin x x, thus Eq. 11.5 reduces to
theor
max

Ex
2x

a0

(11.10)

elliminating x,
theor
max

E
a0 2

(11.11)

theor
max

E
a0

(11.12)

Substituting for from Eq. 11.9, we get

Solution II: For two layers of atoms a0 apart, the strain energy per unit area due to (for linear elastic
systems) is
U

= 1 ao
2
= E

U=

2 ao
2E

(11.13)

If is the surface energy of the solid per unit area, then the total surface energy of two new fracture
surfaces is 2.
For our theoretical strength, U = 2

Victor Saouma

theor
(max )2 a0
2E

theor
= 2 or max = 2

E
a0

Introduction to Continuum Mechanics

Draft

11.2 Theoretical Strength

115

Figure 11.5: Stress Strain Relation at the Atomic Level


Note that here we have assumed that the material obeys Hookes Law up to failure, since this is
seldom the case, we can simplify this approximation to:
E
a0

theor
max =

(11.14)

which is the same as Equation 11.12


J
Example: As an example, let us consider steel which has the following properties: = 1 m2 ; E =
11 N
10
2 10 m2 ; and a0 2 10
m. Thus from Eq. 11.12 we would have:
theor
max

(2 1011 )(1)
2 1010
N
3.16 1010 2
m
E
6

(11.15)
(11.16)
(11.17)

Thus this would be the ideal theoretical strength of steel.

11.2.2

Ideal Strength in Terms of Engineering Parameter

10 We note that the force to separate two atoms drops to zero when the distance between them is a0 + a
where a0 corresponds to the origin and a to . Thus, if we take a = or = 2a, combined with Eq.
2
2
11.11 would yield
E a
theor
(11.18)
max
a0

11

Alternatively combining Eq. 11.9 with = 2a gives


a

theor
max

(11.19)

Combining those two equations will give

12

E
a0

(11.20)

However, since as a rst order approximation a a0 then the surface energy will be

Victor Saouma

Ea0
10

(11.21)
Introduction to Continuum Mechanics

Draft
116

THEORETICAL STRENGTH OF PERFECT CRYSTALS

This equation, combined with Eq. 11.12 will nally give


E
theor
max
10

(11.22)

which is an approximate expression for the theoretical maximum strength in terms of E.

11.3

Size Eect; Grith Theory

13 In his quest for an explanation of the size eect, Grith came across Ingliss paper, and his strike of
genius was to assume that strength is reduced due to the presence of internal aws. Grith postulated
that the theoretical strength can only be reached at the point of highest stress concentration, and
accordingly the far-eld applied stress will be much smaller.

14

Hence, assuming an elliptical imperfection, and from equation 11.2


a

theor
act
max = cr 1 + 2

(11.23)

act
is the stress at the tip of the ellipse which is caused by a (lower) far eld stress cr . Asssuming a0
a
and since 2 a0
1, for an ideal plate under tension with only one single elliptical aw the strength

may be obtained from


a
a0

theor
act
max = 2cr

(11.24)

hence, equating with Eq. 11.12, we obtain


theor
act
max = 2cr

a
=
ao

E
a0

(11.25)

Micro

Macro
From this very important equation, we observe that

1. The left hand side is based on a linear elastic solution of a macroscopic problem solved by Inglis.
2. The right hand side is based on the theoretical strength derived from the sinusoidal stress-strain
assumption of the interatomic forces, and nds its roots in micro-physics.
Finally, this equation would give (at fracture)
act
cr =

E
4a

(11.26)

As an example, let us consider a aw with a size of 2a = 5, 000a0


act
cr

=
=

E
4a
Ea0
10

act
cr
a
a0

=
=

E 2 ao
40 a

2, 500

act
cr

E2
100,000

100 10

(11.27)

Thus if we set a aw size of 2a = 5, 000a0 in Ea0 this is enough to lower the theoretical fracture
10
E

strength from 10 to a critical value of magnitude 100E 10 , or a factor of 100.


As an example

a
theor
act

max = 2cr

ao
106
theor
act
act
6
= 200cr
(11.28)
max = 2cr
a
= 10 m = 1

1010

= = 1010 m
ao
= 1A
Victor Saouma
Introduction to Continuum Mechanics

Draft

11.3 Size Eect; Grith Theory

117

Therefore at failure
act
cr
theor
max

=
=

which can be attained. For instance for steel

Victor Saouma

theor
max
200
E
10

E
2,000

act
cr

30,000
2,000

E
2, 000

(11.29)

= 15 ksi

Introduction to Continuum Mechanics

Draft

THEORETICAL STRENGTH OF PERFECT CRYSTALS

Victor Saouma

Introduction to Continuum Mechanics

118

Draft
Chapter 12

BEAM THEORY
This chapter is adapted from the Authors lecture notes in Structural Analysis.

12.1

Introduction

In the preceding chapters we have focused on the behavior of a continuum, and the 15 equations and
15 variables we introduced, were all derived for an innitesimal element.

In practice, few problems can be solved analytically, and even with computer it is quite dicult to
view every object as a three dimensional one. That is why we introduced the 2D simplication (plane
stress/strain), or 1D for axially symmetric problems. In the preceding chapter we saw a few of those
solutions.

Hence, to widen the scope of application of the fundamental theory developed previously, we could
either resort to numerical methods (such as the nite dierence, nite element, or boundary elements),
or we could further simplify the problem.

Solid bodies, in general, have certain peculiar geometric features amenable to a reduction from three
to fewer dimensions. If one dimension of the structural element1 under consideration is much greater
or smaller than the other three, than we have a beam, or a plate respectively. If the plate is curved,
then we have a shell.
4

For those structural elements, it is customary to consider as internal variables the resultant of the
stresses as was shown in Sect. ??.

Hence, this chapter will focus on a brief introduction to beam theory. This will however be preceded
by an introduction to Statics as the internal forces would also have to be in equilibrium with the external
ones.

Beam theory is perhaps the most successful theory in all of structural mechanics, and it forms the
basis of structural analysis which is so dear to Civil and Mechanical engineers.

12.2

Statics

12.2.1

Equilibrium

Any structural element, or part of it, must satisfy equilibrium.


1 So

far we have restricted ourselves to a continuum, in this chapter we will consider a structural element.

Draft
122

10

BEAM THEORY

Summation of forces and moments, in a static system must be equal to zero2 .


In a 3D cartesian coordinate system there are a total of 6 independent equations of equilibrium:
Fx
Mx

11

=
=

Fy
My

=
=

Fz
Mz

=
=

0
0

(12.1)

In a 2D cartesian coordinate system there are a total of 3 independent equations of equilibrium:


Fx

Fy

Mz

(12.2)

12 All the externally applied forces on a structure must be in equilibrium. Reactions are accordingly
determined.

13

For reaction calculations, the externally applied load may be reduced to an equivalent force3 .

14

Summation of the moments can be taken with respect to any arbitrary point.

15 Whereas forces are represented by a vector, moments are also vectorial quantities and are represented
by a curved arrow or a double arrow vector.

16

Not all equations are applicable to all structures, Table 12.1


Structure Type
Beam, no axial forces
2D Truss, Frame, Beam
Grid
3D Truss, Frame
Beams, no axial Force
2 D Truss, Frame, Beam

Equations
Fx

Fy
Fy

Mz
Mz

Fz
Fx
Fy
Fz
Alternate Set
A
B
Mz
Mz
A
B
Fx
Mz
Mz
A
B
C
Mz
Mz
Mz

Mx
Mx

My
My

Mz

Table 12.1: Equations of Equilibrium


17 The three conventional equations of equilibrium in 2D: Fx , Fy and Mz can be replaced by the
A
B
C
independent moment equations Mz , Mz , Mz provided that A, B, and C are not colinear.
18

It is always preferable to check calculations by another equation of equilibrium.

19

Before you write an equation of equilibrium,


1. Arbitrarily decide which is the +ve direction
2. Assume a direction for the unknown quantities
3. The right hand side of the equation should be zero

If your reaction is negative, then it will be in a direction opposite from the one assumed.
20 Summation of external forces is equal and opposite to the internal ones (more about this below).
Thus the net force/moment is equal to zero.

21

The external forces give rise to the (non-zero) shear and moment diagram.
2 In

a dynamic system F = ma where m is the mass and a is the acceleration.


for internal forces (shear and moment) we must use the actual load distribution.

3 However

Victor Saouma

Introduction to Continuum Mechanics

Draft
12.2 Statics

12.2.2
22

123

Reactions

In the analysis of structures, it is often easier to start by determining the reactions.

23 Once the reactions are determined, internal forces (shear and moment) are determined next; nally,
internal stresses and/or deformations (deections and rotations) are determined last.

24

Depending on the type of structures, there can be dierent types of support conditions, Fig. 12.1.

Figure 12.1: Types of Supports


Roller: provides a restraint in only one direction in a 2D structure, in 3D structures a roller may provide
restraint in one or two directions. A roller will allow rotation.
Hinge: allows rotation but no displacements.
Fixed Support: will prevent rotation and displacements in all directions.

12.2.3

Equations of Conditions

25 If a structure has an internal hinge (which may connect two or more substructures), then this will
provide an additional equation (M = 0 at the hinge) which can be exploited to determine the reactions.
26 Those equations are often exploited in trusses (where each connection is a hinge) to determine reactions.
27 In an inclined roller support with Sx and Sy horizontal and vertical projection, then the reaction
R would have, Fig. 12.2.

Sy
Rx
=
Ry
Sx

Victor Saouma

(12.3)

Introduction to Continuum Mechanics

Draft
124

BEAM THEORY

Figure 12.2: Inclined Roller Support

12.2.4

Static Determinacy

28 In statically determinate structures, reactions depend only on the geometry, boundary conditions and
loads.
29 If the reactions can not be determined simply from the equations of static equilibrium (and equations
of conditions if present), then the reactions of the structure are said to be statically indeterminate.

The degree of static indeterminacy is equal to the dierence between the number of reactions
and the number of equations of equilibrium (plus the number of equations of conditions if applicable),
Fig. 12.3.

30

Figure 12.3: Examples of Static Determinate and Indeterminate Structures


31 Failure of one support in a statically determinate system results in the collapse of the structures.
Thus a statically indeterminate structure is safer than a statically determinate one.
32 For statically indeterminate structures, reactions depend also on the material properties (e.g. Youngs
and/or shear modulus) and element cross sections (e.g. length, area, moment of inertia).

12.2.5

Geometric Instability

33 The stability of a structure is determined not only by the number of reactions but also by their
arrangement.

34

Geometric instability will occur if:


1. All reactions are parallel and a non-parallel load is applied to the structure.
2. All reactions are concurrent, Fig. 12.4.

Victor Saouma

Introduction to Continuum Mechanics

Draft
12.2 Statics

125

Figure 12.4: Geometric Instability Caused by Concurrent Reactions


3. The number of reactions is smaller than the number of equations of equilibrium, that is a mechanism is present in the structure.
35 Mathematically, this can be shown if the determinant of the equations of equilibrium is equal to
zero (or the equations are inter-dependent).

12.2.6

Examples

Example 12-1: Simply Supported Beam


Determine the reactions of the simply supported beam shown below.

Solution:
The beam has 3 reactions, we have 3 equations of static equilibrium, hence it is statically determinate.
(+ - ) Fx = 0; Rax 36 k = 0
)
(+ 6 Fy = 0; Ray + Rdy 60 k (4) k/ft(12) ft = 0

c
(+ ) Mz = 0; 12Ray 6Rdy (60)(6) = 0

or through matrix inversion

1 0
0 1
0 12

(on your

calculator)

Rax 36
Rax 36 k
Ray
Ray
56 k
108
=
=

Rdy
Rdy
52 k
360

Alternatively we could have used another set of equations:

a
(+ ) Mz = 0;

M d = 0;
(+ )

z
Check:

Victor Saouma

(60)(6) + (48)(12) (Rdy )(18) = 0

Rdy = 52 k 6
(Ray )(18) (60)(12) (48)(6) = 0 Ray = 56 k 6

)
(+ 6 Fy = 0; ; 56 52 60 48 = 0

Introduction to Continuum Mechanics

Draft
126

BEAM THEORY

12.3

Shear & Moment Diagrams

12.3.1

Design Sign Conventions

36

Before we derive the Shear-Moment relations, let us arbitrarily dene a sign convention.

4
37 The sign convention adopted here, is the one commonly used for design purposes . With reference to
Fig. 12.5

Figure 12.5: Shear and Moment Sign Conventions for Design


Load Positive along the beams local y axis (assuming a right hand side convention), that is positive
upward.
Axial: tension positive.
Flexure A positive moment is one which causes tension in the lower bers, and compression in the
upper ones. For frame members, a positive moment is one which causes tension along the inner
side.
Shear A positive shear force is one which is up on a negative face, or down on a positive one.
Alternatively, a pair of positive shear forces will cause clockwise rotation.

12.3.2

Load, Shear, Moment Relations

38 Let us derive the basic relations between load, shear and moment. Considering an innitesimal length
dx of a beam subjected to a positive load5 w(x), Fig. 12.6. The innitesimal section must also be in
equilibrium.
39 There are no axial forces, thus we only have two equations of equilibrium to satisfy Fy = 0 and
Mz = 0.
40 Since dx is innitesimally small, the small variation in load along it can be neglected, therefore we
assume w(x) to be constant along dx.

To denote that a small change in shear and moment occurs over the length dx of the element, we add
the dierential quantities dVx and dMx to Vx and Mx on the right face.

41

4 Note
5 In

that this sign convention is the opposite of the one commonly used in Europe!
this derivation, as in all other ones we should assume all quantities to be positive.

Victor Saouma

Introduction to Continuum Mechanics

Draft

12.3 Shear & Moment Diagrams

127

Figure 12.6: Free Body Diagram of an Innitesimal Beam Segment

42

Next considering the rst equation of equilibrium


)
(+ 6 Fy = 0 Vx + wx dx (Vx + dVx ) = 0

or
dV
= w(x)
dx

(12.4)

The slope of the shear curve at any point along the axis of a member is given by
the load curve at that point.
43

Similarly
dx

(Mx + dMx ) = 0
(+ ) Mo = 0 Mx + Vx dx wx dx

Neglecting the dx2 term, this simplies to


dM
= V (x)
dx

(12.5)

The slope of the moment curve at any point along the axis of a member is given
by the shear at that point.
44

Alternative forms of the preceding equations can be obtained by integration


V
V21

=
=

w(x)dx
Vx2 Vx1 =

(12.6)
x2

w(x)dx

(12.7)

x1

The change in shear between 1 and 2, V21 , is equal to the area under the load
between x1 and x2 .
and
M

V (x)dx

M21

M2 M1 =

(12.8)
x2

V (x)dx (12.9)
x1

The change in moment between 1 and 2, M21 , is equal to the area under the
shear curve between x1 and x2 .
Victor Saouma

Introduction to Continuum Mechanics

Draft
128

BEAM THEORY

45

Note that we still need to have V1 and M1 in order to obtain V2 and M2 respectively.

46

It can be shown that the equilibrium of forces and of moments equations are nothing else than

the three dimensional linear momentum


(rb)dV =
V

12.3.3

d
dt

Tij
xj

+ bi = dvi and moment of momentum


dt

(rt)dS +
S

(rv)dV equations satised on the average over the cross section.


V

Examples

Example 12-2: Simple Shear and Moment Diagram


Draw the shear and moment diagram for the beam shown below

Solution:
The free body diagram is drawn below

Victor Saouma

Introduction to Continuum Mechanics

Draft

12.3 Shear & Moment Diagrams

Reactions are determined from


) F
(+  x = 0;

(+ ) MA = 0;

)
(+ 6 Fy = 0;

129

the equilibrium equations


RAx + 6 = 0 RAx = 6 k
(11)(4) + (8)(10) + (4)(2)(14 + 2) RFy (18) = 0 RFy = 14 k
RAy 11 8 (4)(2) + 14 = 0 RAy = 13 k

Shear are determined next.


1.
2.
3.
4.

At A the shear is equal to the reaction and is positive.


At B the shear drops (negative load) by 11 k to 2 k.
At C it drops again by 8 k to 6 k.
It stays constant up to D and then it decreases (constant negative slope since the load is
uniform and negative) by 2 k per linear foot up to 14 k.
5. As a check, 14 k is also the reaction previously determined at F .

Moment is determined last:


1. The moment at A is zero (hinge support).
2. The change in moment between A and B is equal to the area under the corresponding shear
diagram, or MBA = (13)(4) = 52.
3. etc...

Victor Saouma

Introduction to Continuum Mechanics

Draft
1210

BEAM THEORY

12.4

Beam Theory

12.4.1

Basic Kinematic Assumption; Curvature

Fig.12.7 shows portion of an originally straight beam which has been bent to the radius by end
couples M . support conditions, Fig. 12.1. It is assumed that plane cross-sections normal to the

47

O
+ve Curvature, +ve bending

d
-ve Curvature, -ve Bending

Neutral Axis
F

Y
dA

dx

Figure 12.7: Deformation of a Beam under Pure Bending


length of the unbent beam remain plane after the beam is bent.
48 Except for the neutral surface all other longitudinal bers either lengthen or shorten, thereby creating
a longitudinal strain x . Considering a segment EF of length dx at a distance y from the neutral axis,
its original length is
EF = dx = d
(12.10)

and
d =
49

dx

(12.11)

To evaluate this strain, we consider the deformed length E F


E F = ( y)d = d yd = dx y

dx

(12.12)

The strain is now determined from:


x =

dx y dx dx
E F EF

=
EF
dx

(12.13)

or after simplication
x =

(12.14)

where y is measured from the axis of rotation (neutral axis). Thus strains are proportional to the
distance from the neutral axis.
Victor Saouma

Introduction to Continuum Mechanics

Draft

12.4 Beam Theory

1211

50 (Greek letter rho) is the radius of curvature. In some textbook, the curvature (Greek letter
kappa) is also used where
1
(12.15)
=

thus,
x = y

(12.16)

51 It should be noted that Galileo (1564-1642) was the rst one to have made a contribution to beam
theory, yet he failed to make the right assumption for the planar cross section. This crucial assumption
was made later on by Jacob Bernoulli (1654-1705), who did not make it quite right. Later Leonhard
Euler (1707-1783) made signicant contributions to the theory of beam deection, and nally it was
Navier (1785-1836) who claried the issue of the kinematic hypothesis.

12.4.2

Stress-Strain Relations

52 So far we considered the kinematic of the beam, yet later on we will need to consider equilibrium in
terms of the stresses. Hence we need to relate strain to stress.

53

For linear elastic material Hookes law states


(12.17)

x = Ex
where E is Youngs Modulus.
54

Combining Eq. with equation 12.16 we obtain


x = Ey

12.4.3

(12.18)

Internal Equilibrium; Section Properties

55 Just as external forces acting on a structure must be in equilibrium, the internal forces must also
satisfy the equilibrium equations.

The internal forces are determined by slicing the beam. The internal forces on the cut section must
be in equilibrium with the external forces.

56

12.4.3.1
57

Fx = 0; Neutral Axis

The rst equation we consider is the summation of axial forces.

58 Since there are no external axial forces (unlike a column or a beam-column), the internal axial forces
must be in equilibrium.

Fx = 0

x dA = 0

(12.19)

where x was given by Eq. 12.18, substituting we obtain


x dA =
A

Victor Saouma

EydA = 0

(12.20-a)

Introduction to Continuum Mechanics

Draft
1212

BEAM THEORY

But since the curvature and the modulus of elasticity E are constants, we conclude that
(12.21)

ydA = 0
A

or the rst moment of the cross section with respect to the z axis is zero. Hence we conclude that the
neutral axis passes through the centroid of the cross section.
12.4.3.2

M = 0; Moment of Inertia

59 The second equation of internal equilibrium which must be satised is the summation of moments.
However contrarily to the summation of axial forces, we now have an external moment to account for,
the one from the moment diagram at that particular location where the beam was sliced, hence

Mz = 0; +ve; M =

x ydA

(12.22)

Ext.

Int.

where dA is an dierential area a distance y from the neutral axis.


60

Substituting Eq. 12.18


M

x
61

=
=

Ey

x ydA
A

y 2 dA

M = E

(12.23)

We now pause and dene the section moment of inertia with respect to the z axis as
def

y 2 dA

I =

(12.24)

I
c

(12.25)

and section modulus as


def

S =

12.4.4

Beam Formula

62 We now have the ingredients in place to derive one of the most important equations in structures, the
beam formula. This formula will be extensively used for design of structural components.

63

We merely substitute Eq. 12.24 into 12.23,


M

y 2 dA

E
A

y 2 dA

M
EI

==

(12.26)

which shows that the curvature of the longitudinal axis of a beam is proportional to the bending moment
M and inversely proportional to EI which we call exural rigidity.
64

Finally, inserting Eq. 12.18 above, we obtain


x

Victor Saouma

=
=

Ey
M
EI

x = M y
I

(12.27)

Introduction to Continuum Mechanics

Draft

12.4 Beam Theory

1213

Hence, for a positive y (above neutral axis), and a positive moment, we will have compressive stresses
above the neutral axis.
65 Alternatively, the maximum ber stresses can be obtained by combining the preceding equation with
Equation 12.25

x =

M
S

12.4.5

Limitations of the Beam Theory

12.4.6

(12.28)

Example

Example 12-3: Design Example


A 20 ft long, uniformly loaded, beam is simply supported at one end, and rigidly connected at the
other. The beam is composed of a steel tube with thickness t = 0.25 in. Select the radius such that
max 18 ksi, and max L/360.

1 k/ft
r

0.25

20
Solution:
1. Steel has E = 29, 000 ksi, and from above Mmax =

wL2
8 ,

max =

wL4
185EI ,

and I = r3 t.

2. The maximum moment will be


(1) k/ft(20)2 ft2
wL2
=
= 50 k.ft
8
8

Mmax =

(12.29)

3. We next seek a relation between maximum deection and radius


max
I

=
=

wL4
185EI
3

=
=
=

M
S
I
r

r3 t

=
=
=

r t

4. Similarly for the stress

S
I

wL4
185Er 3 t
4
3
3
(1) k/ft(20)4 ft (12)3 in / ft
3 (0.25) in
(185)(29,000) ksi(3.14)r
65.65
r3

=
=

M
r 2 t
(50) k.ft(12) in/ft
(3.14)r 2 (0.25) in
764
r2

(12.30)

(12.31)

5. We now set those two values equal to their respective maximum


max
max

Victor Saouma

(20) ft(12) in/ft


L
65.65
=
= 0.67 in =
r=
360
360
r3
764
764
= 6.51 in
= (18) ksi = 2 r =
r
18
=

65.65
= 4.61 in (12.32-a)
0.67
(12.32-b)

Introduction to Continuum Mechanics

Draft

BEAM THEORY

Victor Saouma

Introduction to Continuum Mechanics

1214

Draft
Chapter 13

VARIATIONAL METHODS
Abridged section from authors lecture notes in finite elements.
Variational methods provide a powerful method to solve complex problems in continuum mechanics
(and other elds as well).

As shown in Appendix C, there is a duality between the strong form, in which a dierential equation
(or Eulers equation) is exactly satised at every point (such as in Finite Dierences), and the weak
form where the equation is satised in an averaged sense (as in nite elements).

Since only few problems in continuum mechanics can be solved analytically, we often have to use
numerical techniques, Finite Elements being one of the most powerful and exible one.

At the core of the nite element formulation are the variational formulations (or energy based methods)
which will be discussed in this chapter.

For illustrative examples, we shall use beams, but the methods is obviously applicable to 3D continuum.

13.1
6

Preliminary Denitions

Work is dened as the product of a force and displacement


W

def

F.ds

(13.1-a)

Fx dx + Fy dy

(13.1-b)

dW

Energy is a quantity representing the ability or capacity to perform work.

The change in energy is proportional to the amount of work performed. Since only the change of
energy is involved, any datum can be used as a basis for measure of energy. Hence energy is neither
created nor consumed.

The rst principle of thermodynamics (Eq. 6.44), states


The time-rate of change of the total energy (i.e., sum of the kinetic energy and the internal
energy) is equal to the sum of the rate of work done by the external forces and the change
of heat content per unit time:
d
dt (K

+ U ) = We + H

(13.2)

Draft
132

VARIATIONAL METHODS

U0
A

U0
A

U0
A

U0
A

Nonlinear

Linear

Figure 13.1: *Strain Energy and Complementary Strain Energy


where K is the kinetic energy, U the internal strain energy, W the external work, and H the heat input
to the system.
10 For an adiabatic system (no heat exchange) and if loads are applied in a quasi static manner (no
kinetic energy), the above relation simplies to:

We = U

13.1.1
11

(13.3)

Internal Strain Energy

The strain energy density of an arbitrary material is dened as, Fig. 13.1

def

U0 =

:d

(13.4)

12

The complementary strain energy density is dened

U0 =

def

13

:d

(13.5)

The strain energy itself is equal to


U

def

U0 d

(13.6)

U0 d

(13.7)

def

14 To obtain a general form of the internal strain energy, we rst dene a stress-strain relationship
accounting for both initial strains and stresses

= D:( 0 ) + 0

(13.8)

where D is the constitutive matrix (Hookes Law); is the strain vector due to the displacements u;
is the initial strain vector; 0 is the initial stress vector; and is the stress vector.

15 The initial strains and stresses are the result of conditions such as heating or cooling of a system or
the presence of pore pressures in a system.

Victor Saouma

Introduction to Continuum Mechanics

Draft

13.1 Preliminary Denitions

16

133

The strain energy U for a linear elastic system is obtained by substituting


= D:

(13.9)

with Eq. 13.4 and 13.8

U=

1
2

T :D:d

T :D:0 d +

T : 0 d

(13.10)

where is the volume of the system.


Considering uniaxial stresses, in the absence of initial strains and stresses, and for linear elastic
systems, Eq. 13.10 reduces to

17

U=

18

1
2

(13.11)

E d

When this relation is applied to various one dimensional structural elements it leads to

Axial Members:

d
2

U=

=P
A
P
= AE
d = Adx

1
2

U=

L
0

P2
dx
AE

(13.12)

Flexural Members:

Mz y

x = Iz
Mz y
U=
= EIz

d = dAdx

2
y dA = Iz

U=

1
2

1
2

L
0

M2
dx
EIz

(13.13)

13.1.2
19

External Work

External work W performed by the applied loads on an arbitrary system is dened as


def

uT bd +

We =

uT
td

(13.14)

where b is the body force vector; is the applied surface traction vector; and t is that portion of the
t
boundary where is applied, and u is the displacement.
t
20

For point loads and moments, the external work is


f

We =
0

Victor Saouma

P d +

M d

(13.15)

Introduction to Continuum Mechanics

Draft
134

21

VARIATIONAL METHODS

For linear elastic systems, (P = K) we have for point loads

P = K

f
1
f
d = K2
We = K
f
2
P d
We =
0

(13.16)

When this last equation is combined with Pf = Kf we obtain


We =

1
Pf f
2

(13.17)

We =

1
Mf f
2

(13.18)

where K is the stiness of the structure.


22

Similarly for an applied moment we have

13.1.3

Virtual Work

23 We dene the virtual work done by the load on a body during a small, admissible (continuous and
satisfying the boundary conditions) change in displacements.

def

Internal Virtual Work Wi

:d

(13.19)

def

External Virtual Work We

tud +

bud

(13.20)

where all the terms have been previously dened and b is the body force vector.
Note that the virtual quantity (displacement or force) is one that we will approximate/guess as long
as it meets some admissibility requirements.

24

13.1.3.1

Internal Virtual Work

Next we shall derive a displacement based expression of U for each type of one dimensional structural
member. It should be noted that the Virtual Force method would yield analogous ones but based on
forces rather than displacements.

25

26 Two sets of solutions will be given, the rst one is independent of the material stress strain relations,
and the other assumes a linear elastic stress strain relation.

Elastic Systems In this set of formulation, we derive expressions of the virtual strain energies which
are independent of the material constitutive laws. Thus U will be left in terms of forces and
displacements.
Axial Members:

U =
0

d = Adx

Victor Saouma

U = A

dx

(13.21)

Introduction to Continuum Mechanics

Draft

13.1 Preliminary Denitions

135

Flexural Members:
U =

x x d

M
=
y
A

= y y =
x ydA

M=

d =

dAdx
0

x dA
U =
A

M dx

(13.22)

Linear Elastic Systems Should we have a linear elastic material ( = E) then:

Axial Members:
U =

x = Ex = E du
dx
= d(u)
dx
d = Adx

U =

E
0

du d(u)
Adx
dx dx

(13.23)

Flexural Members:

My

x = Iz
d v
x = 2 Ey
2
d v
U =
dx
M = dx2 EIz

(v)

x = x = d dx2 y

d = dAdx

U =

x x d

or:
Eq. 13.24

13.1.3.2
27

d2 (v)
d2 v
Ey
ydAdx
2
dx2
A dx

U =

y dA = Iz

EIz
0

(13.24)

d2 v d2 (v)
dx
dx2 dx2

(13.25)

External Virtual Work W

For concentrated forces (and moments):

W =

qdx +

(i )Pi +

(13.26)

(i )Mi

where: i = virtual displacement.

13.1.4

Complementary Virtual Work

28 We dene the complementary virtual work done by the load on a body during a small, admissible
(continuous and satisfying the boundary conditions) change in displacements.

Complementary Internal Virtual Work Wi

def

Complementary External Virtual Work We

def

Victor Saouma

:d

(13.27)

utd

(13.28)

Introduction to Continuum Mechanics

Draft
136

13.1.5
29

VARIATIONAL METHODS

Potential Energy

The potential of external work W in an arbitrary system is dened as


def

We =

uT bd +

uT + uP
td

(13.29)

where u are the displacements, b is the body force vector; is the applied surface traction vector; t is
t
that portion of the boundary where is applied, and P are the applied nodal forces.
t
30

Note that the potential of the external work (W) is dierent from the external work itself (W )

31

The potential energy of a system is dened as

def

U We

(13.30)

U0 d

u + uP
td

ubd +

(13.31)

32 Note that in the potential the full load is always acting, and through the displacements of its points
of application it does work but loses an equivalent amount of potential, this explains the negative sign.

13.2

Principle of Virtual Work and Complementary Virtual


Work

The principles of Virtual Work and Complementary Virtual Work relate force systems which satisfy
the requirements of equilibrium, and deformation systems which satisfy the requirement of compatibility:

33

1. In any application the force system could either be the actual set of external loads dp or some
virtual force system which happens to satisfy the condition of equilibrium p. This set of external
forces will induce internal actual forces d or internal hypothetical forces compatible with the
externally applied load.
2. Similarly the deformation could consist of either the actual joint deections du and compatible
internal deformations d of the structure, or some hypothetical external and internal deformation
u and which satisfy the conditions of compatibility.
34

Thus we may have 2 possible combinations, Table 13.1: where: d corresponds to the actual, and

1
2

Force
External Internal
p

dp
d

Deformation
External Internal
du
d
u

Formulation
U
U

Table 13.1: Possible Combinations of Real and Hypothetical Formulations


(with an overbar) to the hypothetical values.

13.2.1

Principle of Virtual Work

35 Derivation of the principle of virtual work starts with the assumption of that forces are in equilibrium
and satisfaction of the static boundary conditions.

Victor Saouma

Introduction to Continuum Mechanics

Draft

13.2 Principle of Virtual Work and Complementary Virtual Work

36

137

The Equation of equilibrium (Eq. 6.26) which is rewritten as


xy
xx
+
+ bx
x
y
xy
yy
+
+ by
y
x

(13.32)

(13.33)

where b representing the body force. In matrix form, this can be rewritten as

0 y xx
bx
x

=0
+

by
0 y x yy
xy

(13.34)

or
LT + b = 0

(13.35)

Note that this equation can be generalized to 3D.


37 The surface of the solid can be decomposed into two parts t and u where tractions and displacements are respectively specied.

= t + u
t = on t Natural B.C.
t

u = u on u Essential B.C.

(13.36-a)

(13.36-b)
(13.36-c)

Equations 13.35 and 13.36-b constitute a statically admissible stress eld.


38

The principle of virtual work (or more specically of virtual displacement) can be stated as
A deformable system is in equilibrium if the sum of the external virtual work and the internal
virtual work is zero for virtual displacements u which are kinematically admissible.

The major governing equations are summarized


T :d

uT bd

Wi

uT
td

(13.37)

= L:u
u = 0

in
on

(13.38)
(13.39)

t
We

39 Note that the principle is independent of material properties, and that the primary unknowns are the
displacements.

Example 13-1: Tapered Cantiliver Beam, Virtual Displacement


Analyse the problem shown in Fig. 13.2, by the virtual displacement method.
Solution:
1. For this exural problem, we must apply the expression of the virtual internal strain energy as
derived for beams in Eq. 13.25. And the solutions must be expressed in terms of the displacements
which in turn must satisfy the essential boundary conditions.

Victor Saouma

Introduction to Continuum Mechanics

Draft
138

VARIATIONAL METHODS

Figure 13.2: Tapered Cantilivered Beam Analysed by the Vitual Displacement Method
The approximate solutions proposed to this problem are
v
v

x
v2
2l
x
x 2
2
3
L
L
1 cos

=
=

(13.40)
3

v2

(13.41)

2. These equations do indeed satisfy the essential B.C. (i.e kinematic), but for them to also satisfy
equilibrium they must satisfy the principle of virtual work.
3. Using the virtual displacement method we evaluate the displacements v2 from three dierent
combination of virtual and actual displacement:
Solution
1
2
3

Total
Eqn. 13.40
Eqn. 13.40
Eqn. 13.41

Virtual
Eqn. 13.41
Eqn. 13.40
Eqn. 13.41

Where actual and virtual values for the two assumed displacement elds are given below.
Trigonometric (Eqn. 13.40)

Polynomial (Eqn. 13.41)

1 cos

v2

1 cos x v2
2l

x
2l

6
12x
L2 L3
6
12x
L2 L3

x
4L2 cos 2l v2
2
x
4L2 cos 2l v2

v
v

x 2
x 3
2 L
L
x 2
x 3
2 L
L

v2
v2

v2
v2

v EIz v dx

(13.42)

= P2 v2

(13.43)

Solution 1:
L

=
=
=

Victor Saouma

12x
2
6
x
v2
cos
3
4L2
2l
L2
L
0
3EI1
10 16
+ 2 v2 v2
1
2L3

P2 v2

v2 EI1 1

x
dx
2L

(13.44)

Introduction to Continuum Mechanics

Draft

13.2 Principle of Virtual Work and Complementary Virtual Work

139

which yields:
v2 =

P2 L3
2.648EI1

(13.45)

Solution 2:
L

=
=
=

x
4
x
v2 v2 EI1 1
dx
cos2
4
2l
2l
0 16L
4 EI1 3
1
+ 2 v2 v2
3
32L
4
P2 v2

(13.46)

which yields:
v2 =

P2 L3
2.57EI1

(13.47)

Solution 3:
L

=
0

=
=

12x
6
3
L2
L

x
EI1 v2 v2 dx
2l

9EI
v2 v2
L3
P2 v2

(13.48)

which yields:
v2 =

13.2.2

P2 L3
9EI

(13.49)

Principle of Complementary Virtual Work

40 Derivation of the principle of complementary virtual work starts from the assumption of a kinematicaly
admissible displacements and satisfaction of the essential boundary conditions.
41 Whereas we have previously used the vector notation for the principle of virtual work, we will now
use the tensor notation for this derivation.

42

The kinematic condition (strain-displacement):


ij =

43

1
(ui,j + uj,i )
2

The essential boundary conditions are expressed as

ui = u on u

44

(13.50)

(13.51)

The principle of virtual complementary work (or more specically of virtual force) which can be stated

as
A deformable system satises all kinematical requirements if the sum of the external complementary virtual work and the internal complementary virtual work is zero for all statically
admissible virtual stresses ij .
The major governing equations are summarized
Victor Saouma

Introduction to Continuum Mechanics

Draft
1310

VARIATIONAL METHODS

Figure 13.3: Tapered Cantilevered Beam Analysed by the Virtual Force Method

ij ij d

ui ti d

(13.52)

in
on

(13.53)
(13.54)

Wi

We

ij,j = 0
ti = 0

45 Note that the principle is independent of material properties, and that the primary unknowns are the
stresses.

46

Expressions for the complimentary virtual work in beams are given in Table 13.3
Example 13-2: Tapered Cantilivered Beam; Virtual Force
Exact solution of previous problem using principle of virtual work with virtual force.
L

M
0

M
dx =
EIz

Internal

(13.55)

External

Note: This represents the internal virtual strain energy and external virtual work written in terms
of forces and should be compared with the similar expression derived in Eq. 13.25 written in terms of
displacements:
L
d2 v d2 (v)
EIz 2
dx
(13.56)
U =
dx dx2
0

Here: M and P are the virtual forces, and


If P = 1, then M = x and M = P2 x or:

M
EIz
L

(1)

x
0

=
=

and are the actual displacements. See Fig. 13.3


P2 x
EI1 (.5 +

P2
EI1
P2 2L
EI1

x2

x dx
L)

L+x
2l
L
2
0

dx

x
dx
L+x

(13.57)

From Mathematica we note that:


0

1 1
x2
= 3
(a + bx)2 2a(a + bx) + a2 ln(a + bx)
a + bx
b 2
0
Victor Saouma

(13.58)

Introduction to Continuum Mechanics

Draft

13.3 Potential Energy

1311

Thus substituting a = L and b = 1 into Eqn. 13.58, we obtain:

=
=
=
=

2P2 L 1
(L + x)2 2L(L + x) + L2 ln(L + x) |L
0
EI1 2
2P2 L
L2
+ 2L2 + L2 log L
2L2 4L2 + L2 ln 2L
EI1
2
1
2P2 L 2
L (ln 2 )
EI1
2
3
P2 L
2.5887EI1

(13.59)

Similarly:
L

=
=

2M L
M (1)
2M L L 1
=
ln(L + x) |L
=
0
x
EI1 0 L + x
EI1
0 EI1 .5 + L
2M L
2M L
ML
(ln 2L ln L) =
ln 2 =
EI1
EI1
.721EI1

13.3

Potential Energy

13.3.1

(13.60)

Derivation

47

From section ??, if U0 is a potential function, we take its dierential


U0
dij
ij
U0

dij
dU0 =
ij
dU0 =

48

(13.61-a)
(13.61-b)

However, from Eq. 13.4


ij

U0

ij dij

(13.62-a)

dU0

ij dij

(13.62-b)

ij

(13.63)

ij

(13.64)

thus,
U0
ij

U0
ij
49

We now dene the variation of the strain energy density at a point1


U0 =

50

U
ij = ij ij
ij

(13.65)

Applying the principle of virtual work, Eq. 13.37, it can be shown that
1 Note

U =

that the variation of strain energy density is, U0 = ij ij , and the variation of the strain energy itself is
U0 d.

Victor Saouma

Introduction to Continuum Mechanics

Draft
1312

VARIATIONAL METHODS

k= 500 lbf/in

mg= 100 lbf

Figure 13.4: Single DOF Example for Potential Energy

(13.66)

def

U We

(13.67)

U0 d

51

u + uP
td

ubd +

(13.68)

We have thus derived the principle of stationary value of the potential energy:
Of all kinematically admissible deformations (displacements satisfying the essential boundary conditions), the actual deformations (those which correspond to stresses which satisfy
equilibrium) are the ones for which the total potential energy assumes a stationary value.

52

For problems involving multiple degrees of freedom, it results from calculus that

53

1 +
2 + . . . +
n
1
2
n

(13.69)

It can be shown that the minimum potential energy yields a lower bound prediction of displacements.

54 As an illustrative example (adapted from Willam, 1987), let us consider the single dof system shown
in Fig. 13.4. The strain energy U and potential of the external work W are given by

We

1
u(Ku) = 250u2
2
mgu = 100u

(13.70-a)
(13.70-b)

Thus the total potential energy is given by


= 250u2 100u

(13.71)

and will be stationary for


=

Victor Saouma

d
= 0 500u 100 = 0 u = 0.2 in
du

(13.72)

Introduction to Continuum Mechanics

Draft

13.3 Potential Energy

1313

Potential Energy of Single DOF Structure

Energy [lbfin]

20.0

Total Potential Energy


Strain Energy
External Work

0.0

20.0

40.0
0.00

0.10

0.20
Displacement [in]

0.30

Figure 13.5: Graphical Representation of the Potential Energy


Substituting, this would yield
U
W

= 250(0.2)2
= 100(0.2)
= 10 20

= 10 lbf-in
= 20 lbf-in
= 10 lbf-in

(13.73)

Fig. 13.5 illustrates the two components of the potential energy.

13.3.2

Rayleigh-Ritz Method

55 Continuous systems have innite number of degrees of freedom, those are the displacements at every
point within the structure. Their behavior can be described by the Euler Equation, or the partial
dierential equation of equilibrium. However, only the simplest problems have an exact solution which
(satises equilibrium, and the boundary conditions).
56 An approximate method of solution is the Rayleigh-Ritz method which is based on the principle of
virtual displacements. In this method we approximate the displacement eld by a function

u1

c1 1 + 1
i i
0

(13.74-a)

c2 2 + 2
i i
0

(13.74-b)

c3 3 + 3
i i
0

(13.74-c)

i=1
n

u2

i=1
n

u3

i=1

where cj denote undetermined parameters, and are appropriate functions of positions.


i
57

should satisfy three conditions


1. Be continuous.

Victor Saouma

Introduction to Continuum Mechanics

Draft
1314

VARIATIONAL METHODS

2. Must be admissible, i.e. satisfy the essential boundary conditions (the natural boundary conditions
are included already in the variational statement. However, if also satisfy them, then better
results are achieved).
3. Must be independent and complete (which means that the exact displacement and their derivatives
that appear in can be arbitrary matched if enough terms are used. Furthermore, lowest order
terms must also be included).
In general is a polynomial or trigonometric function.
58

We determine the parameters cj by requiring that the principle of virtual work for arbitrary variations
i
or
n
1 2 3
(13.75)
c + 2 ci + 3 ci = 0
(u1 , u2 , u3 ) =
c1 i
ci
ci
i
i=1

cj .
i

for arbitrary and independent variations of c1 , c2 , and c3 , thus it follows that


i
i
i

cj
i

=0

i = 1, 2, , n; j = 1, 2, 3

(13.76)

Thus we obtain a total of 3n linearly independent simultaneous equations. From these displacements,
we can then determine strains and stresses (or internal forces). Hence we have replaced a problem with
an innite number of d.o.f by one with a nite number.
59

Some general observations


1. cj can either be a set of coecients with no physical meanings, or variables associated with nodal
i
generalized displacements (such as deection or displacement).
2. If the coordinate functions satisfy the above requirements, then the solution converges to the
exact one if n increases.
3. For increasing values of n, the previously computed coecients remain unchanged.
4. Since the strains are computed from the approximate displacements, strains and stresses are generally less accurate than the displacements.
5. The equilibrium equations of the problem are satised only in the energy sense = 0 and not in
the dierential equation sense (i.e. in the weak form but not in the strong one). Therefore the displacements obtained from the approximation generally do not satisfy the equations of equilibrium.
6. Since the continuous system is approximated by a nite number of coordinates (or d.o.f.), then the
approximate system is stier than the actual one, and the displacements obtained from the Ritz
method converge to the exact ones from below.

Example 13-3: Uniformly Loaded Simply Supported Beam; Polynomial Approximation


For the uniformly loaded beam shown in Fig. 13.6
let us assume a solution given by the following innite series:
v = a1 x(L x) + a2 x2 (L x)2 + . . .

(13.77)

for this particular solution, let us retain only the rst term:
v = a1 x(L x)

(13.78)

We observe that:
Victor Saouma

Introduction to Continuum Mechanics

Draft

13.3 Potential Energy

1315

Figure 13.6: Uniformly Loaded Simply Supported Beam Analyzed by the Rayleigh-Ritz Method
1. Contrarily to the previous example problem the geometric B.C. are immediately satised at both
x = 0 and x = L.
2. We can keep v in terms of a1 and take

then take both a1 = 0, and a2 = 0 ).

a1

= 0 (If we had left v in terms of a1 and a2 we should

3. Or we can solve for a1 in terms of vmax (@x =


L

=U W =

Recalling that:

M
EIz

d2 v
dx2 ,

L
2)

and take

M2
dx
2EIz

vmax

= 0.

wv(x)dx

(13.79)

the above simplies to:


L

=
0

EIz
2

d2 v
dx2

wv(x) dx

(13.80)

EIz
(2a1 )2 a1 wx(L x) dx
2
EIz 2
L3
L3
4a1 L a1 w
+ a1 w
2
2
3
a1 wL3
2a2 EIz L
1
6
0

=
=
If we now take

a1

(13.81)

= 0, we would obtain:
wL3
6

a1

4a1 EIz l

wL2
24EIz

Having solved the displacement eld in terms of a1 , we now determine vmax at

wL4
24EIz

(13.82)
L
2:

x2
x
2
L L

a1

Victor Saouma

Introduction to Continuum Mechanics

Draft
1316

VARIATIONAL METHODS
wL4
96EIz

(13.83)
4

5
wL
exact
This is to be compared with the exact value of vmax = 384 wLz = 76.8EIz which constitutes 17%
EI
error.
wL2
w
Note: If two terms were retained, then we would have obtained: a1 = 24EIz & a2 = 24EIz and vmax
exact
would be equal to vmax . (Why?)

13.4
60

Summary

Summary of Virtual work methods, Table 13.2.

Virtual Work U
Complimentary Virtual Work U

Starts with
KAD
SAS

Ends with
SAS
KAD

In terms of virtual
Displacement/strains
Forces/Stresses

Solve for
Displacement
Displacement

KAD: Kinematically Admissible Dispacements


SAS: Statically Admissible Stresses
Table 13.2: Comparison of Virtual Work and Complementary Virtual Work

61

A summary of the various methods introduced in this chapter is shown in Fig. 13.7.

62 The duality between the two variational principles is highlighted by Fig.


13.8, where beginning
with kinematically admissible displacements, the principle of virtual work provides statically admissible
solutions. Similarly, for statically admissible stresses, the principle of complementary virtual work leads
to kinematically admissible solutions.
63 Finally, Table 13.3 summarizes some of the major equations associated with one dimensional rod
elements.

U
Axial

Flexure

L
0

1
2

P2
dx
AE

1
2

M
dx
EIz

Virtual Displacement U
General
Linear
L
L
du d(u)
dx
E
Adx
dx dx
0
0
L

M dx
0

d v d (v)
EIz 2
dx
dx dx2

W
1
i 2 Pi i
1
i 2 Mi i

P
M

M dx
0

Virtual Displacement W
i Pi i
i Mi i

w(x)v(x)dx
0

Virtual Force U
General
Linear
L
L
P
dx
P
dx
AE
0
0

Virtual Force W
i Pi i
i Mi i

w(x)v(x)dx
0

M
M
dx
EIz

w(x)v(x)dx
0

Table 13.3: Summary of Variational Terms Associated with One Dimensional Elements

Victor Saouma

Introduction to Continuum Mechanics

Draft
13.4 Summary

1317

+ b = 0

Natural B.C.
Essential B.C.

D:u = 0
u = 0 u

t t = 0 t
def

U0 =

ij

1
2

(ui,j + uj,i ) = 0

u i u = 0 u
def

U0 =

:d

ij,j = 0
ti = 0 t

6
Gauss

Gauss

?
?
Principle of Virtual Work
T :d

uT bd t uT td = 0

Wi We = 0

?
?
Principle of Complementary
Virtual Work
ij ij d u ui ti d = 0

Wi We = 0

?
Principle of Stationary
Potential Energy
= 0
def

= U We
U0 d ( ui bi d +

ui ti d)

?
Rayleigh-Ritz
n

cj j + j
0
i i

uj
i=1

cj
i

=0

i = 1, 2, , n;

j = 1, 2, 3

Figure 13.7: Summary of Variational Methods

Victor Saouma

Introduction to Continuum Mechanics

Draft
1318

VARIATIONAL METHODS

Kinematically Admissible Displacements


Displacements satisfy the kinematic equations
and the the kinematic boundary conditions
6

?
Principle of Stationary
Complementary Energy

Principle of Virtual Work

Principle of Complementary

Principle of Stationary
Potential Energy

Virtual Work
6

?
Statically Admissible Stresses
Stresses satisfy the equilibrium conditions
and the static boundary conditions

Figure 13.8: Duality of Variational Principles

Victor Saouma

Introduction to Continuum Mechanics

Draft
Chapter 14

INELASTICITY (incomplete)
F

t
Creep

t
Relaxation

Figure 14.1: test

Draft
2

INELASTICITY (incomplete)

Strain Hardening

Relaxation

Creep

Creep

Creep

Perfectly Elastic

Relaxation

Viscoelastic

Rigid Perfectly Plastic Elastic Perfectly Plastic

Relaxation

Elastoplastic Hardeing

Figure 14.2: mod1

Figure 14.3: v-kv


Victor Saouma

Introduction to Continuum Mechanics

Draft

E
0

Figure 14.4: vis

E
E1

Ei

En

Figure 14.5: vis

Linear Elasticity

Linear Visosity

.
=

Nonlinear Viscosity

. 1/N

Stress Threshold

Strain Threshold

s < < s
s< <
s

Figure 14.6: comp

0
Figure 14.7: epp

Victor Saouma

Introduction to Continuum Mechanics

Draft
4

INELASTICITY (incomplete)

pi

Si

Ei

Sj

Ej

Em

Figure 14.8: ehs

Victor Saouma

Introduction to Continuum Mechanics

Draft
Appendix A

SHEAR, MOMENT and


DEFLECTION DIAGRAMS for
BEAMS
Adapted from [?] 1) Simple Beam; uniform Load
L

w L
R

R
L / 2

L / 2

Shear

M max.

R
Vx
at center Mmax
Mx
max
x

= V
= w L x
2
2
= wL
8
= wx (L x)
2
4
5
= 384 wL
EI
wx
= 24EI (L3 2Lx2 + x3 )

Moment

2) Simple Beam; Unsymmetric Triangular Load

R1 = V1
Max R2 = V2
Vx
at x = .577L Mmax
Mx
at x = .5193L max
x

=
=
=
=
=
=
=

W
3
2W
3
W
W x2
3 L2

.1283W L
Wx
2
x2
3L2 (L 3 )
.01304 W L
EI
W x3
4
2 2
4
180EIL2 (3x 10L x + 7L )

Draft
A2

SHEAR, MOMENT and DEFLECTION DIAGRAMS for BEAMS

3) Simple Beam; Symmetric Triangular Load

R=V
for x <
Vx
at center Mmax

=
=
=

L
2
L
2

Mx

x
max

=
=

L
2

for x <
for x <

W
2
W
2
2
2L2 (L 4x )
WL
6
x2
W x 1 2 L2
2
3
Wx
2
480EIL2 (5L
W L3
60EI

4x2 )2

4) Simple Beam; Uniform Load Partially Distributed

Max when a < c


Max when a > c
when a < x < a + b
when x < a
when a < x < a + b
when a + b < x
at x = a + R1
w

R1 = V1
R2 = V2
Vx
Mx
Mx
Mx
Mmax

=
=
=
=
=
=
=

wb
2L (2c + b)
wb
2L (2a + b)
R1 w(x

a)
R1 x
R1 x w (x a)2
2
R2 (L x)
R1
R1 a + 2w

5) Simple Beam; Concentrated Load at Center

max
at x =
when x <
whenx <
at x =

L
2
L
2
L
2
L
2

R 1 = V1
R=V
Mmax
Mx
x
max

= wa (2L a)
2L
= 2P
= P4L
= P2x
Px
= 48EI (3L2 4x2 )
P L3
= 48EI

6) Simple Beam; Concentrated Load at Any Point

Victor Saouma

Introduction to Continuum Mechanics

Draft

A3

max when a < b


max when a > b
at x = a
when x < a
at x = a
when x < a
at x =

a(a+2b)
3

R1 = V1
R2 = V2
Mmax
Mx
a
x

& a > b max

=
=
=
=
=
=
=

Pb
L
Pa
L
P ab
L
P bx
L
P a2 b2
3EIL
P bx
2
2
2
6EIL (L b x )
P ab(a+2b) 3a(a+2b)
27EIL

7) Simple Beam; Two Equally Concentrated Symmetric Loads

when x < a
when a < x < L a

R=V
Mmax
max
x
x

= P
= Pa
Pa
= 24EI (3L2 4a2 )
Px
= 6EI (3La 3a2 x2 )
Pa
= 6EI (3Lx 3x2 a2 )

8) Simple Beam; Two Equally Concentrated Unsymmetric Loads

max when a < b


max when b < a
when a < x < L b
max when b < a
max when a < b
when x < a
when a < x < L b

R1 = V1
R 2 = V2
Vx
M1
M2
Mx
Mx

=
=
=
=
=
=
=

P
L (L a + b)
P
L (L b + a)
P
L (b a)
R1 a

R2 b
R1 x
R1 x P (x a)

9) Cantilevered Beam, Uniform Load

Victor Saouma

Introduction to Continuum Mechanics

Draft
A4

SHEAR, MOMENT and DEFLECTION DIAGRAMS for BEAMS

at x = 3 L
8

at x = .4215L

R1 = V1
R2 = V2
Vx
Mmax
M1
Mx
x
max

=
=
=
=
=
=
=
=

3
8 wL
5
8 wL
R1 wx
wL2
8
9
2
128 wL
2
R1 x wx
2
wx
3
48EI (L
wL4
185EI

3Lx+ 2x3 )

10) Propped Cantilever, Concentrated Load at Center

at x = L
when x < L
2
when L < x
2
at x = .4472L

R1 = V1
R2 = V2
Mmax
Mx
Mx
max

=
=
=
=
=
=

5P
16
11P
16
3P L
16
5P x
16
P L
2

11x
163
.009317 P L
EI

11) Propped Cantilever; Concentrated Load

at x = a
at x = L
at x = a
2
L2
when a < .414L at x = L 3L2+a 2
a
when .414L < a at x = L

a
2L+a

R1 = V1
R 2 = V2
M1
M2
a
max

=
=
=
=
=
=

max

P b2
2L3 (a + 2L)
Pa
2
2
2L3 (3L a )
R1 a
P ab
(a
2L2 2 3 + L)
Pa b
(3L + a)
12EIL32
2 3
P a (L a )
3EI (3L2 a2 )2
P ab2
a
6EI
2L+a2

12) Beam Fixed at Both Ends, Uniform Load

Victor Saouma

Introduction to Continuum Mechanics

Draft

A5

at x = 0 and x = L
at x = L
2
at x = L
2

R=V
Vx
Mmax
M
max
x

= wL
2
= w L x
2
2
= wL
122
= wL
24
wL4
= 384EI
wx2
= 24EI (L x)2

13) Beam Fixed at Both Ends; Concentrated Load

at
when
at
when

x=
x<
x=
x<

L
2
L
2
L
2
L
2

R=V
Mmax
Mx
max
x

=
=
=
=
=

P
2
PL
8
P
8 (4x L)
P L3
192EI
P x2
48EI (3L

4x)

14) Cantilever Beam; Triangular Unsymmetric Load

at x = L

at x = 0

R=V
Vx
Mmax
Mx
x
max

= 8W
3
x2
= W L2
= WL
3 2
= W x2
3L
W
= 60EIL2 (x5 5L2 x + 4L5 )
W L3
= 15EI

15) Cantilever Beam; Uniform Load

Victor Saouma

Introduction to Continuum Mechanics

Draft
A6

SHEAR, MOMENT and DEFLECTION DIAGRAMS for BEAMS

at x = L
at x = 0

R=V
Vx
Mx
Mmax
x
max

= wL
= wx
2
= wx
22
= wL
2
w
= 24EI (x4 4L3 x + 3L4 )
wL4
= 8EI

16) Cantilever Beam; Point Load

at x = L
when a < x
at x = 0
at x = a
when x < a
when a < x

R=V
Mmax
Mx
max
a
x
x

= P
= Pb
= P (x a)
P b2
= 6EI (3L b)
P b3
= 3EI
P b2
= 6EI (3L 3x b)
2
= P (Lx) (3b L + x)
6EI

17) Cantilever Beam; Point Load at Free End

at x = L
at x = 0

R=V
Mmax
Mx
max
x

= P
= PL
= Px
3
= PL
3EI
P
= 6EI (2L3 3L2 x + x3 )

18) Cantilever Beam; Concentrated Force and Moment at Free End

Victor Saouma

Introduction to Continuum Mechanics

Draft

A7

at x = 0 and x = L
at x = 0

Victor Saouma

R=V
Mx
Mmax
max
x

= P
= P L x
2
= P2L
P L3
= 12EI
2
= P (Lx) ((L + 2x)
12EI

Introduction to Continuum Mechanics

Draft
Appendix B

SECTION PROPERTIES
Section properties for selected sections are shown in Table B.1.

Draft
B2

SECTION PROPERTIES

Y
x

A
x
y
Ix
Iy

X
y

= bh
b
= 2
h
= 2
3
= bh
12
3
= hb
12

h h

A = bh b h
b
x = 2
h
y = 2
3
3
Ix = bh b h
12 3
3
Iy = hb h b
12

X
y
b
b

A
y

=
=

Ix

h(a+b)
2
h(2a+b)
3(a+b)
h3 (a2 +4ab+b2
36(a+b)

A
x
y
X
Ix
y
Iy

=
=
=
=
=

bh
2
b+c
3
h
3 3
bh
36
bh 2
36 (b

bc + c2 )

r
X

A =
Ix = Iy =

r2 =
r 4
4 =

d2
4
d4
64

A = 2rt = dt
3
Ix = Iy = r3 t = d t
8

b
X
b
a

A
Ix
Iy

= ab
3
= ab
33
= ba
4

Table B.1: Section Properties

Victor Saouma

Introduction to Continuum Mechanics

Draft
Appendix C

MATHEMATICAL
PRELIMINARIES; Part IV
VARIATIONAL METHODS
Abridged section from authors lecture notes in finite elements.

C.1
1

Euler Equation

The fundamental problem of the calculus of variation1 is to nd a function u(x) such that
b

F (x, u, u )dx

(3.1)

is stationary. Or,
= 0

(3.2)

where indicates the variation


2 We dene u(x) to be a function of x in the interval (a, b), and F to be a known function (such as the
energy density).

We dene the domain of a functional as the collection of admissible functions belonging to a class of
functions in function space rather than a region in coordinate space (as is the case for a function).

We seek the function u(x) which extremizes .

Letting u to be a family of neighbouring paths of the extremizing function u(x) and we assume that

at the end points x = a, b they coincide. We dene u as the sum of the extremizing path and some

arbitrary variation, Fig. C.1.

u(x, ) = u(x) + (x) = u(x) + u(x)

(3.3)

where is a small parameter, and u(x) is the variation of u(x)


u = u(x, ) u(x)

= (x)

(3.4-a)
(3.4-b)

1 Dierential calculus involves a function of one or more variable, whereas variational calculus involves a function of a
function, or a functional.

Draft
C2

MATHEMATICAL PRELIMINARIES; Part IV VARIATIONAL METHODS

u, u
u(x)
C
B
u(x)
du
dx

x=a

x=c

x=b

Figure C.1: Variational and Dierential Operators


and (x) is twice dierentiable, has undened amplitude, and (a) = (b) = 0. We note that u coincides

with u if = 0
6 The variational operator and the dierential calculus operator d have clearly dierent meanings. du
is associated with a neighboring point at a distance dx, however u is a small arbitrary change in u for
a given x (there is no associated x).

For boundaries where u is specied, its variation must be zero, and it is arbitrary elsewhere. The
variation u of u is said to undergo a virtual change.

To solve the variational problem of extremizing , we consider


b

(u + ) = () =

F (x, u + , u + )dx

(3.5)

a
9

Since u u as 0, the necessary condition for to be an extremum is

d()
d

10

=0

(3.6)

=0

From Eq. 3.3 and applying the chain rule with = 0, u = u, we obtain

d()
d

=
=0

F
F
+
u
u

dx = 0

(3.7)

11 It can be shown (through integration by part and the fundamental lemma of the calculus of variation)
that this would lead to

d F
F

=0
u
dx u

(3.8)

12 This dierential equation is called the Euler equation associated with and is a necessary condition
for u(x) to extremize .

13

Generalizing for a functional which depends on two eld variables, u = u(x, y) and v = v(x, y)
=

Victor Saouma

F (x, y, u, v, u,x , u,y , v,x , v,y , , v,yy )dxdy

(3.9)

Introduction to Continuum Mechanics

Draft

C.1 Euler Equation

C3

There would be as many Euler equations as dependent eld variables


F
F
F
2 F
u x u,x y u,y + x2 u,xx
F
F
F
2 F
v x v,x y v,y + x2 v,xx

2
F
xy u,xy
F
2
+ xy v,xy

+
+

2 F
y 2 u,yy
2 F
y 2 v,yy

(3.10)

14 We note that the Functional and the corresponding Euler Equations, Eq. 3.1 and 3.8, or Eq. 3.9 and
3.10 describe the same problem.
15 The Euler equations usually correspond to the governing dierential equation and are referred to as
the strong form (or classical form).
16 The functional is referred to as the weak form (or generalized solution). This classication stems
from the fact that equilibrium is enforced in an average sense over the body (and the eld variable is
dierentiated m times in the weak form, and 2m times in the strong form).
17 Euler equations are dierential equations which can not always be solved by exact methods.
An
alternative method consists in bypassing the Euler equations and go directly to the variational statement
of the problem to the solution of the Euler equations.
18 Finite Element formulation are based on the weak form, whereas the formulation of Finite Dierences
are based on the strong form.

19

Finally, we still have to dene


F

=
=

F
F
u u + u
b
F dx
a

F
F
u +
u
u
u

=
a

dx

(3.11)

As above, integration by parts of the second term yields


b

u
a

d F
F

u
dx u

(3.12)

dx

20 We have just shown that nding the stationary value of by setting = 0 is equivalent to nding
the extremal value of by setting d()
equal to zero.
d

=0

Similarly, it can be shown that as with second derivatives in calculus, the second variation 2 can
be used to characterize the extremum as either a minimum or maximum.
21

22

Revisiting the integration by parts of the second term in Eq. 3.7, we obtain
b

F
F
dx =
u
u

d F
dx
dx u

(3.13)

We note that
1. Derivation of the Euler equation required (a) = (b) = 0, thus this equation is a statement of the
essential (or forced) boundary conditions, where u(a) = u(b) = 0.
2. If we left arbitrary, then it would have been necessary to use
the natural boundary conditions.

F
u

= 0 at x = a and b. These are

23 For a problem with, one eld variable, in which the highest derivative in the governing dierential
equation is of order 2m (or simply m in the corresponding functional), then we have

Essential (or Forced, or geometric) boundary conditions, involve derivatives of order zero (the eld
variable itself) through m-1. Trial displacement functions are explicitely required to satisfy this
B.C. Mathematically, this corresponds to Dirichlet boundary-value problems.
Victor Saouma

Introduction to Continuum Mechanics

Draft
C4

MATHEMATICAL PRELIMINARIES; Part IV VARIATIONAL METHODS

Nonessential (or Natural, or static) boundary conditions, involve derivatives of order m and up.
This B.C. is implied by the satisfaction of the variational statement but not explicitly stated in
the functional itself. Mathematically, this corresponds to Neuman boundary-value problems.
These boundary conditions were already introduced, albeit in a less formal way, in Table 9.1.
24

Table C.1 illustrates the boundary conditions associated with some problems
Problem

Axial Member
Distributed load
2
AE d u + q = 0
dx2
1
u

Dierential Equation
m
Essential B.C. [0, m 1]
Natural B.C. [m, 2m 1]

du
dx

or x = Eu,x

Flexural Member
Distributed load
4
EI d w q = 0
dx4
2
w, dw
dx
d2 w
d3 w
dx2 and dx3
or M = EIw,xx and V = EIw,xxx

Table C.1: Essential and Natural Boundary Conditions

Example C-1: Extension of a Bar


The total potential energy of an axial member of length L, modulus of elasticity E, cross sectional
area A, xed at left end and subjected to an axial force P at the right one is given by
L

=
0

EA
2

du
dx

dx P u(L)

(3.14)

Determine the Euler Equation by requiring that be a minimum.


Solution:
Solution I The rst variation of is given by
L

=
0

EA
2
2

du
dx

du
dx

dx P u(L)

(3.15)

Integrating by parts we obtain


L

d
dx

u
0

EA

d
dx

du
udx + EA u P u(L)
dx
0

du
du
EA
P u(L)
dx + EA
dx
dx x=L

EA

du
dx

du
dx

u(0)

(3.16-a)

(3.16-b)

x=0

The last term is zero because of the specied essential boundary condition which implies that
u(0) = 0. Recalling that in an arbitrary operator which can be assigned any value, we set the
coecients of u between (0, L) and those for u at x = L equal to zero separately, and obtain
Euler Equation:

Victor Saouma

d
dx

EA

du
dx

=0

0<x<L

(3.17)

Introduction to Continuum Mechanics

Draft

C.1 Euler Equation

C5

Natural Boundary Condition:


EA

du
P =0
dx

at x = L

Solution II We have
F (x, u, u ) =

EA
2

du
dx

(3.18)

(3.19)

(note that since P is an applied load at the end of the member, it does not appear as part of
F (x, u, u ) To evaluate the Euler Equation from Eq. 3.8, we evaluate
F
F
=0 &
= EAu
u
u

(3.20-a)

Thus, substituting, we obtain


d F
F

u
dx u
du
d
EA
dx
dx

0 Euler Equation

(3.21-a)

0 B.C.

(3.21-b)

Example C-2: Flexure of a Beam


The total potential energy of a beam is given by
L

=
0

1
M pw dx =
2

1
(EIw )w pw dx
2

(3.22)

Derive the rst variational of .


Solution:
Extending Eq. 3.11, and integrating by part twice
L

0
L

F
F
w dx
w +
w
w

F dx =

=
=

(EIw w pw)dx

(3.23-a)
(3.23-b)

=
=

(EIw w )|0

[(EIw ) w pw] dx

(3.23-c)

(EIw w )|0 [(EIw ) w]|0 +

[(EIw ) + p] wdx = 0

(3.23-d)

Or
(EIw ) = p

for all x

which is the governing dierential equation of beams and


Essential
w = 0
w = 0

or
or

Natural
EIw = M = 0
(EIw ) = V = 0

at x = 0 and x = L

Victor Saouma

Introduction to Continuum Mechanics

Draft
C6

MATHEMATICAL PRELIMINARIES; Part IV VARIATIONAL METHODS

Victor Saouma

Introduction to Continuum Mechanics

Draft
Appendix D

MID TERM EXAM


Continuum Mechanics
LMC/DMX/EPFL
Prof. Saouma
Exam I (Closed notes), March 27, 1998
3 Hours
There are 19 problems worth a total of 63 points. Select any problems you want as long as the total
number of corresponding points is equal to or larger than 50.

1. (2 pts) Write in matrix form the following 3rd order tensor Dijk in R2 space. i, j, k range from 1
to 2.
2. (2 pts) Solve for Eij ai in indicial notation.
3. (4 pts) if the stress tensor at point P is given by

10 2 0
= 2 4 1
0
1 6
determine the traction (or stress vector) t on the plane passing through P and parallel to the plane
ABC where A(6, 0, 0), B(0, 4, 0) and C(0, 0, 2).
4. (5 pts) For a plane stress problem charaterized by the following stress tensor
=

6
2

2
4

use Mohrs circle to determine the principal stresses, and show on an appropriate gure the orientation of those principal stresses.
5. (4 pts) The stress tensor throughout a continuum is given with respect to Cartesian axes as

0
3x1 x2 5x2
2
0
2x2
= 5x2
2
3
0
2x2
0
3
(a) Determine the stress vector (or traction) at the point P (2, 1,
to the cylindrical surface x2 + x2 = 4 at P ,
2
3

3) of the plane that is tangent

Draft
D2

MID TERM EXAM


x

3
1

x1

(b) Are the stresses in equlibrium, explain.


2
2
2
6. (2 pts) A displacement eld is given by u = X1 X3 e1 + X1 X2 e2 + X2 X3 e3 , determine the material
deformation gradient F and the material displacement gradient J, and verify that J = F I.

7. (4 pts) A continuum body undergoes the deformation x1 = X1 + AX2, x2 = X2 + AX3 , and


x3 = X3 + AX1 where A is a constant. Determine: 1) Deformation (or Green) tensor C; and 2)
Lagrangian tensor E.
8. (4 pts) Linear and nite strain tensors can be decomposed into the sum or product of two other
tensors.
(a) Which strain tensor can be decomposed into a sum, and which other one into a product.
(b) Why is such a decomposition performed?
9. (2 pts) Why do we have a condition imposed on the strain eld (compatibility equation)?
10. (6 pts) Stress tensors:
(a) When shall we use the Piola-Kircho stress tensors?
(b) What is the dierence between Cauchy, rst and second Piola-Kircho stress tensors?
(c) In which coordinate system is the Cauchy and Piola-Kircho stress tensors expressed?
11. (2 pts) What is the dierence between the tensorial and engineering strain (Eij , ij , i = j) ?
12. (3 pts) In the absence of body forces, does the following stress distribution

2
2x1 x2
0
x2 + (x2 x2 )
x
1

x2 + (x2 x2 )
0
2x1 x2
1
2
1
2
2
0
0
(x1 + x2 )
where is a constant, satisfy equilibrium in the X1 direction?
13. (2 pts) From which principle is the symmetry of the stress tensor derived?
14. (2 pts) How is the First principle obtained from the equation of motion?
15. (4 pts) What are the 1) 15 Equations; and 2) 15 Unknowns in a thermoelastic formulation.
16. (2 pts) What is free energy ?
17. (2 pts) What is the relationship between strain energy and strain?
18. (5 pts) If a plane of elastic symmetry exists in an anisotropic

c1111 c1112 c1133 c1112 c1123


T11

T22
c2222 c2233 c2212 c2223

T33
c3333 c3312 c3323

=
c1212 c1223

T12

T23
SYM.
c2323

T31
Victor Saouma

material,

c1131
E11

c2231
E22

c3331
E33

c1231 2E12 (12 )

c2331 2E23 (23 )

c3131
2E31 (31 )

Introduction to Continuum Mechanics

Draft
then,

D3

1
aj = 0
i
0

0
1
0

0
0
1

show that under these conditions c1131 is equal to zero.


19. (6 pts) The state of stress at a point of structural steel is given by

6 2 0
T = 2 3 0 MP a
0 0 0
with E = 207 GPa, = 80 GPa, and = 0.3.
(a) Determine the engineering strain components
(b) If a ve centimer cube of structural steel is subjected to this stress tensor, what would be the
change in volume?

Victor Saouma

Introduction to Continuum Mechanics

Draft

MID TERM EXAM

Victor Saouma

Introduction to Continuum Mechanics

D4

Draft
Appendix E

MATHEMATICA ASSIGNMENT
and SOLUTION
Connect to Mathematica using the following procedure:
1. login on an HP workstation
2. Open a shell (window)
3. Type xhost+
4. type rlogin mxsg1
5. On the newly opened shell, enter your password rst, and then type setenv DISPLAY xxx:0.0
where xxx is the workstation name which should appear on a small label on the workstation itself.
6. Type mathematica &
and then solve the following problems:
1. The state of stress through a continuum is given with respect to the cartesian axes Ox1 x2 x3 by

3x1 x2 5x2
0
2
0
2x3 MPa
Tij = 5x2
2
0
2x3
0

Determine the stress vector at point P (1, 1, 3) of the plane that is tangent to the cylindrical
surface x2 + x2 = 4 at P .
2
3
2. For the following stress tensor

6 3
Tij = 3 6
0
0

0
0
8

(a) Determine directly the three invariants I , II and III of the following stress tensor
(b) Determine the principal stresses and the principal stress directions.
(c) Show that the transformation tensor of direction cosines transforms the original stress tensor
into the diagonal principal axes stress tensor.
(d) Recompute the three invariants from the principal stresses.
(e) Split the stress tensor into its spherical and deviator parts.
(f) Show that the rst invariant of the deviator is zero.
3. The Lagrangian description of a deformation is given by x1 = X1 +X3 (e2 1), x2 = X2 +X3 (e2 e2 ,
and x3 = e2 X3 where e is a constant. SHow that the Jacobian J does not vanish and determine
the Eulerian equations describing this motion.
2
2
2
4. A displacement eld is given by u = X1 X3 e1 + X1 X2 e2 + X2 X3 e3 . Determine independently the
material deformation gradient F and the material displacement gradient J and verify that J = FI.

Draft
E2

MATHEMATICA ASSIGNMENT and SOLUTION

5. A continuum body undergoes the deformation x1 = X1 , x2 = X2 + AX3 , x3 = X3 + AX2 where A


is a constant. Compute the deformation tensor C and use this to determine the Lagrangian nite
strain tensor E.
6. A continuum body undergoes the deformation x1 = X1 + AX2 , x2 = X2 + AX3 , x3 = X3 + AX2
where A is a constant.
(a) Compute the deformation tensor C
(b) Use the computed C to determine the Lagrangian nite strain tensor E.
(c) COmpute the Eulerian strain tensor E and compare with E for very small values of A.
7. A continuum body undergoes the deformation x1 = X1 + 2X2 , x2 = X2 , x3 = X3
(a)
(b)
(c)
(d)
(e)

Determine
Determine
Determine
Determine
Determine

the
the
the
the
the

Greens deformation tensor C


principal values of C and the corresponding principal directions.
right stretch tensor U and U1 with respect to the principal directions.
right stretch tensor U and U1 with respect to the ei basis.
orthogonal rotation tensor R with respect to the ei basis.

8. A continuum body undergoes the deformation x1


stress tensor for this body is

100
Tij = 0
0

1
= 4X1 , x2 = 2 X2 , x3 = 1 X3 and the Cauchy
2

0 0
0 0 MPa
0 0

(a) Determine the corresponding rst Piola-Kircho stress tensor.


(b) Determine the corresponding second Piola-Kircho stress tensor.
(c) Determine the pseudo stress vector associated with the rst Piola-Kircho stress tensor on the
e1 plane in the deformed state.
(d) Determine the pseudo stress vector associated with the second Piola-Kircho stress tensor on
the e1 plane in the deformed state.
9. Show that in the case of isotropy, the anisotropic stress-strain relation

c1111 c1112 c1133 c1112 c1123 c1131

c2222 c2233 c2212 c2223 c2231

c3333 c3312 c3323 c3331


cAniso =
ijkm

c1212 c1223 c1231

SYM.
c2323 c2331
c3131

reduces to

ciso
ijkm

c1111

c1122
c2222

SYM.

c1133
c2233
c3333

0 0
0 0
0 0
a 0
b

0
0
0
0
0
c

with a = 1 (c1111 c1122 ), b = 1 (c2222 c2233 ), and c = 1 (c3333 c1133 ).


2
2
2
10. Determine the stress tensor at a point where the Lagrangian strain tensor is given by

30 50 20
Eij = 50 40 0 106
20 0 30
and the material is steel with = 119.2 GPa and = 79.2 GPa.

Victor Saouma

Introduction to Continuum Mechanics

Draft

E3

11. Determine the strain tensor at a point where the Cauchy stress tensor is given by

100 42 6
Tij = 42 2 0 MPa
6
0 15
with E = 207 GPa, = 79.2 GPa, and = 0.30
12. Determine the thermally induced stresses in a constrained body for a rise in temerature of 50oF ,
= 5.6 106 / 0F
13. Show that the inverse of

1

xx
0
0
0
xx

yy
0
0
0

1
yy

zz
zz
1
0
0
0
=
(5.1)
0

0
0
1+
0
0
xy (2xy ) E
xy

(2 )
yz
yz yz
0
0
0
0
1+
0

zx (2zx )

is

xx
yy
zz
xy
yz
zx

E
(1+)(12)

zx

1+

1
0
0

0
1
0

0
0
1

xx
yy
zz
xy (2xy )
yz (2yz )
zx (2zx )

(5.2)

and then derive the relations between stresses in terms of strains, and strains in terms of stress, for
plane stress and plane strain.
14. Show that the function = f (r) cos 2 satises the biharmonic equation ( ) = 0 Note: You
must <<CalculusVectorAnalysis, dene , and SetCoordinates[Cylindrical[r,,z]], and
nally use the Laplacian (or Biharmonic) functions.
15. Solve for
Trr
Tr

Tr
T

cos
sin

sin
cos

0
0

0
0

cos
sin

sin
cos

(5.3)

16. If a point load p is applied on a semi-innite medium

p
1
r

p
show that for = r sin we have the following stress tensors:

2p cos
r
0

0
0

2p cos
r

2p sinrcos

2p sinrcos
2

2p sinr cos

(5.4)

Determine the maximum principal stress at an y arbitrary point, (contour) plot the magnitude of
this stress below p. Note that D[,r], D[,{,2}] would give the rst and second derivatives of
with respect to r and respectively.

Victor Saouma

Introduction to Continuum Mechanics

Draft

MATHEMATICA ASSIGNMENT and SOLUTION

Victor Saouma

Introduction to Continuum Mechanics

E4

Das könnte Ihnen auch gefallen