Download as pdf or txt
Download as pdf or txt
You are on page 1of 587

Issam Doghri

Mechanics of Deformable Solids


Springer-Verlag Berlin Heidelberg GmbH

ONLINE LlBRARY
Engineering
http://www.springer.de/engine/
Issam Doghri

Mechanics of
Deformable Solids
Linear, Nonlinear, Analytical and Computational Aspects

with 233 Figures

" Springer
Dr. Issam Doghri
Associate Professor of Applied Mechanics
Universite catholique de Louvain
CESAME
Euler Building
4 Avenue G. Lemaitre
B - 1348 Louvain-la-Neuve
Belgium

EMAIL: doghri@mema.ucl.ac.be

Cataloging-in-Publication Date applied for

Die Deutsche Bibliothek - CIP-Einheitsaufnahme


Doghri, Issam:
Mechanics of deformabIe solids : linear, nonlinear, analytical and
computational aspects 1 Issam Doghri.

(Engineering online library)


ISBN 978-3-642-08629-8 ISBN 978-3-662-04168-0 (eBook)
DOI 10.1007/978-3-662-04168-0

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

© Springer- Verlag Berlin Heidelberg 2000


Originally published by Springer-Verlag Berlin Heidelberg New York in 2000

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

Typesetting by the author with ~TPC


Cover-Design: de'bIik, Berlin

Printed on acid-free paper SPIN: 10710071 62/3020 - 5 4 3 2 1 o


Preface

This book is based on lllY experience in teaching. research, computer software


developlllcnt and consulting.
The audience for this book includes students, engineers or researchers in me-
chanical engineering, civil engineering, applied mathelllatics, materials sci-
ence, and other people who wish or rlPcd to have a good introduction to the
mechanics of deforlllable solids and structures.
SOllle chapters of the book are usually taught at the undergraduate level (e.g.,
beallls and plates in linear elasticity) while others (e.g., large deformations,
numel'ic,l! algoritluns, shells) are usually studied Cît the graduate level. Some
topies are even at the fringe of research (e.g .. computational algorithllls for
nonlinear lllechanics, damage mec:hanics. lllicro-mechanics).
1 belieye however t1wt my presentation of advanced subjects is dear enough
to be accessible to undel'graduate students and prac:ticing engineers if they
are willing to invest SOllle effort. whic:h 1 tried to minimize. ;\Iost mec:hanical
ami civil engineers will be faced in their professiemal life with at least some
problems ]wlonging to the advanced materiaL and even if they use commer-
eial software to solve thell1, a cor rect understanding of the basic ideas will
help thelll to proc:eed properly when setting up a problem, running a code
<Inel analyzing the results.
1 shall now enumerate the reaSOllS whic:h make llle think that this book is
original anei worth publishing .

• Containeel in one textbook are three S'ubjects of major interest in solid


lllechanies:
(1) Theory of linear ebstic:ity (three-elimensional theory, yariational for-
llluiations, two-elilllensionai problellls, torsion. therll1o-elasticity, etc:.)
(2) TIlf'ory of structures (beall1s, plates alld shells) in linear isotropic
elasticity.
- (3) Nonlinear lllechanics (plasticity. viscoplasticity, finite strains, etc.)
ind ud ing c:olllputationai algorithllls.
Although numerous text.books exist on each Olle of the threc subjects, and
eyen an individual parts of each subject (e.g., books on beam theory -often
callecl strength of materials). 1 am not mvare of il book which contains a
comprehensive treatment of ali three subjeets.
VI

• Of course -otherwise the book would be too lengthy- 1 had to make some
choices. Several subjects are omitted completely (e.g., dynamics, vibra-
tions, waves). AIso, each chapter is actually an introduction (or a primer)
to the relevant theme. However, the presentation is given with enough gen-
emlity and depth to achieve two goals:
- (1) Allow the readers to grasp the fundamentals of each subject and
solve basic or most common problems.
- (2) Permit them to read and study more advanced or detailed texts on
the subject if they wish to do so.
• There is a good balance between engineering and mathematics. 1 always
try to introduce a subject via an intuitive approach and then tackle its
formulat ion and analysis in mathematical terms. Many textbooks use a
mathematical equipment which is either too limited or too sophisticated:
I tried to strike the right compromise. I view mathematics as a tool, but
a wonderful one. On the one hand, it should not be so heavy as to render
the purpose obscure, and on the other hand, when used properly, it offers
insight into the engineering problem at hand. AIso, since there is usually
more than one approach to a given subject, 1 have always chosen one which
is simple, but not simplistic. As mentioned above, the presentation always
has the appropriate depth aud generality.
• I use seveml notation systems: tensor (symbolic) notation (e.g., 0"), in-
dex (component) notation (l7ij), matrices ([17ij]) and arrays ({ 171 }). I often
present formulations in at least two different notation systems. The read-
ers can use the notation they feell110re comfortable with: however they are
encouraged to try to understaud and use aU of them. Each notation has
its advantages. Tensor notation provides a neat qualitative understanding
of basic prin cip les and results, index notation is usually what the students
tind easiest to deal with, and matrices and arrays are most useful for com-
puter implementation. However, each notation system has its shortcomings
and cases where it becol11es too cumbersome or should be used with care.
• There is an emphasis on nonlinear material models, includ ing sophisti-
cated ones (e.g., non linear kinematic hardening, ductile damage, micro-
mechanically-based models). In each case, three aspects are examined: basic
experimental facts, mathel11atical formulat ion and numeric al implementa-
tion.
• Computational methods oc:c:upy a good port ion of the book. This is es-
pecially true in the chapters dealing with nonlinear mechanics, where the
emphasis is put on numerical methods, but the correc:t framework is al-
ready introduced in the chapter dealing with variational formulations in
linear elasticity. Usually, analytical and numeric al methods in mechanics
are viewed by the students as completely different worlds: they are taught
in separate courses, by different teachers using separate approaches aud
notations. l\Iy aim is to show the readers that everything stems from basic
principles; there are SOl11e problems which can be solved in closed form
VII

(usually after making sever al reasonable assumptions), and some problems


for which we seek approximate solutions by numerical means.
• In each chapter, there are seve rai problems which are completely worked
out. Some of them cover fundamental issues of each subject and allow the
readers to put the basic theory into practice. Other problems are rather
involved and lengthy, but shed more light on the subject or cover some
aspects which have not been treated in the ba sic theory. AII of them have
been proposed as examination subjects.
• There are sever al appendices which contain useful results and formulae.
• The table of contents is given elsewhere: the logic behind the chapter fiow
is as folIows.
Chapter 1 gives the minimal background information which is needed in
order to start. and the following chapters are arranged according to what 1
found from my teaching experience to be an increasing degree of difficulty.
This has three advantages for the students:
- (1) They can start solving problems very quickly and therefore their
interest re ma ins intact.
- (2) They c:an grasp fundamental notions such as stress, strain, tension,
compression. bending, torsion, energy. etc., rapielly.
- (3) They can better appreciate the simplifying assumptions and sub-
tleties of elifferent theories (e.g., solving a beam problem first by el-
ementary beam theory -strength of materials- and then by the two-
elimensional theory of elasticity).

Acknow ledgments
During my stuelies at College Saeliki (Tunis), Ecolc ."{ ationale eI·Ingenieurs
ele Tunis, Universite Pierre et l'darie Curie (Paris) anei Ecole Normale
Superieure ele Cachan (France), 1 was very luc:ky to have some truly re-
markable anel inspiring teachers who made 111e see and pursuc the beauty in
Science. Also, there are sorue classmates from those years whose frieuelship 1
stiU c:herish today.
In my professional career at the University of California-Santa Barbara,
Centric Engineering Systems (California), Universite catholique ele Louvain
(Belgium) and as a consultant for various companies. 1 hael the privilegc of
working with many really talented people with whom 1 had very interesting
discussions anei interactions.
1 typeel the book rnyself using l5IEX, but 1 benefiteel from the precious
help of two of my graeluate stuelents: Serge ],vlunhoven anei Svetoslav Nikolov.
Serge integrateel alI figures in the source files, formatteel the whole book anei
helpeel with numerous worel-processing problems. Svetoslav prepareel most of
the figures from my hanel-elrawn .. graffiti"' anel enelureel my enelless c:hanges
with patience. \Vithout the help of Serge anei Svetoslav, the book woulcl have
taken much longer to be reaely. 1 am very grateful to them.
When 1 starteel this project, 1 thought that it woulel take me at most
one year of moelerate work in oreler to put together 80me of my papers,
VIII

lecture notes and other handwritten notes. The endeavor ended up taking
alI my free time during two and a half years, including vacations, weekends,
evenings and a four-month sabbaticalleave. 1 am deeply thankful to my wife
for her unwavering support and encouragement, and 1 hope to make up to
her and our two daughters for alI the time that 1 did not spend with them.
Last, but not least, 1 am very grateful to my parents for giving me a
nurturing home and a good education, and to my Creator for blessing me
with good health and overalI luck in life.

Louvain-Ia-Neuve, December 1999.


Table of Contents

Preface....................................................... V

1. Basic mechanics ... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1


1.1 On tensors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Stress................................................. 3
1.3 Strain................................................. 5
1.4 Principal invariants and eigenvalues of stress and strain ..... 6
1.5 l\Iohr·s stress circles .................................... 7
1.6 Equilibrium............................................ 8
1. 7 Local formulation of static problems . . . . . . . . . . . . . . . . . . . . . . 9
1.8 Continuity equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 11
1.9 COl11patibility equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 12
1.10 Strength criteria. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 12
1.11 Linear elasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 16
1.12 Strain energy ... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 21
1.13 Navier equations ....................................... 23
1.14 Beltrami-Mitchell cOl11patibility equations ................. 24
1.15 Saint-Venant·s principle, Uniqueness, Superposition, Special
theories .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 25
1.16 Solved problem: composite cylinder under axialload . . . . . . .. 26

2. Variational formulations, work and energy theorems . . . . .. 29


2.1 Local formulation of static problems ...... . . . . . . . . . . . . . . .. 29
2.2 Virtual work theorel11 (VWT) . . . . . . . . . . . . . . . . . . . . . . . . . . .. 30
2.3 Displacel11ent-based variational forl11ulation . . . . . . . . . . . . . . .. 33
2.4 Potential energy theorel11 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 35
2.5 Stress-based variational formulation. . . . . . . . . . . . . . . . . . . . . .. 37
2.6 Complementary energy theorel11 . . . . . . . . . . . . . . . . . . . . . . . . .. 38
2.7 Energy bounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 39
2.8 Stored energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 40
2.9 Maxwell-Betti reciproc:ity theorem . . . . . . . . . . . . . . . . . . . . . . .. 41
2.10 Castigliano's theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 42
2.11 Introduction to nUl11eric:all11ethods ....................... 45
2.11.1 Method of Ritz .................................. 45
X Table of Contents

2.11.2 Finite element method (FEM) ..................... 45


2.12 Solved problem: volume change of an axially compressed body 47

3. Theory of beams (strength of materials) . . . . . . . . . . . . . . . . .. 49


3.1 Definitions and geometric properties . . . . . . . . . . . . . . . . . . . . .. 49
3.2 Pure bending of a straight beam: 3D elasticity solution . . . . .. 51
3.3 Basic assumptions of beam theory . . . . . . . . . . . . . . . . . . . . . . .. 55
3.3.1 Externalloads................................... 56
3.3.2 Internalloads (stress resultants) . . . . . . . . . . . . . . . . . . .. 57
3.3.3 Equilibrium equations. . . . . . . . . . . . . . . . . . . . . . . . . . . .. 58
3.3.4 Navier-Bernoulli assumption . . . . . . . . . . . . . . . . . . . . . .. 59
3.3.5 Constitutive equations .. . . . . . . . . . . . . . . . . . . . . . . . . .. 59
3.4 Displacement boundary and continuity conditions .......... 60
3.5 Computation of internalloads (stress resultants) ........... 62
3.6 Computation of normal and shear stresses . . . . . . . . . . . . . . . .. 64
3.7 Statically determinate or indeterminate problems .......... 66
3.8 Strain energy ..... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 67
3.9 Work and energy theorems .............................. 68
3.10 Influence lines. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 69
3.11 Solved problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 70
3.11.1 Shear reduced area . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 70
3.11.2 Statically determinate problems . . . . . . . . . . . . . . . . . . .. 71
3.11.3 Statically indeterminate problems .. . . . . . . . . . . . . . . .. 77
3.11.4 l\laxwell-Betti reciprocity theorem . . . . . . . . . . . . . . . . .. 80
3.11.5 Castigliano's theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 85
3.11.6 Virtual work theorem (VWT) . . . . . . . . .. .. . . . . . . . . .. 90

4. Torsion of beams ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 95


4.1 Formulation with a warping function . . . . . . . . . . . . . . . . . . . . .. 95
4.2 Formulation with a conjugate function .... . . . . . . . . . . . . . . .. 98
4.3 Formulation with Prandtl's stress function . . . . . . . . . . . . . . . .. 98
4.4 Contour lines of the stress function ....................... 100
4.5 Maximum tangential stress .............................. 101
4.6 Membrane analogy ..................................... 102
4.7 Multi-connected sections- Particular case .................. 104
4.8 Multi-connected sections- General case .................... 104
4.9 Thin tubes with variable thickness ........................ 107
4.10 Potential energy ........................................ 108
4.11 Cylindrical coordinates .................................. 109
4.12 Solved problems ........................................ 110
4.12.1 Elliptic section ................................... 110
4.12.2 Triangular section ................................ 113
4.12.3 Notched circular section ........................... 114
4.12.4 Thin rectangular section .......................... 116
4.12.5 Ritz's method- Square section ...................... 117
Table of Contents XI

4.12.6 Hollow elliptic section- Special method .............. 118


4.12.7 Hollow elliptic section- General method ............. 119
4.12.8 Thin circular tube ................................ 121
4.12.9 Thin-walled section with multiple voids ............. 121

5. Theory of thin plates ..................................... 123


5.1 Definitions and notation ................................. 123
5.2 Internalloads (stress resultants) .......................... 123
5.3 Equilibrium equations ................................... 126
5.4 Displacements .......................................... 127
5.5 Strains ................................................ 129
5.6 Constitutive equations .................................. 129
5.7 Summary: two un-coupled problems ...................... 130
5.8 Fundamental P.D.E. for bending problem .................. 131
5.9 Boundary conditions .................................... 133
5.10 Contradictions in Kirchhoff-Love theory ................... 135
5.11 Plates with two simply supported opposite edges - Levy's
method ............................................... 136
5.12 Potential energy ........................................ 139
5.13 Influence function ...................................... 140
5.14 Solved problems ........................................ 140
5.14.1 Uniformly loaded rectangular plate with two simply
supported opposite edges and two built-in eelges ...... 140
5.14.2 Uniformly loaeleel rectaugular plate with two simply
supporteel opposite eelges anei two free eelges ......... 142

6. Bending of thin plates in polar coordinates ............... 143


6.1 Change of coordinates ................................... 143
6.2 Axisymmetric problems ................................. 146
6.3 Potential energy ........................................ 148
6.4 Solveel problems ........................................ 149
6.4.1 Uniformly loaeleel plate ............................ 149
6.4.2 Uniform loael along a concentric circle ............... 150
6.4.3 Uniform pressure on a concentric elisk ............... 153
6.4.4 Plate simply supporteel on a number of points ....... 155
6.4.5 Ritz's method .................................... 160

7. Two-dimensional problems in Cartesian coordinates ...... 163


7.1 Plane strain ........................................... 163
7.2 Plane stress .......................................... 164
7.3 Summary: plane strain versus plaue stress ................. 165
7.4 Airy stress function ..................................... 167
7.5 Polynomial solutions ................................... 167
7.6 Solution by Fourier ser ies ................................ 169
7.7 Generalized plane stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
XII Table of Contents

7.8 Solved problems ........................................ 173


7.8.1 Concentrated load at the end of a cantilever beam .... 173
7.8.2 Uniform loads on the upper and lower surfaces of a
simply supported beam ........................... 177
7.8.3 Uniform load on a cantilever beam ................. 182
7.8.4 Uniform load on a beam with two clamped ends ...... 182
7.8.5 Compression of a beam in the height-direction ....... 182
7.8.6 Body forces- Beam under its own weight ............ 187

8. Two-dimensional problems in polar coordinates ........... 193


8.1 Change of coordinates ................................. 193
8.2 Summary: plane strain versus plane stress ................ 195
8.3 Airy stress function ..................................... 196
8.4 Axisymmetric plalle problems ............................ 198
8.5 Periodic Airy stress functions ............................ 200
8.6 Generalized plane strain ................................. 201
8.7 Solved problems ........................................ 201
8.7.1 Hollow circular cylinder under inner and outer pressures201
8.7.2 Composite hollow cylinder under inner and outer pres-
sures ............................................ 205
8.7.3 Coil ,vinding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
8.7.4 Bending of a curved beam . . . . . . . . . . . . . . . . . . . . . . . . . 209
8.7.5 Traction of a circular arch ......................... 214
8.7.6 Rotating disk of uniform thic:kness .................. 215
8.7.7 Rotating disk of variable thickness .................. 218
8.7.8 Stress concentration in a plate with a small circular hole222
8.7.9 Force 011 the straight edge of a semi-infinite plate ..... 224
8.7.10 Pressure on the straight edge of a semi-infinite plate .. 228
8.7.11 Compression of a disk along a diameter ............. 230
8.7.12 Compression of a disk over two opposing arcs ........ 231

9. Thermo-elasticity ......................................... 233


9.1 Constitutive equations .................................. 233
9.2 Heat equation .......................................... 234
9.3 Thermo-mechanical problem ............................. 235
9.3.1 Thermal problem ................................. 236
9.3.2 l'dec:hanic:al problem .............................. 237
9.4 Thermal stresses: some remarks .......................... 237
9.5 Solved problems ........................................ 239
9.5.1 Axisymmetric thermal stresses in a hollow cylinder ... 239
9.5.2 Thermal stresses in a composite cylinder ............ 242
9.5.3 Transient thermal stresses in a thin plate ............ 245
Table of Contents XIII

10. Elastic stability ........................................... 249


10.1 Introduction ........................................... 249
10.2 Direct and energy methods .............................. 250
10.3 Euler's method for axially compressed columns ............. 252
10.3.1 Critical buckling load ............................. 252
10.3.2 Critical buckling stress ............................ 254
10.3.3 Rell1arks ........................................ 256
10.4 Energy-based approximate rnethod ....................... 256
10.5 Non-conservative loads .................................. 258
10.6 Solved problems ........................................ 259
10.6.1 Two rigid bars connec:ted with a spring ............. 259
10.6.2 Colurnn clamped at one end ....................... 260
10.6.3 Column clamped at one end and simply supported at
the other ........................................ 261
10.6.4 Colull1n clall1ped at both ends ..................... 263
10.6.5 Column elastically built-in at one end and sill1ply sup-
ported <It the other ............................... 264
10.6.6 Column with non-uniform properties ................ 266
10.6.7 Ecc:entric: compressive load ........................ 267
10.6.8 Beam-column under compressive and bending forces .. 268
10.6.9 Energy method .................................. 269

11. Theory of thin shells. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273


11.1 Geoll1etry of the mid-surface ............................. 273
11.2 First fundamental fOrIn ................................. 274
11.3 Second fundamental form ............................... 278
11.4 Compatibility conditions of Codazzi and Gauss ............ 280
11.5 Surface of revolution ................................... 280
11.5.1 General case .................................... 280
11.5.2 Conic surfaces ................................... 283
11.6 Gradient of a vector field in curvilinear coordinates ......... 283
11. 7 Kinematics of the mid-surface ........................... 285
11.8 Displacements anei strains outside the mid-surface .......... 287
11.8.1 General theory .................................. 287
11.8.2 Application: plates in rectangular coordinates ....... 289
11.8.3 Application: plates in polar coordinates ............ 290
11. 9 Internal loads (stress resultants) ......................... 290
11.10 Equilibrium equations .................................. 292
11.10.1 General theory .................................. 292
11.10.2 Application: plates in rec:tangular coordinates ....... 295
11.10.3 Application: plates in polar coordinates ............ 295
11.11 Constitutive equations .................................. 296
11.121'dembrane theory ...................................... 298
11.13 Further reading ........................................ 298
XIV Table of Contents

12. Elasto-plasticity .......................................... 301


12.1 One-dimensional model ................................. 301
12.2 Three-dimensional model ................................ 304
12.3 Linear elasticity ........................................ 305
12.4 Equivalent stress ....................................... 306
12.5 Hardening ............................................. 306
12.6 Flow rules ............................................. 307
12.7 Tangent operator, loading/unloading, hardening/softening ... 308
12.8 Elementary examples ................................... 312
12.8.1 Uniaxial tension-compression ...................... 312
12.8.2 Simple shear .................................... 313
12.9 Boundary-value problem ................................ 313
12.10Numerical algorithms ................................... 313
12.10.1 Finite element method (F.E.M.) ................... 313
12.10.2 Return mapping algorithm ........................ 315
12.10.3 Consistent tangent operator ....................... 318
12.11 A general framework for material models .................. 320
12.11.1 State variables .................................. 320
12.11.2 Equations of state ............................... 321
12.11.3 Flow rules ...................................... 322
12.11.4 Rate-independent plasticity ....................... 323
12.11.5 Heat equation ................................... 324
12.12 A class of non-associative plasticity models ................ 325
12.13 Further reading ........................................ 327

13. Elasto-viscoplasticity ..................................... 329


13.1 One-dimensional model ................................. 329
13.2 Three-dimensional model. ............................... 331
13.3 Numerical algorithms ................................... 332
13.3.1 Return mapping algorithm ........................ 333
13.3.2 Consistent tangent operator ....................... 333
13.4 Further reading ........................................ 335

14. Nonlinear continuum mechanics .......................... 337


14.1 Kinematics ............................................ 337
14.1.1 Description of motion ............................ 338
14.1.2 Material time derivative .......................... 338
14.2 Deformation ........................................... 339
14.2.1 Deformation gradient ............................ 339
14.2.2 Polar decomposition ............................. 340
14.2.3 Spectral decompositions .......................... 341
14.2.4 Length variation ................................. 343
14.3 Strain measures ........................................ 344
14.3.1 One-dimensional case ............................ 344
14.3.2 Three-dimensional case ........................... 344
Table of Contents XV

14.4 Strain rates ........................................... 346


14.5 Balance laws .......................................... 348
14.5.1 Transport formula ............................... 348
14.5.2 Conservation of mass ............................. 349
14.5.3 Conservation of linear momentum ................. 350
14.5.4 Conservation of rotational momentum .............. 350
14.5.5 Cauchy stress tensor ............................. 351
14.5.6 Eulerian strong formulation ....................... 351
14.5.7 Eulerian weak formulat ion ........................ 352
14.5.8 Balance of work aud energy rates .................. 353
14.5.9 Nominal stress .................................. 354
14.5.10 Lagrangian weak formulation ...................... 354
14.5.11 Lagrangian strong formulation .................... 355
14.6 Conjugate stress and strain measures ..................... 356
14.6.1 Definition and examples .......................... 356
14.6.2 Interpretation of the second Piola-Kirchhoff stress ... 358
14.6.3 Uniaxial tension/compression ..................... 358
14.7 Objectivity ............................................ 359
14.7.1 Definition ...................................... 360
14.7.2 Examples ....................................... 361
14.8 Objective stress rates ................................... 363
14.8.1 Examples ....................................... 363
14.8.2 A family of objective rates ........................ 365
14.9 Laws of thermodynamics ................................ 366
14.9.1 First law ....................................... 366
14.9.2 Second law ..................................... 366
14.9.3 Clausius-Duhem inequality ....................... 367
14.10 Further reading ........................................ 367

15. Nonlinear elasticity ....................................... 369


15.1 Hyperelasticity and hypoelasticity ........................ 369
15.1.1 Definitions ...................................... 369
15.1.2 Hyperelasticity and material objectivity ............ 369
15.1.3 Elasticity tensors ................................ 371
15.1.4 Incompressibility constraint ....................... 373
15.2 Principal invariants and principal stretches ................ 374
15.3 Isotropic hyperelasticity in principal invariants ............. 375
15.3.1 Formulation .................................... 375
15.3.2 l\'lodified neo-Hookean model ...................... 377
15.3.3 Modified Mooney-Rivlin model .................... 378
15.4 Isotropic hyperelasticity in principal stretches .............. 379
15.4.1 Formulation .................................... 379
15.4.2 Modified Ogden's model .......................... 380
15.5 Examples of homogeneous deformations ................... 381
15.5.1 Homogeneous simple shear ........................ 381
XVI Table of Contents

15.5.2 Uniform extension ............................... 383


15.5.3 Pure dilatation .................................. 384
15.6 Linearization .......................................... 385
15.6.1 Linearization of the deformat ion ................... 386
15.6.2 Linearization of constitutive equations ............. 387
15.6.3 Linearization of the equations of elasto-statics ....... 387
15.6.4 Variational formulations .......................... 388
15.6.5 Linearization of the weak formulation .............. 389
15.7 Mixed variational formulat ion ........................... 390
15.7.1 Formulation .................................... 390
15.7.2 Incompressibility constraint ....................... 392
15.8 Appendices ............................................ 393
15.8.1 The Piola identity ............................... 393
15.8.2 Linearization of a pressure B.C .................... 393
15.8.3 Differentiation of an isotropic function of a second-
order symmetric tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . 394
15.8.4 Elasticity tensors for principal stretch formulation ... 395

16. Finite-strain elasto-plasticity .............................. 397


16.1 First theory ........................................... 397
16.1.1 l\Iultiplicative decomposition of the deformation gra-
dient ........................................... 397
16.1.2 Hyperelastic-plastic constitutive equations .......... 399
16.1.3 Stress-strain relations ............................ 401
16.1.4 Flow rules ...................................... 402
16.1.5 Elastic predictor ................................. 404
16.1.6 Time discretization of the plastic flow rule .......... 404
16.1.7 Return mapping algorithm in principal stresses and
strains ......................................... 406
16.1.8 Algorithmic tangent moduli ....................... 407
16.1.9 Summary of the algorithm ........................ 408
16.1.10 Application: Quadratic logarithmic free energy and
h flow ..................... , ................... 410
16.2 Second theory ......................................... 413
16.2.1 Additive decomposition of the rate of deformation
and hypoelasticity ............................... 413
16.2.2 Computation of the strain increment ............... 415
16.2.3 Polar decomposition algorithm .................... 417
16.2.4 A time-integration algorithm for the rotation matrix . 419
16.2.5 Application: the Jaumann objective stress rate ...... 420
16.2.6 Summary ....................................... 421
16.2.7 Incremental objectivity ........................... 422
Table of Contents XVII

17. Cyclic plasticity .......................................... 423


17.1 One-dimensional model ................................. 423
17.2 Three-dimensional model. ............................... 425
17.3 Dissipation inequality ................................... 428
17.4 Plastic multiplier ...................................... 428
17.5 Tangent operator ...................................... 429
17.6 Hardening modulus .................................... 429
17.7 Retum mapping algorithm .............................. 430
17.8 Consistent tangent operator ............................. 433
17.9 Numerical simulat ion ................................... 435

18. Damage mechanics ....................................... 439


18.1 Damage variable ....................................... 439
18.2 Three-dimensional constitutive model ..................... 442
18.3 Dissipation inequality ................................... 445
18.4 Plastic multiplier ...................................... 445
18.5 Tangent operator ...................................... 445
18.6 Hardening modulus .................................... 446
18.7 Closed-form solutions for loadings with constant triaxiality .. 447
18.8 Return mapping algorithm .............................. 450
18.8.1 Corrections over the elastic predictor ............... 458
18.8.2 Summary of the algorithm ........................ 458
18.8.3 Non-damaged case ............................... 458
18.9 Consistent tangent operator ............................. 459
18.9.1 Non-damaged case ............................... 461
18.10 Numerical simulations .................................. 461
18.10.1 Ductile failure under uniaxial tension .............. 462
18.10.2 Ductile failure under simple shear ................. 462
18.10.3 A post-processor for crack initiation ............... 462
18.11 Further reading ........................................ 465

19. Strain localization ........................................ 469


19.1 l\Iotivation: a one-dimensional example ................... 469
19.2 Uniqueness and ellipticity ............................... 472
19.3 Strain localization ...................................... 474
19.3.1 Continuous bifurcation ........................... 475
19.3.2 Discontinuous bifurcation ......................... 475
19.3.3 Summary ....................................... 477
19.4 Analytical results for initially homogeneous plane problems .. 477
19.4.1 General strain-softening models ................... 477
19.4.2 A ductile damage model .......................... 481
19.5 Numerical results for a ductile damage model .............. 481
19.5.1 Biaxial loadings in plane stress .................... 481
19.5.2 Notched plate with a macro-defect ................. 486
19.6 Nonlocalor intemal-length models ....................... 495
XVIII Table of Contents

19.7 A two-scale homogenization procedure .................... 497


19.8 Numerical algorithms ................................... 502
19.9 Elasticity with damage- Model without threshold .......... 504
19.9.1 Local constitutive equations ...................... 504
19.9.2 Nonlocal macroscopic formulation ................. 505
19.9.3 Numerical simulations ............................ 505
19.10 Elasticity with damage- Model with threshold ............. 508
19.10.1 Local constitutive equations ...................... 508
19.10.2 Nonlocal macroscopic formulation ................. 508
19.10.3 Numerical simulations ............................ 509
19.11 Appendices ............................................ 511
19.11.1 Strain localization criterion in 2D .................. 511
19.11.2 Macroscopic free-energy potential .................. 513
19.11.3 Macroscopic dissipation potential .................. 516

20. Micro-mechanics of materials ............................. 517


20.1 lVIicro/macro approach .................................. 517
20.2 Homogenization schemes ................................ 519
20.2.1 Average strains and stresses ....................... 519
20.2.2 Voigt model .................................... 521
20.2.3 Reuss model .................................... 522
20.2.4 Self-Consistent model ............................ 522
20.2.5 Mori-Tanaka model .............................. 524
20.3 l\ficro/macro constitutive model for semi-crystalline polymers 525
20.3.1 Crystalline phase ................................ 526
20.3.2 Al11orphous phase ............................... 529
20.3.3 Interl11ediate phase .............................. 534
20.3.4 Single inclusion .................................. 536
20.3.5 Overall behavior ................................. 536
20.3.6 Numerical sil11ulations ............................ 536
20.4 Further reading ........................................ 541

A. Cylindrical coordinates ................................... 543

B. Cardan's formulae ........................................ 549

C. Matrices for the representation of second- and fourth-order


tensors ................................................... 551
C.1 Storage ............................................... 551
C.2 Change of coordinates .................................. 554

D. Zero-stress constraints .................................... 557


D.1 Small-strain h elasto-plasticity .......................... 557
D.2 General small-strain models ............................. 559
D.3 General finite-strain models ............................. 560
1. Basic mechanics

In this chapter, we recall some basic equations and definitions for stress,
strain, equilibrium, compatibility, strength criteria, Hooke's law, etc., which
are needed in subsequent chapters. We tried to separate results which are
independent of any constitutive model from those which are specific to lin-
ear elasticity. The reader is assumed to have some knowledge of continuum
mechanics, tensor analysis, linear algebra, etc.

1.1 On tensors
Throughout the book, boldface symbols denote tensors, the order of which
is indicated by the context. Einstein's summation convention over repeated
indices is used unless otherwise indicated:
3

aikbkj == L aikbkj
k=l

Dots and colons are used to indicate tensor products contracted over one and
two indices, respectively:
U· v = UiVi; (a· U)i = aijUj;
(a· b)ij = aikbkj; a : b = aikbki;

(C : a)ij = Cijklalk; (C: D)ijkl = CijmnDnmkl

Tensor products are designated by 0, e.g.,

(U0V)ij =UiVj; (a0b)ijkl =aijbkl

There are numerous references on tensors; for this book an introduction such
as the one given in (Ogden, 1984) or (Segel, 1977) is sufficient. What we need
to remember is that in order for a "mathematical object with indices" to
represent tensor components, it has to obey precise transformation rules. For
example, in a change of coordinates from one orthonormal basis (el,e2,e3)
to another (el, e2, e3), a vector (first-order tensor),
2 1. Basic mechanics

transforms according to:

where Q is a 3 x 3 proper-orthogonal matrix:

QikQjk = QkiQkj = Oij, det Q = 1,

with Oij denoting Kronecker's symbol:

Oij = 1 if i = j; Oij = O if i =1 j
Row (i) of matrix Q contains the components of vector ei in basis (el, e2, e3).
A second-order tensor,

transforms according to:

aij = Qikak/Qjl, aij = QkiaklQlj,

The second- and fourth-order identity tensors (1 and 1) are defined by:
1
l ij = Oij, I ijkl = "2 (OikOjl + OilOjk)
The inverse of a second-order tensor a is denoted by a-l so that:

a . a -1 = a -1 . a = 1, ·l.e. aik (a -1) kj = ( a -1) ilalj = Uij


~

The inverse of a fourth-order tensor C is denoted by C- 1 so that:


C: C- 1 = C- 1 : C = 1, i.e. Cijmn(C-l)nmkl = (C-l)ijpqCqpkl = I ijkl

In several chapters, two lemmas will be useful:


Lemma:
1
Let: C = 1 - Da (8) b and De = a : b,
then: C invertible {::::::} D =1 De,
-1 1
and: C = 1 + D _ De a (8) b, (1.1)

with a and b second-order tensors and C a fourth-order tensor. The proof


for this lemma is straightforward. There is a similar lemma for the case when
a and bare vectors and C a second-order tensor:
Lemma:
1
Let: C = 1- Da (8) b and De = a· b,
then: C invertible {::::::} D =1 De,
1
and: C- l = 1 + D _ De a (8) b (1.2)
1.2 Stress 3

1.2 Stress
Consider an elementary uniaxial tension test, Le. a bar (axis e z ) with uniform
cross section is subjected to forces ±Fez applied at its end sections and along
its axis (Fig. 1.1). In this test, a (uniaxial) stress, i.e. an internal force per
unit area, develops and is defined by
F
O"u = A' (1.3)

where A is the area value of the deformed cross section.

r-===1
ca
Fig. 1.1. Uniaxial tension test on a cylindrical specimen: initial (a) and deformed
(b) states.

This notion is generalized to multi-axial loadings by defining a Cauchy


stress tensor iT which is a second-order symmetric tensor (see Chap. 14 and
references therein for details). Since iT is symmetric, it can be stored as a
3 x 3 symmetric matrix in a given orthonormal basis:

(1.4)

It is essential to remember that in a change of coordinates from an orthonor-


mal basis (8(1» to another (8(2», the stress matrices transform according
to:
(1.5)

where the change of coordinates matrix [Q] is proper-orthogonal:


Q . QT = QT . Q = 1; det Q =1 (1.6)

For an application of Eqs. (1.5,6), see Appendix A.


Similarly to the uniaxial case, the stress components O"ij are interpreted
as internal forces per unit area. If we isolate an elementary parallelepiped
of dimensions dx x dy x dz in a stressed body, then on each face of the
parallelepiped acts an internal force (a vector) whose three components per
unit area in the Cartesian basis (e z , ey, e z ) are plotted in Fig. 1.2.
It is important to notice the convention used in Fig. 1.2. On the facet of
outside normal e z for instance, acts the following force per unit area:
4 1. Basic mechanics

(+)
a zz

1
(-)1

-- ,
a YXff
(-)
a yy (+ I (_
__ ( '" I1 a xz laxz
a(+) ,
xy

(-)
a yz

Fig. 1.2. Stresses acting on the facets of an elementary parallelepiped of dimensions


dx x dy x dz

(1.7)
Upper scripts (+) and (-) in Fig. 1.2 have the following meaning:

(±) _
a xx - a xx ± âx
oaxx (dx)
2'
idem for other stress components. Equation (1.7) is generalized as follows:
the force per unit area act ing on a facet with outside unit normal n is:
(1.8)
In component form, this can be written as:
(1.9)

Note that the so-called "stress vector" (o-T . n) is not in general collinear
with n. The normal stress an is defined by:
an = (o-T . n) . n = ajinjni
The sign convention adopted in this book is that an is positive if it is tensile,
and negative if it compresses the facet (in civil engineering, an opposite sign
convention is usually used).
We shall see later on that it is useful to define a deviatoric stress tensor
8 which is, by design, traceless (tr 8 = O):

1
8 == o- - -(tr 0-)1, i.e. (1.10)
3
1.3 Strain 5

1.3 Strain
Consider the uniaxial tension test of Fig. 1.1 again. If the initial and final
values of the bar length are 10 and 1, respectively, then a strain along the axis
of the bar is defined as:
(1 - 10 )
f:u = (1.11)
10
This notion is generalized to multi-axial loadings by defining a strain ten-
sor, which is defined in the injinitesimal case as the following second-order
symmetric tensor (see Chap. 14 and references therein for details):

(1.12)

where u(x) is the displacement field and " the gradient operator. In Carte-
sian coordinates, Eq. (1.12) can be written under the following form:

f:i. = ~(âUi + âUj) (1.13)


3 2 âXj âXi
Since e is symmetric, it can be stored as a 3 x 3 symmetric matrix in a given
orthonormal basis:

(1.14)

We now give a geometrical interpretation of the strain components in Carte-


sian coordinates (for details, see (Love, 1927)). Consider material line vectors
e x and e y which transform onto e~ and e~ after deformation. The diagonal
(or direct) strain components f: xx and f: yy have the same interpretation as
in the uniaxial stress case (Eq. (1.11)); they represent relative length varia-
tion in the (x) and (y) directions, respectively. As for the shear strain f: xy , it
represents angle variation:
1 I
f: xy = 2ex· e y
I

It is essential to remember that in a change of coordinates from an orthonor-


mal basis to another, the strain matrices obey the same transformation rules
as the stresses:

(1.15)

It can be shown (e.g., Chap. 14) that for infinitesimal strains, (tr e) measures
the variation of volume:
dV dV -dVo
- ~ l+tre i.e. tr e ~ dVo '
dVo '
6 1. Basic mechanics

where dVo and dV are elementary volumes before and after deformation,
respectively. It is useful to define the deviatoric part of E as follows:

e == E - ~(tr E)l, i.e. eij = tij - ~(tmm)Oij (1.16)

It is obvious that -by design- e is traceless (tr e = O), therefore e represents


the isochoric part of the strain tensor.

1.4 Principal invariants and eigenvalues of stress and


strain
Let 7] be a second-order symmetric tensor (rlij = 1/ji); typical examples are
the stress (u) and strain (E) tensors. There are three scalar quantities which
do not change from one orthonormal basis to another. They are called the
principal invariants of 7] and can be expressed as follows:

It = tr 7]= 1/11 +1/22 +'/]33;


12 ~ [Ii - tr (7]2)] = 11/22 1/23 1/11 1/13 1/11 1/12 1;
1/23 1/33 1+1 1/13 1/33 1+1 1/12 1/22
1/11 1/12 1/13
13 det7] = 1/12 1/22 1/23 (1.17)
1/13 1/23 1/33

When changing from an orthonormal basis 8(1) to another 8(2), a necessary


"sanity check" consists in verifying that It, 12 and 13 remain unchanged:

There exists an orthonormal basis (e(1), e(2), e(3») in which 7] is represented


by a diagonal matrix:

~ l'
1/3
(1.18)

where 1/i are the eigenvalues (or principal values) of 7] and e(i) the principal
directions. By definition, 1/i and e(i) are such that:

7]. e(i) = 1/ie(i); i = 1, 2,3; no sum over i (1.19)

Equation (1.19) shows that the eigenvalues are solution of the following equa-
tion:

det(7]- 0:1) = O, (1.20)


1.5 Mohr's stress circles 7

which can be rewritten as follows:


_0: 3 + Il 0: 2 - I 2 0: + h = O, (1.21)
where It, I 2 and I3 are the principal invariants of TI defined by (1.17). Equa-
tion (1.21) can be solved in closed form using Cardan's formulae (see Ap-
pendix B). Equations (1.17) show that the principal invariants and values
are related by:
(1.22)
Principal invariants and values play a central role in modeling isotropic ma-
terials, examples are given in Sect. 1.10 and chapters 12, 15 and 16. For
isotropic and elastic materials, it can be shown that the principal directions
of stress and strain coincide (see Sect. 15.8.3).
Finally, there are cases where TI has the following matrix representation:

'1712
'1722
O
O
1
, (1.23)
O '1733

Examples are stress or strain matrices for plane problems (Chaps. 7 and 8)
and bending moment matrices for plates (Chaps. 5 and 6). In those cases,
solving Eq. (1.20) becomes an easy task: simple algebra gives the principal
values as follows:

"',' = "n ; 'In ± [ ("n ; *')' + 'Ii,j"', * ="" (1.24)

1.5 Mohr's stress circles


A useful graphical representation of normal and shear stresses is due to the
German engineer Otto Mohr (1835-1918). Consider a facet of outward unit
normal n, the "stress vector" u T . n acting on the facet can be decomposed
into normal and shear components as follows:
(1.25)
Designating by an and as the scalar measures of normal and shear stresses:
(1.26)
it is shown -e.g., (Mase, 1970)- that points (an, as) are necessarily situated
within the shaded area of Fig. 1.3. The three Mohr's stress circles of Fig. 1.3
are determined by the principal stresses which are assumed to be distinct and
ordered according to: al < a2 < a3. The radii of the circles are: (a2 - ad/2,
(a3 - a2)/2 and (a3 - ad/2. The latter being the radius of the largest circle,
Fig. 1.3 shows that it is equal to the maximum shear stress. This result is
used for instance in Tresca's yield criterion (Sect. 1.10).
8 1. Basic mechanics

Fig. 1.3. Mohr's stress circles. Normal (O"n) and shear (0".) stresses are necessarily
within the shaded area, with 0"1 < 0"2 < 0"3 being the principal stresses.

1.6 Equilibrium
For a deformed body in static equilibrium, the following vector equation must
be satisfied in each material point:

div (TT +f = O, (1.27)

where li [N/m 3 J are forces per unit volume and (div) designates the di-
vergence operator. In Cartesian coordinates, Eq. (1.27) is equivalent to the
following system of three scalar equations:

oau oa21 oa31 f o,


--+--+--+
OXI OX2 OX3
1

oa12 00"22 00"32 f =


--+--+--+
OXI OX2 OX3
2 0,

oa13 oa23 oa33 f o


--+--+--+
OXI OX2 OX3
3 (1.28)

Using the convention on repeated (or dummy) indices, this system can be
rewritten in the following compact form:

--+
Oaji f i= O (1.29)
OXj

Equilibrium equations are found from the balance of linear momentum (see
Chap. 14). Alternatively, they can be derived directly by writing the equilib-
rium conditions for the elementary parallelepiped of Fig. 1.2, as in (Filonenko-
Borodich, 1958) for instance (or as in Secs. 5.3 for plates and 11.10 for shells).
Indeed, it can be easily shown that equilibrium of forces gives Eqs. (1.28)
while equilibrium of moments gives three scalar equations:

i.e., (T is symmetric (aji = aij).


1.7 Local formulation of static problems 9

1. 7 Local formulation of static problems

Consider a solid body which before deformat ion occupies an open set il of
]R3. The body is subjected to forces per unit volume f in il and to the
following boundary conditions (B.Cs.): forces per unit area F on a part FF
of its boundary and imposed displacements U on a part Fu of the boundary
(Fig. 1.4). We assume that Fu ::j:. 0, FF n Fu = 0, and FF U Fu = F, where
F designates the boundary of the domain. The position vector of a material
partide in a fixed global frame is designated by x.

Fig. 1.4. Loads and boundary conditions

The problem is to find the fields of displacements u(x), strains €(x) and
stresses lT(x) which satisfy the following equations:

u = U on Fu (displacement B. Cs. )
div lTT +f = O in il (equilibrium)
lTT . n = F on FF (force or "traction" B.Cs.)
1
€ = 2(V'u + V'T u) (infinitesimal strains)
"Constitutive equations" (1.30)

In Cartesian coordinates, Eqs. (1.30) become:

Ui = Ui on Fu
âaji f i = O'l l l u
--+ n
âXj
ajinj = Fi on FF
fij = ~
(âUi + âUj)
2 âXj âXi
"Constitutive equations" (1.31)
10 1. Basic mechanics

As a reminder of the notions of open and closed sets, consider the unit
interval, Le., {x E IRjO ~ x ~ 1}, then n =]0, 1[ is the open set, n = [0,1] is
the closed set, and r = {0,1}.
We have not written specific constitutive equations in (1.30e) to empha-
size the point that Eqs. (1.30a-d) are valid for any constitutive model. The
simplest material model is linear elasticity for which Eqs. (1.30e) become (see
Sect. 1.11):
O' = c : €, Le. (Tij = Cijkl€lk,
where the elasticity operator c is named after the British physicist Robert
Hooke (1635-1703).
For simplicity, we have Iimited ourselves to the B.Cs. presented above,
but other B.Cs. are possible (see references in Chap. 2), provided that the
problem is well posed. For mathematicai details, see (Parton and Perline,
1984a), but for the purpose of this book, it suffices to follow the following
"common sense" rules in order to understand this important notion:
- (1) The body must be in static equilibrium, Le. the resultant of alI forces
must vanish, and the resultant moment w.r.t. a fixed point must be zero.
- (2) We cannot impose in a given point of the boundary r and in the same
direction a force and a displacement at the same time. In other words,
we cannot impose Fi and Ui in the same point (we either impose Fi and
compute the corresponding displacement in the i-direction, or impose Ui
and compute the reaction force in the i-direction). What can be imposed
at a given point of rare: three displacement components or three force
components or a displacement component and two force components in the
other two orthogonal directions or a force component and two displacement
components in the other two orthogonal directions.
An important class of problems corresponds to the case where only force
B.Cs. are applied to the boundary r (Le., ru = 0). First, the body must in
static equilibrium (condition (1)) and then the stress solution will be unique
but the displacement field will be defined up to a rigid body displacement.
In several chapters -including this one- we make the small-perturbation
hypothesis (SPH), Le. we assume that strains, displacements and rotations
are "small". Therefore, infinitesimal strains (1.30d) are used as strain mea-
sures. Displacements are assumed to be small compared to a representative
dimension of the body (e.g., for a beam, its length, for a circular plate, its
diameter). An important consequence of SPH is that we write (and solve
for) equilibrium and boundary condition equations on the initial, undeformed
(thus known) configuration of a body. Exceptions to SPH are the search for
possible buckling modes in Chap. 10 and the study of finite-strain problems
in Chaps. 14 to 16.
Finally, problem (1.30) is formulated in a formal way, Le. one which is
not mathematically rigorous. For a mathematical presentation in the case of
linear elasticity, see (Parton and Perline, 1984a-b, 1983) for example.
1.8 Continuity equations 11

1.8 Continuity equations

Consider a body under imposed forces and displacements as described in Sect.


1. 7 and let n be a surface inside the body. Two regions are thus defined,
one on each side of n and are designated by (1) and (2). Their outward
unit normals on n are n(1) and n(2) = _n(l), respectivelYj see Fig. 1.5. The
surface forces acting an side (1) of n are, by definition, uel) T ·n(1) and those
acting on side (2) of n
are: U(2) T . n(2). Equilibrium of forces on n
requires
that:
uel) T . n(l) + u(2)T . n(2) = O on n

Since n(2) = -n(1), this becomes:

u(l)T. n = u(2)T. n on n, (1.32)

with n = n(l) ar n(2). Equation (1.32) requires the continuity of the "stress
vector" u T ·n. Note that continuum mechanics does not require the continuity
of the entire stress tensor u. The field of "stress vectors" can be thought of
as that of cohesive forces holding the body together.
There is a second continuity requirement: the displacement vector must
be continuous across rI :

(1.33)

IT this condition is violated, then the material breaks along rl (e.g., a crack
may appear).
IT the surface n is an interface between two different materials, then
continuity conditions (1.32-33) must hold if the two materials are perfectly
"glued" together at n,
otherwise the two materials may separate (e.g., de-
lamination in laminated composites).
There are many interesting cases where two solids are in contact but
tangential (sliding) ar normal separat ion can occur in some areas. In those
cases, contact conditions other than (1.32-33) must be writtenj see (Johnson,
1987), (Doghri et al., 1998) and references therein.

Fig. 1.5. Continuity conditions across a surface n


12 1. Basic mechanics

1.9 Compatibility equations


Some problems are such that one can guess the form fij(Z) = fji(Z) of the
strain field. The stresses can be computed from the constitutive equations and
equilibrium equations (1.30b) and force B.Cs. (1.3Oc) can then be checked.
However, there remains a question: is it possible to find a displacement field
Ui(Z) such that the strain-displacement relations (1.30d) are satisfied? The
answer to this question is yes if compatibility equations are satisfied (the six
components of a symmetric strain tensor cannot be specified arbitrarily since
a displacement vector has only three components). In Cartesian coordinates,
these equations read (e.g., (Love, 1927)):

8 i'
82
Xi Xk
.
= O; i, j, k, 1 = 1, 2, 3 (1.34)

It can be shown that these equations are reduced to a linear combination of


the following six equations:

8 2 1011 + 8 21023 8 2 1012 8 2 1013 =O


8X28X3 8x18xl 8x18x3 8x18x2
82t22 + 8 2t 13 _ 8 2t21 _ 8 2t23 =O
8x18x3 8x28x2 8X28X3 8x28xl
82t33 + 8 2t12 8 2t31 82t32 =O
8x18x2 8x38x3 8x38x2 8x38xl
2 â 2f12 _ â 2f11 _ â2f22 = O
âx1âx2 âx2âx2 âx1âxl
2 8 2t 13 _ 8 2t11 _ 8 2t 33 = O
âx18x3 8x3âx3 8x18xl
2 â2f23 _ â 2f22 _ â2f33 = O (1.35)
âx2âx3 âx3âx3 âx2âx2
These conditions are necessary and sufficient for a simply connected domain,
otherwise additional conditions must be written in order to ensure that a
single-valued displacement field is found (examples are given in Chaps. 4 and
8). We recall the definition of a simply connected domain: it is such that any
closed curve contained in the domain can be continuously shrunk to a point.
As a counter example, a tube is not simply connected, it is multi-connected.

1.10 Strength criteria


Figure 1.6 shows an idealized stress-strain response of a metallic specimen
under uniaxial tension. It is seen that if the stress exceeds a certain level
u(A) = Uy, then u"'''' is no longer proportional to t",,,,. Also, if the specimen
1.10 Strength criteria 13

is unloaded at any point along (AB), then to a zero stress (O"(C) = O) cor-
responds a non-zero strain (€(C) "1 O), i.e. an irreversible or permanent or
plastic deformation takes place. Elasto-plasticity theory is studied in chapters
12 and 16.

O"(B) ~----=_'9'--

o
€(B)
Fig. 1.6. Uniaxial stress-strain response of a metallic specimen

For a uniaxial tension or compression test in the (x) direction, it is seen


that the response is linear elastic as long as the following yield criterion is
satisfied:

(1.36)

where O"y is a material parameter known as the initial yield stress. For multi-
axialloadings, criterion (1.36) was generalized by von Mises as follows:

(1.37)

where J2 (u) is the von Mises equivalent stress defined by:

J2 (u) = (_S:S)I/2,
3
2 ie
.. J(u)
2 = (s
3
2" ij Sji )1/2 , (1.38)

where sis the deviatoric part ofthe Cauchy stress. Since J2 (u) is the second
invariant of the deviatoric stress, the yield criterion is both isotropic and
pressure insensitive, this last property being generally well verified for metals.
It will also be shown in Sect. 1.12 that the square of J2 (u) is -up to a factor-
equal to the elastic distortion energy. By developing the expression of J2 (u):

h(u) = {~[(0"11 - 0"22)2 + (0"22 - 0"33)2 + (0"33 - 0"11)2]

+3 (0"~2 + 0"~3 + 0"~1)}1/2, (1.39)


14 1. Basic mechanics

we can double-check that it is independent of hydrostatic pressure. It is easy


to check that for uniaxial tension in the (1) direction, we have J2 (u) = lanl,
and this explains the presence of the factor 3/2 in the definition of the van
Mises stress. For simple shear in the (1, 2) plane, we have h(u) = laI2h/3.
The van Mises stress can be "visualized" as being equal -up ta factar-
ta the so-called octahedral shear stress. Assume that the principal stresses
ai and directions e(i), i = 1, 2, 3, are knawn. The principal stress directians
form an orthonormal basis B == (e(l), e(2), etaj). The so-called octahedral
plane makes equal angles with thase directians (Fig. 1.7). The auter unit
normal n ta this plane has the follawing camponents in B:

(1.40)

The stress matrix in B is diaganal and has the follawing expression:

[a] = [~O ~2O aa~ 1 (1.41)

The surface force acting on the octahedral plane has the following components
in B:

(1.42)

It can be decomposed inta normal and tangential (or shear) components as


follows:

(1.43)

The normal stress component is given by:

an = (uT 3
I I
. n) . n = -(al + a2 + aa) = -tr u
3
(1.44)

From Eqs. (1.42-44), the tangential surface force is found to have the following
components in B:

(1.45)

Recalling the definition of the deviatoric stress 8 -Eq. (1.10)- it is found that:
1.10 Strength criteria 15

r
1 [SI
= Vă :: 1 (1.46)

The norm of r is the so-called octahedral shear stress:


1 1
IIrli = J3(s~ + s~ + S~)1/2 = J3(8 : 8)1/2 (1.47)

It is seen that the von Mises equivalent stress J 2 (u) is related to IIrli by:

(1.48)

Fig. 1.7. Orthonormal hasis of principal stress directions (e(l), e(2}, e(3}) and oc-
tahedral plane with outer normal n

An isotropic yield criterion should be either a function of the principal


stress invariants (e.g., von Mises criterion) or a symmetric function of the
principal stresses (e.g., 7resca criterion). The latter criterion is defined as
follows:
(maximum shear stress) ~ a;
IT we know the principal stresses (al, a2, a3), the maximum shear stress is
simply equal to the radius of the largest Mohr's circle (Sect. 1.5); thus the
Tresca yield criterion becomes:
(1.49)
As an application of the von Mises and Tresca yield criteria, consider a
bi-axial stress state where the stress matrix in a Cartesian basis is given by:

al
[ (1.50)
[a]= ~
16 1. Basic mechanics

Using Eq. (1.39), it is found that the von Mises yield criterion takes the
following form:

(1.51)

Introducing the following change of variables:


0"1 0"2 1 0"1 0"2 1
El == (O"Y + O"y) "j'i' E2 == (- O"y + O"y) "j'i' (1.52)

it is seen that inequality (1.51) defines an elliptical sur/ace:

(1.53)

The ellipse is plotted in Fig. 1.8, which can be interpreted easily: if the stress
point (O"d O"y, 0"2/ O"y) is inside the elliptical surface, then the state is linear
elastic. II the stress point is outside the ellipse, then irreversible (plastic)
deformation occurs.
The Tresca criterion gives a hexagonal surface which is also plotted in Fig.
1.8; its interpretation is identical to that of the von Mises elliptical surface.
The Tresca surface is easy to obtain. The stress matrix (1.50) is diagonal and
the principal stresses are simply 0"1, 0"2 and 0"3 = O. Mohr circles for three
different stress regions are plotted in Fig. 1.8.
Figure 1.8 shows that the von Mises and Tresca criteria mostly give the
same predictions but that in general the latter criterion is more conservative.
In practice however, this is not important, because engineers generally use a
safety factor Ks > 1, Le. the yield (or "strength") criteria (1.37, 49) are used
with a right-hand-side equal to O"y / Ks.
The von Mises and Tresca yield criteria are applicable for ductile isotropic
materials. For brittle materials such as ceramics or glass, it is observed that
they break if the tensile stress exceeds a cert ain material parameter O" J or the
tensile strain exceeds a parameter EJ. Such materials remain linear elastic as
long as:

(1.54)

otherwise they break.. In the strength criteria (1.54), 0"1 designates the largest
principal tensile stress and EI the largest principal tensile strain.

1.11 Linear elasticity

The results which were presented so far are independent of any specific ma-
terial behavior. We now restrict our attention to the important case of linear
1.11 Linear elasticity 17

von Mises
Fig. 1.8. For a bi-axial stress state, the von Mises and Tresca yield criteria give
elliptical and hexagonal surfaces, respectively

elasticity (note that nonlinear elastic models do exist, see Chap. 15). By "elas-
tic" we mean "reversible", Le. if the external solicitat ion which is applied to
the body is removed, the latter retrieves its initial shape. For instance, in the
uniaxial tension example of Fig. 1.1, if the tensile loading is brought to zero,
the bar retrieves its initial geometry (length lo and cross section area Ao).
For the same test, a "linear elastic" material is such that the (uniaxial) stress
(Txx is proportional to the axial strain Exx:

(1.55)

where E > O is a material modulus named after the British scientist Thomas
Young (1773-1829). Typical values of E are: 200 GPa (Le., 200 x 109 N/m 2 )
for steel, 70 GPa for aluminum and 0.1 GPa for rubber. For multi-axial
loadings, Eq. (1.55) is generalized as follows:

(T = e : €, Le. (Tij = CijklElk, (1.56)

where e is called Hooke's opemtor. It is a fourth-order tensor which is


positive-definite (see Chap. 2) and has the following symmetries (e.g., (Du-
vaut, 1990)):

Cijkl = Cjikl = Cijlk = Cklij (1.57)

On a first glance, e has 34 = 81 components, but using the symmetries of e,


€ and (T, it can be shown that e has only 21 independent components. An
interesting consequence for computer implementat ion is that e can be stored
18 1. Basic mechanics

as a symmetric 6 x 6 matrix while € and lT can be stored as 6 x 1 arrays; see


Appendix C.
An important particular case is that of isotropic materials, for which
mechanical properties are independent of the loading direction. For those
materials, it can be shown -e.g., (Duvaut, 1990)- that Cijkl has the following
expression:

(1.58)

where the material parameters A and J.I. are named after the French engineer
Gabriel Lame (1795-1870) and are the only material properties needed to de-
scribe isotropic linear elastic behavior (instead of 21 in the general anisotropic
case). Using Eqs. (1.57,58), the stress-strain reIat ion (1.56b) becomes:

(1.59)

where (tr) is the trace operator. We now wish to invert Eqs. (1.59), Le. express
strains in terms of stresses. The "trick" is to relate the trace of lT to that of
€. Computing the trace on each side of (1.59b), it is found that:

tr lT = (3A + 2J.1.)tr € (1.60)

Substitution of this result into Eq. (1.59b) gives:

lT A . Uij A r
€ = -
2J.1.
- 2 (A 2) (tr
J.I. 3 + J.I.
lT)l, l.e. f.ij = -2J.1.- - 2J.1.(3A + 2J.1.) UmmUij
(1.61)

As a first application, consider the same uniaxial tension test as before; the
stress matrix in the Cartesian basis is:

[uJ = [T ~ ~ 1 (1.62)

The corresponding strain components are given by Eqs. (1.61b) as:

A+J.I. A
f. xx = J.I.(3A + 2J.1.) U xx ; f. yy = f.zz = - 2J.1.(3A + 2J.1.) U xx ;
f. xy = f. yz = f.xz = O (1.63)

Usually, the material parameters which are measured experimentally are not
A and J.I. but E and v which can be defined from a uniaxial tension test as
follows:

E = u xx ; v = _~ = _ f.zz , (1.64)
f. xx f. xx f. xx
1.11 Linear elasticity 19

Le. Young's modulus E measures the stiffness in the tension (rudal) direction
and v measures the lateral contract ion. This ratio v is named after the French
engineer S.D. Poisson (1781-1840). Comparing Eqs. (1.63-64), the following
identities are found:

(1.65)

The Lame coefficients are then easily deduced:

A_ Ev . _ E
(1.66)
- (1- 2v)(1 + v)' f-L - 2(1 + v)
Using identities (1.65), it is easy to check that the strain-stress relations (1.61)
can be rewritten under the following simpler form:
l+v v . l+v v
€ = ~u - E(tr u)1, l.e. Eij = ~Uij - EUmmbij (1.67)

As a second application, consider a state of pure shear in the (x, y) plane


(Fig. 1.9). The stress matrix in the Cartesian basis is given by:

(1.68)

The strains are given by Eqs. (1.67) as:

l+v
Exy = ~Uxy; Exx = Eyy = Ezz = Eyz = Exz = O

Using Eq. (1.66b), we have:


U xy
f-L=-
2Exy

This shows that the so-called shear modulus f-L has a similar role to that of E
in a uniaxial tension test, it is the constant slope of the straight line: shear
stress (u xy ) versus engineering shear strain (2E xy ); Fig. 1.9.
As a third application, consider a state of hydrostatic pressure. The stress
matrix in the Cartesian basis is given by:

[uj = [ - :
O
~pO ~ -p
l' (1.69)

where p > O is the applied pressure. The strains are given by Eqs. (1.67) as:
(1 - 2v)
(xx = (yy = Ezz = E (-p); (xy = Eyz = Exz = O
20 1. Basic mechanics

Fig. 1.9. A state of pure shear in the (x, y) plane

A so-called bulk modulus r;, is defined by:


E
3r;,::::: --2- =3A+2J.t, (1.70)
1- v
using Eqs. (1.65). Thus we have €xx = -p/3r;,. Actually, using Eq. (1.60), we
have the very general result:
tr (T
tr €=-- (1.71)
3r;,
Equations (1.70, 71) show that:
1
(v-t'2) =? (r;,-too) =? (tr€-tO) (1.72)

Since for infinitesimal strains, (tr €) measures the variat ion of volume, Eq.
(1.72) shows that for an incompressible material: v = 1/2 or r;, -t 00 (this
means in practice: r;,/J.t ~ 103 ). The larger the value of r;, (w.r.t. J.t), the less
compressible the material. That's why r;, is also called "compressibility mod-
ulus" .

Exercise: show that the stress-strain relations can be written under the
following format using r;, and G ::::: J.t:

(T = 2G€ + (r;, - ~G)(tr €)1, Le. aij = 2G€ij + (r;, - ~G)€mm8ij (1.73)

In summary, for a linear isotropic material, we need two independent material


parameters which are the Lame coefficients A and J.t or Young's modulus E
and Poisson's ratio v or the bulk and shear moduli r;, and G::::: J.t. Those three
pairs are related together by Eqs. (1.65, 66, 70). Parameters E, A, J.t = G and
r;, have the dimension of a stiffness [N/m 2 ], while vis dimensionless. One of
several equivalent stress-strain relations (1.59,67,73) canbe used, depending
on the particular problem at hand.

Exercise: find the strain versus stress relations (1.67) by direct application
of lemma (1.1).
1.12 Strain energy 21

1.12 Strain energy

Consider a linear spring under a tensile force F (Fig. 1.10). The tension
T = F in the spring is related to the displacement U by T = kU, where
k [N/mJ is the stiffness ofthe spring. This relation can be written as follows:

(1.74)

where W(U) is the energy of deformation of the spring.

Fig. 1.10. Linear spring: initial (a) and deformed (b) states.

In linear elasticity, the analogous relation to T = kU is: (T = C : €, which


can be written as:
d 1
(T--(-€·c·€)
- d€ 2 . . , (1.75)
~
W(€)

where, by analogy with Eq. (1.74), W(€) is the strain energy per unit volume
(a better justification will be given in Chap. 2). The strain energy can be
rewritten as follows:
1 1
W(€) = "2€ : (T = "2{ij(Jji (1.76)

For a uniaxial tension test in the (x) direction, W(€) has the simple expres-
sion:

(1.77)

which has a simple interpretation: in uniaxial tension, W(€) is simply the area
under the stress-strain line (the area of the shaded triangle in Fig. 1.11).
22 1. Basic mechanics

Fig. 1.11. In uniaxial tension, the strain energy per unit volume (W) is simply
equal to the area of the shaded triangle

For generalloadings, using the isotropic linear elastic relations (1.59), the
following expression is found:

(1.78)

It is useful to rewrite this expression a couple of times. Using the deviatoric


strain e, Eq. (1.16), simple algebra gives:

W(e) = (~ + ~)(tmm)2 + p.eijeji,


and this becomes -using definit ion (1.70) of the bulk modulus ""-

W(e) = ~(tmm)2 + ~ (1.79)


'--" wd;'(e)
wvol(f",,,,)

It appears that W(e) is the sum of two terms: Wtlol(t mm ) which represents
the part of the energy due to the change of volume, and WdiB(e) which
corresponds to the change of shape. The latter is called the distortion energy.
Note that the interpretations of the bulk and shear moduli given in Sect. 1.11
are consistent with Eq. (1.79): "" is attached to the volume variat ion energy,
while p. appears in the distortion part of the energy.
We now rewrite Eq. (1.79) one more time using the stress tensor. Re-
calling the definit ion of the deviatoric stress tensor s, Eq. (1.10), and using
Eqs. (1.59, 60), it is found that the deviatoric stress and strain tensors are
proportional:
s = 2p.e (1.80)
Finally, substituting into Eq. (1.79) and recalling (1.70),3"" = 3A + 2p., it is
found that the strain energy is the following stress function:

W(e) = W(u) = (u mm )2 + SijSji (1.81)


18"" 4p.
--.....-..- '-v-"
Wvol(u",,,,) Wd;.(S)
1.13 Navier equations 23

Recalling Eq. (1.38), it is seen that the (elastic) distortion energy W di8 (s) is
proportional to the square of the von Mises equivalent stress:

(1.82)

Consequently, the yield criterion (1.37) can be written in terms ofthe (elastic)
distortion energy as follows:
2
WdiB(s) < O'y (1.83)
- 6ţt'

where O'y is the initial yield stress.

1.13 Navier equations


Some problems are such that the form of the displacement field u(x) can
be guessed. The strains are then computed from the strain-displacement re-
lations (1.30d) and the stresses from the constitutive equations of isotropic
linear elasticity (1.59). Equilibrium equations (1.30b) and force B.Cs. (1.30c)
can then be checked. It is desirable to "automate" this process once and for
alI, Le. to express the equilibrium equations in terms of the displacements.
In Cartesian coordinates, using Eqs. (1.31d, 59), Eqs. (1.31b) become:

-
8 [>'~ij -
8um
-
8Ui
+ ţt(- + -8uj )]
+ fi = O
8xj 8xm 8xj 8Xi
We now assume that the Lame coefficients are uniform, i.e. independent of
the position vector x. The previous equations then become:
\ 8 (8u m ) +ţt-
A- --
8 (8ui -) + j i= O
- +8uj
8Xi 8xm 8xj 8xj 8Xi
After renaming dummy indices, the equations can be rewritten as:
8 8u m 82Ui
(>. + ţt)-8 (-8 ) + ţt 8 Xj 8 Xj
Xi X m
+ fi = O (1.84)

The three scalar equations thus obtained are named after the French engineer
Navier (1785-1836). They can be rewritten under the following tensor form
which can be used in other coordinate systems (e.g., cylindrical or spherical):
(>. + ţt)V(div u) + ţtLlu + f = O, (1.85)
where Ll designates the Laplacian operator. Actually, we shall use this pro-
cedure quite often: when we need to differentiate w.r.t. position, we first con-
sider Cartesian coordinates, where the computations are the easiest, then we
try to find a tensor or "intrinsic" form which can be used for other coordinate
systems.
24 1. Basic mechanics

1.14 Beltrami-Mitchell compatibility equations

Some problems are such that one can guess the form Uij(m) = Uji(m) of the
stress solution (several examples are given for plane problems in chapters 7
and 8). This (trial) stress field must satisfy equilibrium equations (1.30b) and
force B.Cs. (1.30c). Strains can be computed from the constitutive equations.
Those strains must satisfy the compatibility equations of Sect. 1.9 in order
to ensure that displacements verifying (1.31d) can be found.
In a stress-based approach, it is desirable to express the compatibility
equations in terms of stresses, so one can check the suitability of a guess
from the beginning. Assuming isotropic linear elasticity, we can substitute
Eqs. (1.61) into (1.35), but we arrive to a simpler representation if we use
the method of Beltrami and Mitchell which proceeds as follows. Define:

This is known (up to a sign) as the hydrostatic stress. Now assume that the
material properties E and v are uniform in space. Substituting Eqs. (1.61)
into (1.34) leads to:

_ 3v (O" EPuH + Okl 8 2uH _ O'k 8 2uH _ 0'1 8 2uH )


E '38xk8Xl 8Xi8Xj '8Xj8Xl 3 8x i 8 x k

1+v (
+__ EPu"'3 + 82Ukl 8 2u'k
___ '_ _ 8 2U'I)
3 = O (1.86)
E 8xk8Xl 8xi 8Xj 8Xj8Xl 8Xi8xk

Setting k = l, summing over the repeated index k = l, and rearranging terms,


we obtain:

..dUij +3 8 2uH _ ~(8Uik) _ ~(8Ujk)


8Xi8xj 8xj 8Xk 8Xi 8Xk

-~ (Oii..dUH + EPuH ) = O, (1.87)


1+v 8Xi8xj

where ..d designates the Laplacian operator. If tr satisfies equilibrium equa-


tions (1.31b), then Eqs. (1.87) take a simpler form:

(1.88)

These are six independent compatibility equations. Further simplification can


be obtained by setting i = k and j = lin Eqs. (1.84). After summation over
repeated indices and rearrangement of terms, we obtain:

~(8Uij) _ 3 (1 - v) ..dUH =O (1.89)


8Xi 8xj 1+v
l.15 Saint-Venant's principle, Uniqueness, Superposition, Special theories 25

Assuming that u(x) satisfies equilibrium equations (1.31b), Eq. (1.89) be-
comes:
l+v
3(1 _ v) div i, (1.90)

where (div) designates the divergence operator. Substitution into Eq. (1.88)
gives:

(1.91)

An important case is when the external forces per unit volume i are uniform
in space (e.g., forces due to gravity, assuming a uniform density). In this case,
Eqs. (1.91) become much simpler:

3 8 2uH
LlUij + -1-- ~ = O, with LlUH = O (1.92)
+ v UXiUXj
Taking the Laplacian again on (1.92a), we obtain:

(1.93)

Le. the stress field must be bi-harmonic. Note that the same result applies
for the strains: LlLltij = O.

1.15 Saint-Venant's principle, Uniqueness,


Superposition, Special theories

There exists a principle due to the French engineer Barre de Saint-Venant


(1797-1886) which is very useful and many interesting problems could not
be solved without it. This principle states that if a system of external loads
applied on a part rA of the surface of a body is replaced by another, statically
equivalent system, then at a sufficient distance from rA, the stresses due
to the two systems will be practically the same. In some simple cases, the
principle can even be demonstrated, e.g. (Parton and Perline, 1984a). The
principle is better understood by considering its applications; several of them
are given throughout the book, e.g. chapters 3, 4, 7 and 8.
For linear elasticity, a fundamental result which is proven in Chap. 2
is that if the basic problem of Sect. 1.7 is well posed, then its solution is
unique. This is a powerful tool for solving problems in linear elasticity: if by
experience or intuition, a solution is found which satisfies aU the equations
(Le., (1.30) or (1.31)), then it is the solution to the problem.
Another powerful result in linear elasticity is the so-called superposition
principle which can be stated as follows.
26 1. Basic mechanics

Consider two different loading systems (i), i = 1,2,


I(i} in il, F(i} on r F and U(i} on rUj solution: u(i}(x), O'(i)(x) and
€(i}(x). Consider now a third loading system (O):
r
[al(I} + {31(2}] in il, [aF(I} + (3F(2)] on F and [aUei} + {3U(2}] on u , r
where a and (3 are given scalars. The "principle" states that the solution of
problem (O) is:
[au(1} (x) + {3U(2} (x)], [aO'(1} (x) + {30'(2} (x)] and [a€(I}(x) + (3€(2) (x)].
Numerous applications of this principle are given throughout the book,
e.g. chapters 3 to 8. One has to remember however that this "principle" stems
from the fact that the basic problem formulated in Sect. 1.7 is linear when
the constitutive model is linear elastic; when the problem becomes nonlinear
because of material or geometric nonlinearities, the superposition "principle"
does not apply.
Finally, in many interesting engineering applications, we do not solve the
basic problem of Sect. 1.7, but modijied and simplijied versions of it. Such
cases arise for so-called structures (beams, plates and shells). Beams are solids
for which one dimension (the length) is much longer than the other two di-
mensions. Plates and shells are solids with one dimension (the thickness)
much smaller than the other two dimensions; when the mid-thickness surface
is planar, the structure is called a plate, otherwise it is a shell. Based on
kinematic assumptions, simplified or special theories are developed for struc-
tures and new variables are defined in order to describe their deformed state.
For beams, internal loads, Le. stresses integrated through the cross section
are defined instead of stresses. For plates and shells, internal loads per unit
length, Le. stresses integrated over the thickness, are used. For bending of
beams, plates and shells, the curvature of the deformed middle fiber or sur-
face is used instead of the usual strain tensor. Special theories offer dramatic
simplifications of the original three-dimensional (3D) problem by reducing it
to a one-dimensional (lD) problem along the middle fiber for beams, and to
a two-dimensional (2D) problem on the middle surface for plates and shells.
For details, see chapters 3, 5, 6 and 11.

1.16 Solved problem: composite cylinder under axial


load
A composite solid is made up of two concentric circular cylinders of length
l: a fiber of radius Rj, Young's modulus Ej and Poisson's ratio Vj, and a
matrix of radii Rj and Rm and elastic properties Em and V m (Fig. 1.12).
Axial and uniform displacements ±(U/2)e z are applied to faces z = ±l/2.
Lateral surface r = Rm is stress-free and perfect adherence at the interface
r = Rj is assumed. We introduce the following notation for convenience:
R U
C == (~)2, e == l' p== -arr(Rj), (1.94)
1.16 Solved problem: composite cylinder under axialload 27

where C represents the fiber volume fraction, e the axial (uniform) strain and
p the (continuous) interface pressure.

matrix : Em, lIm


Fig. 1.12. Composite cylinder under axial tension/compression.

Working with cylindrical coordinates (Appendix A), one can prove that
the stress solution is given as follows:

r
• Fiber: Urr = UOO = -p, Uzz = constant == uzI,
• Matrix: urr,OO = =t= [ ( ~m =t= 1]1 ~ C P'
U zz = constant == U zm , (1.95)
and that the shear stresses vanish everywhere:
UrO = UO z = Urz = O (1.96)
Indeed, the stress expressions satisfy the stress B.Cs., continuity of the stress
vector at the interface, the only non-trivial equilibrium equation:
- O
- - + --'-'----'-'-
dUrr
dr
Urr - UOO
r -,
and the compatibility equations. In order to find the stresses, one can look
for the displacement field under the form:
u(r, 0, z) = u(r)e r + eze z
Actually, we shall see in Chap. 8 that the stresses in the matrix are those of a
Lame's hollow cylinder under internal pressure P and zero external pressure.
Isotropic linear elasticity gives the strain field as follows:
. (1 - vI - 2vJ)

[c;r
• Flber: (rr = (00 = -vIe - EI p,

=t= 1 ± 2vm]

C
x -P' (1.97)
1- C
28 1. Basic mechanics

the other strains being uniform:

€zz = e, €r() = €()z = €rz = O (1.98)

Condition €zz = e ("generalized plane strain") allows to compute the axial


stresses azf and a zm :

(1.99)

The radial displacement field u(r) is simply found from the relations:

du
u = r€()(), dr = €rr

The interface pressure is computed from displacement continuity at the in-


terface (r = Rf). The following expression is found:

v = (1 + vf )(1 - 2vf) (1 + vm) (~ 1_ 2 ) ~


- Ef + Em C + Vm 1- C
(1.100)

Note that if Vm = vf then p = O, azf = eEf' a zm = eEm and arr = a()() = O


everywhere: under axial loading, the composite behaves as a system of two
parallel bars, because there is no mismatch in the lateral contraction.
The equivalent axial stifJness E is defined as follows:

E =
- < a zz > , (1.101)
e
where < a zz > is the stress average:

_1
< a zz >= [2 22]
R2 7rRfazf + 7r(Rm - Rf)azm = Cazf + (1- C)azm
7r m
(1.102)

Substituting the expression of p into those of the axial stresses, the axial
stiffness is found to be:

E = pEf + (1 - C)Em,. +2C (vm ; Vf )2 (1.103)


..
The term under brace is the one given by a simple mixture rule, the additional
term is due to the multi-axial nature of the stress state.
2. Variationa1 formulations, work and energy
theorems

In the previous chapter, we presented the local or "strong" formulation, which


allows to solve various (but rather simple) problems in statics. Often, as we
shall see in subsequent chapters, the formulation is used under approximate
forms (e.g., theories of beams, plates and shells). However, the most powedul
numerical methods (e.g., finite elements) which are used to find approximate
solutions to problems that cannot be solved in closed form (Le., most in-
dustrial problems) are not based on the local formulation but on the global
formulations which will be developed in Secs. 2.2 to 2.6; this constitutes one
of the major interests of this chapter.

2.1 Local formulation of static problems


For simplicity, we consider an orthonormal Cartesian coordinate system
(O, Xl, X2, X3), but the results which we shall find in this chapter can be
easily rewritten for other coordinate systems.
Consider a solid body which before deformat ion occupies an open set il
of]R.3. The body is subjected to forces per unit volume fi [N/m 3 ] in il and
to the following boundary conditions (B.Cs.): forces per unit area Fi [N/m 2 ]
on a part FF of its boundary and imposed displacements Ui [m] on a part
Fu of the boundary (Fig. 2.1). We assume that Fu =1 0, FF n Fu = 0, and
FF U Fu = F, where F designates the boundary of the domain.
We now recall from Sec. 1.7 the local formulat ion (Po) of the problem in
statics (or quasi-statics).

le Formulation (Po):
Find the fields of displacements Ui(X), strains €ij(X) and stresses O"ij(X) which
satisfy the following equations:
30 2. Variational formulations, work and energy theorems

Xa

Fig. 2.1. Loads and boundary conditions

Ui = Ui on r u (displacement B.C.)

--+
aUij f i = o·f i un (equilibrium)
aXj

Uijnj = Fi on rF (force or "traction" B.C.)

fij =~ (~:; + ~i) == U(i,j) (infinitesimal strains)

"Constitutive equations" (2.1)

For now, we do not consider any particular constitutive model, because we


shall see that the global formulations of Sec. 2.2 do not depend on any ma-
terial model.
For simplicity, we have limited ourselves to the B.Cs. presented above,
but of course, other B.Cs. are possible in practice. For slightly more general
B.Cs., see (Hughes, 1987) and for a much more general setting, see (Duvaut,
1990). What we should keep in mind is that the problem under study must
be well posed, as explained in Sec. 1. 7.

2.2 Virtual work theorem (VWT)


A multiplication of the equilibrium equations with (a smooth vector field) Wi
followed by an integrat ion over il give:

r
ln
(aUi j
aXj
+ /i)wi dil = O (2.2)

An integration by parts leads to:


2.2 Virtual work theorem (VWT) 31

We have:

~ (aij + aji) Wi,j (aij being symmetric)


1
'2aij (Wi,j + Wj,i) (permutation of "dummy" indices)
aijw(i,j) (W(i,j): infinitesimal strain associated with Wi )

We also have aijnj = Fi on r F ; Eq. (2.3) can then be rewritten as:

We now introduce the following two sets:

Y {v = {vd I v "sufficiently smooth" and Vi = Ui on ru}


Y* {v* = {vn I v* "sufficiently smooth" and vt = O on ru} (2.5)

In order to avoid discussing mathematical technicalities, we use the vague


condition "sufficiently smooth"; roughly speaking it means that the displace-
ment fields must be such that aH mathematical operations are legitimate
(e.g., Vi, Vi,j, V;' vt,j E L 2 (n), the space of square-integrable functions).
If v E Y, then v is said to be "kinematically admissible" (K.A.) For
instance, the displacement solution u of problem (Po) is K.A. (u E Y). An
element v* of Y* can be viewed as the difference of two K.A. fields. For
instance, we can take v* = v - u, where u is the displacement solution
(u E Y) and v is any K.A. field (v E Y).
If we take w = v (with v E Y) in (2.4), we can reformulate (Po) in the
following way:

"Formulation (Pt}:

Find the displacement field u which satisfies:

uEY

lnr aijv(i,j) dn =
lnr fiVi dn + r
lrF FiVi dr + r
lru aijnjUi dr, "Iv E Y

"Constitutive equations" (2.6)

If we take w = v* (with v* E Y*) in Eq. (2.4), we can reformulate (Po) in


the following way:

" Formulation (P2 ):

Find the displacement field u which satisfies:


32 2. Variational formulations, work and energy theorems

+!
uEY
rO"ijVei,j}
la
dJl = r /ivi
la
dJl
~
Fivi dF, Vv* E Y*

" Constitutive equations" (2.7)

In the literature, either one of formulations (Pt) or (P2) is known as global


or weak formulation of a static problemj they are also known under the name
virtual work theorem (VWT)

le We shall now prove that formulations (Po) and (P2) are equivalent. We
have already proven that (Po) => (P2), we only need to show that (P2 ) =>
(Po). We have:

Using the divergence theorem, Eq. (2.7b) becomes (we assume that u is a
solution of (P2 )):

-!r O"ijvinj dF + rO"ij,jvi


la
dJl + r /ivi
la
dJl + r Fivi
lr
F
dF = O, Vv* E Y*
(2.8)

Since vi = O over Fu, this is equivalent to:

We now need to show that O"ij,j + fi = O in Jl and O"ijnj = Fi on FF' Following


(Hughes, 1987), we first choose vi = (O"ij,j + fi) rP where:
(i) rP > O in Jl' (ii) rP = O on F and (iii) rP is "sufficiently smooth" (the
last two conditions guarantee that v* E Y*). Replacing in Eq. (2.9), we find:

1,(O"ij,j + fi)",(O"ij,j + /i)J~ dJl = O, (2.10)


2':0 >0

which implies that O"ij,j + fi = O in Jl.


We now choose vi = (O"ijnj - Fi) 'ljJ, where:
(i) 'ljJ > O on FF, (ii) 'ljJ = O on Fu and (iii) 'ljJ is "sufficiently smooth" (we
do have v* E Y*). Replacing in Eq. (2.9), we find (using the result that we
obtained in the first step: O"ij,j + /i = O):

(2.11)
2.3 Displacement-based variational formulat ion 33

which implies that (aijnj - Fi) = O on rF.


To conclude, we have proven that if u is solution of (P2 ), it is also a
solution of (Po) .

.. Remarks:

Equation (2.4), at the basis of the VWT is independent of the constitu-


tive model. Therefore, the VWT is written in identical fashion for plasticity,
viscoplasticity, etc., under the small perturbation hypothesis (SPH). Even in
large deformations, it has a similar form (see Chap. 14). It is helpful to read
formulation (P2 ) of the VWT in the following manner:
(The work of the internalloads in the deformations due to the virtual
displacement field v*) = (The work of the externalloads in the virtual
displacement field v*), Vv* E T*.
In reality, the work of the internalloads is defined with an opposite sign,
so the VWT actually reads:
(The work of the internalloads ...) + (The work of the externalloads ... ) = O.
We feeI however that for applications, the first reading is easier.
The VWT is the foundation of powerful numerical methods, in particular the
finite element method (FEM).
The notion of work allows to take into account without difficulties cases of
forces or moments which are concentrated or per unit length, while the local
formulat ion (Po) is not suited for that. Actually, in some textbooks, the VWT
is used as point of departure, and (Po) is found as a consequence (e.g., see
Salenc;on (1988a)).
In formulation (P2 ), we say that (aij and Vii,j»)' (Ii and vi) and (Fi and vi)
are dual or conjugate variables, because the scalar -inner- product of each
pair gives a work.
With an appropriate choice of dual variables, it is possible to construct special
theories (e.g., beams, plates and shells) directly from the VWT (e.g., see
Salenc;on (1988b)). For instance, for the beams studied in Chap. 3, we do not
take (stress and strain) but: (bending moment and curvature), (normal force
and axial strain), (shear force and shear strain).

2.3 Displacement-based variational formulation


From now on, we shall restrict our attention to linear elasticity:
(2.12)

In this chapter, we allow for heterogeneous (Cijkl = Cijkl(Z)) and anisotropic


materials. Hooke's operator Cijkl is:
34 2. Variational formulations, work and energy theorems

-symmetric: Cijkl = Cjikl = Cijlk = Cklij,


-positive definite: '<Ix E fl and '<111 second-order symmetric tensor ("7ij = "7ji),

Cijkl"7ij"7kl 2: O and Cijkl"7ij"7kl = O => "7ij = O


In order to study the "weak" formulation (P2 ) , we will rewrite it with the
help of the following notations:

a(u,w) fa CijkIU(k,I}W(i,j) dfl;

< >
lar J;Wi dfl; < F,w >rF= r FiWi dr
j,w = (2.13)
lrF
It is easy to verify that a(., .), < .,. > and < .,. >rF are symmetric and
bilinear forms. Symmetry means that:

a(u,v*) a(v*,u);
< j,v* > < v*,j >; < F,v* >rF=< v*,F >rF (2.14)

Bilinearity means that we have linearity w.r.t. each argument:

a(o:u+ f3v, v*) o:a(u,v*) + f3a(v,v*),


a( u, o:v* + f3w*) o:a(u,v*) + f3a(u,w*), (2.15)

idem for < .,. > and < .,. >rF' Due to the positiveness of e, a(.,.) is a
positive form: a(w, w) 2: O.
We now present a new formulation (P3 ) of the original problem .

.. Formulation (P3):

Find the displacement field u(x) solution of the following problem:

uEY
a(u,v*) =< j,v* >+< F,v* >rF' '<Iv* E Y* (2.16)

.. We have shown in Sec. 2.2 that formulations (Po) and (P2 ) are equiv-
alent, and therefore (Po) and (P3 ) are equivalent. We shall now prove that
if (P3 ) admits a displacement solution, then this solution is unique. Let us
assume that there exist two displacement fields u and v which are both so-
lutions of (P3 ). We then have:

a(v,v*) - a(u,v*) = O, '<Iv* E Y* (2.17)

And since a(.,.) is bilinear, this is equivalent to:


2.4 Potential energy theorem 35

a(v - u,v*) = O, \lv* E Y* (2.18)


Let us choose v* = v - u (recall that since u and v E Y, then v* E Y*).
We then have: a(v - u, v - u) = O. Since c is positive definite, we obtain:
V(i,j) - U(i,j) = O, and therefore:

v- u = "rigid body displacement" , (2.19)

Le., there exist two vectors a and b such that in each point M of the solid
body we have:
~

v-u= b+OM xa (2.20)

The rest ofthe prooffollows (Duvaut, 1990): if ru contains three non-aligned


points M l , M 2 and M 3 , then since u = U = v on ru, we obtain the equalities:
~ ~ ~

O=b+OMl xa=b+OM2 xa=b+OM3 xa, (2.21)

from which we deduce that:


~ ~ ~

M l M2 xa = O; M 2 M 3 xa = O; M 3 Ml xa = O (2.22)

Since MI, M2 and M 3 are not aligned, the 3 equalities imply that a = O, and
consequently b = O.
We have thus proved the uniqueness of the displacement field solution of
(P3 ) (and since (Po) and (P3 ) are equivalent, we also have uniqueness for
(Po)). Note that we have not demonstrated the existence of a solution; for a
proof see (Duvaut, 1990).

2.4 Potential energy theorem


The potential energy I(w) associated with a displacement field w is defined
as:
1
I(w) == 2a(w,w)- < j,w > - < F,w >rF (2.23)

We now introduce a new formulation (P4 ), which is the potential energy


theorem.

,. Formulation (P4):

Find the displacement field u(a:) solution of the following problem:

uEY
I(u) ::; I(v), \Iv E Y (2.24)
36 2. Variational formulations, work and energy theorems

We are going to prove that problems (P3 ) and (P4 ) are equivalent. Our proofs
of the theorems of potential energy, complementary energy and energy bounds
follow those of Duvaut(1990); for other proofs see, e.g. (Lanczos, 1970), (Lipp-
mann, 1972), (Dym and Shames, 1973), (Oden and Reddy, 1976), (Mason,
1980), (Parton and Perline, 1984b). First, we show that (P3 ) ::::} (P4 ). We
have:
1 1
I(v) - I(u) 2a(v,v)- < j,v > - < F,v >rF -2a(u,u)+ < j,u >
+ < F,u >rF
1
2a(v - u,v - u)
, .... ,
:;:::0
+a(u,v-u)- < j,v -u > - < F,v -u >rF
" . ,
= O, using (P3 ) with v* = v - u
We have used the properties of bilinearity and symmetry of the forms in order
to rearrange the terms as shown.
We now show that (P4 ) ::::} (P3). Let v be any element of Y and a a real
number. Let w = u + a(v - u); we have w E Y. Using (P4 ) with w along
with the bilinearity and symmetry properties, we obtain:

a2
2"f(v - u,v - u),+a[a(u,v - u)- < j,v - u >
....
:;:::0
- < F,v - u >rFl2:: 0, Va E lR (2.25)

The discriminant must be :S O, which implies that:

a(u,v - u)- < j,v - u > - < F,v - u >rF = 0, (2.26)

Le., (P3 ) (recall that (v - u) E Y*) .

.. Mechanical interpretation:

For a K.A. displacement field v, we have:


~a(v,v): strain energy (see Chap. 1 and Sec. 2.8);
< j,v > + < F,v >rF : work of externalloads;
I(v) = ~a(v,v)- < j,v > - < F,v >rF: potential energy.
Theorem (P4) states that among all K.A. displacement fields v, the so-
lut ion u is the one which minimizes the potential energy. This theorem has
important applications; it allows for instance to find approximate solutions
to problems which cannot be solved with the local or "strong" formulation
(Po). Ritz's method (Sec. 2.11.1) is based on (P4 ).
2.5 Stress-based variational formulat ion 37

2.5 Stress-based variational formulat ion

Let us write the constitutive equations as follows:

(2.27)

(D: compliance tensor, c: stiffness tensor). Let '11 be a second-order symmetric


tensor ('T}ij = 'T}ji). We multiply the constitutive equations by 'T}ij and integrate
over n:
(2.28)

Since '11 is symmetric, we have:

(2.29)

Integration by parts gives then:

In Dijkto'kl'T/ij dn = r u(r/ijnjdr - lnr Ui ââ% dn


lr Xj

r
lru
Ui'T}ijnjdr + r
lr
F
Ui'T}ijnjdr

_ r Ui â'T}ij dn (2.30)
ln âXj

We now introduce the following two sets:

E {7' = (Tij) I Tij = Tji, ~~j + li = O in n, Tijnj = Fi on rF}

E* {7'* = (Ti}) ITi} = Tji' ~~; = O in n, Ti}nj = O on r F } (2.31)

If 7' E E, then 7' is said to be "statically admissible" (S.A.) For example,


the stress field u solution of (Po) is S.A. (u EE). Elements of E* are called
self-equilibrated stress fields, i.e. in equilibrium without body or surface forces
(an example from another context: residual or initial stress fields are in self-
equilibrium). AIso, an element 7'* E E* can be viewed as being the difference
of two S.A. fields. For instance, we may consider: 7'* = 7' - u, where u is the
stress solution (u E E) and 7' is any S.A. field (7' EE).
Let us write (2.30) for '11 = 7'*, where 7'* E E*:

(2.32)

We are going to rewrite this result using the following notations:


38 2. Variational formulations, work and energy theorems

A(O", '11) = la DijklUkl'f/ij dn

< U,'11 >ru = ( Ui'f/ijnjdr (2.33)


iru
It can be easily checked that A(.,.) and < .,. >ru are bilinear and that A(.,.)
is symmetric positive definite.
We now introduce a new formulation (Ps ) of the original problem.

It Formulation (Ps):

Find the stress field O"(a:) solution of the following problem:

O"EE
A(O",T*) =< U,T* >ru' VT* E E* (2.34)

We have already shown that (Po) ~ (Ps)j it can be shown that problems
(Po) and (Ps ) are equivalent.

It We are now going to show that the stress field solution of (Ps ) is unique.
Let us assume that there exist two stress fields O" and T which are both
solutions of (Ps). We then have:

A(O",T*) - A(T,T*) = O, VT* E E* (2.35)


Linearity of A(.,.) w.r.t. the first argument implies that:

(2.36)

Let us choose T* = O" - T (recall that T* E E*)j we have:

A(O" - T, O" - T) = O ~ O" - T = 0, (2.37)


due to the positiveness of A(., .).

2.6 Complementary energy theorem

First, we introduce the following notation:

(2.38)

We now introduce a new formulation (P6 ), which is the complementary energy


theorem.
2.7 Energy bounds 39

,. Formulation (P6):

Find the stress field u solution of the following problem:

uEI)
J(U) ~ J(T), "IT E I) (2.39)

,. We are going to show that problems (P5 ) and (P6 ) are equivalent. First,
we show that (P5 ) => (P6). Using the properties of A(.,.) and < .,. >ru ' the
following result is easily established:
1
J(T) - J(u) = 2A(T - U,T - u) + A(U,T - u)- < U,T -u >ru
... , ...
'"
"

>0 = O, using (P5 ) with T* = T - u


(2.40)

We now prove that (P6 ) => (P5 ). Let T be any element of E and a a real
number. Let TI = u + a(T - u); we have TI E E. Using (P6) with TI, and
taking into account the properties of A(., .) and < ., . > ru, we obtain:

2a
2
;1(T - U, T - u), +a [A(U,T - u)- < U,T - u >rul2:: O, Va E IR
'"
~o
(2.41)

The discriminant must be ~ O, which implies that:

A(U,T - u)- < U,T - u >ru= O, (2.42)

Le., (P5) (recall that (T - u) E E*).

,. Mechanical interpretation:

(-J( T)) is called the complementary energy of the S.A. stress field T. The-
orem (P6 ) states that among alI S.A. stress fields T, the solution u is the
one which maximizes the complementary energy. We shall see in the next
section that the displacement and stress fields (u, u) solutions of problem
(Po) satisfy: I(u) = -J(u).

2.7 Energy bounds

If u and u are the displacement and stress fields solutions of problem (Po),
we have:
40 2. Variational formulations, work and energy theorems

u E r, u E E,
-J(T) :S -J(u) = I(u) :S I(v), '<Iv E r and '<IT E E (2.43)

If we examine (P4 ) and (P6 ), we see that the only result that needs to be
proven is the equality I(u) = -J(u). Using the definitions of I(u) and J(u),
we have:

I(u) + J(u)

Since CijkIU(k,1) = aij and DijkLO"kl = U(i,j) , we obtain:

If we use Eq. (2.4) with w = u, we do obtain: I(u) + J(u) = O.

2.8 Stored energy

If u and u are the displacement and stress fields solutions of (Po), Eq. (2.45)
gives:

(2.46)

Which can be read as follows:

(The work of the external loads in the displacement solution u)


+ (The work of the reaction forces in the given displacement U)
= (Twice the stored energy)

However, it must not be thought that only half of the energy which is de-
veloped by the loads is stored in the body. We should not confuse work and
stored energy. For example, and as the VWT or the reciprocity theorem (Sec.
2.9) clearly show, we can very well define the work of a force in a displacement
which is not due to that force.
Let us assume that the final state (u, u) is reached by successive equi-
librium states (au, au), O ~ 1. Since the problem is linear, each state
(au, au) corresponds to given data al in il, aF on r F and aU on r u .
The sum of the elementary works which are developed in order to reach
progressively the final state (u, u) is:
2.9 Maxwell-Betti reciprocity theorem 41

w = 11 {In (O'.fi) (UidO'.) dil + lF (O'.Pi) (uida.) dr

+ lu (O'.Uijnj) (UidO'.) dr}

11 {In fiUi dil lF+ PiUi dr + lu UijnjUi dr} O'. da.

~ {l hUi dil + tF PiUi dr + tu UijnjUi dr}

= ~ 1 UijU(i,j) dil, (2.47)

Le., the stored energy in the body.

2.9 Maxwell-Betti reciprocity theorem

We study the same body, subjected to two different loading systems:


System (1): f(l) in il, p(l) on r F and U on ru; solution: ",(1), 00(1).
System (2): f(2) in il, p(2) on rF and U on ru; solution: ",(2), 00(2).
Note that rF, ru and U are the same in both cases. The VWT in its form
(Pt) gives in each case:

fa 00(1) : e(v) dil = fa f(l) . v dil + tF p(l) . v dr

+ r
iru
(00(1). n)· U dr, Vv E Y

In 00(2) : e(v) dil = fn f(2) • v dil + lF p(2) • v dr

+ r
iru
(00(2). n) . U dr, Vv E Y (2.48)

Application of Eq. (2.48a) with v = ",(2) gives:

in
r f(l). ",(2) dil + r
irF
F(l). ",(2) dr

+
iru
r (00(1). n) . U dr (2.49)

On the other hand, using the constitutive equations, we have:


42 2. Variational formulations, work and energy theorems

la 0'(1) : e(u(2») dn = la c : e(u(1») : e(u(2») dn

= la 0'(2) : e(u(1») dn

r /(2). U(I) dn + r F(2). U(I) dF


in irF
+ r
iru
(0'(2). n)· U dF «2.48b) with v = u(1»)
Combining the last equality with Eq. (2.49), we obtain:

fn /(1) . U(2) dn + r F(1)·


irF
U(2) dF + r iru
(0'(1). n) . U dF =

fn /(2) . U(I) dn + r F(2). U(l) dF + r (0'(2). n) . U dF (2.50)


irF iru
This is Maxwell-Betti, reciprocity theorem, which can be read as follows:

Work of loading (1) in the displacements due to loading (2) =


Work of loading (2) in the displacements due to loading (1)

A major advantage of the theorem is that the solution of a given problem can
be found by solving another, simpler problem. Examples are given in Sec.
2.12 and Chap. 3.

2.10 Castigliano's theorem


In aU this section, we consider the same linear elastic body, and assume that
the imposed displacements on Fu are zero and that each loading case is such
that the whole body is in static equilibrium.
We consider a basic problem where the external loading consists of a
fi n.enumeroJpmn
't b ,1 . t fiorces F1lW (1) , ••• ,L'nW (n) ,werew
h r;l(1) , ... ,W (n) are
unit vectors. The forces are applied on points X(I), ••• , x(n) of the surface.
The displacement solution under each loaded point x(i) is designated by
U(x(i»).
We now consider n intermediate problems (i), where in each one a single
unit force w(i) is applied at point x(i) of the surface. The displacement field
solution of each problem (i) is designated by u(i)(x). Using the superposition
"principle" , the displacements of the basic problem can be written as follows
(see Fig. 2.2 for an illustration):
n
U(x(j») = :EFiU(i)(xU)) (2.51)
i=1
2.10 Castigliano's theorem 43

For the basic problem, the displacement Uj under a force Fjw(j) and in the
direction of the force is defined by:
n
Uj == w(j) . U(x(j») = LFiW(j) . U(i) (x(j») (2.52)
i=l

··· ...... ·· ...... ·1 Fn

Fig. 2.2. Illustration of the superposition principle in the case of a straight beam.
The basic problem is decomposed onto n problems each one consisting of a single
force Fi applied at (x = Xi), i = 1,2, ... , n.

The strain energy of the basic problem is given by:

(2.53)

Differentiation w.r.t. Fk gives:

(2.54)

Note that coefficients w(j) . util (x(j») are independent of the loading Flw(l).
They are constant, Le. they only depend on the elastic properties, the ge-
ometry and the displacement B.Cs. After simplification, Eq. (2.54) can be
rewritten as follows:
44 2. Variational formulations, work and energy theorems

âW =.! t Fi [w(k) . u(i)(x(k») + W(i) • U(k)(X(i»)] (2.55)


âFk 2 i=1

Now, applying Maxwell-Betti reciprocity theorem to any two problems (i) and
(j), we find the following symmetry reiat ion (see Fig. 2.3 for an illustration):

(2.56)

Therefore, (2.55) can be rewritten as:

= L FiW(k) . U(i)(X(k») = Uk,


âW
-
n

âFk i=l

using (2.52). We have thus obtained Castigliano's theorem which gives the
displacement under a concentrated force and in the direction of the force:

(2.57)

The theorem is also applicable for concentrated moments, and gives the ro-
tation under a concentrated moment in the direction of the moment:
âW
(Jk =-- (2.58)
âMk
The theorem is widely used in beam theory (Chap. 3) in cases where we
are interested in computing not the displacement field but only the displace-
ment of a given point; it also allows to compute reaction forces for statically
indeterminate problems.

Fig. 2.3. Illustration of problems (i) and (j) in the case of a straight beam: a unit
force is applied at (x = Xi) and (x = Xj), respectively. Maxwell-Betti theorem gives:
1 x Uj(Xi) = 1 X u'(Xj).
2.11 Intraductian ta numerical methods 45

2.11 Introduction to numerical methods


2.11.1 Method of Ritz

Ritz's numeric al method can be understood as folIows. We limit ourselves to


K.A. displacement fields of the form:

L
n
V(R) (x) = bA1/J(A) (x) (2.59)
A=l

where 1/J(1), . .. ,1/J(n), are given functions of x chosen so that v eR ) (x) is K.A.
Coefficients b1 , ... , bn are computed in such a way as to minimize the poten-
tial energy:

aI(v eR ))
abA = o, A = 1, ... ,n (2.60)

We thus obtain the following linear system of n equations with n unknowns


(b 1 , ... ,bn ):
n
L bBa(1/Je A),1/Je B )) =< f,1/Je A) > + < F,1/Je A) >rF' A = 1, ... ,n (2.61)
B=l

Ritz's method is approximate because it does not consider alI K.A. displace-
ment fields, but only those of the form (2.59). A few applications are given
in Chaps. 4 and 6; many more examples can be found in (Dym and Shames,
1973) and (Dumontet et al., 1994).

2.11.2 Finite element method (FEM)


In this section, we shall give a brief introduction to the finite element method
(FEM), which is perhaps the most powerful numerical method used in solid
mechanics.
In Galerkin's approximation, finite-dimensional sets ye h) and y*(h) are
used instead of Y and Y*. It is assumed that:
ye h) cY (i.e., if v eh ) E ye h) then v eh ) E Y),
y*e h) c Y* (i.e., if v*e h) E y*e h) then v*e h) E Y*) (2.62)

Therefore, instead of solving problem (P2), we seek a solution for the folIowing
problem: Find the displacement field u eh ) which satisfies:
u eh ) E ye h)

In (JijV~i:~? dD = In fiV;e h) dD + lF Fiv;e h) dr, Vv*(h) E y*e h)

"Constitutive equations" (2.63)


46 2. Variational formulations, work and energy theorems

AIso, if v*(h) E y*(h), it is assumed that:

=L
n
v;(h)(x) ~(A)(x)Vi*(A) (2.64)
A=l

where ~(A) are given functions of (x), they are called interpolation or
shape junctions; Vi*(A) are coefficients and (A = 1, ... ,n) are called nodes.
Galerkin's equation (2.63b) now becomes:

Ln
Vi*(A) [ aij
a (A) d{} =
-~-. Ln
Vi*(A)
[
[h~(A) d{} + [ Fi~(A) dr] ,
A=l ln ax] A=l ln lrF
(sums over i, i = 1,2,3) (2.65)
This equation must hold Vv*(h) E y*(h). Since ~(A)(x) are given, Eq. (2.65)
must hold for arbitrary values of Vi*(A). We can set any Vi*(A) =/:- O and alI
other Vi*(B) = O, we thus obtain:

[ aij a~(A) d{} = [ h~(A) d{} + [ Fi~(A) dr, (sum over i = 1,2,3)
ln aXj ln lrF
(2.66)

The FEM solution u(h) to Galerkin's problem is assumed to be in a form


similar to Eq. (2.64):
n
u~h)(x) =L ~(A)(X)Ui(A) (2.67)
A=l

where UitA) is the nodal displacement of node (A) in the direction Xi. In linear
elasticity, the stress corresponding to the displacement u(h) is:

(sums over k, l = 1,2,3) (2.68)

Therefore, Galerkin's formulation can be written as follows: Find the nodal


displacements ulA) which satisfy:

t
~
ui B )
h
[ Cijkl a~(B) a~(A)
~ ~
d{} = [ h~(A) d{} + [ Fi~(A) dr,
h ~
i = 1,2,3, A = 1, ... ,n, (sums over i, k, l = 1,2,3) (2.69)

This is a linear system of 3 x n equations and 3 x n unknowns (UlA»); it can


be written under the form:

[K]{d} = {F} (2.70)


2.12 Solved problem: volume change of an axially compressed body 47

where [K] is called the stiffness matrix, {d} the nodal displacement array and
{F} the nodal force array.

.. Remarks

The finite element method (FEM) involves a number of other steps whieh
are briefly described hereafter.
The space domain is discretized into a finite number of elements; compu-
tations are carried out at the element level and then assembled at the global
level. For instance In. dx ~ Le In. • dx where "e" designates an element.
Integrals over elements are computed using quadrature rules (e.g., Gauss
integration) .
The quality of the finite element approximation depends on the choiee of
the number ofnodes (n) and the interpolation functions (E(A»), which in turn
depends on the type of elements.
In linear elasticity (2D, 3D, beams, plates or shelIs), it can be shown
that starting from the appropriate weak formulation, the FEM always leads
to a linear system similar to (2.70). Actually, even in nonlinear cases (e.g.,
plasticity, finite strains, etc.) we arrive -iteratively- to a similar linear system
(see Chap. 12).
In order to find a detailed information about the above mentioned top-
ies, the reader needs to consult a specialized textbook, e.g. (Hughes, 1987),
(Bathe, 1982), (Zienkiewiez and Taylor, 1989) (there are numerous other
books on finite elements, ranging from the very mathematieal to the "cook-
book").

2.12 Solved problem: volume change of an axially


compressed body

Consider a body of arbitrary shape, subjected to two external forces ±F


acting on material points A and B, respectively, and aligned with the material
segment (AB); see Fig. 2.4-a. The initial volume of the body (i.e., before
application of the forces) is Vo and the initial distance between A and B is 10.
The problem is to compute the volume V after deformation. The material is
homogeneous, isotropic, linear elastic, of Young's modulus E and Poisson's
ratio 1/.
We use the reciprocity theorem to solve this problem, which we designate
by (1). Problem (2) is defined as a uniform pressure (-pn) applied on alI
the external surface an of the body, Fig. 2.4-b, where n is the outward unit
normal to the boundary. Designating the loads in each problem by F(i) and
the displacement solutions by u(i), MaxwelI-Betti theorem gives:
48 2. Variational formulations, work and energy theorems

Fig. 2.4. (a) Problem (1): axially compressed body; (b) Problem (2): uniform
pressure on the surface of the body.

Work (.F(1) , U(2)) Work (.F(2) , u(1)),

i.e., F(lo _1(2)) r (-pn). u(1) dr,


l8n
(2.71)

where 1(2) is the final distance between A and B due to the pressure loading
(2). Using the divergence theorem, we obtain:

F(I(2) -lo) = p fa div (u(1)) dV = p(V(I) - Vo), (2.72)

where V(1) = V designates the final volume due to loading (1). In establishing
(2.72), we used the fact that p = constant, and the following result for the
change of volume which is valid under the SPH (Chap. 1)
dV(I)
- - = 1 + div uel) (2.73)
dV
The stress solution of problem (2) is a hydrostatic uniform state:
(2) _ 1:
aij - -PUij (2.74)
The strains are also uniform and given by isotropic linear elasticity as:

(2.75)

where '" is the bulk modulus, 3", = E/(l - 2v). Under SPH, the relative
length change of material segment (AB) is equal to a direct strain:
1(2) - lo
(2.76)
lo
Substituting this result into Eq. (2.72), the original problem is solved:
Fl o
V-Vo = - - (2.77)
3",
It is immediately seen that if v -* 1/2, then '" -* 00 and V -* Vo: there is no
volume change.
3. Theory of beams (strength of materials)

In this chapter, we present the classical theory of beams, also known as


"strength ofmaterials" (this denomination, although inaccurate and obsolete,
is stiH widely used in practice). A beam is a solid for which one dimension -the
length- is much larger than the other two dimensions. We shall see that beam
theory offers a dramatic simplification of the original three-dimensional (3D)
problem of Chaps. 1 or 2 by reducing it to a one-dimensional (lD) problem
along the so-called middle fiber.

3.1 Definitions and geometric properties


A beam is a solid generated by a planar surface (S) (the cross section) whose
centroid (G) moves along a curve (,C) (the middle fiber), the plane of (S)
being orthogonal to (,C); see Fig. 3.1.

curve ,C

Fig. 3.1. Example of a beam. Notation: (5): cross section, G: centroid of (5) and
L: middle fiber.

The dimensions of (S) must be small in comparison with the length (l)
of,C (typically the largest dimension of (S) is of the order of l/10). The cross
section (S) is either uniform or varies slowly along (C).
If (C) is a planar curve, then the beam is called planar too. A beam is
called straight if (C) is a straight line. Examples of cross sections of beams
possessing a symmetry plane are plotted in Fig. 3.2 (rectangular, circular and
T sections).
50 3. Theory of beams (strength of materials)

y y y
Fig. 3.2. Examples of cross sections symmetrical w.r.t. the (y) axis.

In Fig. 3.3 are plotted a planar surface (S) of centroid (G) and two Carte-
sian frames (O,Y,Z) and (G,y,z). The coordinates of (G) in the "global"
frame (O, Y, Z) are given by:

fs Y dY dZ _ S z fs Z dY dZ _ Sy
Ya = fs dYdZ = A' Za = fs dYdZ = A' (3.1)

where A is the area of (S), and Sy and Sz the first moments of (S) W.r.t.
the axes (OY) and (OZ), respectively.
The moments of inertia of (S) are defined as follows:

w.r.t. (OZ) : Iz = fs y 2 dY dZ, w.r.t. (OY) : /y = fs Z2 dY dZ

mixed : Iyz = fs Y Z dY dZ, polar : Io = fs (y 2 + Z2) dY dZ


(3.2)

O.-____________ Z ~_

S
Y
Y
Fig. 3.3. A planar surface (S) of center (G) and two Cartesian coordinate frames
(O, Y, Z) and (G, y, z).

A matrix of geometric inertia is defined by:

[ Iy I yz ]
Iyz Iz
3.2 Pure bending of a straight beam: 3D elasticity solution 51

Since it is symmetric (ly z = /zy), there exists a basis formed by the so-
called principal axes of inertia in which the matrix is diagonal. If (Gy) is an
axis of symmetry for the cross section (S) (as in Fig. 3.2), then we shall prove
that (Gy) and (G z) are principal axes of inertia. lndeed, the mixed moment
of inertia w.r.t. the (G, y, z) Cartesian frame is given by:

Iyz = 1 S
yz dy dz = l Y=...

y=...
y dy
jb(Y)/2

-b(y)/2
z dz = o,
"--
o '"
where b(y) designates the width of the beam (for a rect angular cross section,
b(y) = constant).
Exercise: show that the centroid of the triangular section of Fig. 3.4 is
given by Ya = h/3 and the moments of inertia of the rect angular section in
the same figure are Iz = bh3 /12 and Iy = hb3 /12.
For more information on geometric properties and many more examples,
see (Pissarenko et al., 1979).

-b/2 o +b/2 Z
- h/2
h/3
-+--4,..--~
G z
2h/3
-b /2 +b/2

Y +h/2
Y
Fig. 3.4. For the triangular section: Yo = h/3, and for the rect angular section:
Iz = bh3 /12 and I y = hb 3 /12.

3.2 Pure bending of a straight beam of uniform cross


section: 3D elasticity solution
Consider a straight beam of uniform cross section which is subjected to bend-
ing moments =FM e z at its ends x = ±1/2 (Fig. 3.5). The cross section is
symmetrical w.r.t. the (Gy) axis, thus (Gy) and (Gz) are principal axes of
inertia. We shall prove that the stress solution is given by:

a xx = ay, aij = O if (i,j)::fi (1,1), (3.3)

where a > O is a positive constant to be determined. According to (3.3),


the axial stress a ,''x is linear along the height (y) of the beam, it is tensile
52 3. Theory of beams (strength of materials)

(positive) in the lower part (y > O) and compressive (negative) in the upper
part (y < O). This corresponds to physical expectationj the "triangular" stress
diagram is plotted in Fig. 3.6.

M M -h/2
( __ ._._._._._._._._._.@.!._._._._._._._._._. , x z

l/2 l/2 +h/2


y
y
Fig. 3.5. Pure bending of a straight beam of uniform cross section

Fig. 3.6. Triangular diagram for the axial stress Uxx

The Beltrami-Mitchell compatibility equations in the absence of body


forces (.:1.:1uij = O, Sect. 1.14) are satisfied since the only non-zero stress
component is linear in y. Equilibrium equations read:
âUij _ O
âXj - ,

and are satisfied. The stress vector on the lateral sudace of the beam is:

It vanishes identically, i.e. the stress free B.C. on the lateral sudace is satis-
fied. The stress vector on the end sections (x = ±l/2) is given by:
(TT . (±ex ) = ±ayex
It has a zero resultant:

±aex fs y dy dz = ±aexAYG = 0,
3.2 Pure bending of a straight beam: 3D elasticity solution 53

because the cross section frame (G, y, z) is centered in G. We now compute


the resultant moment w.r.t. G of the stress vector on the end sections:

fs (ye y + ze z ) x (±ayex ) dydz =faez fs y2 dy dz ± aey fs yz dy dz


=faezlz ± aeylyz
The second term vanishes because Iyz = O, since (Gy) and (Gz) are principal
axes of inertia. Equating the moment of the stress vector with the applied
moment (=fMe z ) at the end sections, we find:

The stress solution is now completely determined:

axx = ~ y, aij =O if (i,j) f. (1,1) (3.4)

According to Saint- Venant's principle (Sect. 1.15), the way the external mo-
ments are actually applied to the end sections x = ±1/2 does not matter as
long as on those sections: (i) the resultant of the stress vector is zero, (ii)
the resultant moment of the stress vector is equal to the resultant moment
which is applied. The stress solution (3.4) will then be a good one "far from
the ends" , a practical requirement being:
l l
x E [-- +h - - h] (3.5)
2 '2 '
where h designates the largest cross section dimension. Isotropic linear elas-
ticity gives the strains as:
M vM
txx = El Y' tyy = tzz = - EI y, txy = t yz = t xz = O, (3.6)

with I == Iz. The direct strains are related to the displacements in the x, y
and z directions by:
AU OV ow
txx = ax' tyy = oy' tzz =oz
-

Integration of these relations using (3.6) gives:

M vM 2 () vM
u= Elyx+F(y,z), V=-2El Y +Gx,z, w=- Elyz+H(x,y)
(3.7)
The shear strains are related to the displacements by:
au OV ov ow au ow
2txy = oy + ax' 2tyz = OZ + 8y' 2t xz = OZ + ax
54 3. Theory of beams (strength of materials)

Since alI shear strains are nil, Eqs. (3.7) give:

M âF âG vM âH + âH = O
El x + ây + âx = O, - EI z + au + âG
âz = O,
âF
âz âx (3.8)

We make the choice: F = H = O (we will comment on this later). Equation


(3.8c) is satisfied, while (3.8b) implies that:
vM 2
G(x, z) = 2Elz + g(x)
Substituting into (3.8a), we obtain:
M 2
g(x) = ___ x + constant
2EI
Therefore, the displacements are given by Eqs. (3.7) as:
M M
u = EIYx, v = 2EI 2
(-x - vy
2 2
+ vz + constant), w = - vM
EI yz (3.9)

Note that the first and second partial derivatives of the vertical displacement
w.r .t. x are given by:

(3.10)

We choose the constant in (3.9b) equal to 12 /4. For the middle fiber (y =z =
O), the displacements are:
M 12
u(x, O, O) = w(x, O, O) = O, v(x, O, O) = 2EI (_x 2 + 4'), (3.11)

from which it is deduced that: v(±l/2, O, O) = O. It appears that the displace-


ment Res. which are plotted in Fig. 3.7 -immovable hinge at (-l/2,0,0)
and simple support at (l /2, 0,0)- are consistent with the results found so
far. Consequently, for these displacement B.Cs., the choices that were made
(F = H = O, constant = 12 /4) are legitimate and Eqs. (3.9) give the dis-
placement solution (recall the uniqueness theorem). Figure 3.7 shows a very
amplified deformed middle fiber (remember that we adopt SPH -the smaH
perturbation hypothesis). The maximum deflection occurs in the middle of
the beam (x = O), as expected because of the symmetry of the problem.
Because of SPH, the curuature of the deformed middle fiber is:
1 â2v
R ~ - âx2 (x,O,O),

from which the foHowing fundamental relations are deduced:

M= EI,
R U xx = ER Y (3.12)
3.3 Basic assumptions of beam theory 55

Consider a point which occupies the positions (xo, Yo, zo) and (x, y, z) before
and after deformat ion, respectively. Equations (3.9, 12a) give:

1 1 2 2 2 12 V
x=xo+ RYoxo, Y=Yo+ 2R (-vYo+vzo -x o +"4)' Z=Zo- RYozo,

from which it is deduced that:

x Y 1 2 2 2 12
= 1 + 2R2
-x - R
o,
(vYo - vZo + X o - -4 )
,
(3.13)
"
~12/R2

An implication of the SPH is that the radius of curvature of the deformed


middle fiber is much larger than the beam's length, Le. 1/ R « 1. Therefore,
the term of order 12 / R2 in Eq. (3.13) can be neglected in front of 1, and the
equation simplifies to:

-Xox - -RY = 1, . Xo
l.e. x - -Y - Xo
R
= ° (3.14)

This is the equation of a plane, which means that a cross section remains
planar afier deformation. Moreover, the normal to the plane is (e x -eyxo/ R),
and this is precisely -see Eq. (3.10a)- the tangent to the deformed middle fiber
which was in (xo, 0, O) before deformation. Therefore, the conclusion is that
a cross section remains planar and normal to the deformed middle fiber.

v(x,O,O)
Y
Fig. 3.7. Middle fiber before and after deformat ion

3.3 Basic assumptions of beam theory

In this chapter, we adopt the SPH and consider isotropic linear elasticity.
The SPH allows to write equilibrium equations on the initial, undeformed
configurat ion (this will not be the case however for buckling problems, see
Chap. 10). In order to develop a beam theory, more assumptions are needed,
and these are described in the following subsections.
56 3. Theory of beams (strength of materials)

3.3.1 External loads

A beam supporting a wall and subjected to its own weight is depicted in


Fig. 3.8a. A first model considers two externalloadings: forces per unit area
are applied to the upper surface of the beam and represent the load due
to the wall, and forces per unit volume represent the (dead) weight of the
beam (Fig. 3.8b). Another, simpler model is used in beam theory: the entire
external load (wall + dead weight) is represented by forces per unit length
applied directly to the middle fiber (Fig. 3.8c). For more on beam modeling
aspects, see (Feodossiev, 1976).

I I
I I

~
~
I I
I I

+ I

»IIIIIIIIIII~
Fig. 3.8. A beam supports a wall and is subjected to its own weight. First model:
forces per unit area on the upper surface and forces per unit volume. Second model
(beam theory): forces per unit length on the middle fiber.

In this chapter, we consider beams (usually straight) possessing a sym-


metry plane and loaded in their plane. For less restrictive assumptions, see
(Friaa, 1982), (Courbon, 1971), (Dellus, 1961), (Timoshenko, 1956a-b) and
(Popov, 1952).
For beams studied in this chapter, (Gy) is an axis of symmetry for the
cross section (8), external forces (including reactions) act in the (G, x, y)
plane and external moments (including reactions) rotate around the (Gz)
3.3 Basic assumptions of beam theory 57

axis. External forces and moments are supposed to be applied directly to the
middle fiber (L) = (Gx), they can be concentrated OI per unit length of (L).
A "general" case is given in Fig. 3.9.
Other types of loads can be accommodated. For example, if a force (He x )
is applied at position (xo, Yo, O), then it can be replaced by a force (H e x )
and a moment (-HYoe z ), both applied directly at the middle fiber, i.e. at
G(xo,O,O). As another example, consider a torsion torque (Ce x ). In this
case, we first study two problems, one of tors ion (Chap. 4) and one of a beam
loaded in its plane (this chapter). Next, we add up the two solutions by virtue
of the superposition "principle".

s
z

x = Xo x = Xo
Fig. 3.9. A beam with a symmetry plane is loaded in its plane. (a): externalloads,
(b) fictitious cut showing the internalloads (stress resultants).

3.3.2 Internalloads (stress resultants)

It is assumed that the dominant stresses are the normal (a xx ) and shear
(a xy ) stresses, ali other stress components are either nil or can be neglected.
Consequently, the stress vector on a cross section of outside normal ±ex is:

The stress resultant over the cross section is:

±ex l a xx dydz ± e y l a xy dydz = ±N(x)ex ± Q(x)e y ,

where N(x) and Q(x) are the normal and shear forces , respectively:
58 3. Theory of beams (strength of materials)

N(x) == fs O"xx dy dz, Q(x) == fs O"xy dy dz (3.15)

The resultant moment ofthe stress vector w.r.t. to the center (G) ofthe cross
section is given by:

M(x)

where we assumed that O"xx and O"xy are independent of the z-coordinate. The
bending moment M(x) in the previous equation is defined by:

(3.16)

Consider a beam with a symmetry plane (G,x,y) which is loaded in that


plane, Fig. 3.9. If we make a jictitious cut along a cross section (x), then
internalloads N(x), Q(x) and M(x) will appear on the left and right sides
as shown in the figure. Note that the simplified beam theory does not require
the continuity of the stress vector (lTT . e x ) as in Chap. 1 but only that of
the internalloads N(x), M(x) and Q(x) (we will see in Sect. 3.5 cases where
these loads can be discontinuous).
The sign conventions for N(x) and Q(x) follow those of u xx and O"xy,
respectively (Chap. 1): N(x) is positive if it is tensile and negative if com-
pressive; if the outer normal is ±ex , then a positive Q(x) is directed along
±ey. The bending moment is positive if it compresses the upper part (y < O)
of the beam.

3.3.3 Equilibrium equations

Consider in Fig. 3.10 an elementary portion of a beam of length (dx) sub-


jected to internalloads and external forces per unit length (-y(x) and p(x».
Equilibrium of forces in the (x) direction gives:

dN
N(x) + dN + 1'(x)dx = N(x), Le. dx = -1'(x)
Equilibrium of forces in the (y) direction gives:

Q(x) + dQ + p(x)dx = Q(x), Le. ~~ = -p(x)


3.3 Basic assumptions of beam theory 59

Equilibrium of moments in the (z) direction W.r.t. G (x, 0, O) gives:

dx dM
M(x) + dM = [Q(x) + dQ]dx + [P(x)dx]2" + M(x), i.e. dx = Q(x),

where second-order terms were neglected. In summary, for a beam with a


symmetry plane loaded in its plane, the differential equations of equilibrium
are not those of Chap. 1 (âUij/âxj + li = O) but the following:

dN dQ dM
- = -')'(x), - = -p(x), - = Q(x) (3.17)
dx dx dx
Note that the last two equations can be combined to give:

J2M
- = -p(x) (3.18)
dx 2

p(x) ')'(x)

Nv--Ţ
_M;ţ~
~_d_x--l~11 TN+
LdM

dN
, Q+dQ
Fig. 3.10. Equilibrium of an elementary portion of a beam of length (dx).

3.3.4 Navier-Bernoulli assumption

For a beam under a general loading, we stiU make the so-called Navier-
Bernoulli assumption: a cross section remains planar and normal to the de-
formed middle fiber. This assumption is valid as long as shear deformations
are smal!. A less restrictive model consists in assuming that a cross section
remains planar after deformation, but not necessarily normal to the deformed
middle fiber. This is known as the Timoshenko beam theory.

3.3.5 Constitutive equations

It is assumed that relation (3.12a) between the bending moment M(x) and
the curvature (1/ R(x)) of the deformed middle fiber holds in a general case:

EI
M(x) = R(x) , (3.19)
60 3. Theory of beams (strength of materials)

where E is Young's modulus and 1 the moment of inertia of the cross section
(8) w.r.t. the (Gz) axis. This means that local pure bending is assumed. The
assumption is valid if shear deformations are negligible. Under the SPH, the
curvature is given by:
1 d2 v
R(x) ~ - dx 2 '
where v(x) represents the deflection of the middle fiber. Consequently, v(x)
is solution of the following differential equation:

~v
-EI dx 2 = M(x), (3.20)

to which displacement B.Cs. must be appended (see Sect. 3.4). ITthe bending
moments are known (and it is usuallY much easier to compute M(x) than
v(x)), then it is possible to plot the correct shape of the deflection v(x)
without computing its expression. Indeed, Eq. (3.20) shows that the sign of
M(x) gives the curvature of the deflection. AIso, an inflection point for v(x)
is obtained at x = Xo if M(x) changes signs and goes through zero at x = Xo.
An illustration is given in Fig. 3.32. Using (3.18), we can transform (3.20)
onto a differential equation where p(x) appears directly:

IT E and 1 are uniform (independent of (x)), then the equation simplifies to:

d 4 v _ p(x)
(3.21)
dx 4 - EI
The rudal (Le., in the (x) direction) displacement u(x) can be computed from:
du _ N(x)
(3.22)
dx - EA
In the presence of bending, the axial displacement is much smaller than the
deflection.

3.4 Displacement boundary and continuity conditions


For a built-in section at (x = O), the B.Cs. are:
dv
u(O) = O, v(O) = O, dx (O) =O (3.23)

A variant of a built-in condition is when a non-zero rotation is possible, and


is proportional to the reaction moment MAi Eqs. (3.23) become in this case:
3.4 Displacement boundary and continuity conditions 61

dv MA
u(O) = O, v(O) = O, dx (O) = K ' (3.24)

where K [Nm/rad] is a stiffness. In both cases, since the axial and transverse
displacements are constrained, there are three reactions: two forces HA and
VA and a moment MA (Fig. 3.11).

=---
---
~~
~~­

(a) (b)
Fig. 3.11. Built-in section at (x = O) in two cases: (a) Rotation prevented, (b)
Elastic rotation.

For a simple support (or hinge on rollers) at (x = O), the B.C. is:
v(O) =O (3.25)
A variant is when the deflection has a non-zero value which is proportional
to the reaction force VAi Eq. (3.25) becomes in this case:
VA
v(O) = k' (3.26)

-
where k [N/m] is a stiffness (this is like having a linear spring as a support).
In both cases, since the deflection is constrained, there is a reaction force VA

r F---
in the y-direction (Fig. 3.12).

-- ----- - - --.

VA VA
ta) (b)
Fig. 3.12. Simple support (hinge on rollers) at (x = O) in two cases: (a) Deflection
prevented, (b) Elastic deflection.

For pinned B.Cs. (or immovable hinge) at (x = O), the conditions are:
u(O) = O, v(O) =O (3.27)
The difference with a simple support is that the rudal displacement is pre-
vented, and the difference with a built-in condition is that the rotation is
free. Since both displacements are constrained, there are two reaction forces
HA and VA (Fig. 3.13).
62 3. Theory of beams (strength of materials)

~f-------
Fig. 3.13. Pinned B.Cs. (immovable hinge) at (x = O).

For a free end at (x = O), since there are no externalloads at that 10-
cation, the internal loads M(O), Q(O) and N(O) also vanish. In terms of
displacements, these conditions can be written as foUows:

tPv d3 v du
dx 2 (O) = O, dx 3 (O) = O, dx (O) =O (3.28)

Quite often, in order to determine the deflection fields, we need to write


continuity conditions for the deflection v(x) and its slope dvjdx. For example,
consider the problem of Fig. 3.14. In that case, the conditions which need to be
written at support Bare: continuity of the slope dv j dx, the bending moment,
and consequently the curvature tPvjdx2 , and a zero deflection (VB = O).

~A ~C
Fig. 3.14. At support E, we need to satisfy the continuity of: dv/dx, d 2 v/dx 2 and
VB =0.

3.5 Computation of internalloads (stress resultants)


In this section, we show how the internalloads (or stress resultants) M(x),
Q(x) and N(x) can be easily computed using the method of jictitious cuts
and writing equilibrium conditions. Consider Fig. 3.9 again where a fictitious
cut is made at position (x = xo) of a loaded beam. We designate by:
p(L)e
x x
+ p(L)e
v v
the resultant of aU external forces (including reaction forces) acting on the
left part of the beam (Le., x ~ xo), and by:
M(L)e z

the resultant moment w.r.t. the center (G) of the cross section (x = xo) of
aU externalloads (including the reactions) which act on the same part. The
3.5 Computation of internalloads (stress resultants) 63

left part of the beam must be in equilibrium under the act ion of aliloads,
external (FJL), FJL), M(L») and internal (N(xo), Q(xo), M(xo)). Taking into
account the sign conventions for internalloads -Fig. 3.9- statics give:

We can also compute the internal loads from the equilibrium conditions of
the right part of the beam. Using similar notations as before, with the upper
script (R) (instead of (L)) designating the right part of the beam (x ~ xo),
we obtain (the sign conventions for internalloads being given by Fig. 3.9):

N(xo) = FJR), Q(xo) = FJR), M(xo) = _M(R) (3.30)

We now check that Eqs. (3.29, 30) are equivalent. Indeed, equilibrium of the
entire un-cut beam gives:
F(L)
x + F(R)
x = O, F(L)
y + F(R)
y = O, M(L) + M(R) =O (3.31)

So it is seen that working with the left -Eqs. (3.29)- or right -Eqs. (3.30)-
parts of the beam leads to identical expressions for the internalloads.
We now study the case of a concentrated force (Voe y ) applied at (x = xo).
We first consider a fictitious cut at xi) = Xo - Idxl. Equilibrium of the left
part of the beam gives:
Q(xi)) = -FJL)
Next, we consider a fictitious cut at xci = Xo + Idxl. Equilibrium of the right
part of the beam gives:
Q(xci) = FJR)
It is seen that Q(x) presents a discontinuity at (x = xo):
Q(xi)) - Q(xci) = _(FJL) + FJRl)
On the other hand, equilibrium of the entire, un-cut beam gives:

F(L)
y
+ v'o + F(Rl
y
= O

This shows that the jump of Q(x) across (x = xo) is equal to:

Q(xi)) - Q(xci) = Vo, (3.32)

=
i.e. the applied concentrated force at (x xo). Similarly, it can be shown that
if a concentrated force (Hoe x ) is applied at (x = xo), then the normal force
N(x) presents a discontinuity at (x = xo) with a jump equal to Ho. Also,
if a concentrated moment (Moe z ) is applied at (x = xo), then the bending
moment M(x) presents a discontinuity at (x = xo) with ajump equal to M o.
64 3. Theory of beams (strength of materials)

3.6 Computation of normal and shear stresses

Once the internalloads M(x), Q(x) and N(x) are known, we often need to
compute the stresses (Le., we need to "go back" from lD to 3D) in order for
instance to apply a strength criterion. For beams studied in this chapter, it
is assumed that the only meaningful stresses are the normal (O"xx) and shear
(O"xy) components; alI other stresses are neglected. We have:

M(x) N(x)
O"xx=-I-Y+~ (3.33)

It is assumed that the expression of the stress due to the bending moment
M(x) is identical to what was found for pure bending in Sect. 3.2. The stresses
given by (3.33) are plotted in Fig. 3.15 along the height of the beam (the Y-
direction). The figure shows that the stress due to M(x) varies linearly with Y
(a "triangular" diagram) while that due to N(x) is uniform along the height
(a "rectangular" diagram). The figure also shows that if N(x) f:. O, the middle
fiber does not coincide with the neutral fiber (Le., O"xx f:. O at Y = O). It is
also seen that the maximum normal stress in a given cross section (x) always
occurs at Y = Ymin or Ymax'

Fig. 3.15. The normal stress along the height of the beam is the sum of two terms:
one due to the bending moment ("trianguiar" diagram) and the other due to the
normal force ("rectangular" diagram).

By definit ion of Q(x), the avemge value of O"xy over a cross section is:

_ Q(x)
O"xy =-y
However, we shall see that the maximum shear stress can be quite large
compared to the average value. Consider an elementary part of a beam con-
tained between the cross sections (x) and (x + dx) and isolate the material
contained between the parallel planes (y) and (y = Ymax), Fig. 3.16. Equi-
librium of forces in the (x)-direction gives (the faces situated on the lateral
surface of the beam are assumed to be stress free):

1( S·
O"xx + 00")
dx
!.l xx
uX
dy* dz -
1

O"xx dy* dz - (dx)
jb(y)/2
-b(y)/2
O"yx dz =O
3.6 Computation of normal and shear stresses 65

This equation simplifies to:

j â(J
â
XX
dy*dz = Jb(Y)/2
(Jyx dz (3.34)
S' X -b(y)/2

Since (J xx = M (x)y / 1, then -assuming the moment of inertia 1 does not


depend on (x)- we obtain:

â(Jxx = ~ dM = ~Q(x)
âx 1 dx 1 '
where we used Eq. (3.17c). We now assume that (Jyx does not depend on the
width coordinate (z). Consequently, Eq. (3.34) becomes:

(Jxy
Q(x)
= Ib(y)
r
ls' y
* d *d
y z (3.35)

The integral in the right-hand-side of the equation is the first moment of the
planar surface S* w.r.t. the (Gz) axis. We now apply formula (3.34) to a

p(x)

S*
!1lLl !1lLl
2 2
dx
y
Fig. 3.16. Equilibrium of a slice of material contained between the cross sections
(x) and (x + dx) and the parallel planes (y) and (y = Ymax).

rectangular cross section of height hand width b, for which we have:

Equation (3.35) then gives the shear stress as:

(3.36)
66 3. Theory of beams (strength of materials)

where A = bh is the area of (8). This parabolic stress profile is plotted in


Fig. 3.17. The maximum shear stress is reached at (y = O) (the centroid of
the cross section) and its value is 1.5 times that of the average shear stress:

Exercise: compute the maximum shear stress in a circular cross section.

Fig. 3.17. Shear stresses along the height of a rect angular cross section: parabolic
profile.

3.7 Statically determinate ar indeterminate prablems

Statics give three scalar equations. If these equations are sufficient to compute
the reaction forces and internalloads M(x), Q(x) and N(x) in every cross
section, then the problem is said to be statically determinate. In this case,
the internal loads and the stresses are independent of material properties.
Examples are given in Sec. 3.11.
When the three scalar equilibrium equations do not allow the computation
of reaction forces and internalloads in every cross section, then the structure
is statically indeterminate.
The problem is solved by a two-step procedure. First, a number of dis-
placement B.Cs. are relaxed so that a statically determinate problem is oh-
tained. Next, the unknown reactions are computed by imposing the B.Cs. of
the original problem. This step involves the use of constitutive equations.
Within this general framework, several specific methods have been de-
veloped. A first method is based on a direct computation of the displace-
ment field. Two other methods use the theorems of Maxwell-Betti or Cas-
tigliano. Several examples are given in Secs. 3.10 and 3.11. More methods
(e.g., forces, displacements, Hardy-Cross, three-moment formulae for contin-
uous beams) are discussed in several textbooks, e.g. (Hibbeler, 1997), (French,
1995) (Roux, 1995), (Paduart et al., 1984), (Courbon, 1971), (Przemieniecki,
1968) and ('fuma, 1988).
3.8 Strain energy 67

3.8 Strain energy


The strain energy per unit volume is given by (Sect. 1.12):
1 II+v v l+v v 2
W = "2l'.ijUji = "2(~Uij - EUmmOij)Uji = ""2EUijUji - 2E(Umm )

For the beams studied in this chapter, it is assumed that the dominant stresses
are U xx and u xy and all other stresses vanish or can be neglected in front of
these two. Consequently, the strain energy simplifies to:
2 2
_ uxx ~
w - 2E + 2p.'

where ţt is the shear modulus. Using the results of Sect. 3.6, we can re-write
the strain energy as:

_...!...
w - 2E
[M(X) N(x)]2
1 Y+ A
~
+ 2p.
[Q(X)
Ib(y)
r
1S' y
* d *d ]2
y z (3.37)

The strain energy of the entire beam is obtained by integrat ion of the expres-
sion of W over the volume:

w= fa wdxdydz (3.38)

The integral of the first term in Eq. (3.37) is given by:

r M2EJ2
2(x) r r r
-------r
2 N2(X)
1.c dx ls Y dydz+ 1.c 2EA2 dx ls dydz
1 '--"'"
A

+ r M(x)N(x)
1.c EIA
dx y dydz,
ls '--v-"
AYG=O

where YG is zero because the cross section frame is (G, Y, z). It is seen that
the contributions of the bending moment M(x) and the normal force N(x)
to the strain energy are uncoupled. The integral related to the shear force
Q(x) -the second term in Eq. (3.37)- can be put under the following form:

r Q2(X) dx
1.c 2p.AI '
where Al is called the shear reduced area. An example is given in Sec. 3.11.
In summary, the strain energy of a beam is given by:

w = r [M 2(X) + N 2
(x) + Q2(X)] dx (3.39)
1.c 2EI 2EA 2ţtAI
68 3. Theory of beams (strength of materials)

Usually, the contribution due to the shear force Q(x) is neglected, and very
often only the energy term corresponding to the bending moment M(x) is kept
in (3.39). If the displacements Ui of the external forces Fi and the rotations
Oj ofthe external moments Mj are known, then the strain energy is also given
by the following formula:

(3.40)

If there are distributed loads, then (3.40) is easily generalized to include

l
integrals such as:
~ p(x)v(x) dx

3.9 Work and energy theorems

We recall some theorems from Chap. 2. Applications are given in Secs. 3.10
and 3.11. Maxwell-Betti reciprocity theorem reads:

Work of loading (1) in the displacements due to loading (2) =


Work of loading (2) in the displacements due to loading (1) (3.41)

A major advantage of the theorem is that the solution of a given problem can
be found by solving another, simpler problem. Castigliano 's theorem gives
the displacement Uk under a concentrated force Fk and in the direction of
the force:

(3.42)

where W is the strain energy of the body. The theorem is useful when we
are interested in computing not the displacement field but only the displace-
ment of a given pointj it also allows to compute reaction forces for statically
indeterminate problems. The theorem is also applicable for concentrated ma-
ments, and gives the rotation Ok under a concentrated moment Mk in the
direction of the moment:

(3.43)

The virtual work theorem (VWT) reads:

[The work of internalloads in the deformations due to a virtual


displacement field v*(x)] = [The work of externalloads in the
virtual displacement field v*(x)], (3.44)
3.10 In:B.uence lines 69

for alI v* which are "sufficiently smooth" and vanish where displacement
B.Cs. are imposed. For beams studied in this chapter, the pairs of dual vari-
ables (internal load, deformation) are the folIowing:
<Fv· M*(x)
M(x) and - dx 2 ="EI (bending moment and curvature),
du* N*(x)
N(x) and dx = EA (normal force and axial strain),
Q*(x)
Q(x) and - A (shear force and shear strain)
ţt l

Consequently, the left-hand-side of (3.44) is given (up to a sign) by:

f [M(X) M*(x) + N(x) N*(x) + Q(x) Q*(X)] dx (3.45)


le EI EA ţtAI

3.10 InHuence lines


Consider the staticalIy indeterminate beam of Fig. 3.18, it is built-in at its
end A (x = O), simply supported at the other end B (x = l) and subjected
to a point force (Pe y) at (x = a). Static equilibrium gives:

HA = O, VA + VB = P, MA - Pa + VBl = O
We choose as unknown reaction VB and define a staticalIy determinate struc-
ture by freeing the simple support at B. The bending moment is given by:

O~ x ~ a: M(x) = -P(a - x) + VB(l- x)j a ~ x ~ l: M(x) = VB(l- x)


Castigliano's theorem gives the deflection at B as:

8W = fi M(x) 8M(x) d = _ Pa2 (l _ ~) VBl3


8VB lo EI 8VB x 2EI 3 + 3EI
Equivalence with the original, statically indeterminate problem, is obtained
by requiring this deflection to be zerOj this gives the reaction at B:

P
VB = -(-) a
2 (
3-- a) (3.46)
2 1 1
Figure 3.19 shows VB/P as a function of a/l, the curve is called the inftuence
line of the reaction force VB. For example, if the beam represents a bridge,
then P can be viewed as a mov ing load (that of a vehicle), and one is inter-
ested in finding the most dangerous position of P w.r.t. VB, MA or v(l/2) for
instance. Another use for influence lines is given hereafter.
Consider the same beam as before but with another loading plotted in Fig.
3.20: a uniformly distributed load q [N/m] between (x = b) and (x = 1 - b).
70 3. Theory of beams (strength of materials)

p
A

a l-a

Fig. 3.18. A beam is built-in at its end A (x = O), simply supported at the other
end B (x = 1) and subjected to a point force (Pe ll ) at (x = a).

1.0
0.8
fu 0.6
p
0.4
0.2
O
O 0.5 1.0

aii
Fig. 3.19. The curve VB/P versus a/l is called the influence line of the reaction
force VB.

The reaction VB at (x = 1) is simply obtained from Eq. (3.46) by substitution


of P with (q da) and integrat ion of a from b to (1 - b):

As an application, for a beam uniformly loaded along alI its length (i.e.,
b = O), the reactions are given by:

TTB = 3ql, VA = 5ql, ql2 (3.48)


Vi 8 8 MA=8

We have illustrated the concept of influence lines with an example, for a more
general discussion of this important topic, see other books, e.g. (Courbon,
1971) or (Hibbeler, 1997).

3.11 Solved problems


3.11.1 Shear reduced area

From Secs. 3.6 and 3.8, the energy contribution WQ ofthe shear force Q(x)
for a rectangular cross section of height hand width b is:
3.11 Solved problems 71

Fig. 3.20. A beam is built-in at (x = O), simply supported at (x = l) and uniformly


loaded between (x = b) and (x = l - b).

r ~ [6Q(X)(~ _ y2)]2 dxdydz


1il 2ţL bh 4 h2
r_18_Q:.. .2. .;. .(x.. :. .) dx jh/2 (~_ y2)2 dy
le ţLbh 2 -h/2 4 h 2
r ~ Q2(X) dx
le 5 ţLA
This energy contribution can be rewritten as:
5
with Al = -A (3.49)
6
Exercise: show that the shear reduced area for a circular cross section is
Al = 9A/IO.

3.11.2 Statically determinate problems

As a first example, consider a uniformly loaded beam which is pinned at


one end (x = O) and simply supported at the other (x = 1), Fig. 3.21. The
method of fictitious cuts gives the shear force and bending moment as:
1 q
Q(x) = q("2 - x), M(x) = "2x(l- x) (3.50)

We do have Q(x) = dM/dx. The diagrams of Q(x) and M(x) are plotted in
Fig. 3.21. It is seen that M(x) is positive everywhere, meaning that aH the
upper part (y < O) of the beam is compressed, and aH the lower part (y > O)
is in tension, and this is physicaHy expected. If the beam were in concrete
(which resists very poorly to tensile stresses), then steel reinforcement must
be added along the lower part of the beam, otherwise cracks will appear in
that region. The bending moment is maximum in the middle of the beam
(x = 1/2) and has the following value:

The beam's deflection is solution of the following differential equation to-


gether with its B.Cs.:
72 3. Theory of beams (strength of materials)

d2 v
-EI dx 2 = M(x); v(O) = O, v(l) =O
Simple integrat ion gives the deflection as:

v(x) = ~(x3
24EI
+ l3 - 2lx2 ) (3.51)

The maximum deflection occurs at (x = l/2) and:


5ql4
V maz = 384EI

O
--~
..
x

ql/2
"Q"
ql/2

"M"

Fig. 3.21. Uniformly loaded beam pinned at one end and simply supported at the
other. Diagrams of shear force and bending moment.

As a second application, consider the so-called three-point bending


problem of Fig. 3.22: a beam pinned at (x = O) and simply supported at
(x = 1) is subjected to a concentrated force (Pey) at (x = ~). Statics give
the vertical reaction forces at the left and right ends as follows:

The bending moment is given by:

The moment is continuous at x = ~:


3.11 Solved problems 73

The shear force however presents a discontinuity at (x = ~) because there is


a concentrated force in the y-direction at that location:

x<~: Q(x) = VA; x>~: Q(x)=-VB

The discontinuity is equal to the applied force:

For the beam's deflection, we need to distinguish two regions and solve the
following differential equations:

d2vi ~V2
x ~ ~: -EI dx 2 = VA x, X ~ ~: -EI dx 2 = VB(I - x),

together with the following boundary and continuity conditions:

Vi °=
( ) 0, V2 (I) = 0, Vi
(~ ) = V2 () dVi (C) __ dV2 (C)
~, dx <, dx <,

The final expressions are given below:

x<~· PI 3 ~ X [~-(2--)-(-)
v(x)=-(l--)- ~ X 2] .
_. 6EI III Il'

x ~ C v(x) PI 3 ~ X [e
x ]
= 6EIT(1- T) (1- [2) - (1- T)2 (3.52)

It is easily checked that the deflection and the slope are indeed continuous
at (x = ~) and their values are:

PI3 [~ ~ ]2 dv Pl2 ~ ~ ~ (3.53)


v(~) = 3EI T(l- T) ; dx(~) = 3EIT(1- T)(l- 2T )

Consider the case (~ = 1/2), i.e. the load is applied at the middle of the
beam (this is one of the tests which are used to determine the strength stress
of brittle materials such as ceramics and glass). The deflection and bending
moment are maximum at (x = l/2) and take the following values:

Pl 3 PI
V max = 48EI; M max =4
Compar ing the two cases where the same total load P is either: (i) concen-
trated at (x = l/2), (ii) uniformly distributed (q = P/l), it is seen that the
former leads to much higher values of V max and M max :
(P)
V max 8 M$!,)x = 2
-v(q)
max
5' M(q)
max
74 3. Theory of beams (strength of materials)

...
x
o
--

"Q" y :1111111111' 111111111111 ~ 11111111111111

Fig. 3.22. Three-point bending: a concentrated force is applied to a beam pinned


at one end and simply supported at the other. Shear force diagram.

A third example, a uniformly loaded cantilever beam, is depicted in


Fig. 3.23. The shear force and bending moment are:

Q(x) = q(l - x), M(x) = -~(l - X)2i (3.54)

It is checked that Q(x) = dM/dx. The diagrams ofQ(x) and M(x) are plotted
in Fig. 3.23. It is seen that M(x) is negative everywhere, which implies that
-as expected- the upper part (y < O) of the beam is in tension, while the
lower part (y > O) is compressed, and this is physically expected. If the beam
were in concrete, then steel reinforcement must be added along the upper
part of the beam (this example may represent a balcony-type of structure).
The bending moment is maximum at the built-in section (x = O) and:

The beam's deflection is solution of the following differential equation to-


gether with its B.Cs.:

- ~V M()
EI dx2= XiV
O dV(O) = O
(O) ='dx

Simple integrat ion gives:

v(x)
q
= 6EI
[(l-x)4
4 -
14 3]
"4 + 1 x (3.55)

The maximum deflection is reached at the tip of the beam (x = l) and:


{j 11111I1111111r~
3.11 Solved problems 75

~
ql2/2. l
ql
"Q"
ql~

ql'/2 ~"'"
"M"
Fig. 3.23. Uniformly loaded cantilever beam. Diagrams of shear force and bending
moment.

As a fourth example, consider a cantilever beam subjected to a con-


centrated force (Pe y) applied at (x = ~), Fig. 3.24. The shear force is given
by:
x < ~: Q(x) = P; x > ~: Q(x) = O
It presents a discontinuity at (x = ~), equal to the applied force:
Q(C) - Q(~+) =P
The bending moment diagram is plotted in Fig. 3.24, we have:
x :::; ~: M(x) = -P(~ - x); x ~ ~: M(x) =O
For x :::; ~, the upper (y < O) and lower (y > O) parts of the beam are in
tension and in compression, respectively. For x ~ ~, the beam is not stressed.
For the beam's deflection, we need to distinguish two regions and solve the
following differential equations:
d2v2
dx 2 = -P(~ - x), x ~ ~ : _El dx 2 -- O'
X :::; ~: -EI d2VI

together with the following boundary and continuity conditions:

VI (O) = O, dVI ( )
d.,x O = O, (~)
VI." = V2 (~)
.",
dVI (~)
dx'"
= dV2
dx'"
(~)

The final expressions are given below:


76 3. Theory of beams (strength of materials)

The deflection and its slope are indeed continuous at x =~, with:

The maximum deflection occurs at the tip (x = l) -which is predictable-


P 2 ~
V max = 2EI~ (l-'3)

Note that solution of point load problems gives that of distributed loads. This
is a consequence of the superposition "principle". For example, in order to
find the deflection v(l) at the tip of the uniformly loaded cantilever beam of
Fig. 3.23, we substitute (P) with (qd~) and integrate ~ from O to l:

v(l) = [1 (q~) e(l-~) = ql4


Jo 2EI 3 8EI'
which is what we found by direct computation.

P
A4------------~~

~ (J.A------.l....-------,I~
P~
P
"Q"
PllllllllllllOOlllllllllll1
p~~ "M"
Fig. 3.24. A cantilever beam is subjected to a concentrated force. Diagrams of
shear force and bending moment.

A fifth problem is depicted in Fig. 3.25: a beam pinned at one end (x =


O) and simply supported at the other (x = l) is subjected to a concentrated
moment (-MBe z ) at (x = l). The internalloads are:

MB
Q(x) =-l
, M(x) = MB-Xl
The diagrams are plotted in Fig. 3.25. The deflection is solution of the fol-
lowing differential equation together with its B.Cs.
3.11 Solved problems 77

d2 v x
-EI dx 2 = MWT; v(O) = O, v(l) =O
The solution is given by:

v(x) = MBI [ X 2]
6El x 1- (T) (3.57)

The maximum deflection is given by:

"Q"
1IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIr MB 1
/

"M"

Fig. 3.25. A beam pinned at one end (x = O) and simply supported at the other
(x = 1) is subjected to a concentrated moment at (x = 1). Diagrams of shear force
and bending moment.

3.11.3 Statically indeterminate problems

As an application ofthe direct method (Sec. 3.7), consider a beam clamped at


both ends and uniformly loaded, Fig. 3.26. There are six reactions (HA, VA,
MA at x = O and HB' VB, MB at x = l) while statics give three equations:
l
HA = HB, VA + VB = ql, MA - MB = (VA - VB)"2
A statically determinate problem can be defined by "liberating" the built-in
link at (x = l) and considering HB, VB and MB as given externalloads. This
78 3. Theory of beams (strength of materials)

problem can be expressed as the sum of four simple problems by virtue of


the superposition principle, see Fig. 3.26. The cantilever beam is uniformly
loaded in problem (1), subjected to a force (- VBetl) at (x = 1) in problem (2),
a moment (MBe z ) at (x = 1) in problem (3) and a force (-HBe x ) at (x = 1)
in problem (4). Most of those problems were solved already. The bending
moment diagrams are plotted in Fig. 3.26 and the displacement fields are:

Now HB, VB and M B are computed by requiring the new problem (which
is the sum of problems (1) to (4)) to be equivalent to the original one, Le.
built-in B.Cs. at (x = 1) need to be imposed. Consequently, the following
three conditions are written:

These conditions give the three equations:

Consequently, the reactions in the original problem are given by:

(3.58)

Results (3.58a, b) are due to the symmetry. The complete solution of the
original problem, Le. internalloads M(x) and Q(x) and deflection v(x) can
be obtained by summing the corresponding solutions of problems (1) to (3).
For instance, it is found that M(I/2) = qI2/24.
Often, the definition of a statically determinate structure is not 1.mique.
For example, for the beam of Fig. 3.26, we may choose to relax the rotations
at (x = O) and (x = 1) and the axial displacement at (x = 1). We thus obtain a
pinned B.C. at (x = O) to which a moment (- MAe z ) is applied, and a simple
support at (x = 1) subjected to a force (-HBe x ) and a moment (MBe z ); see
Fig. 3.27. Using superposition, this problem is decomposed into four simpler
problems which are plotted in Fig. 3.27. Most of those problems were solved
already and their displacements are given hereafter:

V2(X) = -6EI
MA
1 2 (1- -)
X
1
[ X 2
(1- -)
1
-1] .
'
U4(X) = --HB
-x
EA
3.11 Solved problems 79

q12/2 (

ql
+

+
~IIIIIIIIIIIIIIIII~IIIIIIIIIIIIIIIIIIIII)
(a "M3 (x)"
MB MB

+
yf-------~
Fig. 3.26. A statica1ly indeterminate problem: a uniformly loaded beam is clamped
at both ends. Built-in conditions at the (x = 1) end are relaxed and the original
problem is equivalent to the superposition of four statically determinate problems.
80 3. Theory of beams (strength of materials)

The new problem (sum of problems (1) to (4» must be such that the B.Cs. of
the original problem are recovered. This gives the following three conditions:

~ (O) + : (O) + ~ (O) = O, ~~ (1) + ~~ (1) + ~~ (1) = O, u4(1) = O,


which can be rewritten as follows, using the expressions of the displacements,
q12 Jk(11 q12 Jk()l
8" - Jk()l - -2- = O; -8" + -2- + Jk(11 = O; H l1 =O
This gives Jk()l = Jk(11 = q12/12, which is identical to the expres sion found
with the first method.

3.11.4 Maxwell-Betti reciprocity theorem

As a first application, consider a cantilever beam subjected to a distributed


force (qey) between (x = O) and (x = a) and compute the deflection at
(x = 1). In order to solve this problem (1), we introduce problem (2): a unit
force (ley) is applied at (x = 1); Fig. 3.28. The deflection fields of the two
problems are VI (x) and V2(X), respectively. The theorem gives:

la qV2(X) dx = 1 X Vl(l)
The only problem to be solved is (2), which is much simpler than (1). The
deflection of problem (2) is given by (3.56a) after setting ~ = 1 and P = 1,

(3.59)

which gives:

Vl(l) = 10
r qV2(X) dx = 2EI
q [ (1 - a)4
12
14 l2a2
+ 12 + -2- -"3
13a ]
(3.60)

For a uniformly loaded beam (a = l), we obtain: vl(l) = q14/(8EI), which is


what we found with two other methods. Actually, a remarkable fact is that
the solution of problem (2) allows to find directly the tip deflectian v(l) af a
cantilever beam subjected to any kind of loading.
A second application is illustrated in Fig. 3.29: a simply supparted
beam is subjected ta a farce per unit length (qe y) between (x = a) and
(x = 1- a), with a < l/2. The deflection at (x = 1/2) needs ta be computed.
This is prablem (1). We consider a second problem -(2)- in which a unit farce
(le y) is applied at (x = 1/2). The theorem gives:

1,-
a
a
qV2(X) dx = 1 X Vl(-)
1
2
3.11 Solved problems 81

Fig. 3.27. A second method for the beam clamped at both ends. Rotations at
(x = O) and (x = l) and rudal displacement at (x = l) are relaxed, thus creating a
statically determinate problem which is the superposition of four simpler problems.
82 3. Theory of beams (strength of materials)

1IIIIIr
vI(l) =?
q

a l-a
~

1
Fig. 3.28. Problem (1): distributed load between (x = O) and (x = a)j compute
Vi (l). In order to use the reciprocity theorem, we define a problem (2): unit force
at (x = l).

The only problem to be solved is (2), which is simpler than (1). Its solution
is given by (3.52) with P = 1 and ~ = l/2:

(3.61)

and for x ~ l/2, replace x with (l - x). The sought after result is:

r'-
la
a
QV2(X)
q
dx = 2EI
r/2
la' X
(l2 x2)
4" -"3 dx
Q l2 2)( 5 2 2
24EI(4" -a "4 l -a) (3.62)

In the particular case when a = O, Le. a uniform load along the beam, it is
easily checked that the result obtained from (3.51) is retrieved.
As a third application, consider the statically indeterminate problem
(1) of Fig. 3.30: a uniformly loaded beam is elastically built-in at (x = O)
and elastically supported at (x = l) (after (Friaa, 1982)). The two stiffnesses
are designated by K [Nm/rad] and k [N/m], respectively. Static equilibrium
gives the following relations between the reactions HA, VA and MA at (x = O)
and VB at (x = l):

As unknown reaction, we choose force VB at (x = l), and in order to compute


it with the reciprocity theorem, we consider a statically determinate problem
(2) in which the beam is elastically built-in at (x = O) and subjected to a
unit force (ley) at (x = l)- where the beam is no longer supported. The
reciprocity theorem gives the following equality:
3.11 Solved problems 83

o
--~
1/2 1/2
1/2 1/2
Fig. 3.29. Problem (1): for ce per unit length between (x = a) and (x = l-a),
compute Vl (l/2). In order to use the reciprocity theorem, we define problem (2):
unit force at (x = l/2).

which gives the reaction VE as:

(3.63)

It is seen that we only need to solve the statically determinate problem (2)
whose deflection obeys the following differential equation and its B.Cs.

d 2 v2 dV2 1 xl
-EI dx 2 = -1 x (l - x), V2(0) = O, dx (O) =K
The solution is given by:

Elv2(x) =
(l-X)3 I3
- - + (I- + - EI) Ix
6 6 2 K
Computing the integral in Eq. (3.63) and rearranging terms, we obtain:

(3.64)

If the beam is built-in at (x = O), i.e. K -+ 00, and simply supported at


(x = I), i.e. k -+ 00, then the reaction at (x = l) is simply:
84 3. Theory of beams (strength of materials)

prob.(l)
x

~------------------------~ VB

l~l~ prob.(2)

1
Fig. 3.30. Problem (1) (statically indeterminate): a uniformly loaded beam is elas-
tically built-in at (x = O) and elastically supported at (x = l). With the reciprocity
theorem, we only solve problem (2) (statically determinate): unit force at (x = 1)
and no support there.

As a fourth application, consider the continuous beam of Fig. 3.31 with


a pinned end at A (x = O), an elastic support of stiffness k [N /m] at B (x = l)
and a simple support at C (x = 2l). Uniform loads ql and q2 are applied to
the first and second spans, respectively. The reaction force VB at B is chosen
as the unknown in this statically indeterminate problem (1). We now consider
another problem (2) which is statically determinate: pinned end at A, simple
support at C and a unit force at B (where the B.C. is freed); see Fig. 3.3l.
The reciprocity theorem gives:

11 qlV2(X) dx + 1 q2V2(X) dx - VB V2(l)


21
= 1 x V:

This gives the reaction force at B as follows:

ql I~ V2(X) dx + q2 It V2(X) dx
VB = 1 (3.65)
V2(l)+k"
It is seen that we only need to know the defiections of problem (2), and these
are given by Eqs. (3.52) with the substitutions: P -+ 1, l -+ 2l and ~ -+ l.
Computation of the integrals leads to the following expression:
li: _ 5(ql + q2)l
B - 8 + 48EI (3.66)
kl 3

The other reactions forces in the original problem (1) are given by statics as:
VB l VB l
HA = O; VA = -""2 + (3ql + q2)4; Ve = -""2 + (ql + 3q2)4
3.11 Solved problems 85

In the particular case of a rigid simple support at B (Le., k -+ 00) and a


single loaded span (q2 = 0, ql = q), the bending moments M(x) and the
deflection v(x) are plotted in Fig. 3.32. As noted in Sect. 3.3.5, the sign of
M(x) gives the curvature of the deflection and an inflection point for v(x) is
obtained when M(x) changes signs and goes through zero.

Fig. 3.31. Problem (1) (statically indeterminate): pinned end A (x = O), elastic
support at B (x = l) and simple support at C (x = 21). Uniform loads ql and q2
on the first and second spans, respectively. Problem (2) (statically determinate):
pinned end at A, simple support at C, unit force at B and no support there.

3.11.5 Castigliano's theorem

As a first application of Castigliano's theorem, consider a cantilever beam


which is subjected to a force (Pey) and a moment (-Mez)at its end (x = l).
We wish to compute the deflection and rotation at the end section. The
bending moment is given by:

M(x) = -P(l- x) + M
The deflection at (x = l) is given by the theorem as follows:
âW fi M(x) âM(x) 1 fi -
v(l) = âP = io EI8P dx = EI io [-P(l- x) + M][-(l- x)] dx,
where only the contribution of M(x) to W was kept. Computation gives:

l2 Pl M
v(l) = -(-
EI 3
--)
2
(3.67)

The rotation at (x = l) is given by:


86 3. Theory of beams (strength of materials)

"M(x)"

I
"v(x)"

~
Fig. 3.32. Continuous beam on three supports with one span uniformly loaded.
Bending moment diagram M(x) and defiection curve v(x). M(x) changes signs and
goes through zero at x = Xo, which gives an in:ftection point of v(x) at x = Xo
(designated by a dot in the figure).

O(l) = âW = f' M(x) âM(x) dx = ...!....1'[-P(l- x) + M] dx


aM 10 EI aM EI o
The computation of the integral gives:
1 Pl -
O(l) = -(--
EI 2
+ M) (3.68)

We can double-check this result against the defiection which was found in
problems (2) and (3) of Fig. 3.26. Superposition gives:

v(x) = _1_ [P (l- X)3 _ P l3 + Pl2X - Mx 2]


2EI 3 3
Computing dv/dx and setting (x = l) gives dv/dx(l) = -O(l).
As a second application, consider the uniformly loaded beam of Fig.
3.34 and compute the maximum deflection, Le. at (x = l/2). In order to apply
Castigliano's method, we add a fictitious force (Pey) at (x = l/2) which we
shall reset to zero at the end. The bending moment for the new problem is:
x x2 1
M(x) = (ql + P)'2 - q2"' for x ~ 2'
and for x ~ l/2, it suffices to replace x with (l - x). The theorem gives:
3.11 Solved problems 87

Fig. 3.33. Cantilever beam subjected to a force and a moment at (x = 1). Compu-
tation of the deflection and rotation at that location with Castigliano's theorem.

v( i)
2
âW
âP
= 11 M(x)
o
âM(x)
EI âP
dx

~
EI o
11
/2 [(ql + P)~ _
2
qX 2 ]
2
~ dx
2 '

where the factor 2 is due to the symmetry. After computing the integral and
setting P = O, we retrieve what we found already with other methods:

ti I I III ~ 111111 ~ V~I/2) =1


P

O
-- .. 1/2
ql/2 +P/2
Fig. 3.34. In order to compute the deflection at the middle of a uniformly loaded
beam using Castigliano's theorem, we add a fictitious force P at (x = 1/2) which
is reset to zero at the end of the computation.

As a third application, consider the frame of Fig. 3.35 which contains


two vertical beams (properties: h, Eh, h) pinned at A and D and a horizontal
one (I, EI, It) subjected to a uniform load q and a concentrated force P on its
middle F. The beam junctions at B and Care such that relative rotations are
possible and are modeled with rotational springs of stiffness K [N m/rad]. The
whole (statically indeterminate) frame is assumed to behave as a continuous
structure. Statics give the following relations:
88 3. Theory of beams (strength of materials)

1 12
HA = Hv, VA + Vv = ql + P, Vv 1 = P"2 + q2'

from which it is deduced that:


ql+P
VA=VV=--,
2
and the horizontal reactions are equal but unknown. We choose Hv as the
unknown reaction, the horizontal displacement of point D is then "freed",
Le. we consider a statically determinate structure in which we have a simple
support at D with an applied horizontal force Hv. The strain energy is:

(3.69)

The multiplying factor 2 is implied by the symmetry. A local frame (G, x, y)


is defined for each vertical or horizontal beam. The bending moment diagram
is plotted in Fig. 3.36. Along (AB) and (BF), we have:

x E [XA = O, XB = h]: M(x) = -Hvx ~ 8M(x) = -x


8Hv
l qx2 8M(x)
x E [XB = O, XF = "2]: M(x) = -Hv h + VAX - 2"" ~ 8 H v =-h

At the elastic junction B, the bending moment is:


8MB
M B = -Hvh ~ - - = -h
8Hv
The horizontal displacement vv of D is given by the theorem as follows:

vv = _1_ fh M(x) 8M(x) dx + _1_ f l/2 M(x) 8M(x) dx + M B 8MB


2 Eh1h 10 8Hv Etil 10 8Hv K 8Hv
In order for the statically determinate problem (simple support in D sub-
jected to a horizontal force Hv) to be equivalent to the original, statically
indeterminate problem (pinned B.C. at D), we need to set vv = O. After
computing the integrals and rearranging terms, we obtain:

(3.70)

The moments at B, C and F are given by:

(3.71)
3.11 Solved problems 89

p
q

B
El, Il

l/2 l/2

A D

Fig. 3.35. A statically indeterminate frame (pinned at both ends A and D). Rel-
ative rotations at the junction sections B and C between the vertical beams and
the horizontal one are possible.

Fig. 3.36. Bending moment diagram for the statically indeterminate frame of Fig.
3.35.
90 3. Theory of beams (strength of materials)

If the links at B and Care such that no relative rotations are possible, then:
K -4 00, and the horizontal reactions are given by:

(3.72)

Care should be taken in applying Castigliano's theorem when two loads


(or more) are designated by the same symbol. We have to keep in mind
that the derivation of the theorem assumes that a variation of a given load
occurs independently of the other loads. Consider the example of Fig. 3.37 and
compute the deflection under point A. The correct result may be obtained by
applying the superposition principle to the problem of Fig. 3.22 with ~ = a
and ~ = l - aj Eqs. (3.52) then give:

VA = Pa 2 (~_ 2a) (3.73)


EI 2 3
In order to obtain this correct result with Castigliano's method, we need
to first designate one of the forces with another symbol, e.g. F, apply the
theorem and then reset F = P.

P P

A B
a a

Fig. 3.37. In order to compute the deflection of point A with Castigliano's theorem,
we need to designate one of the forces by another symbol, e.g. F, apply the theorem
and then reset F = P.

3.11.6 Virtual work theorem (VWT)


As an application, let us consider the structure of Fig. 3.38 and compute
the horizontal displacement LlA of point A (after Friaa(1982)). We define two
problems: (O), a horizontal unit force is applied in point A, and (1), the initial
problem (see Fig. 3.39). We now apply the VWT to problem (O) taking as
virtual displacements those of problem (1):

l (Mo ~; + No:~ + Qo JL~l ) ds = 1 x LlA (3.74)

The left-hand side of (3.74) corresponds to the work of internalloads -problem


(0)- in the deformations due to a virtual displacement -that of problem (1).
3.11 Solved problems 91

The diagrams of bending moments and normal forces are shown in Fig. 3.40.
Neglecting the term due to shearing in the left-hand side of (3.74), we obtain:

h 1 Pah 2
.1 A = (1 x h)-(Pa)-
2 EI
=- -
2EI
(3.75)

p
A
a b I

T7T------------~

Fig. 3.38. The horizontal part of the frame is loaded by a vertical force at its left
end. Compute the horizontal displacement of the right end (A).
92 3. Theory of beams (strength of materials)

(O): problem (1): virtual displacements

p
A 1
----------~-----------~

Solution: M o, N o, Qo Solution: MI, NI, QI

Fig. 3.39. Use of the virtual work theorem. Problem (O): a unit horizontal force is
applied to point A. Problem (1): the original structure of Fig. 3.38.
3.11 Solved problems 93

Pa
1
---------~

"Mo"

1· h Pa
P P
..-l

"No"

///

Fig. 3.40. Bending moment and normal force diagrams for problems (O) and (1)
of Fig. 3.39.
4. Torsion of beams

In Chap. 3, we studied beams with a symmetry plane subjected to loadings


in that plane. In this chapter we consider another type of loading: torsion. A
straight beam ofaxis (x) and length l is subjected to a torque (Me x ) at its
end (x = l), while the y and z displacements of the other end section (x = O)
are prevented (there is a reaction moment (-Me x ) in the (x = O) section).

4.1 Formulation with a warping function

The beam has a uniform cross section (S) which is delimited by a closed
curve (C), see Fig. 4.1. It is assumed that (S) is simply connected. Hollow
beams (multi-connected sections) will be studied in Secs. 4.7-11.
We assume that the displacement of a material point is the sum of an
infinitesimal rotation in the (y, z) plane and an axial displacement:

u(y,z,x) r(dO)eo + ue x
r(dO)[- sin(O)e y + cos(O)e z ] + ue x
z y
r(dB)( --e y + -e z )+ ue x
r r
-z(dO)e y + y(dO)e z + ue x

Next, we make two assumptions:

(i) The infinitesimal rotation (dO) is proportional to the axial distance (x) of
a cross section from the (x = O) section, i.e. dO = ax, where the constant
a [Iim] is an angle of torsion per unit length.
(ii) The axial displacement (u) is independent of (x), Le. u = aifJ(y,z), where
ifJ(y, z) is a warping junction to be determined.
According to the assumptions which were made, the components of the dis-
placement vector in the Cartesian basis (ex,ey,e z ) are given by:

u = aifJ(y,z); v = -azx; w = ayx (4.1)

The strains are computed by simple differentiation of Eqs. (4.1),


96 4. Torsion of beams

M
z

C
y

Fig. 4.1. Straight beam under torsion: (a) Torque (M), (b) Cross section (S),
boundary (C), outward unit normal (n) and other notations

(4.2)

The stresses are given by isotropic linear elasticity as:

O
[a] = [ a xy O
a xy a xz
O
1; -a xy âA..
'1'
= -z + -â ; -
a xz âA..
= y + -â
'1'
(4.3)
O O ţ.to yţ.to z
a xz

where ţ.t is the shear modulus. Compare with Chap. 3 where the only stresses
were a xx and a xy • Among the three scalar in-bulk equilibrium equations
(without body force), only the following is not trivially satisfied:

âaxy + âaxz =O
ây âz
This requires -using Eqs. (4.3)- r/J(y,z) to be harmonic:
â2 â2
(â y2 + âz 2 )r/J(y,z) =0, Le Llr/J(y,z) =0, in S (4.4)

The boundary C = âS is stress free, thus (n being the outward unit normal
to C):
4.1 Formulation with a warping function 97

[ O
uxy O
u
xy
O
U
xz
1 . [Ony 1= [ 1 O
Uxz O O nz O
This gives a single condition: Fx == uyxny+uzxnz = O. Using (4.3), we obtain:
Fx afjJ afjJ afjJ afjJ
-JLa = (-z + -ay )ny + (y + -az )nz = (-ay n y + -az nz) -zny +ynz
,
~
.. .,

/In

We shall re-write the expres sion of Fx one more time. We designate by s the
curvilinear coordinate along C. The following relations (see Fig. 4.1b) are
easily found and usefuI:
dz dy
ny = -j n z =-- (4.5)
ds ds
Therefore,

Fx = afjJ _ (zdz +ydY ) = afjJ _ ~~(y2 +z2)


JLa an ds ds an 2 ds
This shows that the stress-free B.C. Fx = O on C is equivalent to:
afjJ 1 d 2 2
an = 2ds (y + z ), on C (4.6)

On the end section (x = 1), we shall write that the moment of the stress vector
is equal to the applied torque (M e x ). According to Saint-Venant's principIe,
the solution will be valid far from the ends (Le. in practice for x E [D, 1- D],
where D is the largest dimension of S). Since the stresses are independent of
x, the stress vector on any cross section of outward normal e x is:
(4.7)
We shall show in Sec. 4.3 that its resultant is nil. We are now concerned with
the moment of the stress vector w.r.t. the center of the cross section:
(ye y + ze z ) x (uyxe y + uzxe z ) = (yu zx - zUyx)e x
The resultant ofthis moment should be equal to (Mex)j this gives one scalar
equation:

M = !s(YUzx - zUyx ) dydz (4.8)

In summary, the problem is the following: find a function fjJ(y, z) which sat-
isfies Eqs. (4.4) in S and (4.6) on C. Equation (4.8) is used to find a ifthe
torque M is known (or M if a is given). The stress, strain and displacement
fields are computed from Eqs. (4.3), (4.2) and (4.1). One can easily prove
that the compatibility equations (Chap. 1) are trivially satisfied.
98 4. Torsion of beams

4.2 Formulation with a conjugate function


In this sectiQn, we give an alternative formulation of the torsion problem
using a conjugate function <p(y, z) defined by:
8<p
(4.9)
8y
The function thus defined is harmonic:

..::1<p(y, z) = O, in S (4.10)
This can be easily established:
8 84J 8 84J
..::1<p(y,z) = - ( - - ) + -(-) = O
8y 8z 8z 8y
On C = 8S, we have:

84J = 84J n + 84J n z = 8<p dz _ 8<p (_ dy ) = d<p


8n 8y Y 8z 8z ds 8y ds ds

Therefore, the stress-free B.e. (4.6) written for <p becomes:


d<p 1d 2 2
ds = "2 ds (y + z ), on C (4.11)

Since (S) is simply connected, simple integrat ion allows to rewrite (4.11) as:
1
<p(y,z) = "2(y2 +Z2) + constant, on C (4.12)

The constant can be set to zero without altering the stress or strain expres-
sions. Using Eqs. (4.3) and (4.9), the stresses are found as:
U zx 8<p u yx 8<p
-=--+Yi -=--z (4.13)
J.ta. 8y J.ta. 8z
In summary, the problem is: find a function <p(y, z) which satisfies Eqs. (4.10)
in S and (4.12) on C. Equation (4.8) is used to find a. ifthe torque M is known
(or M if a. is given). The stresses are computed from (4.13).

4.3 Formulation with Prandtl's stress function


In this section, we shall formulate the torsion problem with yet another func-
tion 1jJ(y, z) known as Pmndtl's stress function. It is defined by:

(4.14)
4.3 Formulation with Prandtl's stress function 99

We have t1.1/J(y, z) = t1.cp(y, z) - 2; equation (4.10) implies that:

t1.1/J(y, z) = -2, in S (4.15)

The stress-free B.e. on C -Eq. (4.12)- can be rewritten as:

1/J(y, z) = O, on C (4.16)

From Eqs. (4.13) and (4.14), the stresses are found to be:
(1 zz â1j; (1yz 81/J
-=--; -=- (4.17)
ţta 8y ţta 8z
Since the stresses are obtained by direct derivation of 1/J(y, z) and there are
no additional terms as in (4.13) or (4.3), 1/J(y,z) is called a stress function.
The stress vector resultant on a cross section of outward normal e z is:

ey fs (1yz dydz + e z fs (1zz dydz = ţta [e y fs ~~ dydz - ez fs ~~ dYdZ]


Integration by parts -as in (Germain and Muller, 1980)- shows that the re-
sultant is nil because 1/J(y,z) = O on (C). Equations (4.8) and (4.17) show
that M and a are related by:
M
_ [ (Y 81/J + z 81/J) dy dz
JLa ls 8y 8z
- [
ls [~(Y1/J)
8y
+ ~(z1/J) -
8z
21/J] dydz

-fs div (y1/Je y + z1/Je z ) dy dz + 2 fs 1/J dy dz

-la (y1/Je y + z1/Je z ) •n ds + 2 fs 1/J dy dz,


using the divergence theorem. Since 1/J = O on C, we obtain:

M = 2ţta fs 1/J dy dz (4.18)

The torsional rigidity is defined as (Mia), e.g. for a given M, the smaller the
angle of rotation a, the larger the rigidity (Mia). In summary, the problem
is: find a function 1/J(y,z) which satisfies the Dirichlet problem (4.15)-(4.16):

t1.1/J = -2 in S; 1/J =O on C

Equation (4.18) is used to find a if the torque M is known (or M if a is


given). The stresses are computed from (4.17). The axial displacement of the
cross section is given by (4.1a), but we need first to find (jJ(y,z). Using (4.9)
and (4.14), we obtain:
100 4. Torsion of beams

8ljJ 8'1' 8 y2 z2 81/;


- = - = -(1/;+-+-) = - +z;
8y 8z 8z 2 2 8z
8ljJ 8'1' 8 y2 z2 81/;
- = - - = --(1/;+-+-) = - - - y (4.19)
8z 8y 8y 2 2 8y

4.4 Contour lines of the stress function


In this section, we first establish some important properties of contour lines
of the stress function 1/;; interpretation will be given in the second part of
the section. We consider a contour line CI whose equation is 1/; = kl and
introduce the following notation: 8 1 , of area Al, is the surface contained
inside CI, SI is the curvilinear coordinate along CI and nI is the unit normal
pointing inside 8 1 , see Fig. 4.2. Along CI, we have d1/; = O, Le.
!li!..
81/; dy + 81/; dz = O ===} dz = _..!!JL = a zx ,
8y 8z dy!?lz ayx

using Eqs. (4.17). On the other hand, according to Eq. (4.7), the components
of the stress vector (T • e x on any cross section of outward normal e x are
ty = a yx and t z = a zx . It appears then that we have:

dz a zx tz
-=-=-, onCl (4.20)
dy ayx ty

The stress vector (T • nI along CI is given by:

using (4.5). On the other hand, Eq. (4.20) gives: a zx = ayx(dz/dy). Thus we
conclude that:

(T • nI = O, on CI (4.21)

We now compute the integral of .6.1/; over 8 1 . Firstly, we have:

lsr1.6.1/; dydz = -2 ls1


r dydz = -2A l

Secondly, we can also write:

lsr1.6.1/; dy dz = ls1
r div (V1/;) dy dz = lcr1(V1/;)· (-nddsl,
using the divergence theorem. Equating the two expressions of the integral,
we find the following equality:
4.5 Maximum tangential stress 101

(4.22)

We now interpret the three results (4.20-22) that we found. Equation (4.20)
means that a contour line (tf; = constant) of the stress function is tangent on
each of its points (y, z) to the stress vector t == (T . e", on that point; this is
illustrated in Fig. 4.2. This conclusion will also be used in Sect. 4.5 to find
the location of the maximum tangential stress.
Recalling what stress vector ((T·nl) along CI means (Sec. 1.8), Eq. (4.21)
gives the following important conclusion: if we remove the material contained
inside SI and delimited by CI (alI along the x-axis), then the stress state in
the rest of the beam is undisturbed. This conclusion will be used in Sect.
4.9 to solve the torsion problem of a particular class of hollow beams (Le.,
multi-connected sections) in a simple fashion.
Finally, it will be shown in Sect. 4.8 that Eq. (4.22) fits into a general
theory of hollow beams.

z
C:tf;=o

Fig. 4.2. Contour line CI, equation 1/J = kl. Notation: curvilinear coordinate 81,
normal nI, inside surface 81 with area Al. CI is tangent on each ofits points (y,z)
to the stress vector t == lT . e", on that point

4.5 Maximum tangential stress


We designate by r the tangential stress, Le. the magnitude of the stress vector
t= (T' e",. Using (4.7) and (4.17), we have:

(4.23)

We compute the Laplacian of r 2 :


102 4. Torsion of beams

using (4.15). Thus, we have: .1(7"2) ;::: O. Now consider an arbitrary, simply
connected sudace SI inside S; recalling that .1(7"2) = div ("97"2) and using
the divergence theorem, we find the following result:

where nI is a unit vector normal to (âSd and pointing inside SI (nI is


not to be confused with the outward normal ex to SI or S). By definition
of the gradient, ("97"2) gives the direction of maximum increase of (7"2). If
(7"2)max is reached inside SI, then ("97"2) will point towards the interior of
SI, and this contradicts the inequality found above. In summary, for a simply
connected cross section S, the maximum tangential stress is always reached
at the boundary C.

4.6 Membrane analogy

It is possible in some cases to find a good approximation of Prandtl's stress


function 'if;(y, z) by computing or just guessing the deflection of a membrane.
Consider again the torsion problem of a beam of uniform and simply con-
nected cross section S with boundary C. Now imagine a rigid plate in which
we make a hole whose boundary is a closed curve identical to C, and imagine
that a thin homogeneous membrane is stretched over the plate and fixed aH
along C. Suppose also that the membrane is subjected to uniform tension
per unit length T [N/m] along C and uniform transverse pressure p [N/m 2 J,
see Fig. 4.3. We assume that the deflection u(y, z) of the membrane (i.e., its
x-displacement field) is smaH and that the stresses due to pare negligible
compared to those due to T.
The tension vector T is always tangent to the deformed membrane. In
the (x, y) plane, T has the following directions along the sides (bc) and (da):
âu âu
(bc): e y + (-â) e x ; (da): -e y - (-â) ex
Y (be) Y (da)
Thus the resultant of the forces per unit area acting on those sides is:

âu
T(dz) [ (-â) âu
- (-â) 1 â2 u
e x ~ Tây2dydzex
y (be) Y (da)

Similarly, it is found that the resultant of the forces per unit area acting on
the sides (c d) and (a b) is approximately:
4.6 Membrane analogy 103

Fig. 4.3. Membrane under uniform tension per unit length T along C and uniform
transverse pressure p

l}2u
T âz2dydzex
The equilibrium equation of an element (dy, dz) is then:

( p+T â2u â 2u)


ây2 +T âz 2 dydz =O
Recalling that by construction the membrane is fixed along C, the problem
of the membrane deflection is: find u(y, z) such that:

L1u(y,z)=-; in S, u(y,z) =0 on C (4.24)


This problem is identical to that of torsion with a Prandtl stress function
1fJ(y, z). Since p and Tare constants, 1fJ(y, z) is proportional to u(y, z):
2T
1fJ(y,z) = -u(y,z)
p
The torsional rigidity is given by Eq. (4.18) as:
M=
- 4J1.Tl
- udydz
a: p s
Thus (MIa:) is proportional to the volume contained between the deformed
membrane and the plane of the membrane before deformation.
104 4. Torsion of beams

4.7 Multi-connected sections- Particular case

Consider a solid beam (Le., simply connected section) and let t/J(y, z) be the
Prandtl stress function of its tors ion problem. Now consider a hollow beam
whose (multi-connected) cross section has a single hole of equation t/J = kI,
Le. a contour line of t/J, the stress function of the solid beam. The solution of
the torsion problem of this particular class of hollow beams can be found in
a very simple manner. Indeed, according to Eq. (4.21) and the comments at
the end of Sect. 4.4, the stress state in the hollow beam is identical to that of
the solid beam. The torsional rigidity is reduced by the amount of material
which was removed from the solid beam. An application of these important
results is given in Sec. 4.12.

4.8 Multi-connected sections- General case

In this section, we consider a theory of torsion for hollow beams which is


much more general than that of Sec. 4.7. We mostly follow (Crochet, 1993)
and introduce the following notations (see Fig. 4.4): S designates the surface
of actual material (the dashed region in the figure), C is the outer boundary of
S, there are n voids, the boundary of each is a closed curve Ci (i = 1, ... , n),
Si is the curvilinear coordinate along Ci and ni is the unit normal pointing
inside Si, where Si, of area Ai is the surface enclosed inside Ci.

S
Fig. 4.4. Multi-connected section: notation.

When we repeat the analysis of Sec. 4.1 without assuming that the section
is simply connected, we arrive at the following formulation: find a function
4J(y, z) which satisfies the system of equations:
4.8 Multi-connected sections- General case 105

L1<jJ = 0, in S
8<jJ 18 2 2
8n = '2 8s (y + z), on C

8<jJ !~(y2 + z2),


on Ci (4.25)
8ni28si
The stresses, torque and displacements are derived as follows:
8<jJ 8<jJ
a zz = Ma(y + 8); a yz = Ma( -z + 8y)
M l (ya zz - zayz ) dydz

u a<jJ(y, z); v = -aZXj W = ayx (4.26)


We can also formulate the problem with a conjugate function <p(y, z) defined
as in Sec. 4.2 by: 8<pj8y = -8<jJj8z and 8<pj8z = 8<jJj8y. The problem
becomes: find a function <p(y, z) which satisfies the system of equations:
L1<p = 0, in S
1
<p(y,z) '2(y2 + z2) + k, onC
1 2
<p(y,z) '2(y + z2 ) + k i , on Ci (4.27)

The constant k can be set to zero, but the constants ki cannot; they have
to be computed in such a way that the axial displacement u(y, z) is single-
valued. Let Ci be one of the curves, if we make a complete tour of Ci until
we return to the starting point, the displacement of that point should be the
same at the beginning and at the end of the tour, Le. f Ci du = O. Let us
compute the integral:

f du = f
âU âu
(-dy + -dz)
8y 8z
=a1 â<jJ
(-dy + -dz)
8y
â<jJ
8z

=- f
Ci

f
Ci Ci
8<p 8<p 8<p 8<p
=a (-dy - -dz) a (-8 niz + -8 niy)dsi ,
Ci 8z 8y Ci z Y
using (4.5). Thus it seen that the constants k i are computed by solving the
following system of equations:

1 Ci
8<p
-8 dSi = 0,
ni
(no sum) (4.28)

The stresses, torque and displacements are obtained as follows:


8<p
a zz = Ma(y - ~~); a yz = Ma(-z + 8z)
M l (ya zz - zayz ) dy dz

u = a<jJ(y, z); v = -azx; W = ayx (4.29)


106 4. Torsion of beams

Finally, the torsion problem can be formulated with a Prandtl stress function
defined as in Sec. 4.3 by 1f;(y, z) = <p(y, z) - (y2 +z2)/2. The problem becomes
now: find a function 1f;(y, z) which satisfies the system of equations:

il1f; = -2, in S
1f;(y,z) O, on C
1f;(y,z) ki, on Ci (4.30)

The constants k i are computed in such a way that the axial displacement
u(y,z) is single-valued. Let us rewrite the integrand in Eq. (4.28):

â<p = -â1/J +-V(y


-âni 1 2 +z2) ·ni = -â1f; + (ye +ze z ) ·ni
âni 2 âni Y

Thus the integral in (4.28) can be rewritten as:

[
Ci
â<p dSi

ni
= lâ1f;
-â dSi -
Ci ni

Si
dlV (ye y + ze z ) dy dz
= r âni
lCi
â1f; dSi - 2 r dydz,
lSi
(4.31)
'--v--"
Ai

using the divergence theorem. It appears then that the constants k i are found
from the following system of equations:

1 an .
Ci
81f;
'
dSi = 2Ai, (no sum) (4.32)

Note that for the particular case of a single void (n = 1) whose boundary CI
obeys the condition stated in Sect. 4.7, Eq. (4.32) was already derived by a
simpler method (Eq. (4.22)). Once 1f; is found, the stresses and displacements
are computed from the following relations:
â1f; â1f;
= -p,a ây; a yz = p,a âz
u arjJ(y,z); v = -azx; w = ayx
ârjJ â1f; ârjJ â1f;
ây = âz + z; âz = - ây - y; (4.33)

When we repeat the derivations of Sect. 4.3, we arrive at the following ex-
pression for the torsional rigidity M / a:
4.9 Thin tubes with variable thickness 107

M
J.l.o
= 2 r 'l/JdY dz-lc~
1s
('l/J)(yey+zez)'nds

-t ii ~(yey
o

+ ze z) . (-ni) dS i
k,

= 21 'l/J dydz - t ki 1 div (ye y + zez) dydz,


S i=l , Si ~
....
2Ai

using the divergence theorem. Thus the following formula is found:

(4.34)

4.9 Thin tubes with variable thickness


In this section, we present an approximate method for thin tubes with variable
thickness. We use the following notations (see Fig. 4.5a): the closed curve CI
is the inner boundary of the tube, while the outer boundary is C, 8 is the
material surface (shaded region in Fig. 4.5a) and 8 1 of area Al is the surface
contained within CI. From Sect. 4.8, we know that the Prandtl stress function
'ljJ must satisfy:
'Ij; = O on C; 'Ij; = kl on C1
We make the following assumption: 'l/J varies linearly through the thickness:
k
'l/J(r,O) = h(O)1 (Re(O) - r) (4.35)

where (r, O) are polar coordinates, h(O) is the (variable) thickness and Re(O)
the (variable) outer radius. The expression of'l/J is approximate because it
will not verify in general the condition Ll'l/J(r, O) = -2 in 8.
We designate by h the average thickness of the tube and by 1 its average
perimeter. The constant k 1 is computed from Eq. (4.32):

Thus the constant k1 has the approximate value:

(4.36)

Next, we make the following approximation:


108 4. Torsion of beams

(4.37)

which is found as follows: in order to compute the integral of 'ljJ over the
surface, take its integral over the thickness and multiply it by the average
perimeter l, and since 'ljJ varies linearly through the thickness, its integral over
the thickness is equal to the area kl h/2 of the triangle plotted in Fig. 4.5b.
The torsional rigidity M/a is given by Eq. (4.34) as:
M r
2ţ.ta = ls'ljJ dydz + klAl ~
1
2klhl + klAl
Using Eq. (4.36), we find that M and a are related by:
M h hl
-
2ţ.ta
~ 2Al - (Al
l
+ -)
2
(4.38)

A further simplificat ion can be made by neglecting the area term hl/2 in front
of Al, Another application (a thin-walled section with two voids) is given in
Sec. 4.12.

C:'ljJ=o
S
'ljJ(r, B)

kl

CI : 'ljJ = kl
Fig. 4.5. Thin tube with variable thickness: notation. Assumption: stress function
'I/J varies linearly through the thickness.

4.10 Potential energy


The potential energy for a beam under torsion has the following expression:

I =

=
= (4.39)
4.11 Cylindrical coordinates 109

The work of the torque (M) in the rotation (la:) is given by:

MIa: = Is 1/J(y, z) dy dz
2ţLla: 2 (simply connected),

= 2ţLla: 2 [Is 1/J(y, z) dy dz + ~ kiAi] (multi-connected) (4.40)

Approximations to the solution of a torsion problem can be found by applying


Ritz's method (Chap. 2). An example is given in Sec. 4.12.

4.11 Cylindrical coordinates

It is sometimes easier to use cylindrical (r, 8, x) rather than Cartesian coor-


dinatesj an example is given in Sec. 4.12. We first recall some results from
Appendix A. With y = rcos8 and z = rsin8, we have:

dy = cos8(dr) - r sin 8(d8), dz = sin 8 (dr) + r cos 8(d8)


From which it is deduced that:

dr = cos 8( dy) + sin 8(dz), rdJ) = - sin 8( dy) + cos 8(dz)


Consequently, for 1/J(r,8) = w(y, z), the following identities are deduced:
8w = (cos 8) 8'Ij; _ (sin 8) 81/J, 8w = (sin 8) 81/J + (cos 8) 81/J
8y 8r r 88 8z 8r r 88
The Laplacian of 1/J(r, 8) is given by:

1 8 81/J 1 8 21/J
Ll1/J(r,8) = -:;: 8r (r 8r ) + r2 882 . (4.41)

Finally, the stresses transform according to:

a rx = a yx cos 8 + a zx sin 8, a()x = -ayx sin 8 + a zx cos 8


Using aH the above results, the problem becomes: find a function 1/J(r, 8) SUch
that:

Ll1/J(r,8)=-2 in S, 1/J(r,8) =0 on C, (4.42)

if the section is simply connected, and

Ll1/J(r,8)=-2 in S, 1/J(r,8) =0 on C, 1/J(r,8)=ki on Ci, (4.43)

for a multi-connected section. In this latter case, constants ki are solution to


the following system:
110 4. Torsion of beams

1Ci
ât/J rdO = 2Ai
âr
(4.44)

The torsional rigidity is:

M =21-' r t/J(r,O)rdrdO (simply connected),

t
a ls
= 21-' [Is t/J(r, O) r dr dO + kiAi] (multi-connected) (4.45)

The stresses are given by:

(4.46)

The tangential stress is given by:

(4.47)

The rudal displacement is given by:

u(r, O) = atjJ(r, O),


where the warping function tjJ(r, O) is found by integrating the system:
âtjJ 1 ât/J 1 âtjJ ât/J
-=--, --=-r-- (4.48)
âr r âO r âO âr
The potential energy has the following expression:

w; = J:.ţLla2 [ [(ât/J)2 + (J:. ât/J)2] rdrdO - Mia, (4.49)


p 2 ls âr r âO

where the work of the torque (M) in the rotation (la) is readily computed
from Eq. (4.40).

4.12 Solved problems


4.12.1 Elliptic section

Consider a beam with uniform elliptic cross section S whose boundary C is


defined by the equation:
y2 Z2
a2 + b2 = 1, (y, z) E C
4.12 Solved problems 111

Since a Prandtl stress function 1/J(y, z) must verify: 1/J = Oon C, we are certain
to satisfy this condition if we take 1/J proportional to the equation of C:
y2 z2
1/J(y,z) = k(a2 + b2 -1)

where k is a constant. Now, we need to verify: .t11/J = -2 in Si this gives:


a 2 b2
k=----
a 2 + b2
Thus the Prandtl stress function for the problem is given by:
1
1/J(y, z) = - a 2 + b2 (b 2y2 + a2z2 - a2b2) (4.50)

The method used here to find the stress function works also in other cases
(e.g., a triangular section in Sec. 4.12.2 or a notched circular section in Sec.
4.12.3), but it does not work for rectangular or square sections, for instance.
The torsional rigidity is given by Eq. (4.18) as:

-Ma = 2JL 1s
1/J dy dz = -
a
2
2JL b (b2 Iz + a2 Iy - a 2 b2 A )
+ 2

where Iz and Iy are the moments of inertia of the cross section w.r.t. the z
and y axes, respectively, and A is the cross section area:

Iz == fs y2 dy dz = ~a3bi Iy == fs Z2 dy dz = ~ab3i A == fs dy dz = 7rab


Thus the torsional rigidity can be rewritten as follows:
M a3 b3 JL A4
(4.51)
-;- = JL7r a2 + b2 = 47r2 Ip

where I p is the polar moment of inertia:

I p == fs (y2 + z2) dydz = Iz + Iy = 4(a2 + b2)


Equation (4.51) allows the following interpretation: for the same material -
same JL- and the same cross section area -same A- (thus the same quantity of
material) the torsional rigidity is inversely proportional to the polar moment
of inertia I p , Le. I p ,li==} (MIa) \.c. As an example, consider circular and
elliptic cross sections of the same area: 7r R 2 = 7rab => R = ~, thus

Ip(ellipse) - Ip(circle) = 1 (a 2 + b2 - 2R2) = 1 (a - b)2 >O


Therefore the torsional rigidity ( MI a) for a circular section is larger than
that of an elliptic section of the same area.
112 4. Torsion of beams

The stresses are given by (4.17) as:

b2 M 2M
2J.LQ a2 + b2 Y = 2Iz Y = Aa2 y;
a2 M 2M
-2J.LQa2+b2z=-2Iyz=- Ab2Z (4.52)

The tangential stress is given by (4.23) as:

We have shown in Sect. 4.5 that T max is reached on the boundary C of S.


Thus, we compute T on the boundary:

Assuming that a > b, it is seen that T is maximum when y = 0, thus T max =


T(O, ±b) and it is given by:
2M
T max = Ab (4.53)

Points (O, ±b) in which T max is reached are the points of C which are closest
to the centroid of the section, see Fig. 4.6. This conclusion seems to hold in
general for simply connected sections.

y
-a a

Fig. 4.6. Beam of elliptic cross section. Maximum tangential stresses in the case
a>b

We now compute the axial displacement of the cross section. Equations


(4.19) give:
aq; (b 2 - a2) aq; (b 2 - a 2)
- = Z' - - Y'
ay b +a
2 2 ' az - b2 + a 2 '
Therefore the axial displacement (u = Qq;) is found as:
4.12 Solved problems 113

[- (a2 - fr) ]
u(y,z) =a a2 + b2 yz + D (4.54)

where D is an arbitrary constant corresponding to a rigid translation along


the x-axisj it can be set to zero if such a displacement is prevented. The
contour lines of u(y, z) are plotted in Fig. 4.7. In the case a > b, it is seen
that u < O if yz > O and u > O if yz < O, it is thus dear that warping
will occur. The only exception is a circular section, for which u = O. In
alI other cases, the Navier-Bernoulli assumption, which is valid for beams
under bending, is no longer valid for beams under torsion: a section which
was normal to the middle fiber before deformation is not even planar after
deformation.
For a circular section of radius R, it suffices to set a = b = R in the
previous results,

U zx = /-taYj u xy = -/-taZj T max = /-ta R j M = !!:.~R4


"
a 2

u<O

u>O
Fig. 4.7. Beam of elliptic cross section. Contour lines of the axial displacement
u(y,z)

4.12.2 Triangular section

Consider the triangular section of Fig. 4.8. We choose the Prandtl stress
function 'IjJ to be proportional to the equation of the boundary C of S:

----'
'IjJ(y, z) = k (y - a) (y -
side (1)
zV3 + 2a) (y + zV3 + 2a)
.
side (2) ", .
side (3) "

Thus we have 'IjJ = O on C. In arder to have L1'IjJ = -2 in S, the constant k


should have the value k = -1/(6a), which gives the stress function as:

'IjJ(y, z) = - 6a1 (y3 - 3yz2 + 3ay2 + 3az2 - 4a3) (4.55)

The stresses are given by Eqs. (4.17) as:


114 4. Torsion of beams

o'zx = ţLa 22
2a (y - z + 2aY)j o'"x
ţLa)
= -;-z(y - a (4.56)

The maximum tangential stress is reached on the boundary C. On side (1)


of C (y = a), we have:

Tmax on this side is reached for y = a and z = O (which is the middle of side
(1)) and its value is T.\!-lx = (3ţLaa)/2. On side (2) of C (y - zV3 + 2a = O),
we have:
0'(2) = ţLa (y2 + ay _ 2a 2). 0'(2) = ţLa (y2 + ay _ 2a2)
"X aV3 ' zx 3a

Tmax on this side is reached for y = -a/2 and z = aV3/2 (which is the middle
of side (2)) and its value is T!;lx = (3ţLaa)/2 = T.\!-lx. By symmetry, we have
an identical result for side (3). In conclusion, the maximum tangential stress
is reached in the middle of each side of C, and its value is:
3
Tmax = "2ţLaa
Again, note that the points where T is maximum are those which are closest
to the centroid of the section.
Exercise: Show that M/a = (3ţLlp )/5, where I p = 3V3a4 is the polar mo-
ment of inertia of the section w.r.t. to the centroid.

-2a a y

(1) 2aV3

Fig. 4.8. Beam. of triangular cross section

4.12.3 Notched circular section

A solid beam has a notched circular cross section depicted in Fig. 4.9. The
domain (8) of the section is defined by:
(y _ a)2 + z2 :::; a2, y2 + Z2 ~ b2
4.12 Solved problems 115

It is easier to work with polar coordinates as shown in the figure. With


y = r cos 8 and z = r sin 8, (8) can be redefined as follows:

r2 (1 - 2; cos 8) : :; O, r2 ~ b2 (4.57)

Let us look for a stress function under the following form:

'l/J(r,8) = k(b 2 - r2 ) (1 - 2; cos 8) (4.58)

Equations (4.57) show that 'l/J vanishes on the boundary (C) of (8). Using
formulae from Sec. 4.11, it is found that L1'l/J = -4k. Therefore, in order for
'l/J(r,8) to be a stress function, we need to take k = 1/2.
The stresses are given by Sec. 4.11 as:

JLa(b2 - r 2 ) ~ sin8,
r

U()", -JLa [-r (1- 2ra COS8) + (b 2 - r 2 ) r~ cos 8] (4.59)

We know that the tangential stress T is maximum on the boundary (C). We


distinguish two cases:

(i) r = b (boundary ofthe notch): T = -JLab(l- 2a cos 8/b),


maximum value: T(b, O) = JLa(2a - b).

(ii) r = 2acos8 (large circle): T = J.taa[(b 2 /r 2 ) - 1],


maximum value: T(2a, O) = JLa[a - (~/4a)].

One can check that T(b, O) > T(2a, O), thus the maximum tangential stress is:
T ma", = JLCx(2a - b) (4.60)
It is reached at (r = b, 8 = O), Le. point A of Fig. 4.9. It appears that
if b --t O, then Tma", --t 2JLaa, Le. twice the value for a circular section
without a notch. A vanishingly small notch is a stress concentrator.
The axial displacement u(r,8) is found by integrat ing the following equa-
tions (Sec. 4.11):

-8rjl
8r
= -a (1b
2
- - ) sm
r2
. 8, -8rjl
8e
= -ar (1b
+ -r 22 ) cos 8

The solution is given by:

u(r,8) = -aar (1 + ::) sin 8 + const., (4.61)

where the constant corresponds to a rigid displacement. It appears that there


is warping even if b --t O (u < O for 8 > O and u > O for 8 < O). Recall that
a circular section with no notch does not warp.
116 4. Torsion of beams

Fig. 4.9. Notched circular section.

4.12.4 Thin rectangular section

We apply the membrane analogy method to the torsion of a beam with


uniform and simply connected thin rectangular cross section of dimensions b
and c « b, see Fig. 4.10.

-b/2

-c/2 y -c/2 +c/2 y

+b/2
z
Fig. 4.10. Thin rectangular cross section of dimensions b and c « b

Imagine that in a rigid plate, we cut a hole whose boundary is identical


to the rectangle of Fig. 4.10, then a thin membrane is stretched over the
plate, fixed to the boundary of the hole, and subjected to uniform tension
and pressure. Since the rectangular hole is very thin in the y-direction, we
expect that except near the ends z = ±b/2, the deformed membrane will have
a cylindrical shape whose equation is proportional to [(c 2/4) - y2]. Therefore,
4.12 Solved problems 117

returning to the torsion problem, we look for a Prandtl stress function under
the form:
c2
'IjJ(y,z) = k(4 - y2)

In order to have f1'IjJ = -2 in S, we need to take k = 1. The expression of 'IjJ


satisfies 'IjJ(±c/2, z) = O but for z = ±b/2 and y i ±c/2, we have 'IjJ i O. Thus
'IjJ is an approximate solution because it does not satisfy the stress-free B.C.
'IjJ = O everywhere on the boundary C. The torsional rigidity corresponding
to the approximate solution 'IjJ is:

The stresses corresponding to the approximate solution 'IjJ are:

The maximum tangential stress is reached at y = ±c/2 and its value is


T max= p,ac. We now compute the moment due to the stresses. This is given
by the right-hand-side of Eq. (4.8) as:

r (yu
ls
zx - zU yx ) dydz = 2p,a r y2 dydz = ?:.p,bc
ls 6
3

This is only hal! the value of the applied torque M. The reason for the
discrepancy is that the u yx stresses are completely neglected when we have
assumed that the deformed membrane had a cylindrical shape. Although
those stresses are important only near the edges z = ±b/2, their moment
w.r.t. the centroid of the section is important and gives the "missing" half of
the moment (Timoshenko and Goodier, 1987).

4.12.5 Ritz's method- Square section

In this section, we apply Ritz's method (Chap. 2) to the torsion of a beam


with a simply connected cross section. Consider the following approximation
to Prandtl's stress function:

'IjJ(y,z) ~ bo'IjJo(y,z), (4.62)

where bo is a coefficient to be determined and 'ljJo(y, z) a given function,


"sufficiently smooth" and satisfying:

'ljJo(y, z) = O, on C (4.63)

From Sec. 4.10, it is found that the potential energy I('IjJ) associated with
the stress function (4.62) has the following expression:
118 4. Torsion of beams

~=
J-tla 2
r [b5(â'lj;O)2+b5(â'lj;O)2_2bo'lj;O]
Js 2 ây 2 âz
dydz (4.64)

Constant bo must be such that I is minimum, Le. âI/ âba O (since


â2I/âb5 > O). It is easily found that:

2 Is 'lj;o (y, z) dy dz (4.65)

As an example, consider a square section of side (2a): (y, z) E [-a, al x [-a, al,
and the following expression for 'lj;o(y, z):

(4.66)

The corresponding stress function (4.62) is not exact because it does not
verify i1'1j;(x,y) = -2 in (S). Equation (4.65) gives the constant ba as:
5
bo = -
8a 2
The torsional rigidity is found to be:
M 20 4
-=-J-ta
a 9
Simple algebra gives the maximum tangential stress as:
5
T max = 4J-taa

The value is reached in points (y = ±a, z = O) and (y = O, z = ±a), Le. at


the middle of each side (these again are the points closest to the centroid of
the section).
One can increase the accuracy of Ritz's method by taking more functions
in the approximation, Le. 'Ij;(y, z) ~ Li bi'lj;i(y, z). Results in (Timoshenko
and Goodier, 1987) give:

-
M = 0.1406J-t(2a) ;
4
T max = 0.675J-ta(2a)
a
Relative differences with the single approximation results are 1.21% and
7.41%, respectively. As in Sec. 4.12.4, it is more difficult to approximate
the stresses than a "global" quantity such as the torsional rigidity.

4.12.6 Hollow elliptic section- Special method

Consider a hollow beam whose cross section is represented in Fig. 4.11. The
external (C) and internal (Cd boundaries are two concentric ellipses whose
equations are:
4.12 Solved problems 119

c:

The Prandtl stress function of the solid beam is given by (4.50) as:
22
'ljJ(y, z) = - a2a +b b2 (y2 Z2
a 2 + b2 - 1
)

It appears then that the boundary Cl of the hole is a contour line 'ljJ = kl of
the stress function 'ljJ of the solid beam. Therefore, the stresses in the hollow
beam are equal to those of the solid beam and given by Eqs. (4.52) as:

b2 a2
Uzx = 2J.ta a 2 + b2Y ; u yx = -2J.ta a2 + b2z (4.67)

The torsional rigidity (Mia) ofthe hollow beam is reduced w.r.t. that ofthe
solid beam by the amount of material which was removed. Using Eq. (4.51),
the value of (Mia) for the hollow beam is found as:
M a 3b3 ((a)3((b)3 a 3b3
~ = J.t7r a2 + b2 - J.t7r ((a)2 + ((b)2 = J.t7r a2 + b2 (1 - (4) (4.68)

Finally, note that the actual value kl of the contour line 'ljJ = kl was not
used, although it can be easily computed:

'ljJ(y,z) =

Thus the value of k1 is given by:

4.12.7 Hollow elliptic section- General method

The problem of Fig. 4.11 was already solved using a special approach, we
shall solve it again using the general theory and double-check the results.
Consider the following function:

It vermes:
120 4. Torsion of beams

z
s

Fig. 4.11. Hollow beam of elliptic cross section. The boundary CI of the interior
hole is a contour line 1/J = k l of the stress function 1/J of the solid beam

L1'ljJ = -2 in S, 'ljJ = O on C, 'ljJ = kl on CI,


where the constant kl is given by:
a 2 b2
kl = a2 + b2 (1 - (2)

Therefore, it is concluded that 'ljJ is the Prandtl stress function for the hoUow
beam. The torsional rigidity (MiCi) is given by Eq. (4.34) as:

M =
-2J.LCi s
1
'ljJ dydz + klAI = - 2 1 b2 (b 2 Iz + a2 Iy - a22
a +
b A) + klAI

The area A and the moments of inertia Iy and Iz aU refer to the dashed
surface in Fig. 4.11, Le. the section which contains material, while Al is the
area inside CI. Using formulae given in Sect. 4.12.1, we obtain:
Al 7rab(2; A = 7rab(l _ (2);
7rab3 7ra 3 b
Iy = -4-(1 - (4); Iz = -4-(1 _ (4)

Substituting those results into the expression of (MiCi), it is found that:


M 7ra 3 b3
2J.LCi - 2(a2 + b2) (1 - (4)
which is exactly Eq. (4.68). The stresses are computed from Eqs. (4.33) as:
b2 a2
U zx = 2J.LCi a2 + b2Y ; u yx = -2J.LCi~b2Z
a +
and these are of course Eqs. (4.67), since the same stress function was used
in both methods. Note also that the stresses can be put under the form:
M M
U zx = 2Izy; u yx =-2Iy Z

where again M, Iy and Iz refer to the section which actuaUy contains the ma-
terial, Le. the dashed area in Fig. 4.11. It appears that the stress expressions
thus obtained are form-identical to those for the solid beam, Eqs. (4.52).
Finally, Eq. (4.32) of the general theory is automatically satisfied.
4.12 Solved problems 121

4.12.8 Thin circular tube

Consider the torsion problem of a circular tube of internal and external radii
Il;, and Re, respectively. The torsional rigidity is given by Eq. (4.68) as:

(4.69)

The stresses are given by Eqs. (4.67) as:

a zx = ţtaYi a yx = -ţtaz (4.70)


and they are identical to those for a solid circular section. The tangential
stress is:

= Ja 2yx + a 2zx -
T - aţtr (4.71)

and varies linearly with the thickness. For a very thin tube, Re = Il;, + h,
hj Re « 1, Eq. (4.69) becomes:

M
-a ~ 27rţthR
3
(4.72)
e

Eliminating (aţt) between Eqs. (4.71) and (4.72), we find that:

M r M
T~-----~-- (4.73)
2h7r R~ Re 2hA I

where Al ~ 7rR~ is approximately the area of the surface contained inside


the inner circle (r = Ri) of the tube. Equation (4.73) is a design formula for
thin tubes under torsion.

4.12.9 Thin-walled section with multiple voids

Consider a thin-walled cross section which has two voids, Fig. 4.12. It is
assumed that the stress function t/J has a linear variat ion over the thickness,
see Fig 4.12. Two constants kl and k 2 need to be computed from Eqs. (4.32):

Three lines running through the mid-thickness are defined (Fig. 4.12) and
have the following values of thickness h i and length li:

Making the same assumptions which led to Eq. (4.36), we find that the con-
stants kl and k 2 are solution to the following linear system:
122 4. Torsion of beams

k 1l (k 1 - k2 ) l
h1 1 + h3 3

(k2 - kd l k 2l
h3 3 + h2 2 = (4.74)

The torsional rigidity Mia is given by Eq. (4.34) as:

M
2p.a
= r
Js
'IjJ dydz + k1Al + k 2A 2

Similarly to Sec. 4.9, it is found that M and a are related by:

A further simplificat ion can be made by keeping only the dominant term
(k1Al + ~A2)'

f 8 'ljJdydz terms
Fig. 4.12. Thin tube with variable thickness, a cross section with two voids: nota-
tion. Assumption: stress function t/J varies linearly through the thickness.
5. Theory of thin plates

In this chapter, we present the classical (Kirchhoff-Love) theory ofthin plates.


A plate is a solid with one dimension -the thickness (h)- much smaller than the
other two dimensions, and such that the mid-thickness surface is contained in
a plane. We shall see that plate theory reduces the original three-dimensional
(3D) problem of Chaps. 1 or 2 to a two-dimensional (2D) problem in the
mid-thickness surface.

5.1 Definitions and notation

We designate by (x, y) the mid-thickness surface, it is also the (z = O) plane.


It is assumed that the thickness h is constant. The plate is thus contained
within the (z = ±h/2) planes. Externalloads are represented with a surface
density li(x, y) [N /m 2 ) which is supposed to act directly on the middle surface
(this is analogous to beam theory where it is assumed that externalloads are
transmitted directly to the middle fiber).
Cartesian coordinates w.r.t. a fixed frame are designated by (x, y, z) =
(XI,X2,X3), whatever set ofnotation is more convenient. Greek indices a, (3,
'Y takes values 1 or 2.

5.2 Internal loads (stress resultants)

Plate theory is developed in terms of stress resultants, Le. stresses integrated


through the thickness (recall that in beam theory, internal loads are stress
resultants over the cross section). We consider a "slice" of plate of dimensions
dx x dy x h, see Fig. 5.1. The resultant of contact forces an the facet with
outside normal ea: is given by:

f tI dydz,

where tI is the stress vector on the facet of outer normal ea::


124 5. Theory of thin plates

It appears then that the stress resultant has the following form:

h/2 jh/2 jh/2


Nu == j O'xx dz; N 12 == O'xy dz; Ql == O'xz dz (5.1)
-h/2 -h/2 -h/2

z:r-
dx h

-
z
d.
-
y
I I
x
h
I I

z
Fig. 5.1. A "slice" of plate of dimensions dx x dy x h

The resultant of contact forces along the facet of outside normal €y is:

h/2 jh/2 jh/2


N 22 == j O'yy dz; N 21 == O'yx dz = N 12 ; Q2 == O'yz dz (5.2)
-h/2 -h/2 -h/2

The moment of contact forces on the facet with outside normal €x w.r.t. the
center of the facet is given by:

with:

(5.3)
5.2 Internalloads (stress resultants) 125

The moment of contact forces on the facet with outside normal ey w.r.t. the
center of the facet is given by:

M 22 == ! h~ zayydz; M 21 ==
!h~
zayxdz = M12 (5.4)
-h/2 -h/2

It is important to notice that forces (N11 , N 12 , 01) and moments (Mn ,


M12) are defined per unit length of the y-coordinate line, while forces (N22 ,
N 2b 02) and moments (M22 , M 21 ) are per unit length of the x-line. Stress
resultants (Na {3, Oa) and their moments (Ma {3) together with their positive
directions are plotted in Fig. 5.2.

M(-)
21

(-)
M11 I

....,......~
(-)
M 12 W ----~~----fu I
Mg)
I
I
M n~)
I

M(+)
21
Fig. 5.2. Stress resultants (Na{3, Qa) and their moments (Ma{3) plotted with their
positive directions, for a slice (dx x dy x h).
126 5. Theory of thin plates

5.3 Equilibrium equations


We consider the same slice of plate as in the previous sectionj it is subjected
to the following forces:
dXl
± Nt{Xl ± T,X2) dx2 (outside normal ± et)
dX2
± N 2(Xl,X2 ± T)dxl (outside normal ± e2)
+J(Xl,X2)dxl dx2 (external forces)

A first order Taylor expansion gives:


dXl aNI dXl
NI (Xl + T,X2):::::: N l (Xl,X2) + aXl (Xl,X2)T

Similar results hold for the other terms. It appears then that equilibrium of
forces gives the following vector equation:

which is rewritten as follows, using the definitions of N a,

Therefore, we obtain the following three scalar equations which are the pro-
jections of the vector equation along the el, e2 and e3 directions, respectively:

aN11 aN2l f 1= O
--+--+
aXl a X2
aN12 + aN22 + h = O
aXl a X2
aQl + aQ2 + il = O (5.5)
aXl aX2
We now consider the equilibrium of moments act ing on the slice. Keeping the
notations of Sect. 5.2, it is seen that there are two contributions: one is due
to moments Ma, and the other is due to the moments of the forces N a'
The development for the former term is identical to that for the equilibrium
of forcesj thus the sum of those moments has the following expression:

As for the second contribution (moments of forces NI and N 2 w.r.t. the


center (Xl, X2, X3 = O) of the slice), one obtains the following expression:
5.4 Displacements 127

After a first order Taylor expansion, the sum of moments due to forces N '"
takes the following form:

Adding the contribution of M"" equilibrium of moments gives:


8 8
-8 (-MI2 e l
Xl
+ Mn e 2) + -8X2
(-M22 e l + M21 e 2)
+Q2el - Qle2 + (N12 - N 21 )e3 = O

Projecting the vector equation along directions el, e2 and e3, respectively,
we obtain the following three scalar equations:

_ 8Ml2 _ 8M22 + Q2 O
8XI 8X2
8Mn + 8M21 _ QI O
8XI 8X2
N 12 - N 21 = O (5.6)

In summary, there are six scalar equilibrium equations for plates: Eqs. (5.5)
which express equilibrium of forces and (5.6) which translate equilibrium of
moments, Eq. (5.6c) being trivially satisfied because N",ţJ is symmetric.

5.4 Displacements

Let P(x, y, O) be an arbitrary point of the plate's mid-surface, and Q(x, y, z)


a point in the thickness of the plate such that:

~ = ze z , before deformation.

After deformation, points P and Q transform onto P' and Q' such that:

pp)' = Vl(X, y)e x + V2(X, y)e y + w(x, y)e z , P'Q~ = zn, (5.7)

where n is the unit normal to the deformed mid-surface. Two KirchhoJJ-Love


assumptions are embedded in (5.7b) (and illustrated in Fig. 5.3):

(Hl): A material segment which is initially normal to the mid-surface re-


mains straight and normal to the deformed middle surface.
128 5. Theory of thin plates

:~//:1-·_·_·_·_·_·_·_·_·_·_·_·_·_;t~_·_·_·-1-
)
z
z
(a)

f:::---~
~._._._._._._._._._~
z
(b)
Fig. 5.3. Kirchhoff-Love kinematic assumptions. A material segment before (a)
and after (b) deformation

(H2) The plate is inextensible in the (z) direction (no thickness variation).
We now compute the normal (n) used in (5.7b). Tangent vectors to the
deformed mid-surface at (PI) are given by:

(ex+ ~~ez) in the (z,x) plane,

in the (y, z) plane (5.8)

The cross product of the two vectors gives a normal v,


ow ow
v = ez - oy e y - ox e x
Taking the dot product, it is found that:
OW OW
111.111 2 = 1 + (OX)2 + (Oy)2 ~ 1,
because ofthe small perturbation hypothesis (SPH). Thus the unit normal to
the deformed mid-surface is n ~ v. Consequently, the displacement of point
Q can be determined as follows:
QQ)I ~ -=-=1 t--1
= -P,,!+PP +PQ =-ze z +vl(x,y)e x +v2(x,y)e y +w(x,y)e z

+z (e z - ~; e y - ~~ e x)

In component form, we obtain the following three scalar equations:


OW(Xl, X2)
Ua (Xl,X2,X3)=V o,(Xl,X2)-X3 oX a ; U3(X,y,Z)=w(x,y) (5.9)

Because of the assumption of no-thickness variation (H2), it is found that the


z-displacement does not depend on (X3)'
5.5 Strains 129

5.5 Strains
We shall designate by €(x, y) and e(x, y, z) the strain fields inside and outside
the mid-surface, respectively, Le.

Using Eqs. (5.9), we obtain the following relations:

-1'33 = -1'23 = -1'13 (5.10)

The result (33 = O is a consequence of assumption (H2), while the transverse


shear strains vanish ((23 = (13 = O) because of (H1). Introducing the curva-
ture tensor "'a{3 of the deformed mid-surface, Eqs. (5.10) can be re-written
as follows:

(5.11)

5.6 Constitutive equations

Isotropic linear elasticity gives the following three scalar equations:


_ (1 + v) v
Ea{3 = -E--Ua{3 - E(u" +uzz )Da{3, (5.12)

as well as three others for (33, (23 and (13, which will be considered later.
The following "plane stress" or "thin plate" assumption is adopted:

(H3): The out-of-plane stress (u zz) is negligi bIe in front of the in-plane stresses
(u a {3 ).

As a consequence of (H3), the constitutive equations simplify to:


_ (1 + v) v
Ea{3 ~ -E--Ua{3 - EUl1Da{3

Taking the trace on both sides, it is found that:


_ (1 - v)
1'" = -E--u"

This allows to invert the constitutive equations:

(5.13)
130 5. Theory of thin plates

These equations can be rewritten as follows, using Eqs. (5.11),

The advantage of form (5.14) is that it isolates terms f a t3(Xl,X2) and


1\;-y-y(Xl,X2) which do not depend 01. the thickness coordinate (z) since they
are only related to the deformed mid-surface. Integrating (5.14) through the
thickness gives internal forces per unit length N at3,

N at3 == f
-h/2
h/2
u at3 dz =
Eh
--2
1- v
[(1 - V)f at3 + Vf-y-y Oat3] (5.15)

The bending moments are also obtained from (5.14) by simple integration:

(5.16)

where parameter V [N m] is the bending stiffness or flexural rigidity:

(5.17)

Equations (5.15, 16) are the in-plane and bending constitutive relations, re-
spectively. There is no constitutive model for the shear loads Qa. However, as
we shall see later, from equilibrium, we can express Qa as a function of trans-
verse displacement W(Xl' X2). Knowing internal forces and moments per unit
length, N a t3 and M at3 , the stresses can be found by substitut ion of (5.15-16)
into (5.14),

(5.18)

These equations strongly resemble those of beam theory (Sec. 3.6).

5.7 Summary: two un-coupled problems

When aU the equations found so far are collected together, it is seen that those
for "membrane" forces N a t3 which act in the surface ofthe plate, are not cou-
pled with those for bending moments M at3 and shear (transverse) forces Qa.
More precisely, the following two un-coupled problems are obtained:

Problem {1}: a "membrane" problem in the mid-surface plane,


5.8 Fundamental P.D.E. for bending problem 131

(2 scalar equations)

(3 scalar equations)

(5.19)

There are 5 scalar equations and 5 unknown fields: N n , N 22 , N 12 = N 2 1, VI


and V2, an of which independent ofthe thickness coordinate (z). Problem (1)
is a "generalized plane stress" problem, which will be dealt with in Chap. 7.
In this chapter, we concentrate on the second problem:

Problem (2): a bending problem in the mid-surface plane,

(2 scalar equations)

(1 scalar equation)

(3 scalar equations)

(5.20)

There are 6 scalar equations and 6 unknown fields: M n , M 22 , M 12 = M 2b


Ql, Q2 and w, alI of which functions of (Xl, X2) uniquely. ActualIy, we shall
show in the next section that the problem is reduced to solving one PDE
(partial differential equation) for one single unknown function: w(x, y), under
appropriate B.Cs.
The "internalloads" for bending problem (2), i.e. bending moments M cx {3
and shear forces Qcx acting on an elementary plate slice are represented in
Fig. 5.4. The representation is equivalent to that of Fig. 5.2. Looking at Fig.
5.4, the following observations can be made:

• M u and M 22 act in the (z,x) and (y,z) planes, respectively. They are
bending moments whose sign is positive if they compress the upper part of
the plate (z < O). They play identical roles to that of M(x) in beam theory.

• M 12 and M 2l act in the (y, z) and (z, x) planes, respectively. They act
in the same shearing planes as those of shear forces Ql and Q2.

5.8 Fundamental P.D.E. for bending problem

Using the definition of K. cx {3, the bending moment-curvature constitutive equa-


tions can be rewritten as folIows:
132 5. Theory of thin plates

Fig. 5.4. Sign convention for bending moments Ma(J and shear forces Qa acting
on a slice of plate of dimensions dx x dy x h.

(5.21)

where L1 designates the Laplacian operator. Taking the divergence on both


sides of the previous equality gives:

âMa {3 = -V~(L1w)
âX{3 âXa

Using equilibrium equations (5.20a), it is deduced that:


â
Qa = -V-(L1w) (5.22)
âXa

Taking the divergence on both sides, we find:

âQa = -VL1(L1w)
âXa

Using (5.20b), the following fundamental PDE is found:

(5.23)

where p == h. Therefore, bending problems are reduced to solving a single


PDE where the sole unknown is the transverse displacement W(Xl, X2), under
appropriate B.Cs. (see next section). This result is similar to that of beam
theory where the problem is reduced to solving a differential equation for the
deflection v(x), see Sec. 3.3.5. Once the kinematic solution w(x,y) is found,
all other static variables are easily computed:
-the bending moments from constitutive equations (5.21),
-the shear forces from (5.22),
-the stresses from (5.18): lTa {3(Xl,X2,Z) = (12/h 3 )zMa {3(Xl,X2).
5.9 Boundaxy conditions 133

5.9 Boundary conditions


In this section, we specify the B.Cs. which need to be appended to the PDE
(5.23) in order to solve a given bending problem. For generality, we consider
a curved edge of outward normal (n) and tangent vector (t), parameterized
by a curvilinear coordinate (8); see Fig. 5.5.

Fig. 5.5. Bending moments M nn and M ns and sheax force Qn act ing on a curved
edge of outwaxd unit normal n and curvilineax coordinate s.

Let us check that in the case of straight edges, the sign convention used in
Fig. 5.5 corresponds to that of Figs. 5.2 or 5.4. For an edge (x = a), we have
n = e x , t = ey, and we do retrieve the positive directions of M u and M 12
as plotted in Fig. 5.2. For an edge (y = b), we have: n = ey and t = -e x .
The positive direction of M 22 is obviously correct. As for M 21 , we can see
"heuristically" that Fig. 5.5 gives the direction of" M 2 ( -1)'" whose opposite
direction gives M21 as in Fig. 5.2 .

• Simply supported edge: if the plate is simply supported at an edge, then the
deflection w and the bending moment M nn must vanish:
w = O and Mnn = O, at the edge (5.24)
For a straight edge (x = a), the conditions read: w(a, y) = O and M xx (a, y) =
O. This latter equation can be written as (using Eq. (5.21)):
â2w â2w
âx 2 (a, y) + /J ây2 (a, y) = O

Since w(a, y) = O, 'r/y, simple support conditions at (x = a) simplify to the


following equations:
â2 w
w(a, y) = O; âx 2 (a, y) = O (5.25)

Similarly, for a straight edge (y = b), the B.Cs. are:


â2 w
w(x, b) = O; ây2 (x, b) =O (5.26)
134 5. Theory of thin plates

We may also write both equations (5.25-26) in a unmed form:


w =O and L\w = O, at the edge
• Built-in edge: if the plate is built-in at an edge, then the deflection and the
rotation must vanish there:
8w
w =O and -
8n
= O, at the edge (5.27)

For a straight edge (x = a), those B.Cs. read:


8w
w(a,y) = O; 8x (a,y) =O (5.28)

Similarly, if the plate is built-in at (y = b), the B.Cs. at that edge are:
8w
w(x,b) = O; 8y (x,b) =O (5.29)

• Free edge: if an edge is free, we are inclined to write three conditions


translating the fact that alI internalloads at the edge must vanish: M nn = O,
M ns = O and Qn = O. However, Kirchhoff has shown that the following two
conditions are sufficient:

M nn =O and Qn + ----a;-
8Mns = O, at the edge ( 5.30)

A heuristic proof is given in (Timoshenko and Woinowsky-Krieger, 1982). It


consists in replacing each moment (Mns ds) by a statically equivalent couple
of transverse forces M ns acting in directions ±ez and ds aparl. The principle
of de Saint-Venant implies that at a distance from the edge equal to several
times the thickness, the two systems give the same results. Considering two
adjacent segments of length ds and writing equilibrium of forces of the new
system, one finds Eq. (5.30b). More satisfying proofs of B.Cs. (5.30) can be
found in the literature, e.g. based on a variational formulat ion as in (Crochet,
1994), or on the rate of work of forces and moments acting on an edge, as in
(Green and Zerna, 1968). For a straight edge (x = a), B.Cs. (5.30) read:

M xx (a,y ) -_.
O, Qx
(a,y ) + ----ay-(a,y)
8Mxy -_ O (5.31)

Using Eqs. (5.21-22), the B.Cs. can be rewritten as follows:


&w &w ~w ~w
8x 2 (a, y) + v 8y2 (a, y) = O; 8x3 (a, y) + (2 - v) 8x8y2 (a, y) = O (5.32)

Similarly, for a free edge (y = b), the B.Cs. are:


&w &w ~w ~w
8y2 (x, b) + v 8x 2 (x, b) = O; 8 y3 (x, b) + (2 - v) 8y8x 2 (x, b) =O (5.33)
5.10 Contradictions in Kirchhoff-Love theory 135

• Loaded edge: if an edge is subjected to moments per unit length (mnt) and
forces per unit length (qnez), then by the same argument as for free edges,
the B.Cs. to be satisfied are:
âMns
M nn = mn and Qn + --a;- = qn, at the edge (5.34)

• Remark: corner. The heuristic method alluded to earlier shows that at a


corner (x = a, y = b) between two straight edges, there is a concentrated
force:

(5.35)

For example, for a square plate, uniformly loaded and simply supported on
its four sides, one finds reaction forces at the corners which have the same
direction as the pressure load (thus "retaining" the plate).

5.10 Contradictions in Kirchhoff-Love theory


Recall that in Sect. 5.6, we did not write the constitutive equations corre-
sponding to the out-of-plane strain components. Those equations read:

_ 1( ) _ (1 + v)
lO33 =E 0"33 - vO""!"! ; lO",3 = -E--0""'3
Since -from (5.10b)- the transverse shear strains are nil: f",3 = O, this implies
that the transverse shear stresses also vanish: 0"",3 = O. As a consequence,
using definitions (5.1-2), the shear forces would also vanish: Q", = 0, and this
-except for the simplest case of pure bending- is obviously wrong! Note also
that since by assumption (H3), we have: 0"33 ~ O, it would not be possible
to apply any load on the top or bot tom surfaces of the plate since the stress
vector there is (TT • e3 = 0"3iei = O! We point out that making "plane strain"
assumption (H2): f33 = 0, and the "plane stress" assumption (H3): 0"33 = 0,
simultaneously, leads to a contradiction, in general.
So how to deal with those contractions in the Kirchhoff-Love theory? We give
an engineering pragmatic approach here. As is the case for beams, the KL
theory is a simplified one where we are not trying to solve the 3D equations of
elasticityas presented in Chap. 1, but Eqs. (5.20) instead. Therefore, equilib-
rium requires satisfying (5.20a,b) and not (1.28). Also, constitutive equations
are not those of Chap. 1 relating stresses O"ij to strains fij, but Eqs. (5.20c)
relating bending moments M",{3 and curvature 1'b",{3. In practice, the KL the-
ory is widely used and gives good results for thin plates. Engineers "forget"
how it was derived and just use it!
We need to mention that there exists another theory, due to Reissner and
Mindlin in which kinematic assumption (Hl) is replaced as follows:
136 5. Theory of thin plates

(H1-RM): A material segment which is initially normal to the mid-surface


remains straight but not necessarily normal to the deformed middle surface.
As a consequence, transverse shear strains and stresses do not vanish in the
RM theory. This model is widely used in finite element computations, spe-
cially for "thick" platesj for details see (Hughes, 1987) and references therein.

5.11 Rectangular plates with two simply supported


opposite edges - Levy's method

We shall see in the next chapter that for circular plates under axisymmet-
ric conditions, closed-form solutions can be easily obtained. For rectangular
plates however, finding analytical solutions for even the simplest loadings is
very tedious and involves lengthy calculations with Fourier series. Therefore,
we shall restrict ourselves to a few illustrative examples, and refer to classical
textbooks, such as (Timoshenko and W.-K., 1982) for more solutions. Nowa-
days, good numerical approximations are easily found using finite-element
computer codes.
Among the classical analytical methods for rectangular plates, two are due
to Navier (1820) and Levy (1900), respectively. Navier's method deals with
plates which are simply supported on their four edges. It uses double Fourier
series whose convergence is slow for both M a{3 and Qa. Levy's method -which
is presented in this section- is used for plates with two simply supported and
opposite edgesj the B.Cs. on the two remaining edges being arbitrary. It is
based on a single Fourier series and its convergence rate is "not too bad" .
We consider a rect angular plate whose mid-surface occupies the domain:
(x, y) E [O, a] x [-bJ2, bJ2] which is simply supported at (x = O) and (x = a).
The problem to be solved is the following: find w(x, y) such that:

p(x, y) . [ ] [ b b]
LL1w (x, y ) = -----V-' In 0, a x -'2' '2 '
â2w
w(O, y) = 0, âx2 (O, y) = 0,
w(a,y) = 0,
â2 w
âx2 (a, y) = °
b
B.Cs.on (y = ±'2) (5.36)

Let Wl(X,y) be a particular solution which satisfies the system of equations


(5.36) except the B.Cs. on (y = ±bJ2) (otherwise, by virtue ofthe uniqueness
theorem, Wt (x, y) would be the solution of the problem). The solution of
problem (5.36) is written under the following form:

(5.37)
5.11 Plates with two simply supported opposite edges - Levy's method 137

Since w(x, y) must satisfy alI of Eqs. (5.36), and W2(X, y) satisfies (5.36)
except (5.36d), then: (i) W2(X, y) must verify the folIowing system:

Lt1w2(X,y) =0, in [O,alX[-~,~],


IJ2 W 2
W2(0, y) = O, âx 2 (O, y) = O,
â 2w2
w2(a, y) = O, âx 2 (a, y) = O, (5.38)

and (ii): Wl(X,y) + W2(X, y) must satisfy B.Cs. (5.36d) on edges (y = ±b/2).
Using the method of separat ion of variables, a solution to (5.38) is sought
under the folIowing form:

2: Yn(y) sin(n:x)
00

W2(X, y) = (5.39)
n=l
It is seen that B.Cs. (5-36b, c) are trivially satisfied. Equation (5.38a) leads
to the folIowing differential equation:

d4Yn _ 2( mf )2 ~Yn + (mf )4Yn = O,


dy4 a dy2 a
whose general solution is given by:
mry mry mry mry
Yn(y) = (an + bn -a
) exp(--) + (c n + dn - ) exp(-),
a a a
(5.40)

where an, bn , Cn and dn are constants. Equivalently, Eq. (5.40) can be rewrit-
ten as folIows:
mry . h n7rY )
(A nCOS h -mry
a
+ Bna
- sm -
a
n7ry n7ry . n7rY)
+ ( Cn-cosh-+Dn smh - (5.41)
a a a
An important thing to remember is that either results (5.40) or (5.41) are
valid for all Levy's problems. The choice of one expres sion depends on the
application at hand. The advantage ofEq. (5.41) is that it shows two contribu-
tions which are even and odd functions of (y), respectively. So, ifthe problem
is symmetric w.r.t. the (y = O) axis, (5.41) can be used with Cn = Dn = O.
For an anti-symmetric problem w.r.t. the (y = O) axis, (5.41) is used with
An = Bn = O. As another example, if a plate has a large dimension along
(y), i.e. b» a, then the (y = O) axis is placed such that the edges parallel to
it are (y = O) and (y = b), and Eq. (5.40) is used with Cn = dn = O, because
w(x, y) must remain finite for large values of y > O.
So far, we have found W2(X,y) solution of problem (5.38). It remains
to find Wl(X,y), a particular solution to Eqs. (5.36) except the B.Cs. on
(y = ±b/2). A method is to try a form similar to (5.39), Le.
138 5. Theory of thin plates

00

Wl(X,y) =L km(Y)Sin(m:x) (5.42)


m=l

The B.Cs. (5.36b, c) are thus automatically satisfied. The second step is to
write the pressure loading in a form similar to that of Wl(X, y), Le.
00

p(x,y) = LPm(y)sin(m:x) (5.43)


m=l

Fourier series formulae -e.g., (Tolstov, 1962)- give:

Pm(Y) = -a21 o
G
m7rX dx
p(x, y) sm(-)

a
(5.44)

Substitution of (5.42-43) into (5.36a) leads to the following differential equa-


tion:

(5.45)

for which it suffices to find a particular solution. This depends on the expres-
sion resulting from the integral in (5.44), and hence on the pressure loading.
An important special case is when the pressure is independent of y (e.g., uni-
form or hydrostatic pressures), P = p(x). This implies that Pm = constant. A
particular solution of (5.45) is therefore:

km = (~)4Pm
m7r V
The final solution w(x, y) of the original problem (5.36) is found by summing
the expressions of Wl (x, y) and W2 (x, y), Le.

(5.46)

or equivalently:
00

W(X,y) = L [(~)4Pn + (Ancosh n7ry +Bn n7rY sinh n7rY )


n=1 n7r V a a a
n7rY n7rY . n7rY )]. n7rX
+ ( Cn - - cosh - + Dn smh - sm(--) (5.47)
a a a a
Integration coefficients (An to Dn, or an to d n ) are determined by imposing
the B.Cs. on edges (y = ±b/2). A few applications are given in Sec. 5.14.
5.12 Potential energy 139

5.12 Potential energy


The strain energy per unit volume is given by:
1 -
W == '2O'ijEij 1 -
= '2 O'a{3Ea{3,
since it was assumed that 0'33 = O and la3 = O. Isotropic linear elastic
relations (5.12) allow to write W as the following function of the stresses:

W = 2~ [(1 + v)O'a{3O'a{3 - v(O'-y-y)2]


Since the stresses are related to Ma{3 by: O'a{3 = (12z/h 3 )Ma{3, we can express
W as a function of the bending moments:
1 12z 2 [
W = 2E( h3 ) (1 + v)Ma{3Ma{3 - v(M-y-y) 2]
The bending moment-curvature constitutive relations (5.20c) allow to rewrite
W as follows:

(5.48)
where definition (5.17) ofthe bending stiffness was used. Finally, since /î,a{3 =
-â2 w/âx aâx{3, the following expression of the strain energy W in terms of
the deflection w(x, y) is found:

W = "h,3Z2
2W â 2w (â 2W
6V { (LlW)2 - 2(1 - v) [ââx
2 ây2 - âxây )2] } (5.49)

The strain energy for the entire plate is obtained by integration of W over
the volume. Since the deflection w(x, y) does not depend on the thickness
coordinate (z), the following expression is easily found:

W(w) = 1 -V { (LlW)2 - 2(1- v) [â


s 2
-
2Wâ 2W (â 2W
- - --
âx 2 ây2 âxây
)2]} dxdy,

(5.50)
where S is the mid-surface domain before deformation. For a plate under a
pressure load p(x, y), the potential energy is:

I(w) = W(w) - Lpw dxdy (5.51)

For a plate under a point load (Pe z ) placed at (xo, Yo), we have:
I(w) = W(w) - Pw(XO,yo) (5.52)
As in previous chapters, expressions of I(w) can be used in connection with
numerical methods such as Ritz or the finite element method in order to find
approximate solutions to plate bending problems.
140 5. Theory of thin plates

5.13 Influence function

Consider a plate with arbitrary shape and B.Cs. Inside the domain, define a
rectangle U1 x U2 centered at (711, 712). The only load applied to the plate is
uniform pressure on that rectangle, Le.
• U1 U1 U2 U2
p(x,y) = po 1f (x,y) E [711 - 2,711 + 2] x [712 - 2,712 + 2],
=0 otherwise (5.53)

Assume that the corresponding solution wo(x, y) is found (e.g., by Levy's


method). The solution for a unit load acting at (711,712) can then be easily
deduced. In the expression of wo(x,y), it suffices to set POU1U2 = 1 and then
take the limits U1 ~ O and U2 ~ O. We designate by K(x, y; 711, 712) the end
result, Le. the deflection at (x, y) due to unit load at (711, 712)' Note that as a
consequence of Maxwell-Betti reciprocity theorem, we have: K(x, y; 711,712) =
K(711 , 712; x, y). From the unit load solution, or inftuence function, the solution
of more general loadings can be found by simple integral computation. For
example, for a surface pressure p(x, y), the deflection is:

(5.54)

For a line pressure q(s), the solution is:

w(x, y) = foi K(x, y; 711 (s), 712(S))q(s) ds (5.55)

The technique is similar to that of influence lines for beams (Sec. 3.10); for
applications, see (Timoshenko and W.-K., 1982) or (Crochet, 1994).

5.14 Solved problems

5.14.1 Uniformly loaded rect angular plate with two simply


supported opposite edges and two built-in edges

A rectangular plate is simply supported on edges (x = O) and (x = a), built-


in on edges (y = ±b/2), and subjected to a uniform pressure Pa. Following
the discussion in Sec. 5.11, the deflection is given by (5.47) after setting
Cn =Dn =0,
00
y
'""' [ a 4Pn
w(x,y) = L. (-) V
mry
+ Ancosh- +Bn- smh -mr ].
mry. n7rX
sm(-)
n7r a a a a
n=l
(5.56)
5.14 Solved problems 141

The Fourier series coefficients of the load are found from (5.44) as:

Pn =O 1'f n even num b er, Pn = -4po'f


n7r
1 n odd numb er (5.57)

Boundary conditions at edges (y = ±b/2) read:


b âw b
w(x, ±"2) = O, ây (x, ±"2) =O (5.58)

After some algebra, the foHowing results are found:


-If n is an even number, then An = Bn = O.
-If n is odd, then:

A __ 4poa 4 sinh (3n + (3n cosh (3n Bn = 4poa 4 sinh (3n ,


n- V7r5n5 (3n + cash (3n sinh (3n ' V7r 5n 5 (3n + cosh (3n sinh (3n
where:
(3 = n7rb
n - 2a
Consequently, the final expression of the deflection is:

w(x,y) L -
n5
1 [
1-
sinh (3n + (3n cosh (3n
(3n + cash (3n sinh (3n
h n7ry
cos - -
a
n=1,3,5, ...

+ sinh(3n n7ry . h -
--SIn
n7r- . (n7rx)
Y] S In-- (5.59)
(3n + cosh (3n sinh (3n a a a

From this expression, one can compute aH other variables. For instance, bend-
ing moments M 22 at one of the built-in sides are found to be:

M 22 (X ~) = _ 4poa2 " ~ sinh (3n cosh(3n - (3n sin n7rX (5.60)


,2 7r 3 L...J
n=1,3,5, ...
n 3 (3n + cosh (3n sinh (3n a

We expect these moments to be in the direction of e"" hence negative with


our sign convention; this can be verified. The maximum value of M 22 is found
at the middle of the side:
(_1)(n-l}/2 sinh(3ncosh(3n - (3n
n3 (3n + cosh (3n sinh (3n

Exercise: for a square plate, b = a, show that the maximum values of w and
M 22 are (with v = 0.3):
142 5. Theory of thin plates

5.14.2 Uniformly loaded rectangular plate with two simply


supported opposite edges and two free edges

As another application, we consider the same problem but with edges (y =


±b/2) /ree instead of clamped. The deHection has the general expression
(5.56). We only need to compute new values of coefficients An and Bn by
imposing free edge B.Cs. at (y = ±b/2), Le.

&w b â3 w b
â y 3 (x, ±'2) + (2 - v) âyâx 2 (x,±'2) = O
(5.61)

Exercise: work out alI the computations.


6. Bending of thin plates in polar coordinates

In Chap. 5, we studied bending of thin plates in Cartesian coordinates x and


y. However, there is an important class of problems which are more easily
formulated and solved in polar coordinates r and () (it is easier to define a
circle of center O and radius a by r = a then by x 2 + y2 = a2!)

6.1 Change of coordinates

In polar coordinates, the shear forces are defined by:

hn jhn
Qr =j arz dz, Qo = ao z dz, (6.1)
-hn -hn
and the bending moments are given by:

M rr =j
h/2
zarr dz, Moo =
jh/2
zaoo dz, Mro = MOr = j h/2 zaro dz
-h/2 -h/2 -h/2
(6.2)
In order to obtain the equations of bending of thin plates in polar coordinates,
a first method consists in working directly with a small sector (dr, d9), see
Fig. 6.1. For example, equilibrium of forces in the z-direction reads:

(Qr + â2r d; )(r + ~)dO _ (Qr _ â2r d;)(r _ d;)d9


âQ() dO âQ() d9
+(Q() + - - )dr - (Qo - - - )dr + prdrd9 = O
âO 2 âO 2
This equation simplifies to:

âQr +.!.(Qr+ âQ())+p=O


âr r âO
Equilibrium of forces in the radial and hoop directions gives two other scalar
equations and equilibrium of moments gives two more scalar equations. This
first method is used in other textbooks, e.g. (Timoshenko and Woinowsky-
Krieger, 1982). Another method (my favorite!) proceeds as follows.
144 6. Bending of thin plates in polar coordinates

Using the stress transformation equations -Appendix A- as well as Eqs.


(5.1-2) and (6.1), it is seen that the shear forces transform according to the
following relations:

Qr = Q", cos (O) + Q" sin O; Q() = -Q", sin(O) + Q" cos O (6.3)
We now introduce the following notation:

Q = [ ~: ], Q = [ ~: ] , (6.4)

in Cartesian and polar coordinates, respectively. The reader should not be


confused by notation (6.4): recall from Chap. 5 and Figs. 5.2 and 6.1 that
all shear forces (Q "" Q,,) or (Q r, Q()) act in the z-direction. The important
result shown by Eqs. (6.3) is that Q transforms like a vector.
Using the stress transformat ion equations -Appendix A- again as well
as Eqs. (5.3-4) and (6.2), it is seen that the bending moments transform
according to the following relations:

M rr = M",,,,cos2 0+M,,,,sin 2 0+2M,,,,,sin(0)cos(0)


M(J() M",,,, sin2 0+ M"" cos 2 0- 2M",,, sin(O) cos(O)
M r () (M""-M",,,,)sinOcosO+M,,,,,(cos2 0-sin2 0) (6.5)
We introduce the following notation:

M = [M",,,, M",,,], M = [Mr r Mr()], (6.6)


M",,, M"" Mr() M()(J
in Cartesian and polar coordinates, respectively. Again, the reader should not
be confused by notat ion (6.6). For instance, recall from Chap. 5 and Figs. 5.2
and 6.1 that M",,,, acts in the direction of elI and Mrr in the direction of e(J,
etc. Equations (6.5) show that ii transforms like a second-omer symmetric
tensor. Due to the transformat ion properties of Q and M, the equilibrium
equations (5.20a-b) can be written in the following symbolic or "intrinsic"
form:

div M - Q= O, div Q+ P = O (6.7)


And now, we simply use the formulae for the divergence operator in cylindri-
cal coordinates from Appendix A and directly find the following equations:
8Mrr ~ 8Mr(J Mrr - M(J() _ Q _ O.
8r+r80+ r r-,
8Mr(J + ~Mr() + ~8M(J(J - Q(J = O
8r r r80
8Qr +~(8Q(J +Qr)+P=O (6.8)
8r r 80
6.1 Change of coordinates 145

Note that Eq. (6.8c) is identical to the equation which was found by the
first method. Using intrinsic notations, the curvature relations (5.20d) can
be written as:

'K = -\7(\7w) (6.9)

where w(r, e) is the plate defiection, \7 is the gradient operator and 'K is
represented by symmetric 2 x 2 matrices which are form-identical to those of
(6.6). Using formulae in Appendix A, it is found that:

fPw 1 1 fPw âw â 1 âw
Krr =- âr 2 ; KOO = -;J:;: âe2 + 8r); KrO = KOr = - âr (:;: âe) (6.10)

Constitutive equations (5.20c) can be written as follows:

M ~ ~
15 = (1 - V)K + v tr('K)l = -(1 - v)\7(\7w) - v(.1w)l (6.11)

where .1 is the Laplacian operator and î is the 2 x 2 identity matrix. Using


Eqs. (6.10), it is found that (6.11) gives three scalar relations:

V
Moo
V
Mro
(6.12)
V
Stresses are related to bending moments by:

(j = 12z M (6.13)
h3
where (j is represented by a 2 x 2 matrix which is form-identical to that of
Eqs. (6.6). In component from, Eq. (6.13) gives three scalar relations:
12z 12z 12z
arr = -,;;JMrr ; aoo = -,;;JMoo; aro = -,;;JMro (6.14)

We have established in Chap. 5 that the bending problem obeys fundamental


equation (5.23). Since that equation is written in intrinsic format, it carries
out identically in polar coordinates:

.1.1w(r,e) = p(~e) (+ B.Cs.) (6.15)

From Appendix A, we have:

(6.16)
146 6. Bending of thin plates in polar coordinates

Finally, Eqs. (5.22) can be written under the following intrinsic form:

Q
- - = V(L1w) (6.17)
V
In component form, this gives two scalar equations:
Q
_--!.
a
= -(L1w)' Qo
--
1 a
= --(L1w) (6.18)
V ar ' V ra()

Fig. 6.1. Equilibrium of a small sector (dr, de). Bending moments, shear forces and
other notations.

6.2 Axisymmetric problems


Consider a circular plate of center (O). When the loading and B.Cs. present
an axial symmetry w.r.t. (O, z), the solution depends on the radial distance
(r) only, and aU partial derivatives w.r.t. (() vanish. Equations (6.12c) and
(6.8b) give:

Mro = MOr = O; Qo = O (6.19)

The non-zero internal loads are represented in Fig. 6.2. Among equations
(6.8), only two are not trivially satisfied:

dMrr + Mrr - M oo _ Qr = O; dQr + Qr + p = O (6.20)


dr r dr r
The curvature components -Eqs. (6.10) are given by:
6.2 Axisymmetric problems 147

tFw 1 dw
K.rr =- dr 2 ; K.(J(J = -; dr; K.r(J = K.(Jr =O (6.21)

The constitutive equations (6.12) become:

M rr tFw 1 dw M(J(J 1 dw tFw


-V = dr 2 + /1; dr; -V = ;: dr + /1 dr 2 ; (6.22)

The fundamental bending equation (6.15) becomes:

A Aw(r)
~~ = p(r)
V (+ B .Cs. ) (6.23)

where:

Equation (6.18) gives the non-zero shear force Qr as:

_ Qr = ~(Llw) (6.25)
V dr
In the remainder of the section, we give a general procedure for solving ax-
isymmetric problems. Equation (6.23) can be rewritten as:

~[r~(Llw)] = rp(r)
dr dr V
Simple integrat ion gives:

r~(Llw) =
dr
I r
xp(x) dx
V
+ A'

We divide each side of the equality by r and integrate:

Llw= I I r
d: Y
xp~x) dx+A'lnr+B'
We integrate this equation again, and so ono Finally, the deflection w(r) is
found to have the following general expression:

w(r) = I I
r
du
-;:
U
zdz IZ y I -Vdy Y xp(x) dx + Ar 2 + Blnr + C + Dr2 lnr
(6.26)

where A, B, C and Dare constants to be determined. The general method


consists in "cutting" (by thought only!) the plate along concentric circles
(actually cylinders of height h); in each region thus defined the deflection
has the general form (6.26). The constants of integration are computed from
the B.Cs., the continuity conditions at the interface between two adjacent
148 6. Bending of thin plates in polar coordinates

regions, and if a region contains the origin (r = O), the conditions for finite
solution in (r = O).
The case of uniform pressure p(r) = constant = po over a concentric
region is very common in practice; Eq. (6.26) gives then:

= Por
4
w(r) 641' + Ar2 + Bln r + C + Dr 2 lnr (6.27)

Simple derivation gives the following expressions which will be used later:

Por 3 B
161' + 2Ar + -:;: + Dr(l + 2lnr)
3por 2 B
161' + 2A - r 2 + D(3 + 2lnr)
P r2
.dw(r) :1' +4A+4D(1+lnr) (6.28)

The shear force is given by Eq. (6.25) as:


Qr por 4D
-Ti = 21' +--;- (6.29)

The bending moments are computed from Eqs. (6.22):

por 2 B
(3 + v) 161' + 2(1 + v)A - (1 - v) r 2
+ D[3 + v + 2(1 + v) In r];
por2 B
= (1+ 3v) 161' + 2(1 + v)A + (1 - v) r 2
+ D[l + 3v + 2(1 + v) In r] (6.30)

If the region under consideration contains the origin, then we have:

B = O (w(O) finite); D = O (Mrr(O) and Moo(O) finite) (6.31)

(Recall that when r -t O, In r -t -00 and rIn r -t O). As a consequence, the


previous results take a much simpler form:

) po r2
Por 4
+ Ar2 + C; -V
M rr (
w(r) = 641' = 3 + v 161' + 2(1 + v)A
Moo (1 + 3v )Por2 Q _ _ Por
V 161' + 2(1 + v)A; r - 2 (6.32)

6.3 Potential energy


It was shown in Sec. 5.12 that the strain energy of a plate is:
6.4 Solved problems 149

Fig. 6.2. Axisymmetric problem: non-zero internalloads.

(6.33)

where (S) is the mid-surface domain before deformation. In polar coordinates,


the curvature components K. cr{3(r, O) are obtained from the deflection w(r, O)
via Eqs. (6.10). For axisymmetric problems, they have simple expressions,
leading to a compact formula for W(w),

W(w) = 7r f [( tPw
V -
dr 2
1
+- dW)2 -
-
r dr
2(1- /1)---
1 tPw
r dr 2 dr
dW] rdr (6.34)

The potential energy I(w) is the difference between W(w) and the work of
externa! forces in the deflection field w(r, (}). Numerical methods based on
minimizat ion of I(w) can be used to find approximate solutions to plate
problems. A simple application is given in Sec. 6.4.

6.4 Solved problems

6.4.1 Uniformly loaded plate

Consider a circular plate ofradius (a) which is subjected to uniform pressure


Po. Since the plate contains the origin (r = O), then B = D = O and the
solution is given by Eqs. (6.31). The shear force is Q,. = -por/2. This result
can be double-checked: imagine a cut along a concentric circle of radius r,
equilibrium of forces gives:
2 por
po(7rr ) + Qr(27rr) = O => Qr = -2'
150 6. Bending of thin plates in polar coordinates

which is the same expression. The negative sign (Qr < O) means that Qr
is directed towards the top (Le., -e z ), which is expected. We consider two
cases. When the plate is simply supported at (r = a), the B.Cs. are:

w(a) = O; Mrr(a) =O
Using Eqs. (6.31), the constants A and Care found as:

A = _(3 + v) Po a2 • C = (5 + v) Po a 4
1 + v 32V' 1 + v 64V

Equation (6.31a) then gives the deflection as:

(6.35)

The maximum deflection is


5 + v Poa 4
W max = w(O) = (1 +) 64V
For instance, for steel v = 0.3 and W max ~ 4Poa 4 /(64V). The second case we
consider is that of a built-in plate at (r = a). The B.Cs. are:

dw
w(a) = O; -(a)
dr
=O
The constants A and Care found from Eq. (6.31) as:

Poa 2 Poa4
A = - 32V; C = 64V
This gives the deflection as:

(6.36)

The maximum deflection is

For steel, this value is four times smaller than that for simple support.

6.4.2 Uniform load along a concentric circle

In this section, we consider the problem of a circular plate of radius a, simply


supported at r = a and subjected to a uniform load P/(27fb) along a con-
centric circle of radius b < a; where P [N] designates the total load. This
6.4 Solved problems 151

problem (O) can solved by superposition of two simpler problems: (1') and
(2'), see Fig. 6.3.
Since a load per unit length is applied at r = b, the shear load Qr presents a
discontinuity there (the situation is identical to that of point or concentrated
loads in beams). Indeed, we have:

Qr =O if r < b and Qr = - 2P7rr if r >b


We now solve Problem (1'). Its solution WI' (r) must satisfy the following
equations (see Fig. 6.3):

..:1..:1Wl' (r) = O, r E [O, bl;


WI' (O) and M$~/) (O) finite; M$~/) (b) = M{
The general solution is given by (6.27) with Po = O, Bl' = D 1, = O. The B.C.
on the bending moment at r = b allows to find the following expression of
the deflection:
M{ Z
(6.37)
w1 / (r)=-2(1+v)V r +CI'

where the constants M~ and CII will be computed later. Equations (6.31)
show that the shear force is nil and the bending moments are uniform and
equal to M{:
Q~l/)(r) = O; M$~/)(r) = M~~/)(r) = M{
The solution WZ' (r) of Problem (2') must satisfy the following equations
(see Fig. 6.3):

..:1..:1WZ' (r) = O, rE [b, al;


Qr(2 /)(b) = -~. 27rb'
M(2 /)(b) = M'·
rr l'

Q~2/)(a)=-2:a; M$~/)(a)=O; wz,(a)=O

The general solution is given by (6.27) with Po = O. Constants Az" BZ', CZ',
and Dz, are computed from the B.Cs. written above. After some algebra,
their values are found as follows:
P aZbz [Pa ]
87rV; B ZI=(aZ _bZ)(l_V)V M{-47r(1+v)1o(b) ;
P [ 3 + v + 10]
--- a + (1 - v) B .
2 1

87rV 2(1 + v) 2(1 + v) aZ '

After substitution of these expressions ioto Eq. (6.27), the following expres-
sion for the deflection is found:
152 6. Bending of thin plates in polar coordiJ;lates

The shear load is given by Eq. (6.28) after setting Po = O as:


Q __ 4D 2, V _ -~
r - r - 27rr

We can check that this is correct by another method: imagine a cut along
a concentric circle of radius r E]b, al, static equilibrium of the ring enclosed
between the circles of radii b and r gives:
p
P + Qr(27rr) = O :::} Qr = --2
7rr
< O,
which is the same value. The negative sign means that Qr is directed towards
the top (Le., -e z ), which is expected.
It is seen that in the solutions of problems (1') and (2') there are two as
yet unknown constants: M{ and CI'. They are determined by requiring the
deflection and its slope to be continuous at r = b:

Using Eqs. (6.37-38), it is found that the constants have the following values:

Substitution of these expressions into Eqs. (6.37-38) gives the following ex-
pressions of the deflection for the original problem (O):

rE [O,b]: wo(r) = ~ [(a2 _ b2) (3 + v)a2 - (1- v)r 2 _ (b2 + r2) ln(~)]
87rV 2(1 + v)a2 b

rE [b,a]: wo(r) = ~ [( 2 _ 2) (3 + v)a 2 - (1 - v)b2 _ (b2 2) 1 (~)]


87rV a r 2(1 + v)a2 +r n r
(6.39)

Due to axisymmetry, the maximum deflection is reached in the center (r = O)


of the plate; its value is:
6.4 Solved problems 153

Pb 2 [ a2 (3 + v) a ]
wo(O) = CI' = 87rV (b2 - 1) 2(1+ v) - ln( b) (6.40)

The case of a concentrated load at (r = O) can be solved by setting b = O in


Eq. (6.39b):

P [( 2 2) (3 +
Wo (r ) = 87rV a -r 2(1+v) -r ln(;)
2 v) a] (6.41)

Exercise: Consider a problem identical to the one which was studied in this
section except that the plate is built-in at (r = a).

Pj27rb Pj27rb

a b
Pj27ra t+------~~ Pj27ra

Mi( ) Mi

+
Mi , (1
Pj27ra t Pj27rb! !
Pj27rb tPj27ra

Fig. 6.3. Problem (O): circular plate of radius a subjected to a uniform load along
a circle of radius b < a. The problem is split into two problems (1') and (2').

6.4.3 Uniform pressure on a concentric disk

In this section, we consider the problem of a circular plate of radius a, simply


supported at r = a and subjected to a uniform pressure Po = Pj(7rb2 ) on a
concentric disk of radius b < aj where P [N] designates the totalload. This
problem (1) can be solved by superposition of three simpler problems: (1'),
(2') and (3'), see Fig. 6.4.
Problems (1') and (2') have already been solved in Sect. 6.4.2, their solu-
tions are given by Eqs. (6.37-38). Problem (3') was also solved in Sect. 6.4.1,
154 6. Bending of thin plates in polar coordinates

its solution W3' (r) is given by Eq. (6.35) after making the substitution a ---+ b
and adding an arbitrary constant C3 , because W3' (b) =F O. The deflection
W3' (r) is therefore given by:

W3' ()
r Po- r2[r 2-
= -64V (3 + v) 2]
2---b + C3' (6.42)
1 +v

Superposition gives:

rE [O, b]: WI (r) = W3' (r) + wdr),


rE [b, a]: WI (r) = W2' (r), (6.43)

It appears that there are two constants: M{ and (C3 , + CI'); they are com-
puted from continuity requirement for the deflection and its slope at (r = b):

Exercise: Carry out the computations and give the final expressions of the
deflection WI (r) of the original problem.
We now present another method to solve the original problem (1). The
pressure load can be viewed as the integral of loads per unit length applied
along concentric circles of radius c. For r E [b,a], wI(r) is the integral from
C = O to C = b of the deflection wo(r) of Eq. (6.39b) where we make the
substitutions b ---+ c and P ---+ Po(27rcdc) (Timoshenko and W.-K., 1982):

r 2 ) (3 + v)a2 - (1 - v)c2
rE [b,a]: wI(r) = 10[b Po27rcdc
87rV
[(a 2 _
2(1 + v)a 2
_(c2 +r 2 )ln(;)]

Computing the integral, rearranging terms and setting Pa = Pf(7rb2 ) at the


very end, we obtain:

rE [b,a] : wI(r) ~ [(a 2 _ r 2 ) 2(3 + v)a2 -(1 - v)b2


167rV 2(1 + v)a 2
_(b2 + 2r2 )ln(;)] (6.44)

The case of a concentrated load at (r = O) corresponds to b = O; it can be


checked that Eq. (6.41) is found.
Now for r E [O, b], wI(r) is the sum of two contributions. The first one is
the integral from c = Oto c = r ofthe deflection wo(r) ofEq. (6.39) where we
make the substitutions b ---+ c and P ---+ Po(27rcdc). The second contribution
is the integral from c = r to c = b of the deflection wo(r) of Eq. (6.39a) where
we make the same substitutions b ---+ c and P ---+ Po(27rcdc). Thus, wo(r) is
computed as follows:
6.4 Solved problems 155

l' E [O, b]: wl(1') = r Po27rcdc


10 87rV
[(a 2 _ 1'2) (3 + v)a2 - (1 - v)c2
2(1 + v)a 2
_(c2 + 1'2) ln(;)]
+ r P027rcdc [(a
b

lr 87rV
2 _ c2 ) (3 + v)a2 - (1- v)r 2
2(1 + v)a 2
-(c2 + 1'2) ln(~)]

Exercise: Compute the integral, rearrange terms, set Po = PJ(7rb2 ) at the


end, and give the final expression of the deflection.
It is found that the maximum deflection is:

(6.45)

Equations (6.40) and (6.45) show that Wl(O) > wo(O), Le. for the same total
load P and the same loading zone size, the deflection due to a pressure is
larger then that due to a load distributed along a circle.
For uniform pressure everywhere on the plate, it suffices to set b = a. It
can be checked that Eq. (6.35) is found.
Exercise: Consider a problem identical to the one which was studied in
this section except that the plate is built-in at (1' = a).

6.4.4 Plate simply supported on a number of points

A circular plate of radius a is subjected to transverse and axisymmetric forces


of resultant (Pe z ) and simply supported on a number N ;::: 2 of points along
the boundary. The equi-distant supports are situated at angles O = O, "(, 2"(,
etc. The angle between two neighboring supports is "( = 27rJN. The problem
is not axisymmetricj we shall show that it can be solved by superposition of
three problems: (Po), (Pt) and (P2 ) •

• Problem (Po): the plate is subjected to the same transverse and axisym-
metric forces of resultant (Pe z ), but is simply supported along the whole
boundary. The deflection wo(r) is solution of the following problem

ddwo(r) = (load dependent), l' E [O,a[,


wo(a) = O, M;~)(a) = O, (6.46)

where the right-hand-side of (6.46a) depends on whether the load is dis-


tributed over a surface or a line. We may also need to distinguish several
regions where each one has an associated Eq. (6.46a) and add interface conti-
nuity equations to system (6.46b). Anyways, problem (Po) is an axisymmetric
problem that can be solved by the general technique of Sec. 6.2, and examples
156 6. Bending of thin plates in polar coordinates

M~( ) M~ (1 ')

+
M{) ( M{

Pj21fa t Pj21fb! ! Pj21fb t Pj21fa


(2')

+
(3')

P/2WbFt/:b
Fig. 6.4. Problem (1): circular plate of radius a subjected to a uniform pressure
on a disk of radius b < a. The problem is split into three problems (1'), (2') and
(3').
6.4 Solved problems 157

are given in Secs. 6.4.1 to 6.4.3. The reactions along the boundary (r = a)
are given by statics as:
p
- 2- e z =-Te
- z (6.47)
7ra

• Problem (Pl): the plate is subjected to forces per unit length p(O)e z
along its boundary, defined as follows:
- (i) At and near the supporls, forces per unit length (-Fe z ), F> 0, are
applied along an arc of vanishingly small opening angle X.
- (ii) Along the rest of the boundary (Le., between supports), forces per unit
length (Te z ) are applied.
The system of forces is depicted in Fig. 6.5. Firstly, static equilibrium
gives the following equation

from which the expression of Fis deduced:

F = (1- 1)T (6.48)


X
Secondly, it appears that the load function p(O) is periodic, of period 'Y and
can be studied over the interval [O, 'Y1 where it is defined as follows:
X
p(O) = -F, OE [°'"21;
p(O) = T, X 'Y - _[o
O E1- X
2' 2'
X
p(O) = -F, O Eh - 2,'Y1 (6.49)

It is dear that p(O) is continuous over [O, 'Y1 except on a finite number ofpoints
= =
(O X/2 and O 'Y - (X/2)) and is square-integrable over the interval. From
the above observations, it is conduded -e.g., (Tolstov, 1962)- that p(O) can
be expanded in a Fourier series as follows:

ao ~ 27rkO ~ . 27rkO
p(O) ="2 + LJakcos-- + LJbk sm - - (6.50)
k=l 'Y k=l 'Y

Standard formulae -e.g., (Tolstov, 1962)- give coefficients ao, ak and bk,

ao =
21"(
- p(O) dO; ak = -
21"( 27rkO
p(O)(cos - - ) dO;
'Yo 'Yo 'Y
bk ~ r p(O)(sin 27rkO) dO
'Y lo 'Y
158 6. Bending of thin plates in polar coordinates

Using Eqs. (6.49-50), and recalling that N = 27rh, the following values are
obtained:
sin(Nkx/2)
ao = O; ak = -2T Nkx/2 ; bk =O (6.51)

Because of the symmetry of the problem, it was predictable that bk = O.


The third step in problem (H) consists in taking the tirnit X ----t O. Since
ak ----t -2T, it appears that:
00

p(O) ----t Pl(O) == -2TLcos(NkO) (6.52)


k=l

Finally, we are ready to state and solve problem (Pt): find the deHection
wt{r, O) such that:

-d-dWl (r, O) = O, (r, O) E [O, a[ x [O, 27r[;


M(l)
rr (a' O) = O',
1âM(l)
Q~l)(a,O) + ;'Ţ(a,O) = Pl(O) (6.53)

We look for a solution under the following form:


00

wl(r,O) = LWn(r)cos(nO) (6.54)


n=l

From Eq. (6.53a), the following expression is found:

wl(r,O) =
3
(Alr+Blr +-+Dlrlnr)cosO
Cl
r
00

+ L(Anrn + B n r n+2 + Cnr- n + Dnr-n+2) cos(nO)


n=2

In order to find a finite solution at (r = O), we need to set Cl = Dl = O and


Cn = Dn = O. The expression of the deHection simplifies to:
00

Wl (r, O) = L(Anrn + Bnrn+2) cos(nO)


n=l

Constants An and Bn are found from the B.Cs. (6.53b-c). After some lengthy
algebra, the following final expression of the deHection is found:

Wl (r, O) =
Ta 3
"
V(3 + II) n-N,2N,3N,
_ LJ ... n(n - 1)
[1 2(1+v)
+ ...,.--~.,....-"":""""'-=-
(1 - 1I)(n - 1)n2

_ (rla? ] (:')ncos(nO) (6.55)


n(n + 1) a
6.4 Solved problems 159

The supports are situated at r = a and () = 0, ,,/, ... , (N -1)"{; the deflection
there is

Wl(support) = 2Ta3 E
(3 + II)V n=N,2N,3N, ... n(n -1) n
1 [1+ 1
1 + II ]
+ (1 - lI)n
(6.56)

=f. 0, which is only normal since the only B.Cs.


It is seen that Wl(SUPport)
which were imposed in problem (Pt) are (6.53b-c).

Problem (P2 ): it is such that, when superposed to (Po) and (Pt), the solution
of the original problem is found. In view of the previous remark concerning
Wl (support) =f. 0, we simply take for the solution w2(r, () the following rigid
displacement:

w2(r,() = constant = -Wl(support) (6.57)

• Superposition: the solution of the original problem is obtained as


follows:

w(r, (}) = wo(r) + wl(r, (}) - wdsupport), (6.58)

°
Firstly, since ..1..1Wl (r, () = and wo(r) satisfies -by definition- the in-bulk
equilibrium equation(s) (6.46a), then w(r) satisfies the latter equation(s) also.
Secondly, since wo(a) = 0, then it is clear that w(support) = O. Thirdly, the
bending moment B.C., Mrr(a,{}) = O, being satisfied by wo(r) and wl(r,(), it
is also satisfied by w(r, (}). Finally, the generalized shear force on the boundary
is given by:

Outside the point supports, Pl({}) = T, therefore Vr = O. In conclusion, the


solution ofthe original problem is indeed given by (6.58). Since wl(r = O) = 0,
the deflection at the center of the plate is given by the simple formula:

w(r = O) = wo(O) - Wl(SUPport) (6.59)

• Application: consider a uniformly loaded plate (P = Poll'a 2 ) which is


simply supported on three points (N = 3, "/ = 21l' /3). The solution of problem
(Po) is given by (6.35). The deflection at the center is:

w(r = O) = Pa 2
641l'V
(51 ++ II)
II +
Pa2
(3 + lI)ll'V
['"
n=~9, ... n(n2 -
1 1)
+ (1 -
1+11
II)
'"
n=~9, 1
... n 2(n - 1)
1 (6.60)
160 6. Bending of thin plates in polar coordinates

If we take V = 0.25 and stop at the first five terms of each sum, we find

Pa 2
w(r = O) ~ 0.036""""1) ~ 1.7wo(0)

The result shows once again the importance of the B.Cs.

-F

-F -F

-F
Fig. 6.5. Problem (Pt): forces per unit length (-Fe.) are applied at and near
the supports (small angle X), and (Te.) along the rest of the boundary (between
supports).

6.4.5 Ritz's method

As a simple application of Ritz's method, consider a circular plate of radius


(a), clamped along the boundary and subjected to an axisymmetric load. We
look for an approximate solution under the form:

(6.61)

where A, B and G are constants to be determined. In order to apply Ritz's


method, Wl (r) must be kinematically admissible (K.A.), Le. be "sufficiently
smooth" and satisfy the B.Cs.:

dWl
wl(a) = O; dr (a) =O (6.62)

Simple computation gives A = Ga3 /2 and B = -3Ga/2, thus the approxi-


mate deflection has the following expression

(6.63)
6.4 Solved problems 161

Consider a uniform pressure qo. It can be checked that W1 (r) is not the exact
solution of the problem because

The potential energy can be written as follows:

(6.64)

where the strain energy W(wt) is given by (6.34). Computing the integrals
in (6.64), one finds:

9a 4 3a5
I(w1) =4 7rVC2 - w7rqoC (6.65)

Constant C is computed such I(w1) is minimum, 8I/8C = O. This gives:


C = qoa/30V. The maximum deflection occurs at the center and has the
value: qoa 4 /60V. Compare to the exact expression given by (6.36), qoa4 /64V.
As a second application, consider a point force Fo at (r = O). The only
change is the work of this force, thus (6.64-65) become:

Condition 8I/8C = O gives C = Fo/97raV. Consequently, the deflection at


the center is Foa 2/187rV, compared to the exact expression Foa 2/167rV.
7. Two-dimensional problems in Cartesian
coordinates

In this chapter, we study an important class of problems, namely plane strain


or stress. These are cases such that the original problem can be solved in a
two-dimensional rather than a three-dimensional space.

7.1 Plane strain


Consider a long circular tube submitted to a uniform internal pressure and
whose end sections are not allowed to move in the axial direction (Fig. 7.1). It
is clear that each cross section (except perhaps near the end sections) is in the
same state, Le. the problem is independent of (z). More generally, consider
a solid which is long in one direction (z), subjected to a load which does
not vary with (z) and whose two end sections have a zero z-displacement.
Moreover, the displacement field is assumed to satisfy:
(7.1)
The above-mentioned assumptions define a plane strain problem in the (x, y)
plane. In this chapter, Greek indices (a, (3, 'Y) take values 1 and 2, summation
over repeated indices is assumed unless otherwise indicated, and subscripts
(1,2) are used for the x and y directions, respectively (e.g., Xl = X, U2 = U y ).
Using Eqs. (7.1), the following strain relations are easily found:
(7.2)
Isotropic linear elasticity gives the stresses as:
,\
0"0/{3 = 2ţLf.0/{3 + '\f."("(oO/{3; O"zz = 2('\ + ţL) 0""("(; O"yz = O"xz = O (7.3)

where ,\ and ţL are Lame's coefficients. Using Eqs. (7.2), it is deduced that

O" 0/{3 = O" 0/{3 (Xl, X2)


Equilibrium gives the following equations:

80"0/{3
8x{3
+ f 0/ = O·
,
f Z =
O (7.4)
164 7. Two-dimensional problems in Cartesian coordinates

It is seen that another assumption is necessary: the volumetric forces should


have no z-component.
If we seek a solution in terms of strains, we need to satisfy the compatibil-
ity equations of Sect. 1.9. It is easily found that there is only one non-trivial
compatibility equation to be verified:

2 82 t:12 _ 82 t:11 + B2t:22 (7.5)


8x 1 8x2 - 8x~ 8x~
We shall see later on that for plane problems, it is the stress field which is
usually sought, therefore the Beltrami-Mitchell equations need to be satisfied.
Instead of using the formulae of Sect. 1.14, it is perhaps easier to directly
rewrite Eq. (7.5), using Eqs. (7.3-4). Doing so, the compatibility equation is
found to take the following form (where v is Poisson's ratio):

(7.6)

-x

Fig. 7.1. Circular tube under uniform interna! pressure

7.2 Plane stress


Consider a thin plate (of thickness h) subjected to a loading obeying the
following conditions: the top and bottom surfaces (z = ±hj2) are stress-free,
fz = O and fx and fy are independent of z (see Fig. 7.2). Moreover, the
following assumptions are made on the stress field:
(7.7)

The above-mentioned assumptions define a plane stress problem in the (x, y)


plane (more generalloading conditions will be considered in Sect. 7.7). Two
scalar equilibrium equations need to be satisfied:

8ua {3 + fa = O (7.8)
8x{3
7.3 Summary: plane strain versus plane stress 165

Isotropic linear elasticity gives the strains as:

(7.9)

Using Eqs. (7.7), it is deduced that

If we seek a solution in terms of strains, we need to satisfy the compatibility


equations of Sect. 1.9. It is easily found that there are four non-trivial equa-
tions to be verified, one of them being (7.5). For plane problems, the search
method usually consists in finding the stresses first, therefore, we shall rewrite
Eq. (7.5) in terms of stresses. Using (7.8-9), Eq. (7.5) can be rewritten in the
following form:

(7.10)

In addition to (7.5), the other three compatibility equations which are not
trivially satisfied are the following:

8 2f33 = Oj 8 2f33 = Oj 8 2f33 = O


8Xl 8X2 8x~ 8x~

These equations imply that f33 is a linear function of Xl and X2, which is
false in general. In practice, these three equations are ignored and plane stress
theory, although approximate, compares well with exact solutions (when they
can be found)j see (Timoshenko and Goodier, 1987).

7.3 Summary: plane strain versus plane stress

The equations of plane strain and plane stress found in Sections 7.1 and 7.2
are summarized hereafter.

f23 = f13 = O
0"23 = 0"13 = O
u'" = X2)U", (Xl,

f",{3 = X2)
f",{3 (Xl,

O" ",{3 = O" ",{3 (Xl, X2)


fz=Oj f",=f",(XI,X2)
f
'"
(3 = l(~+~)
2 8Xf3 8x",
811",8 +1 = O
8xf3 '"
166 7. Two-dimensional problems in Cartesian coordinates

Fig. 7.2. Thin plate under a plane stress loading

Plane strain Plane stress


U3 = O U33 = O
f33 = O ~ U33 = VU-Y-Y U33 = O~ f33 = -l~vf-y-y
f Ot{3 =
~(
E u Ot{3 - ~
vU-y-yUOt{3 ) f
_ ~(
Ot {3 - E u Ot{3 - v U -y-yUOt{3
l+v ~)

82 82 l!!li.. 82 82 !!li..
(&f + ~)u'Y'Y = -(1+V)8"'ţ3
_
(8"'~ + ~)U-y-y - -1-v8"'ţ3
It is seen that most of the equations are identical. Actually, if a change of
parameters is made, the last sets of equations (the f Ot{3 - u Ot{3 constitutive
relations and the compatibility equations) become also identical. The reader
can check the following observations:
- (i) IT we solve a plane stress problem, we can find the solution of a corre-
sponding plane strain problem by making the following substitutions:
v --t _v_ E - - t - o
E-
l-v' l-v2
- (ii) When we have the solution of a plane strain problem, we can find that
of plane stress with the following substitutions:
v --t _v_o E --t E(l + 2v)
l+v' (1+v)2
It is remarkable that problems which correspond to completely diJJerent geo-
metric situations in the z direction (very long body for plane strain and very
thin plate for plane stress) obey basically the same equations.
7.4 Airy stress function 167

7.4 Airy stress function


In most of this chapter, body forces are neglected (we shall see in Sect. 7.8.6
how to take them into account). The equilibrium equations for a plane (strain
or stress) problem are then:

âuzz + âuzy = O. âuzy + âuyy = O (7.11)


âx ây 'âx ây
Equation (7.11a) shows that (uzzdy - uzydx) is a total differential, Le.
:3 F(x,y) such that

This implies that:


8F âF
u zy = - 8x j u zz = ây
Similarly, Eq. (7.11b) shows that (uzydy - uyydx) is a total differential, Le.
:3 G(x,y) such that
U yy = -8G
-j
8x
u zy = -8G
ây
Equating the two expressions of U zy shows that (Fdy - Gdx) is a total dif-
ferential, Le. :3 ifJ(x, y) such that

G---' F--
8ifJ 8ifJ
- âx' - 8y
Substitution of these expressions into those of U zz , U yy and uzy gives:

(7.12)

The function ifJ(x, y) is called an Airy stress junctionj Eqs. (7.12) insure that
the equilibrium equations are satisfied. The compatibility equations of Sect.
7.3 become (in the absence of body forces):

82 â2 â 2 ifJ 8 2 ifJ .
(8x 2 + â y2)(âx2 + â y 2)=Oj l.e. LlLlifJ(x,y) =0 (7.13)

This means that the Airy stress function must be bi-harmonic.

7.5 Polynomial solutions


We present hereafter a few Airy functions in the form of homogeneous poly-
nomials. We shall see in Sec. 7.8 that for rectangular domains in the (x, y)
168 7. Two-dimensional problems in Cartesian coordinates

plane, solutions can be found by a judicious combination of polynomial terms.


Let lP2 be a homogeneous polynomial of degree 2:
1 1 2
4J2(x, Y) = "2a2x 2
+ ~xY + "2C2Y (7.14)

This function is always bi-harmonic and the stresses are given by:

(7.15)

This is a state of uniform stress. If the domain is a rectangle (x, y) E [-a, a] x


[-b, b] then the surface forces applied on the sides x = ±a are ±C2ex ~ ~ey
while those applied on the sides Y = ±b are ~~ex ± a2ey.
A homogeneous polynomial lP3 of degree 3,

13 1 2 1 2 1 3
lP3(X, y) = 6a3x + "2b3X Y + "2C3XY + 6d3Y (7.16)

is bi-harmonic and the stresses are given by:

The stresses vary linearly in space. On the sides x = ±a of a rect angular


domain, the surface forces are (C3a ± d3y)e x + (-b3a ~ c3y)ey while the sides
Y = ±b are subjected to (~b3X - c3b)ex + (±a3x + bab)ey.
A homogeneous polynomial lP4 of degree 4,
1 4 1 3 1 22 1 3 1 4
lP4 (x, y) = 12 a4 x + 6 b4X Y + "2 C4X Y + 6 d4XY + 12 e4Y (7.18)

is bi-harmonic if and only if:

(7.19)

The stresses are then given by:

(7.20)

A homogeneous polynomial lP5 of degree 5,


1 5 1 4 1 32 1 23 1 4 1 5
lP5(X,y) = 20a5x + 12b5x Y + 6 C5X Y + 6 d5X Y + 12 e5xy + 20/5Y
(7.21)

is bi-harmonic if and only if:

3a5 + 2C5 + e5 = O; b5 + 2d5 + 3/5 =O (7.22)


7.6 Solution by Fourier series 169

The stresses are then given by:

1
'3csx
3
+ dsx 2 y - (3as + 2cs)xy -
2
'13 (bs + 2ds )y3 ;
1
= asx 3 + bsx 2y + csxy2 + '3 dSy3 ;
1
= 3
-'3bsx - CsX Y - dsxy
2 2
+ '13 (3as + 2cs)y 3 (7.23)

7.6 Solution by Fourier ser ies

Fourier series allow the study of very general loading cases, such as loadings
with discontinuities; see Figs. 7.3 and 7.9 for examples. Let us consider a
rather general case depicted in Fig. 7.3b: a beam of length Z and height c is
subjected on its lower and upper faces (y = ±c/2) to arbitrarily distributed
loadings q+(x) and q-(x), which can be expanded using Fourier series:
± (Xl (Xl 2
q±(x) = q~ + Lq;cosmx + LQ;sinmx, m== ~7r (7.24)
n=l n=l
Equations (7.24) suggest that we look for an Airy function under the form:

<p(x, y) = <Po(x, y) + f:Jn(y) cosmx +


n=l
f:
n=l
Fn(Y) sin mx, m == 2~7r (7.25)

where we have used the method of separation of variables. We now look for
conditions under which the following function:
2n7r
CPn(x, y) = fn(Y) cos mx, m == -Z-

is an Airy function. The computation is identical to the one we carried out


in Sec. 5.11 for Levy's method. Consequently, it is found that:

fn(y) = Qn cosh(my) + .8nmysinh(my) + 'l'nsinh(my) + dnmycosh(my)


(7.26)

where Qn, .8n, 'I'n, and dn are constants. The stresses correspondingto CPn(x, y)
are determined from the usual relations:

8 2CPn 8 2CPn 8 2CPn


(uxx)n = 8y2 ; (uyy)n = 8x 2 ; (uxy)n = - 8x8y
and are found to be:
170 7. Two-dimensional problems in Cartesian coordinates

m 2 (cosmx)[a n cosh my + f3n(2 coshmy + my sinh my) +


'Yn sinh my + on(2 sinh my + my cosh my)]
-m 2 (cosmx)(a n coshmy + ,Bnmy sinh my +
'Yn sinh my + onmy cosh my)
m 2 (sin mx)[an sinh my + f3n(sinh my + my cosh my) +
'Yn coshmy + on(cosh my + mysinhmy)] (7.27)
In a similar manner, the stresses corresponding to the Airy stress function
CPn(x, y) = Fn(Y) sin mx can be found. To do so, replace (an, f3n, "(n, On) with
(An, Bn, C n , D n ), cos(mx) with sin(mx) and sin(mx) with [- cos(mx)] in
Eqs.(7.27). Consequently, stresses (Sxx)n, (Syy)n and (Sxy)n are found. Now
the stresses due to the Airy function (7.25) are given by:
â 2 cpo 00 00

Uxx ~ + ~(Uxx)n+ ~(Sxx)n


y n=l n=l
â 2 cpo 00 00
U yy âx 2 + ~(Uyy)n + ~(Syy)n
n=l n=l
â 2 cpo 00 00
U xy = - âxâ + ~(Uxy)n + ~(Sxy)n (7.28)
y n=l n=l
The constants can be found from the stress B.Cs.

uxy(x, ±~) = O; Uyy(x, ±~) = -q±(x)

An application of the Fourier series method is given in Sec. 7.8.

7.7 Generalized plane stress

Consider a thin plate of uniform thickness (h) whose middle plane is (X3 = O).
When defining plane stress, we made assumptions (7.7) on the stress field and
assumed that the in-plane forces are independent of the thickness coordinate
(X3)' In this section, we make less restrictive assumptions (see Fig. 7.4). Our
presentation mostly follows (Crochet, 1993). The hypotheses are:
- (H1) The upper aud lower surfaces (X3 = ±h/2) are stress-free. This implies
the followiug three scalar equatious:

(7.29)

- (H2) The surface forces acting ou the lateral bouudary are symmetric W.r.t.
the middle plaue (X3 = O) aud have no through-thickness compouent:

(7.30)
7.7 Generalized plane stress 171

x
..

y
Fig. 7.3. Examples of beams under generalloadings

- (H3) The body forces are symmetric w.r.t. the middle plane (X3 = O) and
have no through-thickness component:

(7.31)

We now define the mean value through the thickness of a quantity as follows:

_ 1 jh/2
~(x!, X2) == h ~(Xl, X2, X3) dX3
-h/2
From the symmetry of the problem, we see that the transverse displacements
are anti-symmetric w.r.t. (X3):

This implies that the average transverse displacement is zero:

(7.32)

The transverse shear strains are defined by:

aUa a U3
2Ea3 = --
aX3
+--
aXa
Their mean values are given by:
172 7. Two-dimensional problems in Cartesian coordinates

and they vanish because of symmetry and Eq. (7.32). We now write equilib-
rium in the 3-direction for X3 = ±h/2,

8a33 h 8a31 h 8a32 h


- 8 (Xl,X2'±-)=--8(Xl,X2'±-2)--8 (Xt,X2'±-2) =0,
~ 2 , ~ , , ~
T~
,
o o
because of Eqs. (7.29a). We see that a33 is nil on the top and bottom surfaces
and has zero gradient in the thickness direction near those surfaces. Therefore,
for a thin plate, it is reasonable to make the following assumption:

(7.33)

Taking the average of the constitutive equations, we find:


E Ev
= 1 + v €0I/3 + (1 _ 2v)(1 + v) (tn + (33)00l/3;
E _ Ev __
1 + v t33 + (1 _ 2v)(1 + v) (t..,...,. + (33) = O;
E _
= 1 + V t0l3 =O (7.34)

The last three scalar equations (0'33 = (113 = (123 = O) are form-identical to
the plane stress equations. Since (133 = O, the first four scalar equations in
(7.34) can be rewritten to give:
_ 1 + v(_ v _ ~ ) _ v _
t Ol/3 = - E a 01/3 - -l-- a ..,...,.UOl/3; f33 = --l-- f ..,...,.
+v -v
Again, note that these average stress-strain relations are form-identical to
the plane stress equations of Sect. 7.3. We now turn to the three equilibrium
equations, which can be written as follows:

8aOl/3
--+--+ 8a0l 3 f 01= O; --+--+
8a3/3 8 a 33 f 3= O
8x/3 8X3 8x/3 8X3
Taking the average of these equations, we obtain:

a = 1,2;

Using previous results, it is seen that the third scalar equation is identically
satisfied while the first two are reduced to:
0a0l/3 -
-8-+ 101 =0, (7.35)
X/3
7.8 Solved problems 173

and these are form-identical to the plane stress equilibrium equations. Finally,
compatibility equation (7.5) integrated through the thickness gives:

2 fj2'f.12 = 82(11 + 82(22 (7.36)


8Xl 8X2 8x~ 8x~

In summary, the averages over the thickness of stresses and strains have
to satisfy equations which are identical to those of plane stress. This class
of problems is called generalized plane stress. Note that the "membrane"
problem of Sect. 5.7 is such that the in-plane strains (f a j3) and internalloads
(Na j3) defined there satisfy a generalized plane stress problem, with Na(j/h =
(j a(j·

~_._._._.:.J._._._._._.~
Fig. 7.4. Plate under generalized plane stress conditions. In-plane forces are sym-
mmetric w.r.t. the thickness.

7.8 Solved problems


7.8.1 Concentrated load at the end of a cantilever beam

Consider a straight cantilever beam of rectangular cross section which is built


in at its end (x = O) and subjected to a concentrated force P at the other end
(x = l) (Fig. 7.5). This is a problem we have already solved using beam theory
174 7. Two-dimensional problems in Cartesian coordinates

in Chap. 3, our purpose here is to solve it again using a more sophisticated


theory (namely plane elasticity) and compare the two solutions. The first (and

-c/2

-b/2~
+c/2

y ~
Fig. 7.5. Concentrated load at the end of a cantilever beam

most difficult) step consists in finding an Airy stress function for the problem.
Even though the search is restricted to polynomial functions, the difficulty
consists in choosing the right terms (since no homogeneous polynomial will
do). We start by writing the stress B.Cs. On the lower and upper surfaces
(y = ±c/2), the body is stress free:

(7.37)

On the end section (x = l) we shall write conditions on the stress resultants


and use Saint-Venant's principle. Actually, it is useful to find expressions for
the stress resultants in each cross section (x) with outward unit normal ex ;
the resultant of the stress vector is:
C/2 jb/2
+ uxye y ) dydz
R(x) j
-c/2 -b/2
(uxxe x

b ( j
ex
O/2
U xx dy +ey
lc/2)
dy ,
U xy (7.38)
-c/2 -c/2

since the u a {3's do not depend on z. The moment of the stress vector w.r.t.
the centroid of the cross section is:

M(x)

j C/2 YU
-be z
-c/2 xx dy, (7.39)

since f~~2 z dz = O. On the other hand, static equilibrium of the left part of
the beam gives:
7.8 Solved problems 175

R(x) = Peyj M(x) = P(l- x)e z (7.40)

Equating the expressions in (7.38-39) with those in (7.40), we obtain three


scalar equations:

C/2 jC/2 jC/2


j a xx dy = Oj b a xy dy = Pj -b ya xx dy = P(l - x) (7.41)
-c/2 -c/2 -c/2

These equations put restrictions on the stress expressions. For instance, the
expression ofaxx should contain odd powers of y, which may be multiplied
by any power of x, because the integral in (7.41a) will then be automatically
satisfied. This is written under the following symbolic form:

Using a similar reasoning, we find that Eq. (7.41c) implies that:

ya xx : (yO,y2,y4, etc. ) and (xyO,xy2,xy4, etc. )

Combining the two conditions implies the following form for a xx

a xx : (y, y3, y5, etc. ) and (xy, xy3, xy5, etc. ) (7.42)

Equation (7.41b) gives the following form for the shear stress:

(7.43)

We now need to construct an Airy stress function which will give us stress
expressions of the form (7.42-43). From Sect. 7.5, it is seen that the term
d3y3/6 gives a~3) = d 3y, while the term d4xy3/6 gives a~~ = d 4xy and
a~V = -d4y2/2. In order to satisfy B.Cs. (7.37a), the expression of aX!!
should cont ain a constantj a term ~xy in Airy's function gives a~~ = -b2 •
In summary, the following Airy stress function is proposed for the problem:

(7.44)

The stresses are given by Sect. 7.5 as:

(7.45)

The function in (7.44) is bi-harmonic and the stresses (7.45) satisfy the equi-
librium equations. The constants d3 , d4 and b2 are determined from the B.Cs.
These are Eqs. (7.37) and additional B.Cs. obtained by setting (x = I) in
Eqs. (7.41). Equations (7.37b) and (7.41a) are satisfied, while Eqs. (7.37a)
and (7.41b) imply that:
176 7. Two-dimensional problems in Cartesian coordinates

~ ~e2
d4 = 1; b2 = - 81

where 1 == be3 /12 is the moment of inertia of the cross section w.r.t. the
z-axis. Equation (7.41c) written for (x = l) gives:
~l
d3 =-Ţ

Note that Eqs. (7.41) are verified for any value of x. Substitution of the values
of the constants into Eqs. (7.45) gives the stresses as:

~ ~ e2
Uxx = -I(l- x)y; Uyy = O; uxy = 21(4 - y2) (7.46)

These stress expressions are valid for plane stress (u zz = O) as well as plane
strain (E zz = 0, Uzz = vu"'("'().
In the built-in section (x = O), we have f~~2UXX dy = O, but uxx(O,y) =
-~ly/I i- 0, therefore the solution is not correct at (x = O). On the other
end section (x = l), although uxx(l,y) = O, the solution is correct only ifthe
load ~ is applied according to Eqs. (7.46c). Saint-Venant's principle implies
that Eqs. (7.46) represent the stress solution of the problem far from the
ends (assuming that e > b, a practical requirement is x E [e, l - el, and the
larger the ratio l/e, the better the solution).
It is interesting to compare the stress solution to that found from ele-
mentary beam theory (Chap. 3). The shear force Q(x) and bending moment
M(x) are given by:
Q(x) =~; M(x) = -~(l- x)
The normal and shear stresses are given Sec. 3.6 as:

(O) _ M(x). (O) _ Q(x) (e 2 _ 2)


Uxx - 1 y, uxy - 21 4 Y

Comparison with Eqs. (7.46) shows that for the problem under consideration,
plane elasticity and elementary beam theory give the same stress solutions.
This is not always the case however, as we shall see in the upcoming sections.
In the remainder of this section, we consider plane stress (the thickness b is
assumed to be small). Section 7.3 and Eqs. (7.46) give the strains as:

~ v~ (1 + v)~ e2 2
Exx = - EI (l- x)y; Eyy = EI (l - x)y; Exy = 2EI (4 - y) (7.47)

Integration ofthe relations Exx = âu/âx and Eyy = âv/ây gives the displace-
ments u(x,y) and v(x,y) as:
x2 y2
Elu = -~y(lx - 2") + F(y); Elv = v~(l- x)2" + G(x)
7.8 Solved problems 177

The displacements are related by (8uj8y) + (8vj8x) = 2€xyj using Eq. (7.46c)
and the expressions of u and v, it is found that:

x2 dG c 2 v dF
-P(lx - - )
2
+-
dx
= (1 + v)P-
4
- (1 + _ )py2 - -
2 dy
The left-hand side of the equality is a function of x only, while the right-hand
side is a function of y onlYj for the equality to hold for alI values of x and y,
each side must be equal to the same constant (k, say). Integration of the two
equations thus obtained gives:

Constants k, A and Dare computed from the displacement B.Cs. Let us


choose the folIowing B.Cs. (which correspond to a built-in condition at x = O
in the beam theory sense):

8v
u(O, O) = Oj v(O, O) = Oj 8x (O, O) =O
These conditions imply that k = O, A = D = O. Therefore, the displacement
field is given by:

Elu(x,y)

Elv(x,y)

The displacements of the middle fiber (y = O) are:


P x2 x3
u(x,O) = Oj v(x,O) = EI(l"2 - 6)
The deflection is maximum at the loaded end (x = l) and:

Pl 3
v(l,O) = 3EI
This is exactly the value given by elementary beam theory. The conclusion
however does not hold for other problems, as we shall see in the next section.

7.8.2 Uniform loads on the upper and lower surfaces of a simply


supported beam

Consider a straight beam of rectangular cross section which is subjected


to uniform loads q- and q+ on its upper and lower surfaces (x = ~cj2),
respectively. The beam is simply supported at its ends (x = ±lj2) (see Fig.
178 7. Two-dimensional problems in Cartesian coordinates

q
-c/2

-b/2~
+c/2

y l
Fig. 1.6. Uniform loads on the upper and lower surfaces of a simply supported
beam

7.6). This is a problem we have already solved using beam theory in Chap. 3,
our purpose here is to solve it again using a more sophisticated theory (plane
elasticity) and compare the two solutions.
On the upper and lower surfaces (y = ~c/2), the stress B.Cs. are:

c c q- c q+
a",y(x'~2) =0; a yy (x'-2) = -Ţ; a yy (x'2) =-Ţ (7.49)

Note that q . and q+ are forces per unit length [N/m], they are the integrals
over the thickness b of forces per unit area (q- /b) and (q+ /b).
On the end sections (x = ±l/2) we shall write c~nditions on the stress
resultants and use Saint-Venant's principle. Actually, it is useful to find ex-
pressions for the stress resultants in each cross section (x) with outward unit
normal e",; and these were computed in Sect. 7.8.1.

R(x) = b (e", j C/2 a",,,, dy + e y jC/2)


a",y dy ; M(x) = -be z j C/2 ya",,,, dy
-c/2 -c/2 -c/2
(7.50)

On the other hand, static equilibrium of the left part of the beam gives
q l2
R(x) = -qdXey; M(x) = - ; ("4 - x 2 )e z (7.51)

where qd == q- - q+. Equating the expressions in (7.50-51), we obtain three


scalar equations:

j C/2 a",,,, dy = O; b
jC/2
a",y dy = -qdX; b j C/2 ya",,,, dy = ...!!(-
q l2
- x2 )
-c/2 -c/2 -c/2 2 4
(7.52)

We now use the same procedure as in Sect. 7.8.1 to construct Airy's stress
function. Equation (7.52a) implies the following form for a",,,,:
7.8 Solved problems 179

O'xx.. ( y,y,y, t ) an d/or (n1


3 5 ec. x y,x na y,x
3 ns y, t )
5 ec.

From Eq. (7.52c), we find that:


YO'xx: (yO,y2,y4, etc. ) and (X 2yO,x2y2,x2y4, etc. )
Combining the two conditions implies the following form for O'xx
O'xx : (y,y3,y5, etc. ) and (X2y,x2y3,x2y5, etc. ) (7.53)
Equation (7.52b) gives the following form for the shear stress:
O'xy : (xyo, xy2, xy4, etc. ) (7.54)
We now need to construct an Airy stress function which will give us stress
expressions of the form (7.53-54). From Sect. 7.5, it is found that the following
expression satisfies the requirements:

(7.55)

The corresponding stresses are:


2 1
O'xx = d3y + d5x 2 y - 3
3d5Y ; O'yy = b3y + 3d5Y 3
+ a2; O'xy = -b3x - d5xy
2

(7.56)
Note that, from Sect. 7.5, we need to have b5 + 2d5 + 315 = o. If 15 = O and
b5 = -2d5, then O'xy contains the term -b5x 3 /3, which is unwanted. On the
other hand, if b5 = O and 15 = -2d5/3, then O'xy and O'yy do not contain
additional terms and O'xx contains -2d5y 3/3 which is a possible term. AIso,
if we do not add the constant term a2 to O'yy, then this stress would be an
odd function of y and the B.Cs. on y = ±c/2 could not be satisfied.
The coefficients d 3 , d5 , b3 and a2 are determined from the B.Cs. After
some algebra, the following expressions are found:
qd qd c2
d5 2I; b3 = 81 ;
qs c3 qd c2 l2
a2 = - 241; d 3 = 2I (- 10 + "4 )
where qs == q- + q+ and 1 == bc3 /12 is the moment of inertia of the cross sec-
tion w.r.t. the z-axis. Finally, suhstituting the expressions of the coefficients
into Eqs. (7.56), we find the stresses as follows:

O'xx
qdY (l2
= 2i 2 c2 2 2)
"4 - x - 10 + '3 y
qd l2
(order - - )
b c2
qdY c2 y2 qs c3
O'yy = 2i(4 - 3") - 241 (order q;)
qdX c2 2
(order qd_l)
O'xy = - 2I (4 - y ) bc (7.57)
180 7. Two-dimensiona1 problems in Cartesian coordinates

Again, as in Sect. 7.8.1, these stress expressions are valid for plane stress as
well as plane strain. At the end sections, we have: I~~~2 0'",,,, dy = O, but:

±1 qdY c2
0'",,,,( 2'Y) = 2T(-10 + a2 Y )
2

which means that the normal stresses are zero in average only. As in Sect.
7.8.1, and using Saint-Venant's principle, it is concluded that Eqs. (7.57)
represent the stress solution ofthe problem far from the ends (again, assuming
that c > b, a practical requirement is x E [-(1/2)+c, (1/2)-c], and the larger
the ratio l/c, the better the solution). We now compare the stress solution
to that found from elementary beam theory (Chap. 3). The shear force Q(x)
and bending moment M(x) are:
qd 12
Q(x) = -qdxi M(x) = "2+4' - x 2)
The normal and shear stresses are given by Sec. 3.6 as:
(O) _ M(x). (O) _ Q(x) (c2 _ 2)
O'",x - 1 y, O'",y - 21 4 Y
Comparison with Eqs. (7.57) shows that:
_ (O) qdY _ c2 ~ 2 •
O'xx - 0'",,,,
-...-.,... + , 21 ( 10 + 3Y ),, (7.58)
...

(and recall that O'W = O). II 1 » c, then O'",x ~ 0'10) and O'yy is negligible
when compared to O'",x and O'xyi the elementary solution is therefore valid in
that case.
In the remainder of this section, we consider plane stress (the thickness b
is assumed to be small). Section 7.3 and Eqs. (7.57) give the strains as:

f",,,, qdY (12 _ x2 _ c2 + ~y2 _ v c2 + vy2) + v qsc3


2EI 4 10 3 4 3 24EI
f yy
2 2
qdY (c _ y2 _ V 1 + vx2 + v c _ 2v y2) _ qsc3
2

2EI 4 3 4 10 3 24EI
qdX c2 2
f",y = -(1 + v) 2EI (4' - Y ) (7.59)

Integration ofthe strain/displacement relations and use of Eqs. (7.59a-b) give


the displacements u(x,y) and v(x,y) as follows:
c2
+ ay2 - v4' + v3' + v qs12 x + F(y)
12 X2 2 c2 y2) c3
2Elu qdXY ( '4 -"3 - 10

y2 (c2 y2 12 c2 y2 ) c3
2Elv = qd"2 4' - "6 - v'4 + vx 2 + v 10 - v3' - qs 12 Y + G(x)
7.8 Solved problems 181

We now use the same technique as in Sect. 7.8.1: compute âu/ây and âv/âx,
and substitute into the expression of Exy. This leads to the following equation:

For the equality to hold for alI values of x and y, each side must be equal to
the same constant (k). Integration of the two equations thus obtained gives:

x 2 (l2 2 2 2
F(y) = -ky + A-, G(x) = kx - qd-
2 4
- + -2c5 + -vc4 - -x6 ) + D

Constants k, A and Dare computed from the displacement B.Cs. Let us


choose the following B.Cs. (which correspond to an immovable hinge at the
left end and a simple support at the right in the beam theory sense):

These conditions imply the following values for the constants:

c3l l2 51 2 c2 vc2
k=O; A= qs v 24 ; D=qd"4(48 +5+""8)

Therefore, the displacement field is completely determined. The displace-


ments of the middle fiber (y = O) are:

u(x,O)

v(x,O)

The deflection is maximum at the middle of the beam (x = O, Y = O) and:

(7.60)

The beam theory result (Chap. 3) is:

(O) 5qd l4
vmax = 384EI
It appears then that the 2D theory of elasticity gives a solution (7.60) which
is made up of two terms: one is identical to the beam theory result, and an
additional term which becomes negligible when l » c. As for the horizontal
displacement, it is maximum at the right end (x = l/2) and it is seen that
for long beams it is negligible when compared to the maximum deflection.
182 7. Two-dimensional problems in Cartesian coordinates

7.8.3 Unifonn load on a cantilever beam

Consider a cantilever thin beam which is uniformly loaded on its upper surface
(Fig. 7.7). We could solve this problem directly as in Secs. 7.8.1 and 7.8.2,
but it is much simpler to use the superposition principle as shown in the
figure. Accordingly, the problem is split into two simpler problems which are
themselves statically equivalent to the two problems of Fig. 7.6, which we
have already solved in the previous sections. Consequently, in order to find
the stress field, we set qd = qs = q in Eqs. (7.57), P = ql/2 in (7.46), express
the stresses in those equations in a same frame and then add them up. For
the displacements, the general expressions are written in the same frame and
then added up, but the constants have to be determined according to the
specific displacement B.Cs. we have.
Exercise: work out the details of the computations.

7.8.4 Unifonn load on a beam with two clamped ends

Consider a uniformly loaded beam with two built-in ends (Fig. 7.8). This is a
statically indeterminate problem which can be solved using the superposition
principle as shown in the figure. Accordingly, the problem is split into three
simpler problems. The first one was solved in Sect. 7.8.2, the second one is
a pure bending problem which was solved in Sect. 3.2 and for which the
only non-zero stress is O'~~ = Moy/I. The third problem corresponds to
uniaxial compression where the only non vanishing stress is 0'~3J = -Ho/(bc).
After superposition, the unknowns Mo and Ho are computed by setting the
displacement B.Cs., e.g.

Exercise: work out alI the computations.

7.8.5 Compression of a beam in the height-direction

A beam of length 1 and height c is compressed by forces q- and q+ distributed


over parts of its top and bottom surfaces: y = ±c/2, x E [-a,a] (see Fig.
7.9). We are going to apply the Fourier series technique of Sec. 7.6 to this
problem. The loads are even functions of x, therefore Eq. (7.24) reads:

(7.61)

We look for an Airy function under the form:


r
7.8 Solved problems 183

!! {.d 1111111111111 q

q12/2 fL
ql
1 •.

111111111111111
ql/2t t
ql/2

q12/2
( tL.....-1 -------li l
ql/2 ql/2

1"""11"" r q

+
~f-----------III+ ql/2
Fig. 7.7. Uniform load on a cantilever beam. The superposition principle is used
to obtain a statically equivalent problem.
184 7. Two-dimensional problems in Cartesian coordinates

Mo

Ho
qlJ2 qlJ2

M, (1111111111111 :l~

»11111111111:l
+
MO( A- )
Mo

M
+
Fig. 7.8. Uniform load on a beam with two clamped ends: use of the superposition
principle
7.8 Solved problems 185

00
2mr
ifJ(x,y) = ifJo(x,y) + Lfn(y)cosmx, m- =l - (7.62)
n=l

where fn(Y) is given by (7.26). In order to make ifJ(x,y) bi-harmonic, we need


to have LlLlifJo(x, y) = O, and the stresses will then be given by:

(7.63)

where (uxx)n, (uyy)n and (uxy)n are given by Eqs. (7.27). In order to make
the computations simple, we now consider the particular case where the loads
are equal and piece-wise uniform:

q-(x)=p, if xE[-a,a];

q-(x) =0, if XE[-~,-a[u]a,~] (7.64)

Equations (7.61) are now combined into a single equation:

(7.65)

A standard technique -e.g., (Tolstov, 1962)- gives coefficients qo and qn:

2
-21'/ q(x) dx a
= 4p-;
l -1/2 l
21' / 2 q(x) cos(2mr-
-l
X ) dx = -2p sin(2mr-al ) (7.66)
-1/2
l mr
Since the loading is symmetric w.r.t. the y = O plane, the stress U yy must be
an even function of y; Eq. (7.27) implies then that 'I'n = 6n = O. From Eqs.
(7.63) and (7.27), it is then seen that U yy and u xy are given by:

8 2ifJ 00

U yy = 8x 2o - L m 2 cos(mx) [an cosh(my) + .Bnmy sinh(my)]


n=l

_;::0 + 2I:m Sin(mxHan sinh(my)


y n=1
+
.Bn[sinh(my) + mycosh(my)]} (7.67)
186 7. Two-dimensional problems in Cartesian coordinates

The stress B.es. are:

(7.68)

The simplest choice for 4Jo is cPo(x, y) = -qox2 /4. Using this expression with
(7.67), Eqs. (7.68) yield the following scalar equations:

(cosh JLn)an + (JLn sinh JLn).Bn = mq~ j


(sinh JLn)an + (sinh JLn + JLn cosh JLn).Bn =O
where JLn == me/2 = mre/l. The solution of this linear system is found to be:

an (:~)3 sin(2mry)(sinh JLn + JLn coshJLn)[2JLn + sinh(2JLn)r 1j


.Bn = - (:~3 sin(2mry)(SinhJLn)[2JLn +sinh(2JLn)]-1 (7.69)

Finally, the stresses are given by the following expressions:

L: m
00

U xx = 2 cos(mx) [(an + 2.Bn) cosh(my) + .Bnmysinh(mY)]j


n=l

L: m
00

U yy = -~- 2 cos(mx)[an cosh(my) + .Bnmy sinh(mY)]j


n=l

L: m
00

u xy = 2 sin(mx) [(an + .Bn) sinh(my) + .Bnmy cosh(my)] (7.70)


n=l

The series converge for n ~ 30. The stresses in the (y = O) plane are
represented in Fig. 7.10 for e = l/l0 and a = l/8. It is seen that the
stresses decrease rapidly for x > a, Le. outside the loading zone. AIso,
Uyy(O, O)/p ~ -1.00834.

x

y
Fig. 7.9. Compression of a beam in the height-direction
7.8 Solved problems 187

0.2

O
I
-0.2 I
I
I
-0.4 I
......
Ro
I
I
ti
-0.6 I xx--
I
I yy -----
-0.8 I
I

-1 --" I

-1.2
O 0.1 0.2 0.3 0.4 0.5
x/l
Fig. 7.10. Compression of a beam in the height-direction. Normalized stresses vs.
x/l for y = O, c = 1/10 and a = l/8.

The case of a concentrated load F at (x = 0, y = ±c/2) is obtained by


letting 2a/l --t o. If we assume that the series converge for a value ncv such
that 2ncv1fa/l « 1, then for n E [l,ncv], we can make the approximation
sin(2mra/l) ~ 2mra/l and Eqs. (7.70) give -after simplification- the following
expressions for the stresses:
Uxx(x,O) 4-c ~ (2 x) -sinhţtn + ţtncoshţtn .
L..J cos mr-
c F ~
1 n=l 1 2ţtn + sinh(2ţtn) ,

Uyy(x,O)
F ~ _~ _ 4~ L cos(2mr~) szn. h ţtn + ţtncos hţtn
ncv
(7.71)
C
1 1 n=l 1 2ţtn + sinh(2ţtn)
where F = 2pa represents the concentrated force (per unit thickness). The
series converge for ncv ~ 30. The stresses are plotted in Fig. 7.11 for c = 1/10.
It is found that:
lF F
Uyy(O,O) = -1.839~T = -18.39T
and this is much larger -in absolute values- than (-F/l): the stress state
does not correspond at all to uniaxial compression in the y-direction (so
many students make that mistake!) It is also seen that for x > 1/15, Le.
x/c > 2/3, Uyy(x, O) goes from compression to tension, and then decreases
rapidly before vanishing.
Exercise: compute the stresses at y = ±c/2 for a/l = 0,1/8,1/4 and
c/l = 1/20,1/10,1/5. Plot and comment the results.

7.8.6 Body forces- Beam under its own weight


The reader is reminded that the method of solving plane problems with Airy
functions assumes that there are no body forces. We now consider an example
188 7. Two-dimensional problems in Cartesian coordinates

0.5

-
~
<.)
b
-0.5

-1 xx--
yy ....•......

-1.5

-2
O 0.1 0.2 0.3 0.4 0.5
x/l
Fig. 7.11. Compression of a beam in the height-direction. Normalized stresses vs.
x/l for y = O, c = 1/10 and a/l ~ o.

of a beam of length 1 and height c which is subjected to its own weight (Fig.
7.12); we shall see how to take body forces into account.
The forces per unit volume are: fz = O and fy = pg, where p [kg/m 3 ] is
the mass density and g [N / kg] the acceleration due to gravity. Since externa!
forces are act ing in the y-direction only, we consider the following particular
solution:

u~r:) = u~ = O; u~~ = -pg(y +~) (7.72)

The equilibrium equations read:


âu(p) âu(p) âu(p) âu(p)
â:
z
+ â;Y + fz = O; â:Y + ~Y + fy = O
and they are both identically satisfied. The compatibility equation reads:

and is identically satisfied since uW is linear in y and fy is constant. The


stress B.Cs. on the top surface (y = -c/2) are satisfied:

The stress B.Cs. on the lower surface (y = c/2) give the following equation:
u(p) . e
y
= u(p) (x ~)e + u(p)
zy' 2 z yy
(x '2~)e y = -pgcey ...,l.
r
O
7.8 Solved problems 189

Therefore this B.C. is not satisfied, which means that u(p) is not the solution
of our problem. However, we can use u(p) as follows: this solution is valid
if we apply on the lower surface (y = c/2) a uniform surface force equal
to (-pgce y ). We use the superposition principle in order to correctly solve
our initial problem. This is illustrated in Fig. 7.12 where it is seen that
the final solution is the sum of the solutions of problems (1) and (2). The
stress solution of the former problem is given by Eqs. (7.72), while the latter
problem has already been solved in Sect. 7.8.2 (one only needs to set q- = O
and q+ = -pgcb).

x

pgcb
+

pgcb
Fig. 7.12. Beam under its own weight: use of the superposition principle

As another application, consider the problem depicted in Fig. 7.13: a beam


is subjected to its own weight as well as a uniform load q on its top surface.
Once again, the superposition principle is used, and the stress solution is
the sum of the particular solution (7.72) and the solution of Sect. 7.8.2 after
setting q- = q and q+ = - pgcb.
We now compute the displacements associated with the particular stress
solution (7.72). The strains are given by:
190 7. Two-dimensional problems in Cartesian coordinates

pgcb
+

~--H-H-:-H--:-H-:--~
pgcb
q

Fig. 1.13. Beam under its own weight and a uniform load on its top surface
7.8 Solved problems 191

(7.73)

Integration of Eqs. (7.73a, b) w.r.t. x and y, respectively, leads to the following


expressions for the displacements:

u(p) = vpg (y + ::')x + F(y)' v(p) = -pg (y + c)'#.. + G(x)


E 2 ' E 2
Substituting these expressions into Eq. (7.73c) and using the standard tech-
nique of Sect. 7.8.1, the following expressions for F(y) and G(x) are found:
vpg 2
F(y)=-Ay+D; G(x)=Ax- 2Ex +B

where the constants A, B, and Dare determined from the displacement B.Cs.
8. Two-dimensional problems In polar
coordinates

In Chap. 7, we studied plane problems in Cartesian coordinates x and y.


However, there are several important plane stress or strain problems which
are more easily formulated and solved in polar coordinates r and O.

8.1 Change of coordinates


Polar coordinates are represented in Fig. 8.1 and have already been used in
previous chapters. More useful relations are recalled in this section. Cartesian
and polar coordinates of a point M in the plane are related by:

x = rcosO; y = rsinO

Differentiation gives

dx = -r sin(O)dO + cos(O)dr; dy = r cos(O)dO + sin(O)dr


These relations can be inverted to give:
1
dr = cos(O)dx + sin(O)dy, dO = -[- sin(O)dx + cos(O)dy),
r
from which the following relations which will be often used are obtained:
âr . O
- = cosO; -âr = SIn·
âx ây ,
âO sin O âO cosO
(8.1)
âx r ây r

At each point of the plane, an orthonormal basis (eT) eli) is given by:

e r = cos(O)e x + sin(O)e y ; eli = - sin(O)e x + cos(O)e y


194 8. Two-dimensional problems in polar coordinates

The first row of the matrix contains the components of e r in the Cartesian
basis (e"" e y ) while the second row contains the components of eo in that
basis. The transformation matrix is proper-orthogonal:

QQT = QTQ = 1, detQ = 1,


where 1 is the 2 x 2 identity matrix. Second-order symmetric tensors (e.g.,
stress and strain) transform according to the following rules (Appendix A):

(8.2)

where u(p) and u(c) are 2 x 2 matrices representing the tensors in polar and
Cartesian systems, respectively,

Developing the transformation rule (8.2a), the following relations between


the components are obtained:

a rr a",,,, cos 2 O+ a yy sin 2 O+ 2a",y sin(O) cos (O)


a99 = a",,,, sin 2 0+ ayy cos 2 O- 2a",y sin(O) cos (O)
aro (ayy - a",,,,) sin Ocos 0+ a",y(cos 2 0- sin 2 O) (8.3)

The substitutions O -t (-O), x B r and y B O immediately give the ex-


pressions of a",,,,, a l1y and a",y in terms of a rr , aoo and arD. Note also that
a rr + aoo + a zz = a",,,, + ayy + a zz , because (tr u) is an invariant.
~ ~

eo
y

x
e",
Fig. 8.1. Polar coordinates (r,8)

There are (at least?) three methods for obtaining the equiIibrium equa-
tions in polar coordinates. In the first method, the Cartesian equations are
transformed to polar coordinates, using (8.1) and (8.2). The second method
consists in writing the equilibrium of a smaII sector (dr, dO) -see Fig. 8.2- and
neglecting second-order terms. These two methods are developed in other
textbooks, e.g. (Timoshenko and Goodier, 1987) or (Chou and Pagano, 1967).
8.2 Summary: plane strain versus plane stress 195

In a third method (my favorite!), equilibrium equations are written in tensor


(or symbolic or "intrinsic") notation: div u + j = 0, and then (div u) is
computed in a cylindrical basis; see Appendix A. No matter what method is
used, the following equations are found:
aarr ~ aarO a rr - a(J(J f _ .
ar +r a(} + r + r - O,
2 1 aa(J(J
ar + -:;.a
aaro
rO + -:;'88 + 10 = O (8.4)
For the strain-displacement relations, again several methods can be used. We
prefer to re-write the partial differential eqllations of the Cartesian system
into tensor form: 2€ = V'u + (V'uf and compute V'u in a cylindrical basis;
see Appendix A. The following equations are found:
aUr Ur 1 auo 2c r o = ~ aUr + auo _ Uo
trr = ar; 1'00 = -;: + -:;. a(); L r af) âr r

The constitutive eqllations are form-identical to those in Cartesian coordi-


nates, this is a conseqllence of isotropy; see Appendix A. As for the compat-
ibility eqllations of Sect. 7.3 , once again it is easier to write them in the
symbolic form .1(arr + aoo) = (coeff.)div j, and then express the Laplacian
and divergence operators in polar coordinates; see Appendix A. The plane
strain/stress eqllations in polar coordinates are sllmmarized in the next sec-
tion.

x
Fig. 8.2. Equilibrium of a small sector (dr, dO)

8.2 Summary: plane strain versus plane stress


The equations of plane strain and plane stress are summarized hereafter
(Greek indices a, (3, Î take values r and f)).
196 8. Two-dimensional problems in polar coordinates

EIJz = Erz = O
= O'rz = O
O'IJz
Ua = ua(r, O)
Ea{3 = Ea{3 (r, O)
O'a{3 = O'a{3(r, O)
fz = O; fa = fa(r,O)
c -!t!b:..
Lrr -
c
8r' LIJIJ --.!!z:.
r
+ .!3·
r 8IJ'
2cLrIJ --.!~
r (jIJ
+ 38r _ ~
r
ilir=+.!~+urr-u
8r r 8IJ r
•• +1r=0'
,
~+~~
8r r vrIJ +.!~+J,
r 8IJ (J
=0
Plane strain Plane stress
Uz = O O'zz = O
Ezz = O=? O'zz = vO'-n O'zz = O =? Ezz = -1~IIE~~
Ea{3 = ~(O'a{3 -vO'~~oa(3) Ea{3 -- !±.!!.(
E O'a{3 II O'~~ua{3
- 1+11 ~)

LlO'n = -1~lIdiv J LlO'n = -(1 + v)div J


As observed in Chap. 7, most of the equations are identical, and if a change of
parameters is made, the last sets of equations (the Ea {3-O'a{3 constitutive rela-
tions and the compatibility equations) become also identical. We can go from
plane stress to plane strain solutions by making the following substitutions:
v E
V~--' E~--
1-v' 1 - v2
Vice-versa, when we have the solution of a plane strain problem, we can find
that of plane stress with the following substitutions:
v ~ _v_o E ~ E(1+2v)
1 + v' (1 + V)2

8.3 Airy stress function


For plane problems without body forces, we have seen in Chap. 7 that there
exists a so-called Airy stress function lP(x, y) such that LlLllP(x, y) = O and
(PlP 82lP 82lP
O'zz = 8 y 2; f1 yy = 8x 2 ; O'zy = - 8x8y
One method of writing the counterpart of these equations in polar coordinates
with cjJ(r, O) = lP(x, y), consists in the repeated use of Eqs. (8.1), e.g.
8lP = cos(O) 8cjJ _ sin (O) 8cjJ.
8x 8r r 80'
82lP = (COS(O)~ _ Sin(O)~) (COS(O) 8cjJ _ sin(O) 8cjJ)
8x2 8r r 80 8r r 80
2 ( ) 8 2cjJ . ( 8 1 8cjJ
cos O 8r 2 - cos( O) Slll O) 8r (;: (0)
_ sin(O) ~( (0)8cjJ) sin (O) ~( . (0)8cjJ)
r 80 cos 8r + r 2 80 Slll 80
8.3 Airy stress function 197

and similarly for the other partial derivatives. The second step consists in us-
ing the transformation rules (8.3). This method is rather tedious, see (Timo-
shenko and Goodier, 1987) ar (Sechler, 1952) for the complete computations.
We present hereafter a second method of derivation.
We write the Cartesian relations in the matrix form:

[
Uyy
-uxy
-ux y ] _
U xx
- [~8x2
~
8x8y
~l
8x8y
~
8y2

Using (8.2) and (8.3), this equality can be rewritten as follows:

where the upper-script (c) refers to the Cartesian system. Multiplying by


Q(O) at the left and QT (O) at the right, we obtain:

where the upper-script (P) refers ta the polar system. Nating that

Q(O)Q(!:) = [-sinO cos.O ] = Q(!:)Q(O) (= Q(!: +0)),


2 -cosO -smO 2 2

(actually, we have the more general reIat ion Q(01)Q(02) = Q(02)Q(0t} =


Q(OI + ( 2 )), we obtain:

which can be written in component form as:

The expression of (\l\lif;) in polar coordinates (Appendix A) gives:

1 1 â2 if; âif; â 2 if; â 1 âif;


Urr = ;-C;:· â02 + âr); U(J(J = âr2; Ur(J = - âr (;- âO) (8.5)

The Airy function if;(r, O) has to be bi-harmonic: LlLlif;(r, O) = O. Using the


expression of the Laplacian (Appendix A), this is written as:
198 8. Two-dimensional problems in polar coordinates

8.4 Axisymmetric plane problems

We first study problems such that t/J = t/J(r), Le. the Airy stress function is
radial only and does not depend on the angular position O. Several interesting
problems are covered by this case; we shall even see in Sect. 8.6 a problem
(bending of a circular beam) which does not possess an axisymmetric geome-
try but can be solved nevertheless with a radial only Airy function. The latter
must be bi-harmonic: .::1.::1t/J(r) = O. Recall that in Sec. 6.2, the deflection w(r)
of a plate in axisymmetric problems obeys the equation .::1.::1w(r) = p(r)/V.
The general solution of the homogeneous equation was worked out and a
simple substitution w(r) ~ t/J(r) gives

t/J(r) = Alnr + Br2 1nr + Cr 2 + D (8.7)


where A, B, C, Dare constants to be determined. Using Eqs. (8.5), the in-
plane diagonal stresses are found as
A
2"
r
+ B(l + 21n r) + 2C;
A
-"2 + B(3 + 21nr) + 2C; (8.8)
r

and the shear stress is zero (ar/J = O) as expected.


If the physical domain contains the origin (r = O) -e.g. a plate without
a hole at the center- then we must have A = B = O in order to have finite
stresses at (r = O). In this case, we see that a rr = al)I) = 2C, i.e. there is a
state of uniform in-plane compression or traction.
If the domain does not contain the origin but is a complete ring, then we
shall see that continuity of the displacement field imposes that B = O.
We now compute the stmins assuming plane stress. Using the constitutive
equations of Sect 8.2, we find the in-plane diagonal strains given by:
A
(1 + v) 2"
r
+ 2(1 - v)B In r + (1 - 3v)B + 2(1 - v)C;
A
-(1 + v)2" + 2(1 - v)B In r + (3 - v)B + 2(1 - v)C (8.9)
r

while the shear strain is zero (Er/J = O). In order to find the displacement
field, we first integrate the strain/displacement relation Err = aUr/ar w.r.t.
r; we obtain:

EU r = -(1 + v)-Ar + 2(1 - v)Br In r - (1 + v)Br + 2(1 - v)Cr + F(O)

where F(O) is an arbitrary function of O. Next, the relation


Ur
101)1)=-+--
1 aUI)
r r 00
8.4 Axisymmetric plane problems 199

is integrated w.r.t. 9 ruter replacing Ur with its expression. We obtain:

Euo = 4Br9 - f F(9) d() + G(r)

where G(r) is an arbitrary function of r. We now substitute Ur and Uo into


the expression of the shear strain:

1 aUr aUD
2t: r o = - ( - -
r auo Uo) + -
ar = O
and find:
dF
d9 + f dG
F(9) d() = G(r) - r dr

Since the left-hand side of the equation depends on 9 only, and the right-
hand side on r only, each side must be equal to the same constant L, say
(the technique is identical to the one we used in Chap. 7 for plane problems
in Cartesian coordinates). Further, we introduce the notation

H(9) == f F(9) d9

Therefore, we have the following pair of differential equations:

rPH dG
- + H(9) = L' -r- + G(r) = L
d9 2 ' dr
The general solutions of these equations are:

H(9) = L + âsin9 - bcos9; G(r) = L + Cr


where â, b and care constants to be determined. F(9) is obtained from:

F(9) = ~~ = âcos9 + bsin9

In summary, the in-plane displacement field for plane stress problems is given
by:
A
= -(1 + v)-
r
+ 2(1- v)Br lnr - (1 + v)Br + 2(1- v)Cr
+ âcos9 + bsin9
Euo = 4Br9 - âsin9 + bcos9 + Cr (8.10)

For a plane strain problem, the displacement field is obtained from (8.10) by
a simple change of material parameters as defined in Sect. 8.2. Constants â, b
and c can be interpreted after splitting (8.10) as follows:
200 8. Two-dimensional problems in polar coordinates

EU)
( EU(J
r = (...... ) +â (cos O ) +b- ( sin O ) +er ( O )
- sin O cos O 1

Thus it is seen that constants â and b correspond to rigid displacements


along the (x) and (y) axis, respectively, while c corresponds to infinitesimal
rigid rotation around the (O, z) axis. If the displacement B.Cs. prevent rigid
motion, then constants â, b and c will vanish.
Finally, we consider the case when the physical domain does not contain
the origin (r = O). For a complete ring, B = O, otherwise uli(r, 0+ 27r) =fi
Uli (r, O) because of the term (4BrO). If the domain is a sector, then we can
have B =fi O. Examples are given in Sect. 8.7.

8.5 Periodic Airy stress functions

There are cases when it is interesting to consider Airy functions of the form:

4Jm(r,O) = Im(r) cos(mO) (8.11)

An example will be given in Sect. 8.7 (a plate with a small circular hole in
its center is submitted to traction at its ends). For 4Jm(r, O) of Eq. (8.11) to
be an Airy function, it must verify: LlLl4Jm(r, O) = O. This implies that

Solutions of this differential equation are sought under the form Im = rP •


Substitution gives the following equation

(p - m)(p + m)(p - 2 + m)(p - 2 - m) = O

IT m 2 =fi 1 and m =fi O, then Im(r) is a linear combination of r m, r- m , r 2 - m


and r2+ m . If m 2 = 1 than p = 1 is a double root and p = -1 and p = 3
are simple roots, therefore Im(r) is a linear combination of r, rlnr, l/r
and r 3 • The case m = O corresponds to axisymmetric problems, where we
have already seen that 10 (r) is a linear combinat ion of r 2 , r 2 In r, In r and a
constant. For generalloading cases -and similarly to Sect. 7.10- we may want
to look for an Airy function 4J(r, O) as a Fourier series

L Im(r) cos(mO) + L Fm(r) sin(mO)


00 00

4J(r,O) = lo(r) +
m=l m=l
8.6 Generalized plane strain 201

8.6 Generalized plane strain

Generalized plane strain problems are such that an assumptions of Sect. 7.1
are verified except one: the end sections are free to move in the (z) direction;
therefore tzz :f. O. Those problems can be solved by a direct method (e.g.,
Secs. 1.16 or 9.5.2) or with the superposition and Saint Venant's principles.
As an illustration, consider the plane-strain Lame problem of Sect. 8.7.1. The
stress vector acting on each end section is:

0-(1) . (±e Z ) = ±u(1)e


zz z,

with u~~) = YU~~ = constant. The resultant of the stress vector is simply

and the resultant moment w.r.t. to the centroid of each end section is zero.
Lame problem (1) is the sum of two problems (2) and (3), see Fig. 8.3. In
problem (2), the end sections are not constrained in the (z) direction; this
is a generalized plane strain problem that we wish to solve. Problem (3) is a
simple uniaxial tension where the only non-zero stress is:

Thus the stress solution of problem (2) (0-(2) = 0-(1) - 0-(3») is simply:

u(2)
rr
= u(l).
rr'
u(2)
IJIJ
= u(1)·
IJIJ'
u(2)
zz
=O
The strains for the uniaxial tension problem (3) are
(3) (1) (1)
t(3) = U zz = yU".f'Y. t(3) = t(3) = _Yt(3) = _y2 U n
zz EE' rr IJIJ zz E

Thus the strain solution of problem (2) (€(2) = €(1) - €(3») is:
(1) (1) (1)
t(2) = t(l) + y2 U'Y'Y. t(2) = t(l) + y2 U'Y'Y. t(2) = _y U'Y'Y
rr rr E' IJIJ IJIJ E' zz E

8.7 Solved problems

8.7.1 Hollow circular cylinder under inner and outer pressures

A hollow circular cylinder is subjected to uniform pressures Pa on its inner


surface (r = a) and Pb on the outer surface (r = b); see Fig. 8.4. The solution
of this classical axisymmetric problem was found by the French scientist Lame
202 8. Two-dimensional problems in polar coordinates

-x

+
R R

Fig. 8.3. The plane-strain Lame problem is the sum of problems (2) -generalized
plane strain- and (3)-uniaxial tension.
8.7 Solved problems 203

around 1852 and is stiH widely used. As discussed in Sect. 8.4, the in-plane
stresses are given by (8.8) after setting B = O,
A A
(J'rr = -2
r
+2Cj (J'rHJ = --+2Cj
r2
(J'rll =0 (8.12)

Constants B and C are found from the stress B.Cs.,

These give two non-trivial scalar conditions:

Therefore, the constants A and Care solution of the linear system


A A
a2 + 2C = -Paj lJ2 + 2C = -Pb

With the following notation for the pressure difference and the radius ratio
. b
P == Pa - Pb > O (assumptlOn), ", == -a > 1,
the constants are found to be:
b2p -
A- ___ o 2C= -Pb+-P-
- ",2-1' rp-l

Finally, the diagonal in-plane stresses are:

(8.13)

Note that (J'rr :::; O, 'rIr E [a, b], Le. the radial stresses are compressive ev-
erywhere. We also have ((J'rr + Pb) :::; O and ((J'IIII + Pb) > O everywhere. For
instance, if the outer pressure is zero (Pb = O) then the hoop stress is tensile
-and the radial stress compressive- in the entire body. Assuming plane strain
(long cylinder in the (z)-direction and €zz = O), the rudal stress is:

(J'zz = 2V(-pb + ~IP)


", -

and is uniform. Each stress added to Pb is plotted in Fig. 8.5, which shows
that the largest stress -in absolute value- is always the hoop stress. IT Pb = O,
this stress is always tensile, thus for a brittle material, if the maximum hoop
stress is large enaugh ta cause fracture, then radial cracks will appear at
(r = a). IT the material is ductile (e.g., a metal), then plasticity may develap.
The 'fresca criterion (Sec. 1.10) takes the fallawing farm ((J'rn (J'IIII and (J'zz
are principal stresses because the shear stresses vanish)
204 8. Two-dimensional problems in polar coordinates

where ay is the initial yield stress. Figure 8.5 immediately shows that the
criterion gives the following inequality

O"IJIJ(a) - O"rr(a) < O"y


This is equivalent to the following design formula

- (1 - 1 ) O"y (8.14)
P<
rp- - 2

which can be easily interpreted: if the pressure difference (fi = Pa - Pb) reaches
a value given by the right-hand side of the relation, then plastic deformations
will develop at the inner surface of the cylinder (r = a). So if plasticity is not
desired, than p and "l = b/a must be designed such that p never reaches the
value given by Eq. (8.13). In the case of a very thin tube of thickness h,
h
b = a + h; -«
a
1

we have 7]2 ~ 1 + 2h/a, and therefore

d - h
O"IJIJ ~ -Pb + il
a-
an P < ~O"y

Thus the hoop stress is uniform and the design formula takes a simple (and
very used) form. For a very thick tube (a «b, i.e. "l» 1), we have:

O"IJIJ(a) ~ -Pb + p= Pa - 2Pb


The displacement field is obtained form the general plane stress solution
(8.10) after setting B = O and â = b = c = O and making the substitutions
v E
v -+ 1 _ v; E -+ 1 _ v2 .

It is found that the tangential displacement UIJ is zero while the radial dis-
placement is given by

(1 2p
+ v) [b--:;:- 2 ]
Ur = E("l2 _ 1) + (1 - 2v)(Pa - 7] Pb)r (8.15)

From the expression of the axial stress O"zz, it is seen that if the inner and
outer pressures are such that Pa = 7]2Pb, then we have at the same time
plane stress (O"zz = O) and plane strain (E zz = O), and in this case the radial
displacement takes the simple form U r = (1 + v)b2Pb/(Er).
As a final note, let us remark that the Lame's problem can be solved
without using the Airy stress function technique at aU. Since the problem is
axisymmetric, the in-plane displacements have the following form:
8.7 Solved problems 205

Ur = u(r)j Un = O
The strains are computed, then the stresses are found from the constitutive
equations and plugged into the equilibrium equations. Consequently, only one
non-trivial differential equation is found:

(equivalently, this differential equation can be found by substituting the as-


sumed form of the displacement field directly into the Navier equations of
Sect. 1.13). This differential equation has the general solution

(3
u(r)=ar+-
r
where the constants a and (3 are found from the stress B.Cs.
Exercise: Work out alI the computations and check that the displacement
and stress solutions are identical to those of (8.13) and (8.15).

Fig. 8.4. Hollow circular cylinder under inner and outer pressures

8.7.2 Composite hollow cylinder under inner and outer pressures

A composite solid is made up of two concentric hollow circular cylinders (1)


and (2). The inner and outer radii of (1) are a and e, respectively, and those
of (2) are e > a and b > e, respectively. The composite is subject to pressures
Pa at its inner surface (r = a) and Pb at the outer surface (r = b). Assuming
that (1) and (2) are perfectly "glued" together at their interface (r = e),
the problem can be solved by considering two sub-problems: (i) cylinder (1)
with pressures Pa at r = a and Pc (to be determined later) at r = e, and (ii)
cylinder (2) with pressures Pc at r = e and Pb at r = b. The solution of each
ofthese two Lame problems is given in Sec. 8.7.1. Now, the interface pressure
Pc is computed by imposing the displacement continuity at r = e. Using Eq.
(8.15), Pc is found from the following equality:
206 8. Two-dimensional problems in polar coordinates

2 -
.,.,LIP
O'zz + Pb

o 1 T/a

-fi

Fig. 8.5. Diagonal stresses (plus Pb for each) vs. normalized radial distance rla

8.7.3 CoiI winding


Aflat strip (e.g., aluminum sheet) is cold rolled under tension over a man-
drel (Fig. 8.6). The latter is a hollow circular cylinder of inner and outer
radii Ro and Rt. respectivelYi its inner surface (r = Ro) is stress-free. The
final coil contains N wraps of thickness h each. Coil winding is modeled
here as a "shrink-ring" process (this will become clear later). Each wrap J
(J = 1,2, ... , N) is considered as a circular tube of inner and outer radii RJ
and RJH = RJ + h, respectively. The wraps are thin, h/ RI « 1, and their
number is large, N» 1. The winding tension is designated by O'T •

• Stresses in the mandrel:


At the winding of each wrap K ~ 1, the stresses in the mandrel are those
of a hollow cylinder with zero pressure at its inner radius Ro and a pressure
PIK at the outer radius RI' Each new wrap adds new stresses to the existing
ones in the mandrel. Using (8.13), the final stresses in the mandrel are:
Mandrel, rE [Ro,RI]:
(M) ( ) _ 1 =f (Ro/r)2 ;-.
O'rr (J(J r - 1 _ (R /R )2 ~ PIK (8.17)
• o 1 K=l
8.7 Solved problems 207

• Stresses in the coiI:


We now turn our attention to the stresses inside the coil, i.e. inside each
wrap J. First we consider the case when J is the last wrap, Le. wrap (J + 1)
has not been wound yet. Since J is a thin wrap, the hoop stress inside it is
nearly uniform. A fictitious cut gives: O"(J(J ~ O"T. On the other hand, wrap
J is considered as a thin Lame's tube under inner pressure [-O"rr(RJ)] and
zero outer pressure at RJ+1' Section 8.7.1 then gives:

O"(J(J ~ hRJ [-O"rr(RJ)]


In summary, the stresses inside wrap J when it is the last one are given by:

(8.18)

We now consider the successive winding of additional wraps. At the winding


of each wrap K > J, the stresses inside the coil are those of a hollow cylinder
with pressures PIK at its inner radius RI and (O"Th/ R K ) at the outer radius
RK = RI +(K -1)h. Each new wrap K > J adds new stresses to the existing
ones in wrap J. Using (8.13) and (8.18), the final stresses in the coil are:
Coil, wrap J 2: 1, rE [RJ, RJ+1] :

(J) _ ;...[ h (RK /r)2 + 1 ( h )] .


O"(J(J(r)-O"T+ L.J -O"TR+(R /R)2-1 PIK-UT R '
K=J+I K K 1 K

(8.19)

It is easy to check that the radial stress is continuous between two consecu-
tlve J +)
. wraps J and ( 1 ,'l.e. O"rr
(J)( RJ+l ) = O"rr(J+l)(RJ+1 ) .
• Incremental mandrei/coil pressure:
The stresses in the mandrel and the coil, Eqs. (8.17, 19), involve a pressure
PIK which we need to compute. Recall that this is the pressure at the interface
mandrel/coil (r = Rt} due to the winding of wrap K 2: 1. The situation is
that of a composite hollow cylinder where the mandrel has radii Ro and RI
and elastic properties EM and /lM, and the coil has radii RI and RK and
properties Eo and ilO. Surface r = Ro is stress-free while there is a pressure
uTh/RK at r = R K . This problem is solved in Sec. 8.7.2. The solution given
by (8.16) becomes, after some algebraic manipulation:

{ 1 + /lM 1- (RdRK)2 [(1- 2 )(RI)2 + 1]


PIK EM (RdRo)2 -1 /lM Ro

+ 1 + /10 [(1 _ 2110)( RI )2 + 1]} = 2(1- /1&) O"T..!!:..- (8.20)


Eo RK EoRK
208 8. Two-dimensional problems in polar coordinates

• Final pressure on the mandrel:


Taking r = Rl in Eq. (8.17), it is easily found that:
N
(1~~) (Rd = - L PIK
K=l

On the other hand, taking J = 1 and r = Rl in Eq. (8.19b), we find that:

h
L PIK
N
(1~~)(Rd = -(1T"R -
1 K=2

FinalIy, setting K = 1 in Eq. (8.20) gives:

h
Pl1 = (1T Rl

In summary, we have checked that the final pressure at the interface man-
drel/ coil is continuous and has the folIowing value:
N
Pl == -(1rr(Rt) = L PIK (8.21)
K=l

• Removal of the mandrel:


Once aH N wraps are wound, the mandrel is removed. This induces a
stress re-distribution in the coiL The new stress Reld is found as follows:
to the stresses which existed when the mandrel was present, we add new
stresses which correspond to a holIow cylinder with a traction (-pd at its
in ner surface r = Rl and zero pressure at the outer surface r = RN+l. The
expressions ofthose additional stresses are given by (8.13). Consequently, the
stress Reld inside the coil after removing the mandrei is the following:

( ) _ ()(before) (RN+1/r)2 TI ( ) (8.22)


(1rr,()() r - (1rr,()() r T (RN+1/Rl)2 -1 -Pl ,

where Pl is given by (8.21) and "before" refers to stresses (8.19) which existed
before removing the mandreL It is easy to check that (1rr (Rt) = -Pl - (-Pl) =
0, i.e. surface r = Rl is stress free indeed.

• Discussion:
The stresses before and after removing the mandrel are plotted in Fig. 8.7.
The results coincide with those reported in (de Vathaire and Faessel, 1981).
See aiso that reference for more informat ion and results with Iess restric-
tive assumptions. Indeed, the coil winding model presented in this section
has some embedded assumptions: (1) isotropic linear elasticity, (2) perfect
adherence at alI interfaces, (3) solids are Iong in the axial (z) direction, (4)
8.7 Solved problems 209

plane strain (zz-strain = O). The real situation may differ significantly from
the assumed model: the strip may be anisotropic, having different transverse
and mid-surface properties; slip or separation may occur between two adja-
cent wraps; the length of the coiI may be of the same order as its diameter;
the axial displacement of the coiI is not constrained. Nevertheless, the as-
sumptions make for simple calculations, which do not compare too badly
with experimentally-measured coiI stresses.

Fig. 8.6. Aflat strip is cold rolled under tension UT over a mandrel.

8.7.4 Bending of a curved beam

A circular beam of in ner radius (a) and outer radius (b) is subjected to
bending moments (±Mez ) at its end sections (O = ±a); see Fig. 8.8. We
designate by ro the radius of the middle "fiber", ro == (a + b)j2. Since the
bending moment is constant in any cross section (O = constant) of the beam,
we expect the stress distribution not to depend on the angle (O).Therefore
we shalllook for an Airy stress function of the form </J = </J(r), although the
geometry of the beam is not axisymmetric.
Note that the choice of the axes is important. Since the tangential dis-
placement has the form (see Eq. (8.10)) Ue = 4BrO j E, then symmetry im-
poses that we define O such that ue(r, O) = O (in the middle section) and
ue(r,a) = -ue(r, -a)(in the end sections).
Since the domain does not contain (r = O) but is a sector, then constants
A and B of the general solution (8.8) can take non-zero values; thus
A A
U rr = 2'
r
+ B(l + 2lnr) + 2C, ueo = -2'
r
+ B(3 + 2lnr) + 2C, (8.23)
210 8. Two-dimensional problems in polar coordinates

, ",,"
...... ... --
- -- --
-_-. ....,,~_.--_ .
..... ..;Z:'~:·~·:::::·::·:···
--
h
b I 99 before ---
b
" 99 after -----
I
I rr before .......... .
I rr after _. - . -
I
I
I
I
-3~~~~~~~~~~~~

0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1


r[m]
Fig. 8.1. Stresses in the coil before and after removing the mandrel. Numerical
parameters: Ro = 0.2 m, Rl = 0.25 m, N = 600, h = 1.25 mm, EM = 200 GPa,
VM = 0.3, Ea = 200 GPa, Va = 0.3.

Fig. 8.8. Bending of a curved beam


8.7 Solved problems 211

and O"r(J = O. The constants A, B and Care determined from the stress B.Cs.
The inner and outer radii are stress free

u . (-e r ) = O, at r = a; u· e r = O, at r = b.
These give two non-trivial scalar conditions:

(8.24)

which are equivalent to the following linear system


A A
a2 +B(1+2Ina)+2C=O; b2 +B(1+2Inb)+2C=O (8.25)

We now turn to the stress B.Cs. at the end sections (O = ±a). Since con-
centrated moments are applied there, we are going to write that -at each of
those sections- the stress vector (t) is such that its resultant is zero, and the
resultant of its moment is equal to the concentrated moment:

l b
t dr = O; l b
o7p xt dr = ±Mez (8.26)

where we have assumed a unit thickness. According to Saint-Venant's prin-


ciple, the solution which we shall find will be exact far from the end sections
(this implies that the height of the beam (b - a) should be small compared
to the length (2aro)). The stress vector (t) act ing on each end section is

t = u(r)· [±e(J(±a)]
±O"r(J(r) er(±a) ± O"o(J(r)e(J(±a)
-....-..-
o
Thus the moment of (t) w.r.t. the centroid (o±) of each end section is
---t
O±P xt = (r - ro)er(±a) x [±O"(J(J(r)eo(±a)]
±(r - rO)O"(J(J(r)e z

Since eo on each end section (O = ±a) is a constant vector, and ez is constant


too, the B.Cs. (8.26) give the following scalar equations:

l b
O"(J() dr = O; l b
rO"(J(J dr - ro l
~
b
O"(J(J dr =M (8.27)

o
We now follow the procedure of (Timoshenko and Goodier, 1987) to show that
these two scalar equations provide only one condition to be satisfied. Recall
that for axisymmetric problems, the stresses are derived from the Airy stress
function according to:
212 8. Two-dimensional problems in polar coordinates

Thus Eq. (8.27a) can be written as

laba(J(J dr = lba cfJt/J dt/J b b


d dr = [-d 1 = [rarr la = 0,
2
r r a
and is satisfied if the stress-free conditions (8.24) are enforced. Now, the
left-hand side of (8.27b) can be written as follows

labra(J(J dr = lba r cfJt/J


2
dt/J b l b
dr dr = [r-
dra 1- dt/J
a r
2 b b
-d dr = [r arrla -[t/Jla'
---......-..-
O

using integration by parts. Thus, expression (8.7) of t/J(r) shows that Eq.
(8.27b) gives the following condition:

To summarize, the three constants A, B and Care solution to the linear


system formed by Eqs. (8.25) and (8.28). After some algebra, the solution of
this system is found to be:

where'TI and V are dimensionless parameters given by:

Finally, plugging the expressions of A, B and C into Eqs. (8.23), the in-plane
diagonal stresses are found to be

(8.29)

The stress solution does not involve the opening angle of the beam (2a).
We should not forget however that the solution is valid far from the end
sections and that the height (b - a) must be small compared to the length
(2aro = a(a + b)). The stresses are plotted in Fig. 8.9 in the case b = 2a.
Intuitive expectations concerning the signs of radial and hoop stresses are
confirmed. The figures show that the radial stress is compressive everywhere
8.7 Solved problems 213

and Iurrl reaches a maximum value ~ 1.07M/a2 at r ~ 1.36a. The hoop


stress is zero at r ~ 1.445a ("neutral fiber", it does not coincide with the
"middle fiber" r = ro = 1.5a). AIso, U(J(J is compressive in the lower part of
the beam (r < 1.445a), tensile in the upper part (r > 1.445a) aud takes the
following values at the inner aud outer radii
M M
u(J(J(a) ~ -7.7552"
a
= -lu(J(Jlmaxi u(J(J(b) ~ 4.9172"
a
It is seen that the radial stresses -which are neglected in elementary beam
theory- are small compared to the hoop stresses according to 2D elasticity.
However, the elementary theory predicts a linear variation of the hoop stress
along the height (r) of the beam, while Fig. 8.9 shows that the variation is
not linear and that uIJ(J(b) 1- -u(JIJ(a).
Let us now check auother assumption of the elementary theory. Since the
tangential displacement (u(J) is proportional to (r) for a given (O), then each
cross section remains planar aIter deformation.

6~--~----~--~----~--~
4
2

O
-2

-4
-6

1.2 1.4 1.6 1.8 2


rla
Fig. 8.9. Normalized stresses (U88a2/M) and (u rr a 2/M) vs. normalized radial
distance (r la) in the case b = 2a

An application of the results found in this section is given by the problem


depicted in Fig. 8.10: what are the (residual) stresses generated by the closure
(by welding for instance) of a ring which has a small cut of angle (2{3)? In
order to answer this question, we first compute the bending moment (M)
needed to close the ring. The tangential displacement required in order to
put the two ends of the ring in contact is ~ 2{3r. On the other hand, this
relative displacement is:
4Br
u(J(r,7r - (3) - u(J(r, -7r + (3) = E(27r - 2(3)
214 8. Two-dimensional problems in polar coordinates

Therefore, equating the twa displacements, we find

B = {3E ~ (3E,
4('IT - (3) 4'IT
because (3 « 'IT. However, we have previously found that (B) is given by
B = 2M(rr - 1)/(a2V), thus the bending moment able to close the ring is:

M ~ (3Ea 2 V (8.30)
8'IT(1J2 - 1)
The residual stresses are then given by (8.29) after substitution of the ex-
pression of (M) that we just found.

Fig. 8.10. Closure of a ring having a small cut of angle (2{:J)

8.7.5 Traction of a circular arch


A semi-circular beam of inner and outer radii a and b, respectively, is sub-
jected to traction forces (±Fey) at its end sections O = ±'IT/2; see Fig. 8.11.
The croşs section is rectangular of thickness t. Plane stress is assumed. One
can prove that Airy's stress function is given as follows:

4J(r, O) = (Ar 3 + -Br + Cr + Dr In r) cos O (8.31)

In order to find this expression, we can proceed as follows. A fictitious cut at


an angle O gives the normal force N(O), the shear force Q(O) and the bending
moment M (O) -see Chap. 3- as follows:

N(O) = FcosO; Q(O) = F sin O; M(O) = F(a; b) casO;


8.7 Solved problems 215

Since N(O) and Q(O) are the stress resultants of U99 and Ur 9, respectively,
then looking at (8.5), it appears that a stress function under the following
form would work: r/J(r, O) = f(r) coso. Imposing that LlLlr/J(r, O) = O, Eq.
(8.31) is found. The stresses do not involve constant C; the other constants
are found from the stress B.Cs. which are:

r = a: u· (-e r ) = O;
r = b: u ·e r = O;

0= ±- :
2
7r
t l b
u· e9dr = ±Fey ; (8.32)

The following final stress expressions are found:

U rr

with: V (8.33)

It can be checked that the largest stress is the hoop stress computed at r = a
(the inner surface) and O = O (the cross section where the bending moment
M(O) is maximum). The value of the largest stress is:

(8.34)

The largest stress given by elementary beam theory (Chap. 3) is the sum of
two terms, one due to M(O) and the other to N(O),

U(O) = M(O) (b - a) + N(O) = 3F(a + b) + F (835)


max (b - a)t 3/12 2 (b - a)t t(b - a)2 t(b - a) .
Consider the following application: F = 5 kN, a = 7.5 cm, b = 10 cm, t =
2.5 cm. These data give: U max = 194.4 M Pa and u~~x = 176 M Pa, which
means that elementary beam theory underestimates the maximum stress by
a large margin (9.5 %); this is due to the fact that in this example the radius
of curvature (a + b)/2 is small compared to the length (7ra).

8.7.6 Rotating disk of uniform thickness

A circular disk of outer radius (b) and uniform thickness, rotates with constant
angular velocity (w) [rad S-I] around the (O, z) axis. Consider a material
216 8. Two-dimensional problems in polar coordinates

F
Fig. 8.11. Traction forces (±Fey) are applied to the end sections (8 = ±1f/2) of a
semi-circular beam of inner and outer radii a and b.

---t
point (P) in the disk whose position vector is op= re r , the velocity (v)
w.r.t. a fixed frame is given by
~ .
v = op = rer = r(}e/J = rwe/J,
where a superposed dot means "â/ât at fixed r", t designating "time". The
acceleration ("Y) is given by:

"Y = v. = rwe/J. = -rw ().e r = -rw2 e r


Since we can reasonably assume that ar/J = O and a rr and a/J/J do not depend
on ((}), it is found that the dynamics equations div (T = P"Y (p [kg/m 3 ] being
the mass density) give one non-trivial equation only:

darr arr - a/J/J 2


--+
dr r
=-prw (8.36)

Now, consider an axis system which is linked to the disk, i.e. rotates with
it at the same angular velocity. If we move the term (prw 2 ) to the left-hand
side of (8.36), then the latter looks like the equilibrium equation of a static
problem with a force per unit volume fr = prw2 [N/m 3 ] (or a body force
rw 2 [N/kg]). Since there are body forces, we cannot apply the Airy stress
function technique. We proceed as in the method suggested at the end of
8.7 Solved problems 217

Sect. 8.7.1. Symmetry suggests displacements of the form Ur = u(r) and


Uo = O. We compute the strains, then the stresses from the constitutive
model, we substitute the expressions into (8.36) and find:

cFu 1 du u
-+ ----= (8.37)
dr
2 r dr r2
The general solution of Eq. (8.37) is given by

A (1 - v 2 ) 2 3
Eu = 2(1 - v)Cr - (1 + v)-
r
-
8
pw r (8.38)

where A and Care constants to be determined (the multiplying factors 2(1-


v) and (1 + v) are only there to obtain stress expressions similar to (8.8)).
Computing the strains and then the stresses from the constitutive model, we
find

We now consider two situations. If the disk contains the origin (r = O), then
we must set (A = O) in order to have finite stresses at (r = O). The stress
B.Cs. at the boundary of the disk (r = b) allow the computation of the
constant (C). For instance, if the boundary is stress free, then

(3 + v) 2
urr(b) = O ====? 2C = - 8 - P(wb)

Therefore the stresses are given by:

(8.40)

The stresses reach their maximum values at the center of the disk

) (3+v) 22
urr(O) = uoo(O = -8-PW b

Note that the maximum stresses increase with the square of the outer radius
(b) and the square of the angular velocity (w).
A second situation we consider is a disk with a concentric circular hole
of radius (a). Constants A and Care computed from the stress B.Cs. on the
inner and outer radii, (r = a) and (r = b). For instance, if those surfaces are
stress free, then: urr(a) = urr(b) = O, and the constants are

The stresses are then given by


218 8. Two-dimensional problems in polar coordinates

(8.41)

In studying the variations of O'rr and 0'00, it is found that the radial stress is
maximum at r = v;;b where

(O'rr)max = (3 +8 v) pw2( b - a
)2

The hoop stress is maximum at (r = a) where


( ) max
O'o() (3 + v) 2 [b 2+ -
= ---pw (1-- -v)a2]
4 3+v
It is easy to check that (uoo)max > (O'rr)max. If a -t 0, then it is seen that
the maximum stress is double the value which was found for a disk with no
hole. This is an example of stress concentration, a very important problem
in practice. The stresses are plotted in Fig. 8.12 for two cases: a disk with no
hole, and a disk with a concentric hole of radius a = b/5.

2.5 .....---r---r---,--r--...,
2 \ 88 (i) --
\ 88 (ii) -----
1.5 \. rr (i) .......... .
" .... rr
...... (ii) -. - .-
1 r---.....,~_ ....... _-
I ~ ~.~.::.::.:.~.:.::::~:--
...
0.5
I .........
o
-0.5 ....._ ......_ - - 1 ._ _.&....._........_ - - '
O 0.2 0.4 0.6 0.8 1
r/b
Fig. 8.12. Stresses in a rotating disk of uniform thickness and outer radius (b) in
two cases: (i) No hole, (ii) A concentric hole of radius a = b/5.

8.7.7 Rotating disk of variable thickness

Figures 8.12 show that the stresses are localized in the central region of the
disk, especially when the disk has a perforation at the center. This means
that when the thickness of the disk is uniform, the material is not used in
8.7 Solved problems 219

an optimal fashion, this is why disks with variable thickness h(r) are used in
practice. Those "disks" are actually solids of revolution. Both the geometry
and the loading possess an ruda! symmetry around (O,z), see Fig. 8.13. Since
we do not have a plane problem anymore, we use cylindrical (r, O, z) instead
of polar coordinates (r,O). Our presentation mostly follows (Crochet, 1993).

h(r) r
o

Fig. 8.13. Disk of variable thickness

In a plane (O = constant), the tangent T and outer normal N to the


outer surface z = ±h(r) are given by

T = (dr)e r ± (dh)ezj N = -(dh)e r ± (dr)e z

which gives liNII = (drh/1 + (dhJdr)2. Assuming that the thickness varies
slowly (ldhJdrl « 1), the unit vector normal to the outersurface z = ±h(r)
and pointing outside the surface is
dh
n = -(dr)er ±ez
The outer surface z = ±h(r) is stress free, thus the stress vector (u T . n)
acting on that surface is nil
dh
At (r, z = ±h(r)) : urrnr + uzrnz = -Urr dr ± U zr = Oj
dh
urznr + uzznz = -Urz dr ± Uzz =O (8.42)

Because of the ruda! symmetry, the stresses are independent of O and:

Urr(r, -h) = urr(r, h)j uzAr, -h) = Uzz (r, h)j


Equation (8.42a) implies that:
220 8. Two-dimensional problems in polar coordinates

Therefore, the following equality is obtained:

azr(r, -h) = -azr(r, h)

Equilibrium in the radial direction gives (see Appendix A, the stresses not
depending on O)

aarr aazr
Tr+az+
a rr - a()()
r
2
+pwr=
°
This equation is integrated in the thickness direction:

h(r) aarr
j - a dz + azr(r, h) - azr(r, -h)
-h(r) r ' '" J

217 zr (r,h)

1 jh(r)
+-
r -h(r)
(arr - a()()) dz + 2pw2r h(r) = °
We recall Leibnitz's formula
d
-d
lb (a)
f(a,x)dx =
l af
b (a)
-a dx + f(a,b(a))-d
db da
- f(a,a(a))-d
a ata) ala) a a a
We use this formula in order to obtain the following result (where r plays the
role of a and z that of x)

d jh(r) jh(r) âarr dh dh


-d arr(r, z) dz = - a dz + arr(r, h)-d - arr(r, -h)( --d )
r -h(r) -h(r) r ,r ... r J

2I7 rr (r,h)dh/dr

The equilibrium equation integrated in the thickness thus becomes:

d jh(r} dh ljh(r}
d a rr dz + 2 [azr(r, h) - arr(r, h)-d 1+- (a rr - a()()) dz
r -h(r) , r r -h(r) J

'"

We now define mean values (through the thickness) of the stresses:

1 jh(r) 1 jh(r)
O'rr(r) == 2h( ) a rr dz; O'()()(r) == 2h( ) a()() dz
r -h(r) r -h(r)

With these notations, the equilibrium equation can be written as follows

ddr (hO'rr) + !!.(arr - O'()()) + pw2r h = 0,


r
or equivalently
8.7 Solved problems 221

(8.43)

As an application, we consider a disk without a hole and we wish to compute


the thickness h(r) such that we have a uniform average stress:

O'rr = O'()() = constant = ao; Vr

This gives the following differential equation

dh 2 2
aor- = -pw r h
dr
whose solution is given by

pw 2 r2
h = ho exp( - - - ) (8.44)
2ao
This result -found in Sweden around 1900- can be rewritten as

-hho = exp[-(-)];
r 2
>.
>. == fff,ao
- 2
pw
(8.45)

(>. [m] has a dimension of a length). The thickness profile is plotted in Fig.
8.14. Note that the curve h(r) is not unique because it involves an arbitrary
constant (h o). The shape of the disk is also dictated by the constant stress ao,
therefore we may design a suitable disk geometry by tuning the values of h o
and ao, subjected to some design constraints. There is a long and interesting
discussion of the subject as well as several examples in (Den Hartog, 1952).

1
0.9
0.8
0.7
o
0.6
..c:: 0.5
---
..c::
0.4
0.3
0.2
0.1
O
O 0.5 1 1.5 2
rl>'
Fig. 8.14. Normalized thickness h(r)lho versus normalized radial distance rl>' for
a disk with uniform average stresses
222 8. Two-dimensional problems in polar coordinates

8.7.8 Stress concentration in a plate with a small circular hole

A rectangular plate has a small hole of radius (a) at the center and is sub-
jected to uniform traction (S) in the x-direction, see Fig. 8.15. We assume
that (a) is very small compared to the in-plane dimensions of the plate (21 in
the x-direction and 2b in the y-direction). By comparison to a non-perforated
plate (problem (O)), we expect the stress solutions to be different near the
hole, but basically identical far from the hole. The stress solution in Cartesian
coordinates for problem (O) is obviously given by

0'(0)
xx
= S·' y0'(0)
y
= o·' x0'(0)
y
= O

Using (8.3), these stresses become in polar coordinates:

Scos 2 () = %[1 + cos(20)]; a~~) = Ssin2 0= %[1- cos(2())];


= -Ssin(())cos() = -%sin(2())

We now return to the problem of a plate with a hole and isolate a ring of
inner and outer radii (a) and (b), respectively. Since b» a, then as we wrote
earlier, the stress vector (o- . e r ) act ing on the surface (r = b) is equal to
(0-(0) . e r ) of problem (O):

at r = b: 0-. er 0-(0) . er
a rr e r + a ro eo
(O) (O)

%e r + %[cos(2())er - sin(20)eo]
'"'-v-" '" '
problem (1) problem (2)

It is seen that the solution in the domain (a :::; r :::; b) can be obtained
by superposition (addition) of the solutions of two problems: (1) and (2).
Problem (1) is a Lame "cylinder" which is stress free at the inner radius
(Pa = O at r = a) and subjected to a traction S/2 at the outer radius
(Pb = -S/2 at r = b). The solution of problem (1) is given by (8.13),

(1) _
arr,OO -
S
2"
[1=f (a/r)2]
1- (a/b)2 :::::!
S [
2"
2]
1 =f (a/r) , (8.46)

and a~~) = o. As for problem (2), the expression of the stress vector at (r = b)
suggests the following periodic Airy function:

cp2(r,0) = fz(r) cos (2())


This corresponds to the case (m = 2) of Sect. 8.9. Consequently, cp2(r, O) is
bi-harmonic if:
8.7 Solved problems 223

C
ifJ2 (r, O) = (Ar 2 + Br 4 + 2" + D) cos(20)
r
where A, B, C and Dare constants to be determined. The stress field of
problem (2) is obtained from ifJ2(r, O) by application of Eqs. (8.5):
3C 2D
-2(A + -4
r
+ -r2 ) cos(20);
3C
2(A + 6Br 2 + 4) cos(20);
r
3C D .2
2(A + 3Br - - - - ) slll(20) (8.47)
r4 r2
We now write the stress B.Cs. for problem (2). The inner radius (r = a) is
stress free while the stress vector at (r = b) was already computed,

at r = a: /7(2) . (-e r ) = 0,
at r = b: /7(2) . er = %[cos(20)e r - sin(20)eoJ (8.48)

The stress B.Cs. give four scalar equations:

(Ţ(2)
rr o·,
(a' O) = (Ţ~~) (a, O) = O;
S
a(2) (b O) = - cos(20)·
rr' 2 ' a~~) (b, O) = -% sin(20)
Thus, the constants are solution of the following linear system
3C
+ 3Ba2 - -3C4
2D D
4 +-2 =0,
A+- A - - = 0,
a a a a2
3C 2D S 2 3C D S
A + b4 + b2 = -4' A+ 3Bb - - - - = - -
b4 b2 4'
which is solved with the assumption (alb« 1). The solution is:

A=-~·
4'
B=O·
,
C=-a4~.
4'
D=a2~2
Now, as we explained earlier, the stress solution of the original problem is
+ /7(2). Putting together the different results, it is found that:
/7 = /7(1)

a 2+ [1 + 3(-;:.)
"2S { 1- (-;:-) a 4- a 2] cos(20) } ;
4(-;:-)

(ŢOO %{1 + (;)2 - [1 + 3(;)4] COS(20)};


a 4+2(-)
S [1-3(-)
-- a 2] slll(20)
. (8.49)
2 r r
This solution was first found by Kirsh in 1898. Although (b) does not appear
in the stress expressions, in order for the solution to be valid, the hole radius
224 8. Two-dimensional problems in polar coordinates

(a) must be very small compared to the in-plane dimensions of the plate
(a « b and a « 1). As initially assumed, it is checked that if r -t 00, then
Urr -t u~~), uee -t u~~) and ure -t u;~). This means that far from the hole
(Le., for r/a» 1), the solution is identical to that for a non-perforated plate.
At the hole boundary (r = a), the stresses are

Urr(a,O) = ure(a,O) = O, and uee(a,O) = S[l- 2cos(20)]

It is seen that Uee is maximum for 0= ±7f/2 , i.e. at points (m) and (n) of
Fig. 8.15. On those points, we have

(uee)max = 3S, (8.50)

Le. three times the value of the traction which is applied at the ends of the
plate. This is another instance of stress concentration. If the maximum stress
is large enough to cause fracture, then cracks will appear at points (m) and
rut
(n) and propagate in the vertical directions m---nţl and l .
At points (P) and (q) (Fig. 8.15), we have uee = -S, Le. compression in the
tangential (vertical) direction. Along the y-axis, O = ±7f/2, and

7f S [2+(-)
a 2 a 4]
uee(r,±-)=- +3(-) =uxx(x=O,y)
2 2 r r
These stresses are plotted in Fig. 8.15. It is seen that the perturbation due
to the hole is localized at its vicinity and that far from the hole the stress is
equal to the far end solution U xx = S.
Exercise: Compute the displacement field corresponding to the stress
solution (8.49). Hint: same technique as in Sec. 8.4.
Exercise: Consider a problem identical to the one which was solved in
this section except that the plate is subjected to uniform traction (S) at
its ends x = ±l and uniform traction (T) at its other ends y = ±b. Hint:
superposition principle and change of coordinates formulae.

8.7.9 Force on the straight edge of a semi-infinite plate

A semi-infinite plate is subjected on its straight edge to a force (Pe x ), which is


uniform in the thickness direction; Fig. 8.16a. Actually P [N/m] is a force per
unit thickness. We shall consider a unit thickness for simplicity. It is assumed
that the B.Cs. on the remote border (at "infinity") are such that the plate
is in equilibrium. A 3D version of the problem was solved by Boussinesq in
1885, and Flamant in 1892 found the solution of the 2D problem we consider
here.
Let us isolate a half-disk of "center" (A) and radius (1'), Fig. 8.16a. Intu-
itively, we imagine that on the circular border ofthe half-disk act compressive
stresses in the radial directions. Moreover we expect those stresses (u rr ) to
be proportional to PcosO/(7f1') because PcosO = Pe x . er is the projection
8.7 Solved problems 225

/ ......

/
/ 2a
x

s s

y
Fig. 8.15. Plate with a small circular hole subjected to traction at its ends

P P
y A y A

x x
Fig. 8.16. Concentrated force on the straight edge of a semi-infinite plate
226 8. Two-dimensional problems in polar coordinates

of the external force in the radial direction and (7rr) the perimeter of the
halI-disk. Therefore, we look for a stress solution under the following form:
kP
U rr = - - cos 8j Ur(J = U(J(J = O
7rr
where k is a constant to be determin~d. Static equilibrium of the forces acting
on the half-disk gives the following condition
7r/2
j (O" er)r d8 + Pe", = O,
-7r/2
which can be rewritten as

Using the expressions of the assumed stress field and that of e r in terms of
e",and e y , we obtain two scalar equations

kP j7r /2 j7r /2
-- cos 2 (8) d8 + P = O; sin(8) cos (8) d8 =O (8.51)
7r -7r /2 -7r /2
The second equation is identica1ly satisfied and the first equation gives k = 2.
Therefore, we need to prove that the following stress field is the solution we
are looking for
2P
U rr = --cos8j Ur(J = U(J(J = O (8.52)
7rr
The radial stresses acting on the circular boundary have a symmetric dis-
tribution w.r.t. the (x) axis. It is easy to check (e.g., graphically) that their
resultant has no horizontal component (the vertical component is by construc-
tion equal to P), and that the moment of alI forces acting on the half-disk
w.r.t. (A) is zero. The radial stresses present a singularity at (r = O), the
application point of (P) (if the material is ductiIe, this means that plasticity
will develop in a small region around (A) with finite values of plastic stresses).
On the horizontal border (8 = ±7r/2), urr(r, ±7r/2) = O (except for r = O).
The only in-bulk non-trivial equilibrium equation is

8urr + U rr = O
ar r
and is satisfied. Finally, it can be easily checked that the following compati-
bility equation is satisfied

1 8urr 8 2 urr 1 8 2 u rr
+ ~ = ;. ar +
UlIII )
Ll
(
U rr 8r2 + r2 88 2 =O
o
8.7 Solved problems 227

Exercise: Find an Airy stress function 4>(r, O) for the problem. Hint: integrate
Eqs. (8.5) after using (8.52). Answer: 4>(r,O) = -PrO(sinO)/7r.
The stresses acting on a horizontal plane (x = a > O), Fig. 8.16a, are
given by the transformation rules (8.3) as:

_ 2P cos3 O = _ 2P cos 4 O < O


7rr 7ra
_ 2P cos (O) sin (0) = _ 2P (sin(O) cos 0)2 < O
2
7rr 7ra
- 2P cos2 (O) sin O = - 2P sin( O) cos3 O (8.53)
7rr 7ra
The stress distribution is shown in Fig. 8.17. Recall that the (O = O) axis
corresponds to the direction of the externalload. The horizontal (ayy ) and
vertical (a",,,,) stresses are compressive and symmetric w.r.t. the vertical (O =
O) axis. For (x = a) and (O = O), U yy is zero and a",,,, maximum.

0.4
.......................
...........

-
0.2 ..........
d ••.. yy -----
~ O .... -_.":.-- .... xy .......... .
Q., ,. .. ....
,,'" .........,
N
-0.2 -'""'" ..... ...... ........
E .....................
'tl
CI)
-0.4
:'Sl
> -0.6
:a
b
-0.8

-1
-1 -0.5 O 0.5 1
() [rad]
Fig. 8.17. Stresses U:z:x, Uyy and Uxy at (x = a), versus the polar angle «(})

Isotropic linear elasticity and Eqs. (8.51) give the strain field as:
2P 2vP
trr = --E
7r r
cos O; t(J(J = --
7rEr
cos O; tr(J =O (8.54)

Exercise: Find the displacement field. Hint: Same technique as in Sect. 8.4.
Answer:

Ur = - !~ cos(O) In r - (1 :;)P Osin(O) - B sin (O) + C cos O;


U(J = !~ sin (O) (v + lnr) - :;)P
(1 (Ocos(O) - sin O) - Bcos(O)
- C sinO + Dr (8.55)
228 8. Two-dimensiona1 problems in polar coordinates

In order to determine the three constants (B, C and D), we can assume, for
instance, the following conditions:
(i) The points on the (x) axis have no horizontal displacement: U(J = O for
() = O. This implies that B = D = O.
(ii) There exists a point (x = d, y = O) which is situated "far enough" from
the load (P) so that its vertical di:;;~lacement is zero: ur(r = d, () = O) = o.
This gives the displacement of the (x) axis points as
2P d
ur(r,O) = -ln(-)
7fE r
(8.56)

The solution found in this section has several applications. For example in
soil mechanics, it can be used for a pile (foundation) carrying a load P. Other,
less apparent applications (stresses in compressed disks) are studied in Secs.
8.7.11 and 8.7.12 and are based on results which we establish hereafter.
Consider in Fig. 8.12b a (full) disk of diameter d and center O, which is
tangent to the straight edge at point A. Let M(r, () be a point on the circular
boundary and n the outward unit normal at M. It can be shown that:

n = cos(()e r + sin(()e(J (8.57)

The following equalities are clear:


--t --t --t d --t d
OM=OA + AM= -"2e", + rer; OM="2n (8.58)

U sing the identities

ey = cos(()e r - sin(()e(J, cos() = ~,


equality (8.57) immediately follows. The stress vector at M is

u· n = cos(()u· e r - sin(()u· e(J = O"rrCOS(()er

Since cos8 = rid at point M, it appears, using (8.51) that:


u· n = -2P
- cos(8)e r
7fd
(8.59)

This means that the boundary of the disk is subjected ta nonuniform com-
--t
pression 2Pcos81(7fd) directed along -e r =MA Ir.

8.7.10 Pressure on the straight edge of a semi-infinite plate

A semi-infinite plate is subjected an its straight edge ta a load p which is


uniformly distributed over segment [01, O2 ]; Fig. 8.18.
Exercise: show that the stress solution is obtained by superposition of
the stresses derived from the following two Airy functions:
8.7 Solved problems 229

where A is a positive constant and (rl' lh), (r2, ( 2) determine the position of
a point M as indicated in Fig. 8.18. One can easily prove that in order for
the stress vectors derived from the two Airy functions to sum up to the given
pressure on the straight edge, we must have:

A=E. (8.61)
27r
Transforming the stress components due to each Airy function to the same
Cartesian system (O, x, y) depicted in Fig. 8.18, and adding them up, the
following stress solution is found:

axx,yy(X, y) _P..{3 =f ..!!...[sin(202 ) - sin(20l )),


7r 27r
= E.[cos(20
27r
2 ) - cos(20 l )), (8.62)

where {3 == O2 - 01. It is found that the principal stresses are given by the
simple formula:

(8.63)

In particular, along the (Ox) axis, we have {3 =o - (-o) and thus

al
,2(X, O) = E[-20
7r ± sin(20))
Finally, along any circle passing containing 01 and O2 , angle {3 is constant,
and thus the principal stresses on any such circle are constant too.

Fig. 8.18. Pressure on the straight edge of a semi-infinite plate.


230 8. Two-dimensional problems in polar coordinates

8.7.11 Compression of a disk along a diameter

A circular disk of diameter d = 2a is subjected to opposite forces (=t=Pe y )


act ing at (x = 0, y = ±a), Fig. 8.19a. Let M be a point on the circular
---t
boundary and n the outward unit normal at M, n =0 M / a. Adding the
stress vectors at M computed in (8.58) for a force (-Pe y ) act ing at A(x =
0, y = a) and a force (+Pe y ) acting at B(x = 0, y = -a), it is found that:

(8.64)

We point out that there are three polar coordinate systems that should not
be confused: one is attached to point A as defined in Fig. 8.16, a second one
is attached to point B in a similar fashion and a third one is linked with the
(O, x, y) Cartesian system of Fig. 8.19a. The latter figure shows that:

e~B) = e~A) j O(B) = O(A) - i-


Substituting in Eq. (8.63), we obtain:

O'(A) . n + O'(B) . n = _.!...(cos O(A)e(A) + sinO(A)e(A)) = -.!...n, (8.65)


7ra r (J 7ra

using (8.56). This corresponds to uniform radial compression P/(7ra) applied


on the boundary. It is then concluded that the stress solution of the problem
at hand is obtained by superposition of three stress states corresponding to:
(a) force (-Pe y) acting at A(x = 0, y = a),
(b) force (+Pe y ) acting at B(x = 0, y = -a),
(c) uniform radial traction P/(7ra) on the boundary.
Transforming the stress components from the local systems defined above
to the Cartesian system of Fig. 8.19a, the final expressions are found:

.!... _ 2P (a - y)x 2 2P (a + y)x 2


7ra 7r [x + (a - y)2J2 - -;- [x 2 + (a + y)2J2'
2

P 2P (a_y)3 2P (a+y)3
= 7ra - -;- [x + (a - y)2J2 - -;- [x + (a + y)2J2 j
2 2

2P (a - y)2x 2P (a + y)2X
(8.66)
7r [x 2 + (a - y)2J2 7r [x 2 + (a + y)212

The stresses at the center of the disk are:


2P -6P
O'xx(O, O) = 7rd j O'yy(O, O) = 7rd j O'xy(O, O) =O (8.67)

It is seen that -contrary to many students' initial belief- the stress state at the
center does not correspond at alI to a uniaxial stress in the y-direction equal
8.7 Solved problems 231

to (-P / d). Actually, the latter value corresponds to the avemge yy-stress
along the (y = O) diameter:

1 jd/2 _p
< lTyy(x, O) >= d lTyy(x, O) dy =d (8.68)
-d/2
The result can be established by direct integral computation, or simply by
making a fictitious cut along the (y = O) diameter and writing static equilib-
rium of half the disk.

O! : O!

~
, I

d d
2' "2
x x

.4 d
2 "2

B
p
O! O!

Fig. 8.19. Compression of a disk: (a) along a diameter, (b) over two opposing arC8.

8.7.12 Compression of a disk over two opposing arcs

A circular disk of diameter d = 2a is compressed by radial forces q uniformly


distributed along two opposing arcs of opening angle 20: eachj Fig. 8.19b.
The stresses in the center of the disk are obtained by application of the
superposition "principle" to (8.67). Each pair offorces acting on the opposite
ends of a diameter can be parameterized by an angle (). The stresses in the
local system (O, X9,Y9) are given by (8.67). We first transform those stresses
to the global system (O,x,y) of Fig. 8.19b. Next, we replace P with (qadO).
Finally, we integrate () between (1f /2) - o: and (1f /2) + a. Consequently, the
stresses at the center are:
(11" /2}+0 1
[ -[-1 =f 2 cos(2())]qa dO
lTxx,yy(O, O)
(1I"/2}-0 1fa
(11"/2)+0 -2
[ - sin(2())qa d()
lTxy(O,O)
(11"/2)-0 1fa
232 8. Two-dimensional problems in polar coordinates

Simple integral computations give:

axx,yy(O, O) = - :: [a =f sin(2a)]; axy(O, O) = O (8.69)

Due to the symmetry, result (8.68b) was expected. The following extreme
case is obvious:
1f
a = "2 ===} aXX,yy(O, O) = -q,
these are indeed the stresses corresponding to uniform radial compression q
along the entire boundary. Setting P = 2aaq, we can rewrite (8.69a) as:

-
axx,yy (O ,O)- -~
1fa
[1 =f 2 sin2(2a)]
a

Another extreme case is then clear:


P
a ---t O===} a xx yy(O, O) = --(1 =f 2),
, 1fa

and these are the stresses found in (8.67) for a pair of concentrated forces P
compressing the disk along a diameter.
9. Thermo-elasticity

In an previous chapters, we have considered isothermal problems. This chap-


ter deals with the computation of thermal stresses.

9.1 Constitutive equations


Consider a bar of length l which is fixed at one end (x = O) and subjected
to a uniform change in temperature from TreI to T, see Fig. 9.la. The strain
in the bar is purely thermal and its value is fth = a(T - TreI), where a
([lrCJ or [1/ K]) is the coefficient of thermal expansion. TreI is a reference
temperature for which by definition the thermal strain is zero (fth(Tre/) = O).
Since the bar can expand or retract freely and there is no mechanical load,
the stress is zero (O" = O). Now, assume that in addition to uniform heating
or cooling (T - TreI)' the bar is also subjected to an axial force F at its end
(x = l), Fig. 9.lb. The key idea is that the total strain in the bar is the sum
of mechanical and thermal strains. This can be written as follows:

(9.1)

whereE is Young's modulus and O" = F/A the stress, A being the cross section
area. Idea (9.1) is extended from lD to 3D in a straightforward fashion:
(9.2)

where c is the fourth-order Hooke's elasticity tensor. Note that Eqs. (9.2) are
form-identical to those of isothermal elasticity, the total strain.: in the latter
case being replaced with the mechanical strain (.: - .:th). In the anisotropic
case, f~J = (T - TreI )aij, while in the isotropic case, f~J = (T - TreI )a8ij
and Eq. (9.2a) becomes:
l+v V •
--yu- E(tru)l+a(T-Tre/)l, l.e.
l+v V
= --YO"ij - EO"mm 8ij + a(T - Tre /)8 ij , (9.3)
234 9. Thermo-elasticity

where v is Poisson's ratio. Relations (9.3) can be easily inverted by using Eq.
(9.2b) which gives in the isotropic case:

where A and J.L are Lame's coefficients. Using relations from Sec. 1.11, the
previous equations can be rewritten as follows:

Ev E E
(1_2v)(1+v)(tr€)1+ 1+v€-1_2va(T-Tre/)lj Le.
Ev E E
= (1 - 2v)(1 + v) EmmtSij + 1 + v Eij - 1 _ 2v a (T - Trei )tSij (9.4)

Note that in general, the material properties (E, v,a) may depend on
temperature (e.g., E(T) decreases when T increases).
As an application, we shall compute the increase in temperature (tST)
needed in order to make a ring of inner radius a at To = Trei fit around a
disk of radius e > a. We assume that (tST) is uniform. This may be difficult
to realize in practice. For instance, if we heat the inner face (r = a) and leave
the outside face at room temperature, then tST is not uniform but varies with
In r (see Sect. 9.5). We also assume that the displacement is radial: U r = u(r),
U8 = O. This gives the strains as (rr = du/dr and (88 = u/r. Since the strains
in this problem are purely thermal, then

(rr = (~~ = atST, and (88 = 4i = atST


The displacement is then simply u = ratST. The ring can be fitted around
the disk if a + u(a) ~ e, Le. a + aatST ~ e. This gives:

tST>~
- aa
For (e - a)/a = 10- 3 and a = 1O- 5 / o C, we find tST ~ 100°C.

9.2 Heat equation

Starting from the first law of thermodynamics, it is shown in Sec. 12.11.5


that the heat equation can be written in the following local form:

8q· .
-div q + r = per, Le. - _3 + r = peT (9.5)
8xj

where p [kg/m 3 ] is the mass density, e [J/kg/K] the specific heat, T [K] the
absolute temperature, a superposed dot means a time derivative, r [W/m 3 ]
is a specific heat supply, Le. a volumetric heat source and q the heat flux
9.3 Thermo-mechanical problem 235

Fig. 9.1. Bar fixed at one end (x = O) and subjected to uniform heating or cooling
from Trej to T; (a) no mechanical load is applied, (b) an axial force F is applied
at the end (x = l).

vector. Fourier's heat con duct ion law relates q linearly to the temperature
gradient:

q = -k· 'VT, i.e. (9.6)

or, in the isotropic case:

q = -k'VT, Le. (9.7)

where k [WImi K] is a thermal conductivity coefficient. Note that the heat


equation and Fourier's law look like the equations of mot ion and the consti-
tutive equations, respectively. If k does not vary with spatial position (x),
then using (9.7), Eq. (9.5) becomes:

kL1T +r = peT (9.8)

(L1 being the Laplacian operator). In the particular cases when we have a
steady state regime (T = O) and no volumetric heat source (r = O), the heat
equation (9.8) takes a very simple form:

(9.9)

9.3 Thermo-mechanical problem

In linear thermo-elasticity, it can be shown (see Sec. 12.11.5) that the thermo-
mechanical problem can be solved in an uncoupled way, i.e. we first sol ve
a thermal problem in order to compute the temperature field, and then a
mechanical problem in order to compute the stress, strain and displacement
fields with the temperature field as given data.
236 9. Thermo-elasticity

Consider a solid body which -before deformat ion- occupies an open set
nof JR3 and is subjected to mechanical and thermalloadings (Fig. 9.2). We
designate by r the boundary of n and n the union of n and r (li is a closed
set of lR3 ). If the problem depends on time, then the solution is sought over
a time interval [O, tI l.

9.3.1 Thermal problem

The data are:

r : nxlO, tl[~ lR (specific heat source r in n);


T g : rTxlo, tl[~ lR (given temperatures Tg on rT);
Q: rQxlO,t/[~ lR (given heat flux Q on rQ);
T o : n ~ lR (initial temperature field To in n);
p, e, k or k ij (material properties).

It is assumed that r T and rQ do not vary with temperature.


n
The problem is: find T : x [O, t/l ~ lR such that:

-divq+r=pcT, in nxlo, tl[ (heat equation)


T=Tg , on rTxlo, t/[ (temperature B.C.)
-q·n = Q, on rQxlo, t/[ (heat flux B.C.)
T(x, O) = To(x), for xEn (initial temperature condition)
q = -k· \lT, in nx]O, t/[ (heat conduction law).

If we have a steady state regime, then the temperature field does not
depend on time, and the thermal problem has a much simpler formulation.
The data are:

r:n~lR (specific heat source r in n);


T g :rT ~ lR (given temperatures Tg on rT);
Q:rQ ~ lR (given heat flux Q on rQ);
k or k ij (material properties)

The problem is: find T : n ~ lR such that:

-div q+r =0, in n (heat equation)


T=Tg , on r T (temperature B.C.)
-q·n=Q, on rQ (heat flux B.C.)
q = - k· \lT, in n (Fourier's heat conduction law).

Note that, in both cases, it is possible to find a variational formulation


for the thermal problem, e.g. (Hughes, 1987).
9.4 Thermal stresses: some remarks 237

9.3.2 Mechanical problem

In both cases (transient or steady state), once the temperature field (T(::c, t)
or T(::c)) is found, we move on to the mechanical problem, with the tempera-
ture field as given data. The other data are (we assume a quasi-static problem
for simplicity):

h:[}~lR. (forces per unit volume f in [});


Ui: ru ~ lR. (given displacements U on ru);
Fi: rF ~ lR. (given forces per unit area F on rF ).
E, v, a (material properties in the isotropic case).

The problem is: find the displacement, strain and stress fields (u, €, 0") such
that:

div O" +f=0, in [} (in-bulk equilibrium equations)


u=U, on ru (displacement B.Cs.)
O"·n =F, on r F (stress B.Cs.)
O" =c : (€ _ eth ) m [} (linear thermo-elasticity)
1
€ ="2[Vu + (VU)T] in [} (strain / displacement relations).

In the isotropic case, the linear thermoelastic relations are Eqs. (9.3) or (9.4).
The only difference between the mechanical problem thus defined and
the classical isothermal mechanical problem of Chap. 1 resides in the linear
thermoelastic constitutive model.
Let us remark that if we use a strain-based approach to solve the prob-
lem, then we need to satisfy the compatibility equations, and these are exactly
those of Sec. 1.9. If we seek a solution in terms of stresses, then we also have to
satisfy compatibility equations, but the Beltrami-Mitchell equations of Sec.
1.14 are no longer valid, because their derivat ion was based on isothermal
constitutive equations. In general, the results which were found in the previ-
ous chapters remain valid in the thermo-mechanical case if the constitutive
model was not used to derive them. For instance, for plane problems without
body forces, the stresses can stiU be derived from an Airy stress function f/J
according to Eqs. (7.12) or (8.5), but compatibility is no longer satisfied by
requiring f/J to be bi-harmonic.

9.4 Thermal stresses: some remarks


(1) Consider a solid body which is subjected to a temperature field T(::c) but
is not constrained by external forces or imposed displacements. The solution
aij = O satisfies equilibrium in every point as well as the stress B.Cs. For tij =
238 9. Thermo-elasticity

Tg
Fig. 9.2. Body under thermo-mechanicalloadings

(T - Trei )o:eSij to be the strain solution, it has to satisfy the compatibility


equations of Sec. 1.9, which can be written in this case as:

2
-88
8€ O·· = 1, 2, 3 ;
=;~, J f == fU = €22 = f33
Xi Xj

If the material is homogeneous, then o: and Trei do not depend on X and the
compatibility equations become:

8:Jxj = O; i.e. T = A(t) + B(t)x + C(t)y + D(t)z


It appears then that only a temperature field which is linear in space does
not give thermal stresses. Two elementary examples (a bar and a ring) were
given in Sec. 9.1. However, we shall keep in mind the assumptions under
which the conclusion was found, and examine the following two examples.
(2) Consider a bar clamped at both ends and subjected to a uniform
change in temperature (eST), Fig. 9.3. Although eST is uniform and the ma-
terial is homogeneous, the bar will have a thermal stress because it is con-
strained in its displacement. Equation (9.1) gives the total axial strain (which
is zero) as:
U
€ = O= - + o:eST =::} u = -Eo:eST
E
Therefore, if eST > O, there is a compressive stress in the bar given by the
formula above; the reactions at the built-in ends are H = uA, with A the
cross section area.
(3) As another example, consider a heterogeneous material which is not
constrained. Even a uniform temperature change (eST) will cause thermal
stresses. This problem is treated in detail in Sec. 9.5.2.
9.5 Solved problems 239

il----fJ-T >_0 _~~I---H


Fig. 9.3. Bar clamped at both ends and subjected to a uniform change in temper-
ature (5T)

9.5 Solved problems

9.5.1 Axisymmetric thermal stresses in a hollow cylinder

In this section, we consider a long hollow circular cylinder, with inner and
outer radii a and b, which is subjected to a radial temperature field T(r).
Firstly, we shall solve the mechanical problem, assuming that T(r) is given.
Due to the axial symmetry of the problem, we assume that the displacement
field is radial only: U r = u(r), U9 = O, and assuming plane strain, Uz = O.
The total strains are {rr = dujdr, {99 = ujr and alI other strains are nil. The
stresses are given by (9.4); it is found that alI shear stresses vanish. There is
only one non-trivial scalar equilibrium equation:

darr arr - a99


--+
dr r
=O
After some algebra, it is found that this is equivalent to the folIowing differ-
ential equation:
d[ld ] (l+v)dT
dr -;. dr (ru) = a (1 -v) dr

whose general solution is:


B (1 + v) 1
u(r)=Ar+-;:-+a(l_v)-;'la Tpdp
r (9.10)

Constants A and Bare found from the stress B.Cs. Assuming that the inner
and outer surfaces are stress free, two non-trivial scalar equations are found:
arr(a) = Oand arr(b) = O. Using (9.4), the strainjdisplacement relations and
Eq. (9.10), the folIowing expressions for A and Bare found:

A -a(l
(1 + v)
+ v)Tref + a (1 _ v)
(1 - 2v) rb
(b2 _ a2) la Tp dp;

B =
(1 + v) a2 b

a(1-v)(b2 -a2)la Tpdp


r (9.11)

Substituting these expressions into u(r), {rn {99 and then into Eqs. (9.4), the
following stress expressions are obtained:
240 9. Thermo-elasticity

aE 1
(1- /1) r 2
[(rlJ2 _ a2
2
a
-
2
) l b
a Tp dp -
l r
a Tp dp ;
1
= [(rb2 +a l +l 1
2 2 b r
aE 1
a Tp dp - Tr 2 ;
a(J(J )
(1- /1) r 2 _ a2 a Tp dp

a zz aE ( 2/1 b
(1 _ /1) lJ2 _ a2 la Tp dp - T
r
)
+ aETrel (9.12)

We have assumed plane strain so far. IT the end sections of the cylinder are
not constrained in their rudal displacements, then this new problem can be
solved by a procedure identical to that of Sect. 8.6 (generalized plane strain).
As far as the stresses are concerned, it is found that only the rudal stress
changes; its new expression is:

where a zz is given by (9.12c). It is easily found that:

R
27raE =- l b
a T p dp + TreI
(b2 -
2
a2 )

Therefore, the rudal stress in a cylinder with /ree ends is given by:

2
aE ( b2-a2
o-zz=l_/I a Tpdp-T ) l b
(9.13)

f:
It is seen that at the ends of the cylinder, 27r o-zzr dr = 0, but o-u f:. 0,
which means -using Saint Venant's principle- that the solution is valid far
from the ends. AIso, note that o-zz = a rr + a(J(J, while in plane strain a zz =
/I( a rr + a(J(J) - Ea(T - TreI). Finally, in plane strain, only a zz depends on the
reference temperature TreI, while in the free ends case, none of the stresses
depends on TreI.
The stress solutions (9.12) and (9.13) are valid for any given radial tem-
perature field T(r). Let us compute T(r) in the following case: the inner
surface (r = a) of the cylinder is maintained at a uniform (in space) and
constant (in time) temperature Ta, while the external surface is at tempera-
ture Tb (for instance, Ta corresponds to a fluid temperature and n to room
temperature). Assuming a steady state regime, the problem to be solved is:

LlT(r) = O; with T(a) = Ta and T(b) = Tb


Since rLlT(r) = djdr[r(dTjdr)], it is easily found that
b b r
ln(-)T(r) = Taln(-) + Tb ln(-) (9.14)
a r a
9.5 Solved problems 241

Simple integration then gives the expression of the following integral:

ln(-)
b
a
l
a
r
Tpdp =

Substitution into Eqs. (9.12a,b) gives the radial and hoop stresses as:

c
U(J(J
(9.15)
C
The axial stresses are given by Eqs. (9.13) and (9.12c) as:

b 2a 2 b
1- 21n- - b2 21n- (free ends);
c r -a a
v [1 - 2ln ~ _ 2a 2 In ~]
C r b2 - a2 a

_ Ea [Ta ln(b/r) + Tb ln(r/a) _ T ] (plane strain) (9.16)


C ln(b/a) ref

In aH cases, C is a constant given by

C = Ea(Ta -n)
-
2(1 - v) In a
b

Stresses in the case of free ends are plotted in Fig. 9.4 for b = 2a. The stresses
at the inner and outer surfaces are given by:

If Ta > Tb (Le., the interior is warmer than the exterior) then the inner surface
is prevented from expanding as much as it would "like" to, therefore we expect
the stresses at the inner surface to be compressive. This is confirmed by Fig.
9.4 which shows that u(J(J(a) < O.
For a brittle material, Fig. 9.4 shows that cracks may appear at the outer
surface (r = b) because it is subjected to tensile stresses. For a ductile ma-
terial, we may use Tresca's criterion (Sec. 1.10), and Fig. 9.5 shows that it
gives in this case (uy being the initial yield stress):

-u(J(J(a) < uy
242 9. Thermo-elasticity

Recalling the expressions of uee(a) and e, this is equivalent to the following


condition (assuming that Ta > T b ):

T. 2(1 - v)uy ln(b/a)


< --'--='"--;:---""""'"--''--'----:;-
'T'
- .L b (9.17)
a aE [-1 + i~:2In(b/a)]
For a ductile material, Tresca's criterion shows that the cylinder remains
elastic as long as this inequality is satisfied. Otherwise, plasticity will develop
at the inner surface (r = a).
For a very thin tube (b = a + h, h/a« 1), the following expressions are
found for the stresses:
aE(Ta - Tb ) • (b) _ - (b) '" aE(Ta - T b )
2(1 - v) ,U(J(J - U zz '" 2(1- v)

The yield condition (9.17) takes a simple form:

T. 'T' 2(1 - v)uy


a -.Lb < aE

0.8
0.6
0.4
b 0.2
'"O
CI.)
O
N
-0.2 ~/ .....
~ ,' ...... rr--
-0.4
...Ei -0.6 , .."
,,"..... (}() -----
",..' .'
O
Z -0.8
zz .......... .
1/
-1 ~."
(o"
-1.2
-1.4 L..._....&..._-'-_--I......_ " ' - - _.....
0.5 0.6 0.7 0.8 0.9 1
r/b
Fig. 9.4. Tube with free ends: stresses, divided by aE(Ta - n)/[2(1 - v)], vs.
normalized radial distance r /b in the case b = 2a

9.5.2 Thermal stresses in a composite cylinder

A composite solid is made up of two concentric circular cylinders of length l: a


fiber of radius RI and properties E" VI and ali and a matrix of radii RI and
Rm and properties Em, Vm and am (Fig. 9.5). The composite is subjected to
9.5 Solved problems 243

a uniform change in temperature, oT. Lateral surface (r = Rm) is stress-free


and perfect adherence at the interface (r = R,) is assumed. A generalized
plane strain assumption is made: uniform axial strain in the entire composite.
It is useful to notice that an isothermal version of the problem (where axial
and uniform displacements ±(U/2)e z are applied to faces z = ±1/2) was
solved in Sec. 1.16. We shall make repeated reference to that solution. As in
Sec. 1.16, we introduce the following notation for convenience:

C == (::m)2, Ezz = constant == e, p == -C1rr (R,), (9.18)

where C represents the fiber volume fraction, e the axial (uniform) strain and
p the (continuous) interface pressure. One can prove that the stress solution
is form-identical to that of Sec. 1.16,
u(r) = u(r)(iSO), (9.19)
where "iso" refers to the isothermal expressions (1.95). Equations (9.3) give
the strain field as follows ("iso" designating the isothermal Eqs. (1.97)):

• Fiber: Err,(J(J = Err,(J(J


(iso)
+ ( 1 + v, ) alu~T
• Matrix: Err,(J(J = Err,(J(J
(iso)
+ ( 1 + Vm ) amu~T (9.20)

the other strains being uniform:

Ezz = e, Er(J = E(Jz = Erz = O (9.21)

Condition Ezz = e allows to compute the axial stresses C1 z' and C1 zm :

(9.22)
where "iso" refers to the isothermal Eqs. (1.99). Strain/displacement relations
give the radial displacements as:

u, = ujiBO) + (1 + v,)a,(oT)r, Um = U~BO) + (1 + vm)am(oT)r (9.23)

Imposing displacement continuity at r = R, allows to compute the fiber/ma-


trix interface pressure:
(vm-v,)e [(1 + vm)a m - (1 + v, )a,l oT
p = V V
~
p(i.o)

V ==
(1 + v,)(l - 2vâ
E, +
(1 + v m)
Em
(i.C + 1 _ 2 m) ~C
V 1-
(9.24)

Note that the expression of Vis identical to that of the isothermal case, Eq.
(1.100b). Also, if we set oT = O, we obviously retrieve the isothermal pressure
p(iso). As in Sec. 1.16, the axial stress average is defined as follows:
244 9. Thermo-elasticity

< Uzz >= R1 2 [7fRJUzl + 7f(R~ - RJ)uzm ] = CUzl + (1- C)uzm


7f m

Setting < U zz >= O allows to compute the axial strain. Using Eqs. (9.22, 24)
the final expression is found from the following equality:

[CEI + (1- C)Em + (2CjV)(vm - VI)2] e = {CElal + (1- C)Emam


+(2C jV)(vm - vI )[(1 + vm)a m - (1 + vI )a/]) 8T (9.25)

Therefore, an the stress, strain and displacement fields are now completely
defined.
Under certain conditions (periodicity, axisymmetry, no fiber interaction)
the two-material cylinder studied so far can be considered as a unit cell of
a composite material containing many fibers (of volume fraction C) in a
matrix. As an application, consider ceramic fibers and a metallic matrix with
the following properties:

Fiber: SiC (SCS6) : EI= 360 GPa, vI= 0.17, = 4.9 x 1O- 6 rC,
al
Em = 75.2 GPa, Vm = 0.25, am = 11.7 x 10- 6 rC

With RI = 70/Lm and Rm = 110/Lm,


we obtain: C = 0.405. For 8T =
-800°C (cooling down from processing to room temperatures), the following
stresses are found (in MPa):

Fiber: U rr = -187.3, U()() = -187.3, U zz = -546.5,


Matrix: urr(RI) = -187.3, u()()(RI) = 422.2, U zz = 372.0

Since al < am, the matrix would like to contract more than the fibers would
allow it to, therefore the matrix will subject the fibers to compressive stresses.
It is seen that tensile hoop stresses appear in the matrix, which may lead to
radial cracking in the matrix. A solution is to put a coating layer between
each fiber and the matrix, in order to reduce tensile stresses in the latter. The
problem is to find the right properties for this additional layer. For details,
see (Doghri et al., 1994) and (Doghri and Leckie, 1994).

fiber (ceramic)
EOt, Va, al

Fig. 9.5. Composite material: ceramic fiber in a metallic matrix


9.5 Solved problems 245

9.5.3 Transient thermal stresses in a thin plate

In this section, we consider a thin plate of uniform thickness h (middle plane:


z = O) which is not subjected to imposed forces or displacements (Fig. 9.6).
We assume that there is a temperature field in the plate of the form T =
T(z, t), Le. T varies with transverse position and time.
We shall first compute the thermal stresses due to T(z, t). We look for
a stress solution under the form aij = aij (z, t). The three scalar in-bulk
equilibrium equations give in this case:

8axz = 8a yz = 8azz = O
8z 8z 8z
Stress-free B.Cs. on z = ±h/2 give the conditions:
h h h
a xz (±'2,t) = a yz (±"2,t) = a zz (±'2,t) = O
Equilibrium and B.C. equations imply that alI out-of-plane stresses vanish:

Thus we have a membrane problem, Le. we only have in-plane stresses. More-
over, since T = T(z, t), there is no privileged direction in the (x,y) plane.
Therefore, we shalIlook for a stress field under the following form:

a xy = O; a xx = a yy = a(z, t) (9.26)

Constitutive equations (9.3) give the strains as:

(1 - v)
E a(z, t) + a(T - TreI) = f.yyj

2v
-Ea(z, t) + a(T - Tre/)j f. xy = f. yz = f. zx = O
Compatibility -Sec. 1.9- gives two non-trivial equations:

8 2 f.xx = 8 2 f.yy = O
8z 2 8z 2
They imply the folIowing form for the strains:
(1 - v)
f. xx = f.yy = [zM(t) + N(t)]-E-
where M(t) are N(t) are functions which need to be computed. Equating the
strain expressions with those from the constitutive model, it is found that:
E
a(z, t) = zM(t) + N(t) - 1 _ va(T - TreI) (9.27)
246 9. Thermo-elasticity

The border of the plate (i.e, with outward normal in the (x, y) plane) is stress
free. Imposing this condition point-wise implies a(z, t) = O; this is impossible
to realize if the temperature distribution T(z, t) is nonlinear in z. Therefore,
we shall use Saint-Venant 's principle: we only ask for the resultants of the
stress vector and its moment to vanish, and the solution will be valid inside
the plate, far enough from the border. The two conditions are translated as
follows:
h/2 jh/2
j a(z, t) dz = O; a(z, t)z dz = O
-h/2 -h/2
Since f~'~2 z dz = O and f~'~2 z2 dz = h 3 /12, it is found that:

nE jh/2
N(t) = (1 - v )h -h/2 [T(z,t)-Tre/]dz;

12nE jh/2
M(t) = ( )h3 [T(z,t)-Tre/]zdz; (9.28)
1- v -h/2
The mechanical problem is thus fully solved. We now determine the tem-
perature field T(z, t) in the following case: initially, the plate is at uniform
temperature To, then surfaces z = ±h/2 are cooled down at a constant rate,
h
T(z, O) = To; T(±2"' t) = To - (t (9.29)

where the positive constant ( ([K/s] or [OC/s]) is the cooling rate. The heat
equation (9.8) takes the simple form:

â2 T âT
k--
âz 2
= pc-
ât

----
An approximate solution is given by:

(pc h 2
T(z, t) = T o - (t+ 2/;("4 - Z2) (9.30)
'"--"'
It is seen that T(z, t) is made up of two terms: a time-dependent linear term
and a time-independent parabolic term in z.
Since T(z, t) is an even function of z, we have M(t) = O, and N(t) is
found as (assuming that T o = TreI for simplicity):

N(t) = Ea( (pch 2 _ t)


1- v 12k
The stresses are then given by Eqs. (9.26) and (9.27) as:

En(pc 2 h2
axx=ayy=a(z,t)= 2(1-v)k(z -12) (9.31)
9.5 Solved problems 247

Note that the stresses are independent of time and have a parabolic profile
through the thickness. The stresses at (z = O) -middle plane- and (z = ±h/2)
are given by:

Ea(pch 2 h Ea(pch2
u(O, t) = - 24(1 _ v)k < O; u(±'2' t) = 12(1 _ v)k > O

It is seen that the stresses are compressive at (z = O) and tensile at (z =


±h/2). These results make sense: since the interior (z = O) is warmer that
the outer surfaces (z = ±h/2), it is prevented from expanding freely, and
thus it is compressed.
As a numerical application, we consider two brittle materials: ordinary
(sodium) glass and Pyrex (borosilicate glass). The following data are taken
from (Ashby and Jones, 1980):
-Ordinary glass: E = 74 X 109 N m- 2 , v = 0.2, a = 8.5 x 10- 6 K-t, p =
2480 kgm- 3 , c = 990 Jkg- 1 K- 1 , k = 1 Wm- 1 K-l
-Pyrex: E = 65 X 109 Nm- 2 , v = 0.2, a = 4 x 10-6 K-t, p =
2230 kgm- 3 , c = 800 J kg- 1 K-l, k = 1 W m- 1 K- 1
The stresses -divided by (h 2 10 11 _ are plotted in Fig. 9.7. It is seen that for the
same cooling rate ((), ordinary glass is subjected to higher thermal stresses
then Pyrex and is therefore more likely to break (the principal reason is that
the coefficient of thermal expansion a of Pyrex is much smaller than that of
ordinary glass, and the thermal stresses are proportional to a).
In this chapter, we studied only a few examples ofthermoelastic problems.
The reader can find many more problems in other textbooks, e.g. (Boley and
Weiner, 1960).

Fig. 9.6. Plate of thickness h whose external surfaces z = ±h/2 are cooled at
constant rate.
248 9. Thermo-elasticity

1.5 Glass - -
Pyrex -----
1

0.5

o
-0.5

-1~~~~--~--~--~--~
-0.6 -0.4 -0.2 O 0.2 0.4 0.6
z/h
Fig. 9.7. Thermal stresses in a thin plate. Normalized stresses vs. normalized
transverse position for two kinds of glass: (a) ordinary, (b) special (Pyrex).
10. Elastic stability

In alI previous chapters, we have implicitly considered structures which are


in stable equilibrium. There are cases where an equilibrium configurat ion
is (or becomes) unstable. In this chapter, we give some notions about this
important topic by solving some simple problems (although many of them
have practical interest).

10.1 Introduction
Instability phenomena appear mainly in long beams or thin plates or shells;
such structures may become unstable in the direction where their rigidity is
minimum. Let us give a few examples.
A long beam (column) may buckle under a compressive axialload (Fig.
10.1).

Fig. 10.1. Buckling of a column under a compressive axialload

A thin plate may buckle under a compressive load act ing in its middle
surface (Fig. lO.2).
250 10. Elastic stability

Fig. 10.2. Buckling of a thin plate under a compressive load acting in its middle
surface

A thin circular cylindrical tube subjected to a uniform pressure on its


outer surface may become unstable: it flattens and its cross section becomes
elliptic (Fig. 10.3)

Fig. 10.3. A thin circular cylindrical tube subjected to a uniform pressure on its
outer surface may become unstable

A thin circular cylindrical tube subjected to axial compression may


buckle: materiallines initially parallel to the axis are no longer straight (Fig.
IOA)

10.2 Direct and energy methods


Consider a deformed structure which occupies an equilibrium configuration
(S) in a fixed reference frame. We shall use two methods for studying the
stability of (S): direct and energy methods. Let us illustrate these methods
by an elementary example (Fig. 10.5).
In all cases (a), (b) and (c) of Fig. 10.5, (1) is an equilibrium position. In
order to study its stability, we perturb it slightly to a new position (2). In
Fig. 1O.5a, it is obvious that the balI will go back from (2) to (1), thus (1)
is a stable equilibrium position. In Fig. 1O.5b, it is also obvious that the ball
will not go back from (2) to (1), thus (1) is unstable. In Fig. 1O.5c, positions
(1) and (2) are equivalent, but the ball will not go back form (2) to (1), thus
(1) is unstable. This example illustrates the direct method: in order to study
the stability of an equilibrium position (S), we perturb the structure slightly
from that positionj if it "would like" to go back to (S), in other words if
lO.2 Direct and energy methods 251

Fig. 10.4. Buckling of a thin circular cylindrical tube subjected to a compressive


axialload

2
"
\ .... ;: . ,

a
Fig. 10.5. Illustration of stability definitions with an elementary example. The
equilibrium position (1) is: (a) stable, (b) unstable and (c) unstable.
252 10. Elastic stability

the recall forces overcome the applied forces, then (8) is a stable equilibrium
configuration.
We now go back to Fig. 10.5 in order to illustrate another definit ion for
stability: the energy method. In the transition (1) ~ (2), the potential energy
increases in Fig. 10.5a, decreases in Fig. 1O.5b and remains unchanged in
Fig. 10.5c. Recalling that only Fig. 1O.5a corresponds to a stable equilibrium
for (1), we conclude that (1) is a stable equilibrium position if the potential
energy increases in any perturbation from (1) to (2). More generally, in order
to study the stability of an equilibrium position (8), we apply a kinematically
admissible (K.A.) virtual displacement (Le. "smooth enough" and satisfies the
displacement B.Cs., see Chap. 2). Let 6W be the variation ofthe deformat ion
energy in this virtual displacement, and 67 the variat ion of the work of
externalloads in the same displacement. The variation of the potential energy
is defined as (see Chap. 2): 6& = 6W - 67. Equilibrium position (8) is stable
if 6& > O far any K.A. virtual displacement. This is known as the Lejeune-
Dirichlet theorem.
In general, the direct method leads to the resolution of differential equa-
tions whiIe the energy method leads to the minimizat ion of a functional.
There are cases where both methods may faiI (see Sec. 10.5).

10.3 Euler's method for axially compressed columns


10.3.1 Critical buckling load

Consider the column of Fig. 1O.6a with the displacement B.Cs. u(O) = v(O) =
O and v(l) = O (various other B.Cs. will be considered in Sec. 10.6). The
structure is perturbed slightly from its vertical equilibrium positionj the new
position is depicted in Fig. 1O.6b.
In this deformed configuration, equilibrium of forces is identically satisfied,
and equilibrium of moments requires that M(x) = Qv(x), where M(x) is the
internal bending moment in the cross section (x), see Chap. 3. On the other
hand, Hooke's law for beams gives (Chap. 3) M(x) = -Elv"(x), where
E is Young's modulus, 1 the moment of inertia of the cross section and
v" (x) == rPv I dx 2 • Therefore, writing the equilibrium of moments together
with the displacement B.Cs., the following equations are obtained:

-Elv"(x) = Qv(x)j v(O) = Oj v(l) =O


We introduce a constant k [Iim] defined by:

k 2 == .!I.
EI
(10.1)

With this notation, the equilibrium problem of Fig. 10.6b becomes: find v(x)
which satisfies the following equations:
10.3 Euler's method for axially compressed columns 253

EI

Fig. 10.6. Column under a compressive axial load. (a) Vertical equilibrium posi-
tion; (b) assumed buckling mode.

v" (x) + k 2 v(x) = O; v(O) = O; v(l) =O


The general solution is given by: v(x) = Acos(kx) + B sin(kx). The B.C.
v(O) = O implies that A = O, while v(l) = O implies that B sin(kl) = o.
There are two cases. If kl =1 nn (where n is an integer) then B = O
and v(x) = O, Vx E [O, ll. This is the trivial solution which corresponds to
a vertical position of the bar. We are interested in cases where equilibrium
positions other than the vertical one -i.e. such as in Fig. 10.6b- are possible.
The second case for which B sin(kl) = O is kl = nn (and B is arbitrary,
in particular it can take any non-zero value). In this case we have

(10.2)

Different buckling modes corresponding to different values of n in Eq. (10.2a)


are plotted in Fig. 10.7. The smallest value in Eq. (10.2b) corresponds to
n = 1 and is called Euler's critical buckling load, its value Qe is given by:

(10.3)

80 far, we have proved the following results:


-If Q < Qe, then the only possible solution is v(x) = O everywhere,
corresponding to the vertical equilibrium position, which is therefore stable.
-If Q ~ Qel then two solutions are possible: v(x) = O everywhere (verti-
cal position) and another solution: v(x) = Bsin(nxjl), where B =1 O is an
arbitrary constant.
254 10. Elastic stability

n=3
n=2
n=1

Fig. 10.1. Column under a compressive axial load. First three buckling modes:
n = 1,2,3 in Eq. (10.2a).

We shall admit the following results which are experimentally confirmed:


-If Q ~ Qc, then the vertical equilibrium position is unstable and buckling
will occur (the first buckling mode corresponds to v(x) = B sin(7rx/l) , B -:j:. O).
Engineers are usually not interested in studying the stability of buckling
modes, they are much more interested in preventing them.
Now, a word about the moment of inertia 1. Buckling occurs in the plane
where the inertia is the smallest (an elementary experiment consists in com-
pressing a ruler which has a thin rectangular cross section). Thus, designating
the centroid of the cross section by (G) and the principal axes of inertia by
(Gy) and (Gz) such that Iz ~ Iy, we have: 1 = Iz. Consider for instance a
rectangular cross section of dimensions hand b > h. The (Gz) axis is parallel
to the long side and 1 = Iz = bh3 /12. For a circular cross section of diameter
d, the moment of inertia 1 is identical for aU axes passing through the center
(G); its value is 7rd4 /64.

10.3.2 Critica} buckling stress

The radius of gyration l' [m] and the slenderness ratio A are defined with the
following formulae:

(IOA)

The ,critical buckling stress is: ac == Qc/A, where A is the cross section
area. Using Eqs. (10.3-4), ac can be expressed as:

(10.5)
10.3 Euler's method for axially compressed columns 255

In order to increase ac, and thus the resistance to buckling, Eq. (10.5)
shows that we can increase E (the material stiffness) or decrease A (the
slenderness ratio). Equation (1OAb) shows that we can decrease A in two
ways: by decreasing 1 (the length of the beam) or increasing r (the radius of
gyration). Note that both r and A are purely geometric quantities. Let us take
an example. For a circular cross section of diameter d, we have 1 = 7rdf /64
and A = 7rd2 /4, thus r = d/4. For a rectangular cross section of dimensions
hand b > h, we have 1 = bh3 /12 and A = bh, thus r = h/(2v'3). Assume
that the circular and rectangular cross sections have the same area, which
one resists better to buckling (for the same values of E and l)? Well, we can
rewrite the two values of r as follows:

.
r(clrcle) = v'A
c; r(rectangle) = v'A
j3f
2y 7r 2 31
h

It appears then that if b/h > 7r/3, a circular section resists better than a
rectangular one of the same area.
The curve ac(A) ofEq. (10.5) is plotted in Fig. 10.8 (in that figure ay des-
ignates the initial yield stress in compression, a material parameter). RecaU
that in the derivations which led to the expression of ac(A), linear elasticity
was assumed aU along. Therefore, Eq. (10.5) is only valid for a ::; ay, thus
in Fig. 10.8 only the (Be) part of the curve is valid. In other words, Euler's
formula (10.3) is only valid if A ~ Amin, where Amin is obtained by setting
ac = ay in Eq. (10.5),

Amin = 7r
v-;;;
[E (10.6)

Equation (10.6) is interpreted as follows. If A (a geometric quantity) is larger


than the value of Amin (a material parameter) then buckling occurs if the
compressive stress a reaches or exceeds the value of ac given by Eq. (10.5),
or equivalently if the applied load Q reaches or exceeds the value of Qc given
by Eq. (10.3). If A < Amin, buckling may stiH occur, but the formulae of Qc
and ac are no longer given by Eqs. (10.3) and (10.5); we have so-called plastic
buckling, see (Hutchinson, 1974) and references therein.
As a numerical application, consider a steel beam with E = 210GPa,
ay = 230 M Pa and a cross rectangular section 4 cm x 5 cm. What is the
minimum length lmin of the beam for which Euler's formula is valid? Firstly,
Eq. (10.6) gives Amin ~ 95. Secondly, the cross section properties are A =
20cm2 , I = (80/3)cm 4 and r = (2/V3)cm. Therefore, lmin = rAmin ~
1.10m.
We now compute Qc for 1 = 1.75 m. Since 1 > lmin, it is legitimate to use
Eq. (10.3) which then gives Qc ~ 180047 kN; the corresponding compressive
stress is ac = Qc/A ~ 90.2 MPa, and it is smaller than ay.
256 10. Elastic stability

Amin A
Fig. 10.8. Critical buckling stress Uc vs. the slenderness ratio A. Euler's formula
is valid for A ;::: Amin.

10.3.3 Remarks

In this chapter, as in the previous chapters, we make use ofthe small pertur-
bation hypothesis (SPH). However, and unlike previous chapters, we write
equilibrium equations on the deformed configurat ion and not on the initial,
undeformed one. An important consequence is that the problems thus for-
mulated are nonlinear. Indeed, the reader can check that for ali problems
solved in this chapter, the relation between input (applied load) and out put
(deflection) is not linear. Therefore, the superposition "principle" cannot be
applied when there is instability.
Another thing to keep in mind is that the critical buckling load Q c is very
sensitive to the displacement B.Cs. Indeed, the results of Secs. 10.3 and 10.6
show that Qc can be written under the form:
2 EI
Qc = 7r (al)2' where:

- (a) a = 1 in Eq. (10.3), Fig. 10.6, B.Cs. v(O) = O, v(l) = O.


- (b) a = 2 in Eq. (10.10), Fig. 10.12, B.Cs. v(O) = vl(O) = O.
- (c) a ~ 0.7 in Eq. (10.11), Fig. 10.13, B.Cs. v(O) = vl(O) = O, v(l) = O.
- (d) a = 1/2 in Eq. (10.12), Fig. 10.14, B.Cs. v(O) = vl(O) = O, v(l)
vl(l) = O.
As expected, if the degrees of freedom are reduced, then Qc increases (a
decreases) j compare for instances cases (b), (c) and (d).

10.4 Energy-based approximate method


In Sec. 10.3, we applied the direct method to find the exact solution. In this
section, we present a procedure based on the energy approach which allows
to obtain approximate values of the buckling load when an exact solution
cannot be found.
10.4 Energy-based approximate method 257

Consider a beam of length (1) which is axially compressed, Fig. 10.9. The
B.Cs. at the ends (x = O) and (x = 1) are arbitrary. The expression of the
deformation energy 8W is given in Chap. 3. If we only keep the contribution
due to bending, we obtain the following expression:

8W ~ -
111
2 o
M2(X)
- - dx = -
EI
111
2 o
EI[v"(x)]2 dx

where we used Hooke's law for beams (M(x) = -Elv"(x)).


The work of externalloads is: 8T = Q(l - AB), where

AB = l XB
XA
dx = lsB (cos a) ds,
o
where s is the curvilinear coordinate along the middle fiber of the deformed
beam (origin: SA = O). Since tana = v'(x), we have:

where we used the assumption Iv'(x)1 « 1. The assembly of the different


results gives:
8T ~ Q (l- SB + ~ Io BB [V'(X)]2 dS)
We now make two approximations which are justified by the SPH:

- (i) J;B [v' (x)]2 ds ~ J~[v'(xW dx (a similar assumption was also implicitly
made in the derivation of 8W);
- (ii) SB ~l.
As a consequence of (i) and (ii), the following expression of 8T is obtained:

8T~Q~ (/[v'(X)]2dx
2 10
The energy method states that the equilibrium position v(x) = O is stable if
(8W - 8T) > O for any K.A. displacement field v(x). In our case, this reads:

~ t
2 10
EI[v"(x)]2 dx - Q~
2 10
t [v'(x)f dx > O, V v K.A.

Thus, the following value of the buckling load Qc is obtained:

(10.7)

For instance, K.A. displacement fields v(x) for the problem of Sect. 10.3 must
satisfy v(O) = v(l) = o.
258 10. Elastic stability

Now the approximate method consists in computing Qc with just one


K.A. displacement field v(x) which is chosen to be close to the presumed
(but unknown) buckling mode, thus:

Qc ~
- = "'-'!..-:/-.::...----'....:....:.--
Qc
I~ EI[ii" (x)]2 dx
(10.8)
10 [v'(x)J2 dx

A x B Bo

y
Fig. 10.9. Computation of the buckling load by the energy method: notations.

10.5 Non-conservative loads

There are cases where both the direct and energy methods lail in computing
the critical buckling load. An example is given in Fig. 10.10 for which the
compressive load Q remains always directed along the middle fiber of the
column. It can be shown that the work of Q is non-conservative, Le. it depends
on the deformation path. In other examples considered in this chapter, the
applied load keeps a fixed direction and its work is path-independent.
When we study the problem with the direct method, we find that there is
no value for Q for which there exist deformed configurations of the column,
and this conclusion is obviously not physically acceptable. It can be shown
that the problem can be solved by using a so-called dynamic method which
can be described as follows.
It is assumed that the column is subjected to an initial disturbance which
produces small vibrations. If their amplitude decreases with time, and the
vibrations eventually die out, then the vertical position is stable. However,
if the applied loads are such that the amplitude of the vibrations begins to
grow without limit, then the vertical position is unstable. For details, see
(Timoshenko and Gere, 1961).
10.6 Solved problems 259

Fig. 10.10. A case where both direct and energy methods fail: a load Q remains
in the direction of the middle fiber of the column

10.6 Solved problems


10.6.1 Two rigid bars connected with a spring
Two rigid bars ofthe same length (l/2) are connected with a rotational spring
of stiffness K [Nm/rad]. A compressive axial force Q is applied at (x = l), the
x and y displacements are prevented at (x = O), while only x-displacement
is allowed at (x = l); see Fig. 1O.11a.
We study the stability of the vertical equilibrium position of Fig. 1O.11a
with the two methods of Sect. 10.2. First we start with the direct method. We
perturb the structure slightly from its initial equilibrium position, we obtain
a configurat ion depicted in Fig. 10.l1b. In order for the structure to retum
to the vertical position, the moment in the spring due to the relative rotation
(} of its ends must overcome the moment due to the extemalload Q, Le.

(10.9)

where we have used the small angle approximation: sin(() /2) ~ () /2. Equation
(10.9) is interpreted as follows: the vertical equilibrium position of Fig. 1O.11a
is stable if the extemalload Q is smaller than 4K/l, otherwise it is not.
We now solve the same problem with the energy method. To the vertical
equilibrium position of Fig. 10.11a, we apply a K.A. virtual displacement, we
obtain the configuration of Fig. 1O.11b. Since the bars are rigid, the variat ion
of the deformat ion energy is solely due to that of the spring whose value is:

8W = .!.K(}2
2
The variat ion of the work of the extemalload Q is given by:
260 10. Elastic stability

where we used the small angle approximation: cos(fJ /2) ~ 1 - (fJ2/8). It


appears that the variation of the potential energy 6& = 6W - 67 is given by:
fJ2 1
6&~-(K-Q-)
2 4
According to the theorem of Sect. 10.2, the vertical equilibrium position
of Fig. 1O.11a is stable if 6& > O for any K.A. virtual rotation fJ, thus if:
Q < Kl/4, which is the same condition we found with the direct method.

l/2
EI=oo

l/2

Fig. 10.11. Two rigid bars connected with a rotational spring and subjected to
a compressive axial load. (a) Vertical equilibrium positionj (b) the structure is
perturbed slightly around the vertical position.

10.6.2 Column clamped at one end

A column of length 1 is clamped at (x = O) and axially compressed with a


force Q at (x = l), Fig. 1O.12a. It can be seen that the buckling mode in Fig.
1O.12b is equivalent to that of a bar of length (2l) in Fig. 10.6b. Thus Qc is
obtained by replacing 1 with (2l) in Eq. 10.3:

2 EI
Qc = 7[ (2l)2 (10.10)

As an exercise, we shall recompute Qc using Euler's direct method. Equilib-


rium of moments gives M(x) = Q[v(x) - v(l)]. Hooke's law for beams gives
10.6 Solved problems 261

M(x) = -Elv"(x). Combining the equations and writing the displacement


B.Cs., we obtain:

-Elv" (x) = Q[v(x) - v(I)]; v(O) = O; v'(O) =O


with v'(x) = dv/dx and v"(x) = rPv/dx 2 . A constant k is defined by k 2 ==
Q / EI. With this notation, the equilibrium problem of Fig. 10.12b becomes:
find v(x) which satisfies the following equations:

v" (x) + k 2 v(x) = k 2 v(l); v(O) = O; v' (O) =O


The general solution is given by

v(x) = A cos(kx) + B sin(kx) + v(l)


The B.Cs. give the two equations:

A + v(l) = O; Bk = O

Thus B = O and: v(l) = -A = Acos(kl) + v(I), i.e Acos(kl) = O. It is seen


that if cos(kl) :f:. O, then A = O and v(x) = O everywhere: this is the (trivial)
vertical equilibrium position. However, if cos(kl) = O, i.e. ki = (2n + 1)7r/2,
then it is possible to have A :f:. O and thus a non trivial equilibrium solution.
The lowest buckling load correspond to n = O, i.e k = 7r / (21) and is given by:
Qc = EI(7r/21)2, which is exactly Eq. (10.10).

10.6.3 Column clamped at one end and simply supported at the


other

A column of length I is clamped at (x = O). At the other end (x = l) there is an


axial compressive load Q and the lateral displacement is prevented (v(l) = O),
see Fig. 10.13. Equilibrium of moments gives: M(x) = Qv(x) - H(I- x) and
Hooke's law gives M(x) = -Elv"(x). Combining the equations and writing
the displacement B.Cs., we obtain:

v" (x) + k 2 v(x) = .! (I - x); v(O) = O; v' (O) = O; v(l) =O


where again a constant k is defined by k 2 == Q/EI. The general solution is
given by
. H
v(x) = Acos(kx) + Bsm(kx) + Q (I- x)
The B.Cs. give the three equations:

H H
A + QI = O; Bk - Q = O; Acos(kl) + Bsin(kl) = O
262 10. Elastic stability

..
y y

Mo = Qv(l)
Q

\
\
\
\

Fig. 10.12. Buckling of a column of length (l) cIamped at (x = O) and axially


compressed at (x = l).
10.6 Solved problems 263

Since k i- O, substitution of the expressions of A and B into the third equation


gives:
H.
kQ [sm(kl) - kl cos(kl)] =O
If cos(kl) = O, then: B = O, H = O and A = O, i.e. we have the (trivial) equi-
librium solution v(x) = O everywhere. A non-trivial solution can be obtained
if tan(kl) = kl. The smallest buckling load is obtained for kl ~ 4.493rad
which gives:

EI
Qc ~ 20.187"[2 (10.11)

Fig. 10.13. Buckling of a column of length (1) clamped at (x = O) and simply


supported at (x = 1), v(l) = O.

10.6.4 Colmnn clamped at both ends

A column of length l is clamped at both ends (x = O) and (x = l), see Fig.


10.14. The solution for this problem can be found very simply by an analogy
argument (Timoshenko and Gere, 1961), but as an exercise, we compute the
buckling load using the direct method. Equilibrium of moments and Hooke's
law give: M(x) = Qv(x) - Hx + M o = -Elv"(x). Combining the equations
and writing the displacement B.Cs., we obtain:

2 H Mo
v" (x) + k v(x) = - x - - ' v(O) = v(l) = Oi d(O) = v'(l) = O
EI EI'
264 10. Elastic stability

where again k is such that k 2 == Q/EI. The general solution is given by:

v(x) = Acos(kx) + Bsin(kx) + ~x _ ~o


The B.Cs. give the four equations:

A- Mo
Q
= Oj A cos(kl) + B sin(kl) + ~ 1 - ~o = O
H
Bk+- =Oj
Q
-Aksin(kl) + Bkcos(kl) + ~ =O

Since k i= O, substitution of the expressions of A and B from the first and


third equations into the second and fourth gives the following linear system
of two equations and two unknowns Mo and H:

- ~o k[l - cos(kl)] + ~ [kl - sin(kl)] = Oj

- ~o k sin(kl) + ~ [1 - cos(kl)] =O
The determinant of this system can be written as follows:

det = -2k[1- cos(kl)] + k 2lsin(kl) = -4ksin(~l)[sin(~l) _ ~l cos(~l)]


Non-trivial solutions are found when det = O. A first solution is given by
sin(kl/2) = O, thus kl/2 = mfj the smallest buekling load is Qc1 = 41[2 EI/l2.
A second non-trivial solution is obtained when sin(kl/2) i= Oand sin(kl/2) =
(kl/2) cos(kl/2). We found in the previous section that buckling is possible
for kl/2 = 4.493 rad, thus: Qc2 = (8.986)2EI/l2. Since Qc2 > Qc1, buckling
first oecurs when Q = Qc!' Le.

(10.12)

10.6.5 Column elastically built-in at one end and simply


supported at the other

In this section, we compute the buckling load Qc of a column oflength 1which


is elastically built-in at one end (angular spring of stiffness K) and simply
supported at the otherj see Fig. 10.15. Static equilibrium on a deformed
configuration gives the reaction at (x = l) as (Ve y) and those at (x = O)
as (Pe x ), (-Ve y) and (-Vle z )' The direct method leads to the following
differential equation for the deflection (with k 2 == P/(EI)):
V
v" (x) + k2v(x) = EI (l - x),
10.6 Solved problems 265

y
..

Fig. 10.14. Buckling of a column of length (l) clamped at both ends (x = O) and
(x = l).

whose general solution is:

v(x) = Acos(kx) + Bsin(kx) + ~(l- x) (10.13)

The constants are determined from the B.Cs.


dv Vl
v(O) = O, v(l) = O, dx (O) = - K '

which translate onto the following three equations:

A +
vpl = O, Acos(kl)
.
+ Bsm(kl)
V Vl
= O, kB - P = - K

A non-trivial solution (corresponding to V "1- O and k "I-O) may exist if the


following condition is satisfied:

(K -l) sin(kl) = K cos(kl) (10.14)


P kl P
The reader can check the following two limiting cases:
- (i) K = O. The expression of the critical buckling load Pc is identical to Eq.
(10.3). lndeed, in this case the column can freely rotate at x = O, which is
equivalent to having a hinge there.
- (ii) K -t 00. The expression of Pc is identical to Eq. (10.11). lndeed, in this
case, no rotation is allowed at x = O, meaning that the column is clamped
there.
266 10. Elastic stability

__E_l______~~p_
~_K____~LV
~ JIh
x=O x=l
Fig. 10.15. Column elastically built-in at one end and simply supported at the
other.

10.6.6 Column with non-uniform properties

A column of Iength (l = l2 + it) is made up of two parls with properties


(~, 12 ) for x E [0,l2] and (El, Il) for x E]12,l]; see Fig. 10.16. The column
is built-in at one end (x = O) and subjected to a compressive force P at the
other (x = l). We use the notation: k~ == P/(Ellt} and ki == P/(E2h).
Static equilibrium gives the reactions at (x = O) as (Pe x ) and (-PVI (l)e z ).
Using the direct method, we find two differential equations whose general so-
Iutions are:

x E [O, l2] : V2(X) = A 2cos(k2x) + B 2 sin(~x) + VI (l);


x E [12,1]: VI(X) = Al cos(klx) + BI sin(klx) + VI(l) (10.15)

The constants are determined from the B.Cs. at (x = O) and continuity


conditions at (x = l2),

These translate onto the following system:

A 2 +VI(l) O; B 2 k 2 = O;
-VI (l) cos( k 2 l 2 ) = Al cos(kll 2 ) + BI sin(kl l2 )j
vI(l)k2sin(k2l 2) = -klAI sin(kl l2 ) + Blk l cos(kll 2 )

The Iast two equations are solved for Al and B I , then the values found are
substituted into the following equation

Al cos(kll) + BI sin(kll) = O,
which is simply Eq. (1O.15b) applied to x = l. After some algebra, the equa-
tion can be written as follows:

(10.16)

From now on, we consider it = l2 = l/2 for simplicity, and study three cases:
- (i) E 212 = Ellt. The reader can check that the expression of the critical
Ioad is Pc = 7r 2 E l lt/(2l)2, Le. Eq. (10.10). Indeed this case corresponds to
a column with uniform properlies.
10.6 Solved problems 267

- (ii) (E212j Eild -+ 00. The reader can check that Pc = 7f2 El Itfl2. Indeed,
it is as if the flexible part (1) is built-in at (x = lj2), thus Pc can be
obtained simply by setting (2lj2) in case (i).
- (iii) E 2 12 = 4El ft. After some algebraic manipulation, it is found that Eq.
(10.16) can be written in this case as follows:

(10.17)

The smallest criticalload is given by: Pc ~ 0.647f2 El Itfl2 . As expected,


the value is Iar ger than that of case (i) and smaller than that of (ii).

p 1_________________________ _

""T'-r-r-T~-r-7_.------- ------------

Fig. 10.16. Column with non-uniform properties

10.6.7 Eccentric compressive load

A column, pinned at one end (x = O) and simply supported at the other


(x = l) is subjected to a compressive force P which is parallel to the mid-
fiber axis (x) and eccentric with a distance e; see Fig. 10.17. Because of the
eccentricity, the beam will bend no matter what the value of P is. But we
shall see that the deflection may become too large if the load reaches a critical
value Pc.
Working on a deformed configurat ion v(x), one arrives at the following
differential equation together with its B.Cs.
268 10. Elastic stability

V" (x) + k 2 v(x) = -k2 e, v(O) = v(l) = O, (10.18)


where k 2 == P/(EI). It is found that the solution is given by:
1 - cos(kl) .
v(x) = -[1 - cos(kx)]e + sin(kl) e sm(kx) (10.19)

Using trigonometric formulae, the maximum deflection Vma", = v(l/2) can be


written under the following simple form:

Vma", = (COS(!l/2) - 1) e (10.20)

Let us first check the result for small values of the applied load P. A Taylor
expansion gives:

Le., a linear load-deflection relation is obtained for small values of P. We are


interested however in what happens for larger values of the load. Equation
(10.20) shows that if cos(kl/2) -7 O then Vma", -7 00. The smallest value of
k when this happens corresponds to kl/2 = 7r/2, which gives the following
critical load:
EI7r2
Pc = -l-2-' (10.21)

which is identical to Euler's solution (10.3), Le. the no-eccentricity case. It


can be shown however (Den Hartog, 1949) that the maximum stress rapidly
exceeds the axial compressive stress (P/A) even for small values of the ec-
centricity. Since in practice, a load cannot be perfectly directed along the
mid-fiber axis and an eccentricity always exists, it is important to use safety
factors when using formulae such as those of Secs. 10.3 and 10.6.

p
;rE,!
4 •
Fig. 10.17. Eccentric compressive load

10.6.8 Beam-column under compressive and bending forces

A so-called beam-column, pinned at one end (x = O) and simply supported


at the other (x = l) is subjected to a compressive force (-Pe",) at (x = l)
10.6 Solved problems 269

acting along the mid-fiber axis (x), and a bending force (Fe y) at (x = l/2),
Fig. 10.18. Because of the presence of F the beam will bend no matter what
the value of P is, but the axial force influences the deflection of the beam.
Working on a deformed configuration, one arrives at the following dif-
ferential equation for the deflection v(x) of the left part of the beam (with
k 2 == P/(EI)):

whose general solution is:

v(x) = Acos(kx) + Bsin(kx) - :;x, x E [O,~]


The constants are found from the two equations:
1
v(O) = O; vl ( - ) = O,
2
which follow from B.C. at (x = O) and symmetry at (x = l/2), respectively.
Substituting the values of A (= O) and B, the deflection becomes:

F [Sin(kX)] 1 (10.22)
v(x) = 2P kcos(kl/2) -x ,x E [0'"2]

After some algebraic manipulation, the maximum deflection V max = v(l/2)


can be expres sed as follows:
Fl 3 3(tan K - K)
(10.23)
V max = 48EI K3 '

------
with K == kl/2. The term under brace is what is obtained in Chap. 3 for
P = O. Equation (10.23) shows that V max may become "infinitely" (or un-
controllably, or arbitrarily) large when cos K ~ O. The smallest value when
this happens is K = 7r /2. The corresponding critical load is:

7r 2 EI
Pc = -l-2-' (10.24)

and this is exactly (10.3). Actually, this is another way to define or understand
buckling: when the axial compressive force reaches the value Pc, even the
smallest lateral force F can cause an important deflection.

10.6.9 Energy method

As a first application of the energy method of Sec. IOA, we consider the


problem of Sect. 10.3 again and try two expressions for v(x). First, we consider
the following K.A. field:
270 10. Elastic stability

. f/2
.. f/2
Fig. 10.18. Beam-column under compressive and bending forces

ii(x) = sin(7fy)
It is smooth and verifies the B.Cs. ii(O) = ii(l) = O. Applying formula (10.8)
is this case gives: Qc = 7f2EI/l2, which is the exact result which was found
using Euler's direct method, Eq. (10.3). This is only normal since the trial
field ii (x) coincides exactly with the real buckling mode. Now consider another
trial buckling mode for the same problem:

ii(x) = x(l - x)
This field is also smooth and satisfies the B.Cs. ii(O) = ii(l) = O. Application
of formula (10.8) gives Qc = 12EI/l2, which is a bad approximation to the
exact solution (10.3). We could have predicted that ii(x) is a poor trial mode
because ii" (x) = -2 = constant, thus M(x) = constant, which is a bad
approximation.
It is seen that the accuracy of the approximate value of Qc depends on
how close ii(x) is to the real (but unknown) buckling mode. The issue is even
more crucial because the energy-based method gives approximate values of Qc
which are higher than the correct ones. So the question is: how to choose v(x)
in practice? A good method consists in considering a beam which is subjected
to an adequately chosen transverse loading and taking the deflection of the
beam as the trial field ii(x).
As an example, consider the problem of Fig. 10.12 and suppose that the
exact solution is not known (otherwise, we will not be using an approximate
method). Now consider another problem: a transverse load (Pe y) is applied
at the end (x = l) of a cantilever beam which is built-in at (x = O). This
latter problem -(2)- was solved in Chap. 3 and the deflection was found as:
2 _ P v2(l)
V2(X) = Cx (3l- x), C = 6EI = 2i3
Obviously, V2(X) is K.A. (it is smooth enough and satisfies the B.Cs. V2(0) =
v~(O) = O). AIso, we expect the deformed shape of the cantilever beam of
problem (2) to be close to that of the buckling mode of the original problem of
Fig. 10.12. Thus it seems appropriate to choose ii(x) = V2(X). Formula (10.8)
gives the corresponding approximation of the buckling load: Qc = 5EI/(2l 2).
The correct solution is given by (10.10) as: Qc = 7f2EI/(2l)2 ~ 2.467EI/l 2.
Thus it is seen that the approximate method gives a value of Qc to within
1.33%, which is a good precision.
10.6 Solved problems 271

We can obtain an even better accuracy if we use the method recommended


in (Timoshenko and Gere, 1961). It consists in using M(x) in formula (10.8)
and not replacing it with (-Elv"(x)). In other words, we need to use the
following formula instead of (10.8):
fI Jif2(x)
Qc '" Q- c -_ Jo EI
1
dx (10.25)
Io [V'(X)]2 dx
~

For the example studied here (Fig. 10.12), we have:

M(x) Q[v(x) - v(l)] = Qv(l) [X2(~l; x) - 1] ;


3v(l)
v'(x) = 2i3x(2l- x)

Equation (10.25) then gives Qc = 42Elj(171 2) ~ 2.470Eljl2. This time


the accuracy of the approximation is 0.12%, which is much better than the
previous value (1.33%).
11. Theory of thin shells

Plates and shells are solids with one dimension -the thickness (h)- much
smaller than the other two dimensions. When the mid-thickness surface (8)
is contained in a plane, such solids are called plates, otherwise they are shells.
Plates were studied in chapters 5 and 6 for bending and 7 and 8 for in-plane
loadings (Le. plane stress). This chapter is concerned with thin shells. Most
of the time, it is assumed that the thickness is constant.

11.1 Geometry of the mid-surface


A point P in the mid-surface (8) of a shell has coordinates X, Y and Z in
a fixed Cartesian frame of origin O. (8) is represented by the following three
equations:

(11.1 )
where F, G and H are continuous and single-valued functions of the curui-
linear coordinates Xl and X2 (remark: there is an infinite number of possible
representations (11.1) for a given surface 8). The position vector r of point
P is:

(11.2)
A coordinate line Xl is a line in 8 such that X2 = constantj similarly along
a coordinate line X2 we have Xl = constant. As an example, consider that
8 is a circular cylinder. The easiest choice for the curvilinear coordinates is
Xl = () and X2 = z, where () is the polar angle and z the axial distancej see
Fig. 11.1.
We now return to the general case. Let II be a plane containing P and
the normal to 8 at Pj the intersection of II and 8 is a planar curve whose
radius of curvature at P we designate by R. If we rotate II about the normal,
there exist two orthogonal directions such -ţ,hat R is maximum or minimum,
respectively. Those directions at every point P of 8 define two orthogonal
networks of principal curuature linesj see (Dreyfuss, 1962) for proofs.
In this chapter, we shall always choose the curvilinear coordinates Xl and
X2 such that their coordinate lines are principal curvature lines. For a more
274 11. Theory of thin shells

z -line

() -line
y

Fig. 11.1. Example: Sis a circular cylinder. Curvilinear coordinates: Xl = () and


X2 = z.

general presentation, see (Green and Zerna, 1968), (Marsden and Hughes,
1983).

11.2 First fundamental form


This section is concerned with the measure of lengths on the mid-surface.
Consider on the surface S a point P(XI,X2) and two coordinate lines Xl and
X2 intersecting at P , the tangent vectors to each line at Pare designated by
Al and A2' respectively (see Fig. 11.2), and are defined by:
âr âr
AI=--j A 2 = - (11.3)
âXI âX2
These vectors are orthogonal following the choice we made at the end of the
previous section.

Xl -line

Fig. 11.2. Since line coordinates Xl and X2 are principal curvature lines, tangent
vectors Al and A 2 to these lines at P are orthogonal.

Let p i (Xl + dxt, X2 + dX2) be a point infinitesimally close to P(XI' X2)'


The difference in position between P and p i can be written as:
11.2 First fundamental form 275

ds r(xl + dXI, X2 + dx 2) - r(xI' X2)


8r 8r 2
r(XI,X2)+-8 dXI+-8 dX2+0(dxi)-r(XI,X2)
Xl X2
~ Al dXI + A2dx2

The norm of (ds) is the length (ds) of (PP'):

ds 2 = ds . ds = Al . Al dx~ + A 2 . A 2 dx~ (11.4)


~ ~

If Al and A 2 were not orthogonal, we would have (A I ·A2 ) termsj their differ-
entiation leads to the Christoffel symbols. Equation (11.4) which is rewritten
hereafter is called the first fundamental form:

(11.5)

This equation allows the computation of distances along S when Xl and


X2 vary. It is assumed that ai > O. Unit vectors el and e2 tangent to the
coordinate lines at Pare defined by:
Al A2
el =-j e2=- (11.6)
al a2

A normal vector to S at P is defined by:

(11.7)

Coordinates Xl and X2 will be chosen such that the normal e3 points away
from the centers of principal curvature at P, when those centers are on the
same side of the surface S.
In order to compute the elements of the first fundamental form, there are
two techniques: (i) differentiate the position vector r(xI' X2)j (ii) compute
lengths along the Xl and X2 coordinate lines separately. We illustrate these
methods in four cases.
The first example is a circular cylinder of radius R, Fig. 11.1. The curvi-
linear coordinates are Xl = O and X2 = z. The position vector r(O, z) is:

r = Re r + ze z = R(e x cos O+ ey sin O) + ze z


Differentiation w.r.t. the curvilinear coordinates gives:

~~ = R( -e x sin O+ e y cos O) = Reo j


8r
- =e z
8z
Since e() and e z are unit vectors along the O and z coordinate lines, respec-
tively, then using Eq. (11.4) we obtain:
276 Il. Theory of thin shells

(11.8)
We can find these results much more easily using the second technique.
Indeed, it is obvious that the lengths along the () and z lines are (RdiJ) and
(dz), respectively, the sum of their squares gives Eq. (11.8c).
The normal is e3 = elJ x e z = er, and points outside the centers of
curvature (if we have chosen Xl = z and X2 = (), we would not have found an
outward normal).
As a second example, consider a spherical surface of radius R, Fig. 11.3.
The curvilinear coordinates are chosen as Xl = 4J and X2 = ().

z 4J -line

Fig. 11.3. Spherical surface. Curvilinear coordinates: Xl = <p and X2 = ().

The position vector r(4J, ()) is given by:


r = (R sin 4J)( e x cos () + e y sin ()) + (R cos 4J )e z
Differentiation w.r.t. the curvilinear coordinates gives:

Al ~~ = (R cos 4J )(e x cos () + e y sin ()) - (R sin 4J )e z

A2 ~~ = (Rsin 4J)( -e x sin () + e y cos ())


The norms of these vectors are al and a2, respectively,
ai = Al . Al = R2; a~ = A 2 . A 2 = R 2 sin2 4J
Equation (11.4) then gives:
ds 2 = R 2d4J2 + R 2 sin2 4Jd()2 (11.9)
Unit vectors el and e2 along the Xl and X2 lines are given by el = AI! R
and e2 = A 2/(Rsin 4J). The normal is e3 = el x e2 and points outside the
centers of curvature.
11.2 First fundamental form 277

We can find Eq. (11.9) much more easily using the second technique.
Indeed, the "B = constant" line is a "vertical" circle of radius R contained
in a plane which includes the z axisj the infinitesimallength along this circle
is (RdcjJ). The "cjJ = constant" line is a "horizontal" circle ofradius (R sin cjJ)
contained in a plane orthogonal to the z axisj the infinitesimal length along
this circle is (RsincjJdB). The sum of the squares of the lengths gives Eq.
(11.9c).
As a third example, we consider a surface S in the (r, B) plane, Fig. 11.4.
The curvilinear coordinates are Xl = r and X2 = B. The position vector is:
r = re r = r(e", cos B + e y sinB)
Differentiation w.r.t. the curvilinear coordinates gives:
ar
-=erj
ar
~~ = r( -e", sin B + e y cos B) = ree

This gives the elements of the first fundamental form as:


al = lj a2 = rj ds 2 = dr 2 + r 2diJ2 (11.10)
Unit vectors tangent to the coordinate lines at (r, B) are: el = e r and e2 = ee.
The outside normal is e3 = el x e2 = e z •
Equation (11.10) can also be found using the second technique. Indeed,
the "B = constant" line is a straight line along which the infinitesimallength
is (dr). The "r = constant" line is a circle of radius r along which the in-
finitesimallength is (rdB). The sum of the squares of the lengths gives Eq.
(11.l0c).

r -line

Fig. 11.4. Surface in the (r,9) plane.

As a fourth example, we consider a surface S in the (x, y) plane. The


position vector is r = xe", + ye y • Simple differentiation gives:

Al = -
ar = e",j A2 = -
ar = e y
ax ay
278 11. Theory of thin shells

The elements of the first fundamental form are:

(11.11)

11.3 Second fundamental form


In this section, we study the variat ion of vectors el, e2 and e3 between two
infinitesimally close points P(Xl, X2) and P'(Xl + dXl' X2 + dX2)' Let Rl and
R2 be the principal radii of curvature at P of the Xl and X2 lines, respectively.
The following notations and computations are illustrated in Fig. 11.5:

1;
dS l al
7r
cos(- - dfh)
2
= sm(d8l

) ~ dlJ l = -R = -R dXl
1 1

Similarly, we have: e~ . e2 = a2dx2/ R2. It appears then that:


,
e3 = e3 + (Rall dxt}el + (Ra22dX2)e2
Comparison with a first order Taylor expansion of e~(xl + dXl' X2 + dX2)
shows that:

Fig. 11.5. Points P(Xl,X2) and Pl(Xl +dxl, X2) of the Xl coordinate line: tangent
vectors el and e~, normal vectors e3 and e~, principal radius of curvature Rl

Differentiation of the orthogonality condition el . e3 = O gives:


ael
-·e3
aXl
ael
-'e3
aX2
Differentiation of the unit vector conditions el . el = 1 and e2 . e2 = 1 gives:
11.3 Second fundamental form 279

where Q = 1,2. So far, we have projected âeI/âxa, along e3 and el. It


remains to project along e2; we have:

Dot products with el and e2, respectively, give two equations:

âa2 âal âel


- - e2' el +- el' el +al--' el
âx l " - - ' " âX2 "--'" â X2
o l "-,,--'
o
âa2 âal âel
--e2' e2 + - e l ' e2 +al--' e2
âx l " - - ' " âX2 "--'" âX2
l o

These simplify to:

âe2 1 âal âel 1 âa2


-·el=--; --·e2=--
âXI a2 âX2 âX2 al âXI
Differentiation of the orthogonality condition el . e2 = O gives:

-el' -âe2 - ____


1 âal o

âXI - a2 âX2'
âe2 1 âa2
-el'-=---
âX2 al âXI
An terms relative to el are now known. The missing terms for e2 can be
found from symmetry consideration as follows:

ae2 ae2 a2
- . e3 = O' - . e3 = - -
aXI ' âX2 R2
In summary, the terms aei/ âXa, i = 1,2,3, CI! = 1,2 are computed as follows:

âel 1 âal al âel 1 âa2


--
âXI
=- - - e 2 - -e3;
a2 â X2 RI
-=--e2
âX2 al âXI
âe2 1 âal âe2 1 âa2 a2
-=---el; - = - - - e l - -e3
âXI a2 âX2 âX2 al aXI R2
âe3 al âe3 a2
âXI =Rl el ;
- = -e2 (11.12)
âX2 R2
280 11. Theory of thin shells

11.4 Compatibility conditions of Codazzi and Gauss

The expressions of al, a2, Rl and R 2 in terms of the curvilinear coordinates Xl


and X2 cannot be chosen independently; they have to satisfy the compatibility
conditions of Codazzi and Gauss which can be written as follows (Dreyfuss,
1962):

~(~) = J.... âa2 ; ~(~) = J.... âa l ;


âXl R 2 Rl âXl â X2 Rl R 2 âX2
~(~ âa2) + ~(~ âal) = _ ala2 (11.13)
âXl al âXl âX2 a2 âX2 RlR2

We recall here that Gauss' curvature '" = 1/(Rl R 2 ) is not necessarily positive.
An example of a bell curve is given in Fig. 11.6, it is seen that '" > O in the
elliptic region, '" = O in the parabolic port ion and '" < O in the hyperbolic
part of the curve.

I l \ ",>0
j :

! ",=0
I
!I ",<O
I
I
I

Fig. 11.6. Bell curve: Gauss' curvature K, is positive in the elliptic region, zero in

the paxabolic port ion and negative in the hyperbolic part.

11.5 Surface of revolut ion

In this section, we consider a surface of revolution (S) and compute the ele-
ments of the first fundamental form as well as the principal radii of curvature
Rl and R 2 which are needed for the second fundamental farm.

11.5.1 General case

Surface (S) is generated by the revolution araund the z-axis of a curve whose
equation is r = f(z), Fig. 11.7. The curvilinear coordinates we choose are
Xl = ifJ and X2 = (). Consequently, an orthonormal basis (eq" e(), e3) is defined,
Fig. 11.7.
This choice of curvilinear coordinates fails in some cases. For example, if
(S) is a cylinder or a cone then ifJ = constant, such cases will be examined
11.5 Surface of revolution 281

in the next subsection. Another example is that of a bell curve where two
points on concave and convex parts of the curve, respectively, may have
parallel tangent vectors and therefore the same angle 1J.
Angle 1J and axial distance z play equivalent roles, but not r. Indeed, in
general r and () do not define a point P of S in a unique manner.

o .-----...,....--,L--__

1J -line

Oq,
z

Fig. 11.7. Surface ofrevolution: curvilinear coordinates c/J and 8, normal vector e3,
principal radii of curvature R,p and Re.

With r(z) : : : : dJ /dz, Fig. 11.7 shows that:

~ -1J = arctanr(z) ===} cos(~ -1J) = {1 + [tan(arctan f(z)]2} -1/2


2 2
From which the following useful identity is deduced:

sin1J = {1 + [t(Z)]2} -1/2 (11.14)

Since the 1J coordinate line is defined by: r = J(z), the principal radius of
curvature R 1 = Rq, of this line is given by:

{1 + [f'(Z)]2P/2
(11.15)
IRq,1 = II" (z)1
where I"(z)::::::: d2 f/dz 2 •
We now compute R 2 = Ro, the principal radius of curvature of the ()
coordinate line. It can be shown that -for instance by using Eq. (1l.12f):
282 11. Theory of thin shells

âe3/â() = e()a()/R()- that R() = O",P, where Dt/> is the intersection of the
normal to 8 at P (direction e3) with the axis of revolution, see Fig. 11.7. The
radius of curvature is therefore: Ro = j(z)/ sin ifJ, which -using Eq. (11.14)-
can be expressed as a function of z as follows:

R() = j(z){1 + [!'(z)fF/ 2 (11.16)

The position vector of a point P on the surface 8 is:

r = ze z + re r = ze z + j(z)(ey cos() + e", sin ())


Differentiation gives the tangent vectors as:

âr dz
= difJ[e z + j
I
A", = âifJ (z)erJ

A() = : = j(z)(-eysin() + e", cos()) = j(z)e()


Computing the norms of these vectors, we obtain:

These expressions can be simplified as follows:

at/> = Rt/>; a() = R()sinifJ (11.17)

The second expression is obvious, the first one is obtained by noting that
differentiation of the reiat ion tan(~ - ifJ) = f'(z) gives

dz 1 + [f'(z)J2
difJ 1" (z)
and then using Eq. (11.15).
Equations (11.17) can be found with a second and simpler method. Com-
puting infinitesimallengths along the ifJ and () coordinate lines and summing
their squares, it is easily found that:

Using Eq. (11.4), Eqs. (11.17) immediately follow.


Exercise: Re-derive the results of this section when the curve generating
the surface of revolution (8) is parameterized with: z = j(ifJ) and r = g(ifJ).
Exercise: Write the compatibility conditions of Sec. 11.4.
11.6 Gradient of a vector field in curvilinear coordinates 283

11.5.2 Conie surfaees

If S is a conic surface of half angle o, we cannot choose cjJ as a curvilinear


coordinate because it is constant everywhere on S, Fig. 11.8.
We may choose Xl = Z (the axial distance) and X2 = () (the polar angle).
The summit of the cone corresponds to Z = O. Note that tangent vector el
is not in the Z direction but is along the generator, see Fig. 11.8.
The principal radii of curvature are such that 1/ Rz = O and R() =
Z tan 0/ sin cjJ. Since cjJ = 1f /2 - o, it is found that:

sin o 1
R()=z--j - = 0 (11.18)
cos o Rz
2

Computing infinitesimallengths along the coordinate lines and summing their


squares, it is easily found that:

Since R()sincjJ = ztano, then using Eq. (11.4), we have:


1
a z = - - j a()
coso
= ztano (11.19)

A second possible choice for the curvilinear coordinate Xl is the distance


along the generator (origin: Xl = O at the summit). Since z = Xl cos cjJ, the
principal radii of curvature are given by:
1
R() = Xl tan Oj - =O (11.20)
RI
Squares of infinitesimallengths along the surface are given by:

Since R() sincjJ = ztano = Xl sin o, then using Eq. (11.4), we have:

al = 1j a() = Xl sin o (11.21)

11.6 Gradient of a vector field in curvilinear coordinates


In this section we compute the gradient of a vector field in the orthonormal
basis (el, e2, e3). The results will be useful for the computation of the strain
field later in this chapter.
Let Q(XI, X2, X3) be a point outside the mid-surface S of the shell and
P( Xl, X2) a point in S such that the position vectors are related by:
284 11. Theory of thin shells

z
Fig. 11.8. Conie surface, ~ = constant. Curvilinear coordinates: Xa = 9 and Xl =
axial distance Z or Xl = distance along the generator.

r( Q) = r(P) + X3e3 = r(xI' X2) + x3e3


Let f be a vector field (e.g., a displacement field) defined by:

f = F(XI,X2, x3)el + G(XI' X2, x3)e2 + H(XI,x2,x3)e3


Our purpose is to compute the gradient of f, which is designated by (Vf)
and defined by: df = Vf· CR;. First, we compute~:

~ =

aleldxl + a2e2dx2 + e3dx3 + X3 (~: eldxl + ~: e2dX2)


X3 X3
= (1 + Rl )aldxlel + (1 + R 2 )a2 dx 2e 2 + dX3e3
where we used Eqs. (11.12). We now compute (df):

df = (dF)el + Fdel + (dG)e2 + Gde2 + (dH)e3 + Hde3

After computation and rearrangement, the components of (df) in the or-


thonormal basis (el, e2, e3) are found as follows:

(dfh =
(dfh
11. 7 Kinematics of the mid-surface 285

Since we now know aH the components of the (3 x 1) arrays (df) and (dQ),
those of the (3 ~ 3) matrix ('\7 f) are computed thanks to the linear relation:
df = '\7f· dQ, l.e.

Identification with the dx 1 , dX2 and dXg terms in (df)i allows to compute
('\7 f)ij, i.e. the ith row in the matrix '\7 f. The final results are given below:

'\7f =

ax;
8F
8G
8X3
1 (11.22)
8H
ax;

11.7 Kinematics of the mid-surface


In this section, we compute displacement and strain fields of the mid-surface
S due to the application of loads and B.Cs. A point P of S transforms onto
a point pI; their position vectors are related by:

(11.23)
U(P)

where u is the displacement vector of P, with components VI, V2 and Vg in


the orthonormal basis (eI, e2, eg).
The displacement gradient '\7u(P) is given by Eq. (11.22) with Xg = O,
F = VI (XI, X2), G = V2(Xl' X2), H = Vg(Xl' X2) and 8/8x3 = O. The foHowing
matrix is easily found:

'\7u(P)

=
~1 (11.24)

The meaning ofthe notations fU, f22, Wl, W2, f31 and fh will become clear
later.
286 11. Theory of thin shells

Tangent vectors Al and A 2 to the coordinate lines at P transform onto


vectors A~ and A; according to:

A~ (1 + V'u) . (aleI) = ad(1 + EU)el + Wle2 - Jhe3]


A; (1 + V'u) . (a2e2) = a2[w2el + (1 + (22)e2 - fJ2e3] (11.25)

where 1 is the second order identity tensor and (1 + V'u) is the deformation
gradient (which will be studied further in Chap. 14). Vectors A~ and A; are
tangent to the deformed coordinate lines.
The foliowing dot products are easily computed:

A~ ·A; ala2[(1 + EU)W2 + (1 + En)WI + fJlfJ2]


A'·A'
l l ai[(1 + (11)2 + w;
+ fJ;]
A; ·A; a~[(1 + (22)2 + w~ + fJ~] (11.26)

These results show that A~ and A; are not orthogonal. By setting al = a2 =


1, it also appears that unit vectors el and e2 do not transform onto unit
vectors.
So far, the computations are valid even in finite strains; we now turn
to the infinitesimal case. The infinitesimal strain € is given by the following
matrix in the orthonormal basis (el, e2, e3):

(WI+W2)/2 -fJI/2]
E22 -fJ2/2
-fJ2/2 O

In the infinitesimal case, the absolute values of EU, E22, WI, W2, fJI and fJ2
are ali « 1. In this case, we have:

We have a similar result for !lA;II. Thus the significance of EU and E22 is
dear: they represent the change in length along the Xl and X2 coordinate
lines, respectively (they are "nominal" strains, see Chap. 14).
A unit vector along the deformed Xl coordinate line is defined by:

A similar result holds for e~. A normal to the deformed surface at P' is
defined by:
e~ = e~ x e~ ~ e3 + fJ2e2 + fJlel
In summary, for infinitesimal strains, unit vectors e~ and e~ tangent to the
deformed coordinate lines at P', and e~ which is normal to the deformed
surface are defined by:
11.8 Displacements and strains outside the mid-surface 287

e~ = el + Wle2 - (3le3; e; = W2el + e2 - (32e3

e~ (31 el + (32e2 + e3 (11.27)

We have seen the interpretation of tu and t22; as for (31 and (32, we immedi-
ately have:
(31 = e~ . el and (32 = e~ . e2.
Therefore (301. is the projection of the normal e~ to the deformed mid-surface
along eOl.' the coordinate line direction before deformation. It remains to
interpret Wl and W2. By definition, the infinitesimal shear strain is:

which means that the average value (Wl + w2)/2 is equal to t12'
In summary, infinitesimal strains in the mid-surface are given by:

1 8Vl V2 8al V3
tu - - + -al-a2-8X2
al 8XI
- +RI
-
1 8V2 VI 8a2 V3
--+--+-
a2 8X2 al a2 8XI R2
1 8VI V2 8a2 1 8V2 VI 8al
-- - - - +-- --- (11.28)
a28x2 ala28xI a18xl ala2 8 x 2

These strains tOl.ţ3 depend on the displacements of the mid-surface (VI, V2 and
V3) and its geometry before deformation (through aI, a2, RI and R 2 ).

11.8 Displacements and strains outside the mid-surface

11.8.1 General theory

In the previous sections, we computed displacements and strains for points


situated on the mid-surface (for which X3 = O). In this section, we turn our
attention to points outside the mid-surface (X3 ::j:. O). Classical thin shell
theory uses the Kirchhoff-Love assumptions which are identical to those of
thin plate theory (Chap. 5) and are recalled hereafter:
(H1) A material segment (or fiber) which is normal to the mid-surface
before deformation remains normal to the deformed mid-surface.
(H2) A normal material segment (or fiber) does not witness any change
in length.
Let P and Q be two material points situated inside and outside the mid-
surface, respectively, such that before deformation: ~ = X3e3. After defor-
mation, those points transform onto P' and Q'. According to assumptions
(H1) and (H2), we have: P'Q~ = x3e~. The displacement of Q is given by:
288 11. Theory of thin shells

()
uQ _ ~ ~ =%
QQ=Qr+PP+P 'Q~ =-x3e3+vlxl,x2el+v2XllX2e2
() ( )
+V3(xI,x2)e3 + x3(e3 + f3lel + (he2)

where we used Eq. (11.27c). It appears then that the components Ub U2 and
U3 of the displacement vector of Q in the orthonormal basis (el, e2, e3) are
defined as follows:

UI (Xl, X2, X3) = VI (Xl, X2) + X3f31 (Xl, X2)j


U2(XI,X2,X3) = V2(X1,X2) +X3{h(X1,X2)j
U3(XI,X2,X3) = V3(XI,X2) (11.29)

We introduce the following notation which will be needed later:

(11.30)

In order to compute the strains outside the mid-surface, we first compute


Vu(Q). This is given by the general formulae (11.22) with: F = Ub G = U2
and H = U3, where Ut, U2 and U3 are given by Eq. (11.29). Using notations
(11.30), the following result is found:

(1+~)-1(fn+x3#1:n) (1+~)-1(W2+X3'Y2) 131]


Vu(Q)= [ (1+~)-I(WI+X3'YI) (1+~)-1(f22+X3#1:22) 132
-~ -{h O
(11.31)

Using Eqs. (11.30-31), the components of the infinitesimal strain tensor € =


(V u + V T u) /2 have the following expressions:
X3 1
(1 + R 1 )- (IOn + X3#1:n)
X3 1
(1 + R2 )- ( 1022 + X3#1:22)

= (1 + ~: )-1 (WI + X3'Yt} + (1 + ~: )-I(W2 + X3'Y2)


t23 = t13 = O (11.32)

where fat), the strains in the mid-surface, are given by Eqs. (11.28), and
WoI and 1301 by (11.24). The transverse shear strains t23 and t13 vanish as a
consequence of assumption (Hl), and the normal strain t33 is zero because
of (H2). IT X3 is set to zero, it is easy to check that tOl{3 = f Ol{3.
Exercise (Crochet, 1994) With:
11.8 Displacements and strains outside the mid-surface 289

1\;12 - -1- [o2 V3


- 1 oal
- - OV3 1 oa2
-- - - OV3]
-
al a2 OXIOX2 al OX2 OXI a2 OXl OX2

1 [lOVI VI oal ] 1 [1 OV2 V2 oa2 ]


+ Rl a2ox2 - ala2 OX2 +R2 alOXI - ala2 OXI '

show that:

Terms lio:{3 are the components of the curvature tensor of the deformed mid-
surface, see (Dreyfuss, 1962).

11.8.2 Application: plates in rectangular coordinates

For a plate, the general formulae are greatly simplified because the radii of
curvature of the mid-surface before deformation are infinite, thus: 1/ Rl
1/ R2 = O. In Cartesian coordinates, Eqs. (11.10) and (11.24) give:

OV3 OV3
al = a2 = 1; (31 = - - ; (32 = - -
OXI OX2

Section 11.8.1 gives the components of the curvature tensor as:

The mid-surface strains are given by Eqs. (11.32) as:

OVI OV2 OVI OV2


En = -o
Xl
; E22 =~;
UX2
2E12 = -
OX2
+ -OXI
The strains outside the mid-surface are given by Eqs. (11.32) as:

OVI 02V3 _ OV2 02V3


~ - X3"""î2; E22 = ~ - X3"""î2;
UXI uX I UX2 uX 2

~(OVI + OV2) _ X3 0 2V3 .


2 OX2 OXI OXIOX2 '

The expressions for the curvature and the strains are identical to those which
were developed in Chap. 5.
290 11. Theory of thin shells

11.8.3 Application: plates in polar coordinates

As was mentioned before, for a plate we have 1/RI = 1/R 2 = O. In polar


coordinates (Xl = r, X2 = O), Eqs. (11.10) and (11.24) give:

aV3 1 aV3
al = 1; a2 = r; (3.. = - - ; (30 = - - -
ar r ao
The components ofthe curvature tensor are given in Sec. 11.8.1 as:

a2V3 1 &V3 1 aV3 1 a2V3 1 8v3


K. rr = - 8r 2 ; K.OO = - r2 a02 - ; : ar; K.rO = -;: arao + r2 ao

The mid-surface strains are given by Eqs. (11.28) as:

The strains outside the mid-surface are given by Eqs. (11.32) as:

Err Err + X3K.rr ; EOO = EOO + X3K.00;


ErO ErO + X3K.r O

The expressions for the curvature and the strains are identical to those which
were developed in Chap. 6.

11.9 Internal loads (stress resultants)

Similarly to plate theory (chapters 5 and 6), shell theory is developed in terms
of stress resultants, Le. stresses integrated through the thickness. We consider
a "slice" of shell of thickness h, see Fig. 11.9. The resultant of contact forces
on the facet with outside normal el is given by:

where tI is the stress vector on the facet of outer normal el:

and (dA 2 ) is an infinitesimal area measure on the facet:

Where R 2 is the radius of curvature of the X2 line and dS 2 = R 2 d(}2 is the


length of the X2 line within the opening angle d(}2. It appears then that the
stress resultant has the following form:
11.9 Internalloads (stress resultants) 291

j h/2
au(l
X3
+ -R )dX 3; N 12 ::::::
j h/2
a12(1
X3
+ -R )dX 3;
-h/2 2 -h/2 2

j h/2

-h/2
a13(1
X
+ R3 )dX3
2
(11.33)

h
2"

Fig. 11.9. A "slice" of shell of thickness h. Facet with outside normal el.

The resultant of contact forces along the facet of outside normal e2 is:

i1N 2 = (N22e2
,
+ N21el + Q2e3) ds l , .J
with:
'"
N2

j h/2

-h/2
a22(1
X3
+ -R )dX 3;
l

j h/2
a23(l
x
+ R3 )dX3 (11.34)
-h/2 l

The moment of contact forces on the facet with outside normal el w.r.t.
the center of the facet is given by:

i1M I :::::: / X3e3 x t l dA2

Using the expressions of tI and (dA 2), it is found that:


292 11. Theory of thin shells

h/2 X jh/2 X
M n == j x3 a n(1 + R3 )dX3; M 12 == X3a12(1 + R3 )dX3 (11.35)
-h/2 2 -h/2 2
The moment of contact forces on the facet with outside normal e2 w.r.t.
the center of the facet is given by:
L1M2 = ,(-M22 e 1 + M2le2)dsl;
, with:
...
M2

h/2 X jh/2 X
M22 == j X3a22(1 + R3 )dX3; M2l == X3a2l(1 + R3 )dX3 (11.36)
-h/2 1 -h/2 1

It is important to remember that -as in plate theory- N n , N 12 and Ql


are forces per unit length of the X2 coordinate line, while N 22 , N 2l and Q2
are forces per unit length of the Xl line. AIso, M n and M12 are moments per
unit length of the X2 coordinate line, and M 22 and M2l are moments per unit
length of the Xl line.
A difference with plate theory is that N 12 ;j:. N 2l and M 12 ;j:. M 21 (unless
the curvature radii Rl and R 2 are equal).
Stress resultants (Na{3, Qa) and their moments (Ma{3) together with their
positive directions are plotted in Fig. 11.10.

11.10 Equilibrium equations


11.10.1 General theory
We consider the same slice of shell as in the previous section; it is subjected
to the following forces (the notations are those of Sec. 11.9):
dXl dXl dx l
Nl(Xl + 2' X2)a2(Xl + 2,X2)dx2 (outside normal el (Xl + 2,X2))
dXl dXl
-Ndxl - 2,X2)a2(Xl - 2,X2)dx2 (opposite facet)
dX2 dX2 dx 2
N 2(XI,X2 + 2)al(Xl,X2 + 2 )dxl (outside normal e2(xl,x2 + 2))
- N 2 (Xl, X2 - 2dX2 )al (Xl, X2 - 2dX2 )dXl (opposite facet)
+ f(Xl, X2)al (Xl, X2)(dxl)a2(Xl, X2)(dx2) (external forces)
where f represents the external forces per unit area of the mid-sudace before
deformation. A first-order Taylor expansion gives:
dXl dXl
Nl(Xl + 2,x2)a2(xl + 2,X2) ~ N l (xl,x2)a2(xl,x2)
8 dXl
+-8 [N l (xl,x2)a2(xl, X2)]-
Xl 2
11.10 Equilibrium equations 293

N,(-
22

-M(+)
11
M(+)/
12

Fig. 11.10. Stress resultants (N"'f3' Q",) and their moments (M"'f3) plotted with
their positive directions.
294 11. Theory of thin shells

Similar results hold for the other terms. It appears then that equilibrium of
forces gives the following vector equation:
a a
-a [N I(xt. x2)a2(xI,x2)] + -a [N 2(xt, X2)al(XI,X2)]
Xl X2
+ f(xt. X2)al (Xt. X2)a2(XI, X2) = o
Using Eqs. (11.33), this can be rewritten as follows:
a a
-a [(Nnel +NI2e2+Qle3)a2]+-a [(N2Iel +N22e2+Q2e3)al]+ f ala2 =o
Xl X2
Using the expressions of aei/axa given by Eqs. (11.12), we obtain the follow-
ing three scalar equations which are the projections of the vector equation
along the el, e2 and e3 directions, respectively:
a a aal aa2 ala2
-a (Nl1 a2) + -a (N2I a l) + N12 -a - N22 -a + QI-R + h a la2 = o
Xl X2 X2 Xl 1
a a aal aa2 ala2
-a (NI2 a2)
Xl
+ -a (N22at) - Nl1 -a
X2 X2 Xl R + hal a2 = O
+ N21 -a + Q2-
2
a a
-a (Qla2) + -a (Q2 at) -
~~ ~~
Nl1 - R - N22 - R + !Jala2 = O (11.37)
Xl X2 1 2

We now consider the equilibrium of moments acting on the slice. Keeping


the notations of Sect. 11.9, it is seen that there are two contributions: one
is due to moments Ma and the other to the moments of forces N a' The
development for the former term is identical to that for the equilibrium of
forces; thus the sum of those moments has the following expression:

[a:l (M l a2) + a:2 (M2a d] dx l dx2


As for the second contribution (moments of forces N a w.r.t. P(XI,X2, X3 =
O)), one refers to the direction convention of Fig. 11.10 and obtains the fol-
lowing expressions:
dXl [ dx l dx l
al(xl,x2)2el x Nl(Xl + 2,x2)a2(xl + 2,X2)dx2
dx l
+ Nl(Xl - 2,x2)a2(xl dx l
- 2,X2)dx2]

dx 2 x [N 2(Xl,X2
+a2(xl,x2)Te2 dx 2
+ T)al(xl,x2 dx2
+ T)dx l

dX2
+ N 2(xt. X2 - 2)al(xl,x2 dx2)dxl ]
- 2

Similarly to the equilibrium of forces, a first-order Taylor expansion is per-


formed. Consequently, the sum of moments due to forces N a takes the fol-
lowing form:
11.10 Equilibrium equations 295

(el x NI +e2 x N2)ala2dxldx2 = (N12e3 -Qle2-N2Ie3+Q2el)ala2dxldx2


where we used Eqs. (11.33-34). When we add the contribution of moments
M <'" we find that equilibrium of moments gives:
a a
-8 [(-MI2e l + M U e 2)a2] + -a [(-M22e l + M2Ie 2)al]
Xl X2
+ [Q2el - Qle2 + (N12 - N2t}e3] ala2 = o
where we used Eqs. (11.35-36). Using the expressions of aei/axa given by
Eqs. (11.12), and projecting the vector equation along directions e2, el and
e3, respectively, we obtain the following three scalar equations:
8 a aal aa2
-a (Mll a2)+-a (M21at)+MI2-a -M22 -a -Qlala2=0
Xl X2 X2 Xl
8 a aal aa2
-a (MI2 a2) + -a (M22a l) - M l l -a + M21 -8 - Q2al a2 = o
Xl X2 X2 Xl
M l2 M
N l2 - N 21 + - - - 21 = o (11.38)
RI R2
In summary, there are six scalar equations: (11.37) and (11.38), which express
equilibrium of forces and moments, respectively. To the eontrary of plate
theory (Chaps. 5 and 6), membrane and bending equations are fully eoupled
in general, Le. equilibrium equations for "membrane" loads Na{J which act
in the surface of the shell, are coupled with those for bending moments Ma{J
and shear loads Qa. AIso, since Na{J and Ma{J are not symmetric in general,
Eq. (l1.38c) is not trivially satisfied.

11.10.2 Application: plates in rect angular coordinates


For a plate, we have: 1/RI = 1/ R 2 = O. In Cartesian coordinates, Eqs. (11.11)
give al = a2 = 1 and (11.37-38) take very simple forms:

aNll aN21 1 - o· 8Nl2 + aN22 + 12 = O·,


aXI + aX2 + 1 - , aXI a X2
aQI aQ2 f - o.
a Xl +aX2 + 3 - ,
aMl l + 8M21 = QI; aMl2 + aM22 = Q2;
8XI aX2 aXI aX2
N l2 = N 21 (11.39)
The reader can check that these equations are identical to those of Chap. 5.

11.10.3 Application: plates in polar coordinates


For a plate in polar coordinates, in addition to 1/RI = 1/R2 = O, we have
Xl = T, X2 = O, al = 1 and a2 = T. Equations (11.37-38) become in this case:
296 11. Theory of thin shells

âNrr 1 âNor Nrr - Noo .ţ O


--+---+
âr r âfJ r
+Jr= ;
âNro 1 âNoo 2Nro .ţ O
--+---+
âr râO
-+Jo= ;
r
1 â âQ()
;[âr(rQ r) + âfJ] + fa = O;
âMrr ! âM()r Mrr - M()o _ Q .
âr+râfJ + r -r,

âMr() + ! âM()() + 2 Mr() = Q();


ar r âfJ r
Nr() = N()r (11.40)

These equations are identica! to those of Chap. 6.

11.11 Constitutive equations

The results obtained so far are valid for any constitutive model. We now con-
sider the particular case of isotropic linear elasticity which gives the following
six scalar equations:

(1 + v) v
E U a {3 - E(u-y-y + u zz )t5a {3;
1 _ (1 + v)
E[U33 - vUn]; f a3 = -E--Ua3 (11.41)

As in Chap. 5, we make the following "plane stress" assumption:

(H3): The u33-stress is negligible in front of the "membrane" stresses (u a {3).

As a consequence of (H3), the first three constitutive equations simplify to:

Similarly to plate theory, the kinematic assumptions (H1) and (H2) of


KirchhofI-Love imply that f33 = f a 3 = O, which leads to the same con-
tradictions as in Chap. 5. Inverting the constitutive equations, we obtain:

(11.42)

Now, we should do the following:


(i) Replace fll, f22 and f12 with their expressions as a function of f a {3, /î,a{3,
X3, and Ra; Eqs. (11.32).
(ii) Compute the integrals w.r.t. X3 which give the expressions of internal
11.11 Constitutive equations 297

loads N a{3 and M a{3; Eqs. (11.33-36).


For example, the (Ju-stress is given by:

Internal force per unit length Nu is obtained as follows:

E jh/2 [ X3 X3 1
Nu -1--2 (l+-
R )(l+-R)~ (tU+ X3IbU)
- v ~h/2 2 1

+V(t22 + X31b22)] dX3 (11.44)

Because of the presence of terms (X3 / Ra), which we do not have for plates,
the integrat ion w.r.t. X3 is not easy in general (the case Rl = R 2 is an excep-
tion). In order to proceed further, we consider shells for which the following
assumption applies:

(H4): We consider thin shells such that max(h/ Rl, h/ R 2) « 1 and we neglect
all terms (X3 / Ra) in front of 1.

Consequently, it can be easily verified that internal forces per unit length
are given by:
Eh
N a{3 = 1 _ v2 [(1 - v)ta{3 + Vt"Oa{3] , (11.45)

while internal moments per unit length have the following expressions:

M a{3 = V [(1 - V)lb a{3 + VIb"Oa{3] , (11.46)


where V [N m] is the bending stiffness:
Eh 3
V=----
- 12(1 - v 2 )
Equations (11.45-46) are form-identical to those obtained for thin plates
(Chap. 5), however some major differences do exist:

(i) Mid-surface strain and curvature components, ta{3 and lb a(3, are related
to the displacements of the mid-surface via equations which are much more
complicated than those of plate theory.
(ii) Equilibrium equations are more complicated too, and -unlike plate theory-
it is not possible to un-couple bending from membrane equations.
As a final comment, it is noted that assumption (H4) implies that N 12 = N 21
and M 12 = M 21 . Therefore if Rl =1- R2' then equilibrium equation (1l.38c)
cannot be satisfied. For a membrane (next section), Eq. (l1.38c) is trivially
satisfied.
298 11. Theory of thin shells

11.12 Membrane theory

An important simplification to the theory developed so far is obtained


in the case of a membmne, which by definition can only transmit ten-
sionjcompression loadings in its surface and cannot resist any bending (in
the conventional meaning of the word, a membrane can only resist tension,
a typical example being a balloon).
Assuming that M a {3 = O, equilibrium equations (l1.38a, b) then give
Qa = O, and Eq. (l1.38c) reduces to N l2 = N 21 , which is in agreement with
assumption (H4) aud Eq. (11.45). Equilibrium equations (11.37) become:

â â âal âa2
- â (NU a2) + - â (N2I aI) + N 12 -â - N 22 -â + !I a la2 = O
Xl X2 X2 Xl
â â âal âa2
- â (NI2 a 2)
Xl
+ -â
X2
(N22 aI) - Nu - â
X2
+ N 21 -â Xl + !2ala2 = O
1 1
--Nu - - N22 + h =O (11.47)
RI R2
This is a system of three equations for three unknowns: Nu, N 22 and N l2 =
N 21 . Once N a {3 are known, shell displacements cau be found as explained
hereafter. First, constitutive equations (11.45) are inverted to give:

(11.48)

Next, tcx{3 are expressed in terms of the displacements and the geometry of
the mid-surface, Eqs. (11.28),

1 âVI V2 âal V3
--+----+-
al âXI ala2 âX2 RI
1 âV2 VI âa2 V3
--+----+-
a2 âX2 al a2 âXI R2
1 âVI V2 âa2 1 âV2 VI âal
- - - ---- + - - - - - - -
a2 âX2 al a2 âXI al âXI al a2 âX2

Let us remark that for simple problems (e.g., cases presenting axial symme-
try) internal forces per unit length N a {3 are computed not by solving system
(11.47) but simply by using a method of fictitious cuts and writing static
equilibrium conditions as in Chaps. 3 or 6; see next section for references.

11.13 Further reading

Solving thin shell problems often involves lengthy calculations. For membrane
theory, the amount of work can be manageable for rather simple problems.
11.13 Further reading 299

For bending problems however, even the simplest cases involve tedious com-
putations. For closed-form solutions for membrane or bending problems, we
refer to (Timoshenko and Woinowsky-Krieger, 1982), (Flugge, 1973), (Green
and Zerna, 1968), (Crochet, 1994). Nowadays, good numerical approxima-
tions can be found for many interesting shell problems using the finite ele-
ment method. For the linear case, see (Hughes, 1987) and references therein.
For shells with material or geometric nonlinearities, we point to the series
of articles by Simo and his co-workers -e.g. (Simo and Fox, 1989)- and their
lists of references.
12. Elasto-plasticity

In all previous chapters, we have considered linear elasticity, which is the sim-
plest -although most widely used- material model. Most engineering materials
have a linear elastic behavior at the early stages of deformation. However,
when certain criteria are reached, several materials (e.g., metals) undergo ir-
reversible, or permanent or plastic deformations. In this chapter, we present
constitutive equations and computational algorithms for rate-independent
elasto-plasticity.

12.1 One-dimensional model


Figure 12.1 shows an idealized uniaxial stress-strain response of a metallic
specimen.

B
a(B) ~----=---

of--------,f----i----

~(B)

Fig. 12.1. Uniaxial stress-strain response of a metallic specimen

The path (OAB) represents the response under increased loading; it con-
sists oftwo parts: (OA) and (AB). (OA) is a straight line of slope E, Young's
modulus of the material. If we unload anywhere along (OA), then the stress-
strain response will be along A -? O, if we reload again,the response will be
along O -? A. Point O is a state of zero strain and zero stress. Along (OA)
302 12. Elasto-plasticity

or (AO) the response is linear elastic, a = EL Along (AB) the response is


nonlinear, stress and strain are no longer proportional. The stress at point A
is ay == a(A) and called the initial yield stress. If starting from point B, we
unload, then the stress-strain states will be along B -+ C, a straight line of
slope E. Point C corresponds to zero stress, but non-zero strain: there is a
permanent, irreversible, or plastic strain fP == f(C). If we reload again, then
the stress-strain points will be along the line C -+ B. Looking at the figure,
it is seen that the strain in point B is made up of plastie and elastic parts:

f(B) = fP +f e
'-v-"
«O)

If we look at the triangle (6BD), we see that: a(B) = Efe. In summary, we


have the following relations:

f = fe + fP, a = Efe, lai ~ ay + R(P) (12.1)


"-v-'
~o

The inequality defines a yield criterion; R(P) is ealled a hardening stress (in
monotonie tension, R(P) is equal to (a - ay), Le., the differenee between the
eurrent stress and the initial yield stress); p is the aeeumulated plastic strain
defined by:

(12.2)

Exeept for monotonie loading, p(t) is not equal in general to IfP(t)l. The fol-
lowing example helps explain the differenee and the model in general. Con-
sider a eyclie tension/compression test under the total strain history shown
in Fig. 12.2a. The stress-strain response is shown in Fig. 12.2b (linear hard-
ening is assumed: R(P) = kp). If we examine the first cycle, we ean make the
following observations:
(A~Al): fP > O, p(A 1 ) = fP(Al)'
(AIBD: fP = O, p(BD = fP(AI)'
(Bi CI): fP < O, p(Cd = fP(Al) - (fP(Cd - fP(Al)) = 2fP(Ad - fP(CI).
So, we see that p(Cd # IfP(CI)I. Actually, in this example p(t) will always
increase from one cycle to the next (P(A 1 ) < p(A 2 ) < p(A 3 ), etc.), while we
will always have IfP(t)1 < i.
The model is such that the yield stresses in tension and compression
are identical (in absolute values), e.g. a(BD = -a(AI). Therefore, the model
cannot describe the so-called Baushinger effect. Finally, Fig. 12.2b shows that
la(t)1 keeps increasing from one cycle to the next (a(Al) < a(A 2 ) < a(A 3 ),
etc.), therefore the model cannot reach the experimentally observed stabilized
cycle. The model is primarily intended for the description of metal plasticity
under monotonie (and mainly proportional) loadings, and is widely used for
that purpose. Modeling of cyclic plasticity is addressed in Chap. 17.
12.1 One-dimensional model 303

-t

I
a
--------
I

-t I It

-------1
I

Fig. 12.2. Cyclic tension/compression test: (a) Strain history, (b) Stress-strain
response
304 12. Elasto-plasticity

12.2 Three-dimensional model


In this section, the (incremental) 3D theory of elasto-plasticity is formulated
by generalizing the ideas that we developed in lD in the previous section.
The total strain E is assumed to be the sum of two parts: an elastic strain Ee
and a plastic (inelastic, permanent or irreversible) strain EP :

(12.3)

The Cauchy stress u and the elastic strain Ee are related with a linear elastic
constitutive model:

(12.4)

A yield function I(u, R) defines a yield surface (f(u, R) = O) and an elasticity


domain (f(u, R) ~ O):

l(u,R) == aeq - ay - R(P) ~ O, (12.5)

where a eq is the von Mises equivalent stress defined by:

(12.6)

where 8 is the deviatoric part of the Cauchy stress:

(12.7)

The plastic ftow rule governs the evolution of the plastic strain:

'p .
E = "(-,
al .l.e., .p
fi'
.
= "(--
al (12.8)
au 3 aaij

The scalar l' is called the plastic multiplier. We have l' ~ O and the precise
sign is determined by the following conditions:

1'>0 if I =O (plasticity) and j = O (yielding)


1'=0 if I < O (elasticity ) or I = O and j < O (unloading)
(12.9)

The condition j = O is called consistency condition, it means that if we have


plastic yielding during a time interval, the solution should always remain on
the yield surface during alI that time. The scalar p is the accumulated plastic
strain, defined by:

(12.10)
12.3 Linear elasticity 305

with

. . _ (~.p .p )1/2 (12.11)


l.e., p - 3 ti/'-ji

The scalar R(P) is called the hardening stress. The model presented here is
often called J 2 elasto-plasticity or J2 flow theory, where J2 (lT) == (Jeq. The
flow rule (12.8) is also known as Prandtl-Reuss equation.

12.3 Linear elasticity

Linear elasticity has been discussed with great detail in Chaps. 1 and 2,
we only recall here some results which are useful in this chapter. Hooke's
operator E has the following symmetries:

(12.12)

If the elastic behavior is isotropic, then

E = 2Gl dev + Kl ® 1, Le., Eijkl = 2GIt'kl + KOijOkl


E- 1 = 1+v 1 _.!:I®1 . -1 1+v v
l.e., (E )ijkl = ~Iijkl - EOijOkl
EE'
(12.13)

where G and K are the shear and bulk moduli, respectively; they are given
in terms of Young's modulus E and Poisson's ratio v by:

E E
2G=-- and 3 K = - - (12.14)
1 + v' 1- 2v

The fourth-order tensor I dev is the deviatoric part of the fourth-order identity
tensor 1:

I dev = 1 - ~1 ® 1, i.e.,It"'l = ~(OikOjl + OilOjk) - ~OijOkl (12.15)

It is easy to verify that for a second-order tensor a, we have:

I dev
:a = deva, .l.e., I ijkla1k
dev
= aij - 1
aammUij
J:
(12.16)

which means that (I dev a) gives the deviatoric part of a. Other useful
results are:

(12.17)
306 12. Elasto-plasticity

12.4 Equivalent stress

Since the von Mises equivalent stress (1eq is the second invariant of the devi-
atoric stress, it is both isotropic and pressure insensitive, this last property
being generally well verified for metals. Some plasticity and strength crite-
ria have been discussed in detail in Chap. 1. Recall that an isotropic stress
criterion should be either a function of the principal stress invariants (e.g.,
von Mises criterion) or a symmetric function of the principal stresses (e.g.,
Tresca criterion). By developing the expression of (1eq, we can double-check
that it is independent of hydrostatic pressure; it is easy to show that:

=
(12.18)

There are other plasticity criteria which are isotropic but depend on the other
stress invariants. For example, geologic materials are pressure-sensitive, so
their plasticity criteria depend not only on (1 eq, but also on the first stress
invariant (1mm (e.g., the "cap" model) and even on the third invariant (det u)
(e.g., Lade's model); see (Desai and Siriwardane, 1984) and references therein.
There are plasticity criteria which are anisotropic, among them Hill's
criterion is probably the best known; it is written as (see Hill (1950) for
details):

(12.19)

where A, B, e, L, M, N are material parameters. We see that Hill's criterion is


pressure-insensitive (it was initially developed to model the behavior of metal
sheets), its expression is similar to that of von Mises and the two expressions
become identical if:

A = B = e =!2 and L = M = N = ~2 (12.20)

12.5 Hardening

The most commonly used forms for the hardening stress R(P) are:

R(P) = kp, R(P) = kpm, R(P) =R oo [l-exp(-mp)] (12.21)

They correspond to linear hardening, a power law and an exponential law,


respectively. The last model gives a saturation of hardening at large strains
(R(P) -t Roo when p -t 00).
12.6 Flow rules 307

Yield function (12.5) uses isotropic hardening. This has the following
meaning. In the space of principal stresses, the yield surface is a cylinder
whose radius is [oy + R(P)]. With increasing values of p, the radius increases,
but nothing else changes: the yield surface remains a cylinder whose axis is
fixed.
As mentioned in Sec. 12.1, isotropic hardening cannot model the exper-
imentally observed cyclic behavior of metals. In order to represent cyclic
plasticity, other models have been developed, such as kinematic hardening,
where, with increasing values of p, the radius of the cylinder representing the
yield surface increases, but also the cylinder moves in the stress space (that's
the reason behind using the word "kinematic"). One such model is described
in Chap. 17.

12.6 Flow rules

Since the yield surface is based on a von Mises criterion, the plastic flow rule
reads:
. af .3 s
.p
f:: = "{-
au = "{--
2 0eq
(12.22)

We see that, by design, i:P is incompressible (€~m = O), which translates the
experimental observat ionthat in general plastic flow of metals occurs without
volume change. Using (12.22) together with the definit ion of 0eq, it is easy
to check that:

(12.23)

So, the plastic multiplier l' is equal to the accumulated plastic strain rate p.
The plastic flow rule can be derived from Hill's maximum dissipation
principle which can be stated as follows.
Let 5 designate the elasticity domain, a set of admissible states of stress
u* and hardening stress R*:

5 = {(u*, R*), oij = aii I f(u*, R*) ::; O} (12.24)

It is shown in Sec. 12.11.3 that the mechanical dissipation has the following
expression for given values of (i;P,p)

v (u*, R*; i:P,p) = u* : i:P - R*p (12.25)

Hill's principle states that among all possible admissible states (u*, R*), and
for given values of (i:P,p), the solution (u,R) is the one for which the me-
chanical dissipation is maximum:
308 12. Elasto-plasticity

V (O', R; €p,p) = max {V (0'*, R*; €p,p)} (12.26)


(O'*,R*)ES
By a mere change of sign, this becomes a minimizat ion problem under the
inequality constraint f(O'*, R*) ~ O; it is solved using a standard technique
in optimizat ion theory, see (Luenberger, 1989) for example. Accordingly, a
Lagrangian functional is defined as:

-V (0'*, R*; €p,p) + "y* f(O'*, R*)


-0'* : €p + R*p + "y* f(O'*, R*) (12.27)

where "y* 2: O is a Lagrangian multiplier. The solution to the problem is then


given by (O', R) and "y which satisfy the following conditions:

~
ac( 0', R·.·p·)-,p .af (0', R)-O
",€ ,p =-€ +,~ -

-----
uO'* u~

ac ( R·'·P·) _. . of ( R) - o
aR* 0', ",€ ,p =p+, aR* 0', -
-1

"y 2: O, f(O',R) ~ O, "yf(O',R) = O (12.28)

Therefore, the normality (or associative) flow rules -(12.28a-b)- and the so-
called Kuhn-Tucker conditions -(12.28c)- can be derived from Hill's maximum
dissipation principle; it can also be shown that convexity of the elasticity
domain S is also implied by Hill's principle, see (Simo and Hughes, 1998) for
instance.
The plastic power (per unit volume) is given by:

(12.29)

Therefore, the mechanical dissipation is simply ayp 2: O.

12.7 Tangent operator, loading/unloading,


hardening/ softening

Taking the time derivatives of Eqs. (12.3-4), one finds:

(12.30)

II the transformat ion is elastic, then €p = 0, and iT = E : €, otherwise we use


the plastic flow rule (12.8) and the result "y = p to write:

O'. = E : (.€ - p f)
.8- (12.31)
80'
The time derivative ofthe yield function f (O', R) is:
12.7 Tangent operator, loading/ unloading, hardening/ softening 309

al .
1· = au dR. al E . h·
:u - dp P = au: :€ - p, (12.32)

where:
_ 81 81 dR
h= au : E: au + dp (12.33)

When 1 = o (plasticity), we can distinguish two cases:

(i) j < o (elastic unloading): then according to Eq. (12.9b), P= O and


therefore from Eq. (12.32):

(12.34)

(ii) j = O (plastic loading, consistency condition), then Eqs. (12.9a,32) give:

. 1 al E·
P="hau: :€>
o (12.35)

We will always assume that h is positivej we will make some comments later
in this section.
If a stress state is on the yield surface (f = O), the distinction between
elastic unloading (j < O) and plastic loading (j = O) has a simple geometric
interpretation. Because we assume that h is positive, Eqs. (12.34-35) show
that it suffices to study the sign of the cosine of the "angle" between (al/au)
-the normal to the yield surface- and (E : E)j see Fig. 12.3 .

.!lL
{JU

1=0 E:E
plastic
yielding

Fig. 12.3. Geometric interpretation of loading/unloading conditions

Now, rewriting Eq. (12.31) using (12.35), we find:

u = E : E- (E : ~~) ~ ~~ : E : E (12.36)

This gives:

iT = H: E, (12.37)
310 12. Elasto-plasticity

The fourth-order tensor H is called the tangent operator (it is the multi-axial
generalization ofthe tangent modulus in uniaxial tension aala€). H has the
same symmetries as E but, unlike it, it is not constant; it depends on the
deviatoric stress 8 and (dRldp).
For computations, if we store (al Iau) as a 6 x 1 array and E and H as
6 x 6 -symmetric- matrices (see Appendix C), then:

(12.38)

IT the elasticity is isotropic, then the expressions that we found become much
simpler. Since (al Iau) is deviatoric, Eqs. (12.13a) and (12.16) imply that:

E: al = 2G al (12.39)
au au
Also, using Eqs. (12.6, 22), we find:

al al 3
(12.40)
au : au = '2
Therefore, we obtain:

p = h2Gal
au :€
.

h 3G+ dR
dp
E _ (2G)2 al ® al
H = h au au
(12.41)

For computations, we obtain the simple matrix form:

[H] = [E] _ (2G)2 {al }{ al }T (12.42)


h au au
IT we have strain hardening (dRldp > O), then we see that h will always be
positive. IT there is strain softening (dRldp < O), h will also remain positive
in general; e.g., for steel, 3G ~ 225 GPa while IdRldpl is of the order of
10 GPa. This idea is generalized by Hill (1983) who writes that since the
plastic modulus is softer than the elastic one, we have:

e: (E - H) : e = i (e : E: ~~) ~ O
2 (12.43)

and therefore h should be positive. Another interpretation is given by Nguyen


and Bui (1974) who show that the positiveness of h prevents the stress-strain
response from presenting a "snap-back" .
We now return to the general case and write H as:
12.7 Tangent operator, loading/unloading, hardening/softening 311

181 81 ]
H = E: [1 - h 8u ® (8u : E) (12.44)

A hardening modulus is defined as follows:

h* == h _ 81 : E: 81 = dR (12.45)
8u 8u dp
Using lemma (1.1), it is easily checked that H is invertible if and only if
(dRjdp =1- O), and in this case:

-1 -1 1 81 81
H = E + (dRjdp) 8u ® 8u (12.46)

So for the material model we study in this section, it suffices to examine the
value of (dRjdp) to find out whether the tangent operator H (a fourth-order
tensor) is invertible or not. AIso, the expression of H- 1 is simple and useful
for numerical analysis.
Another use for (dRj dp) is that it gives a geometric interpretation of
strain hardening and strain softening. From Eq. (12.32), we see that the
consistency condition j = O implies that:
dR 181 .
- =--:u (12.47)
dp p8u
Since p > O (plastic loading), we see that it suffices to study the sign of the
cosine of the "angle" between the normal to the yield surface (8 fi 8u) and
0-; see Fig. 12.4.

.EL
8U

1=0
strain
hardening
strain
softening

Fig. 12.4. Geometric interpretation of strain hardening/softening conditions

In strain softening, the yield surface "shrinks" (O- is directed towards the
interior of the yield surface).
Since O- = H : € in plastic loading, we see that the loadingjunloading
and hardeningjsoftening criteria are very similar (the only difference is that
the first case involves Hooke's operator E while it is the tangent operator H
which appears in the second criterion).
Using the plastic flow rule, Eq. (12.47) can be rewritten as:
312 12. Elasto-plasticity

dR 1. 'p
dp = p2 (T : €. (12.48)

It is seen that iT : io P > O for a strain hardening material (dR/dp > O); this
is one form of Drucker's definition of material stability (see (Lubliner, 1990)
for more details).

12.8 Elementary examples

12.8.1 Uniaxial tension-compression

For a uniaxial stress state of tension or compression in the 1-direction, for


example, one has:

O
(12.49)

The general formulae of Sec. 12.2 take the following simple forms:

s (12.50)

We retrieve the equations of Sec. 12.1 and one more, related to the plastic
flow rule. It is seen that a eq = lai, and this is the reason behind using the
factor (3/2) is the original definition of a eq • AIso, p = WI which explains the
factor (2/3) in the definition of p.
The tangent operator in the plastic loading regime can be found by ap-
plying Eq. (12.46):

.
E =
(H- 1 )
uu a
.
=
( 1+ 1) .
E (dR/dp) a (12.51)

which gives:

âa E
-- E (12.52)
âE 1+ (dRjdp)

This result can be found directly from Eqs. (12.50) and the consistency con-
dition (j = O).
12.9 Boundary-value problem 313

12.8.2 Simple shear


Consider simple shear such that U12 is the only non-zero stress component.
It is easily found that:
U eq = V3jud
Exercise: Show that the stress and strain rates are related by:
G(dRjdp) .
0- 12 = 3G + (dRjdp) (2<:12) (12.53)

12.9 Boundary-value problem


First of all, we recall the statement of a boundary-value problem (BVP) in
quasi-statics, as seen in chapters 1 and 2. Consider a solid body which before
deformation occupies an open set n of ntJ. The body is subjected to forces per
unit volume li [Njm 3 ] in n and to the following boundary conditions (B.Cs.):
forces per unit area Fi [N jm2 ] on a part rF of the boundary and imposed
displacements Ui [m] on the remaining part ru. We define the following two
sets of virtual displacement fields:
Y {v = Iv "sufficiently smooth" and Vi = Ui on r u }
°
{Vi}

Y* {v* = {vi} I v* "sufficiently smooth" and V; = on ru}


(12.54)
It was shown in Chap. 2 that the problem can be formulated under the
following "weak" form:
Find the displacement field 'U which satisfies:

tF
'UEY

fa UijV(i,j) dn = fa fi V ; dn + Fi v; dr, Vv* E Y*


"Constitutive equations" (12.55)
A major difference with elasticity is that the constitutive model here is
history-dependent, therefore the imposed forces and displacements (fi, Fi, Ui)
as well as the solution are functions of time, or a time-like parameter. If the
problem is studied in a time interval [O, ti], then Eq. (12.55b) has to be sat-
isfied 'rit E [O, ti]' For a mathematical study of the BVP in plasticity theory,
see (Temam, 1985).

12.10 Numerical algorithms


12.10.1 Finite element method (F.E.M.)
For most elasto-plastic problems, closed-form solutions are not possible to
obtain and numerical solutions are sought via the finite element method
314 12. Elasto-plasticity

(F.E.M.), which was introduced in Chap. 2. We recall the main steps hereafter
(a significant difference is that the solution here may depend on time).
Finite-dimensional sets y(h} and y*(h} are used instead of Y and Y*. It
is assumed that:
y(h) C y, and y*(h) C Y* (12.56)

For v*(h} E y*(h}, and for the F.E.M. solution u(h} E y(h), it is assumed
that:

L {(A) (x)v:*(A) (t), L {(A) (x)Ui(A)(t)


n n
v;(h}(x, t) = u~h}(x, t) = (12.57)
A=l A=l

where {(A)(x) are given interpolation or shape functionsj v:*(A)(t) are coeffi-
cientsj (A = 1, ... ,n) are nodes and Ui(A}(t) nodal displacements.
It was shown in Chap. 2 that the F.E.M. approximation of formulation
(12.55) can be written as:
Find the displacement field U(h) which satisfies:
u(h) E y(h}

r aij â{(A)
la âXj
dil = r 1i{(A) dil + r Fi{(A) dF,
la lrF
(sum over j = 1,2,3)

" Constitutive equations" (12.58)

The time interval of interest is subdivided into a finite number of time inter-
vals [tn, tn+I]' The equilibrium equations (12.58b) are enforced at the discrete
times tn, tn+I' The problem can be stated as follows: knowing the solution
(displacements, strains, stresses) at tn, and given the applied load at tn+I,
find the solution at tn+I' As we shall see in Sec. 12.10.2, if e(x, tn+l) is
given, then u(x, tn+I) can be computed. For that reason, u(x, tn+I) can be
viewed as a -nonlinear- function of e(x, tn+d and therefore of u(h} (x, tn+I)'
Consequently, Eqs. (12.58b) are considered as a nonlinear system where the
unknowns are UitA) (tn+l)' This nonlinear problem is solved iteratively. Those
iterations are called global because they are defined at the level of the entire
body or structure (while iterations of the return mapping algorithm -Sec.
12.10.2- will be called local because they are defined point-wise). Equations
(12.58b) can be recast in the following form (the nodal displacements UitA)
are stored as UI, 1 = 1, ... ,3 x n):

(12.59)

Newton's method applied to this nonlinear system gives at each iteration (r):

(12.60)
12.10 Numerical algorithms 315

Using the chain rule, we see that in order to evaluate the Jacobian matrix
(EJ:FI/8UJ), we need to compute (8aijj8tkt)j this is explained in Sec. 12.10.3.
Here is a summary of the algorithmic procedure:
An approximation of the nodal displacements at t n +1 is proposed, from
which the strain field at t n +1 is computed (actually, since the integrals are
evaluated by numerical quadrature -e.g., method of Gauss- only values at a
finite number of integration points are needed):

(12.61)

The constitutive equations are then integrated in time to find the stresses at
t n +1 (Sec. 12.10.2). If the weak form of equilibrium at t n +1 is satisfied -Eqs.
(12.58b) or (12.59)- then the solution at t n +1 has been found, and the process
can move on to the next time interval [t n +1, t n +2]' Otherwise, a new iteration
-within the same time interval [t n , t n +11- starts: a new approximation to the
nodal displacements at t n +1 is proposed by solving the linear system (12.60).
We only deal with aspects related to the constitutive equations in this
section. For a more detailed presentation of the F.E.M. in nonlinear mechan-
ies, see (Owen and Hinton, 1980), (Zienkiewicz and Taylor, 1991), (Crisfield,
1997, 1998), (Ferencz and Hughes, 1998), (Kleiber, 1998) and (Simo, 1998).

12.10.2 Return mapping algorithm

The purpose of the return mapping algorithm is the following: the values
of the variables at tn being known and the total strain increment Lle being
given, find the values of the variables at tn+l'
An increment of a variable (stress, strain, etc.) will be denoted by:

(12.62)

The stress-elastic strain relation (12.4) written at time tn+l is:

U n +1 = E: [.cen ~ Lle), - ,(e~ +", Lle


P ),] (12.63)

The first step in the return mapping algorithm is called " elastic predictor":
the increment is assumed to be entirely elastic (Llp = O, LleP = O). The stress
at t n +1 is then simply:
trial
Un+l = E .. (en+l - eP)
n (12.64)

If this trial stress satisfies the yield condition:

trial - (a )trial
j n+l - eq n+l
- a Y - R(pn ) <
_
O (12.65)
316 12. Elasto-plasticity

then the basic assumption is true and the elastic predictor is indeed the
solution to the constitutive equations:

(T n+l = (T trial
n+l ,
P
€n+l = €n,P Pn+l = Pn (12.66)

If f~:':'~l > Othen the assumption is incorrect: plasticity has developed during
the increment, and we must find the correct state at tn+l such that it satisfies
fn+l = O and the rest of the constitutive equations. This is the second step
in the return mapping algorithm: the plastic corrector phase.
Combining Eqs. (12.63-64), we can write:
~
vn+l -
- ~trial
Vn+l
_ E'. A cP
.u", (12.67)

From now on, we will assume that the elasticity is isotropic (it would not make
much sense to use an isotropic yield criterion with anisotropic elasticity).
Since Ll€p is deviatoric, using (12.13), Eq. (12.67) can be rewritten as:
~ -
vn+l-vn+l
~trial _ 2G .u",
AcP
(12.68)

Note that the trace of (T n+l is known:


t r (Tn+l -- t r (T triat
n+l -- 3Ktr €n+l (12.69)

Therefore, only the deviatoric part 8 of the stress needs to be computed.


Combining Eqs. (12.68-69), it is found that:
8 -
n+l -
8trial -
n+l
2G .uA€p (12.70)

For simplicity, we now introduce the following notation for the normal to the
yield surface in stress space:

N= Of =~..!.. (12.71)
O(T 20'eq

The rate equations are discretized in time according to a fully implicit (back-
ward Euler) integrat ion scheme:

(12.72)

Ortiz and Popov (1985) have studied integration algorithms generalizing the
trapezoidal and midpoint rules and found in particular that the backward
Euler scheme offers unconditional stability and good accuracy. Using How
rule (12.8), the plastic strain increment is approximated as follows:

(12.73)

From now on, the subscript (n + 1) is omitted for simplicity and an variables
which do not contain this subscript are evaluated at tn+l' Equation (12.70)
is now rewritten as:
12.10 Numerical algorithms 317

s = strial - 2G N .:1p ~ ~3 N a eq = ~3 Ntrial atrial


eq - 2G N .:1p (12.74)

This implies that:

N = Ntrial, an
d 2G A
"-1p 2
+ 3aeq 2 triat
= '3aeq (12.75)

The equality N = Ntrial means that we have a radial return algorithm,


originally proposed by Wilkins (1964). The second equality in (12.75) means
that the problem is reduced to solving a single scalar equation with the sole
unknown p:

k(P) == 3G.:1p + R(P) + ay _a!~ial = O (12.76)


'--""

If the hardening is nonlinear, this is a nonlinear equation which can be solved


iteratively with Newton's method. Those iterations are called local because
they are defined at the (Gauss) point level. At each iteration, one solves:

(12.77)

where cp is a "correction" to p{s}, Le.

(12.78)

Note that the denominator of cp is simply h of Eqs. (12.41). For other elasto-
plastic models, this is not always the case. The starting iteration corresponds
to the elastic predictor:

(12.79)

The iterations are stopped once a convergence criterion is satisfied:

Ik(p{S})1 ~ tolerance, e.g., 10-6 (12.80)


ay
The solution at tnH can then be completely computed:

aeq ay + R(P)
s = ~a Ntrial
3 eq

a- s + ~(tr a-trial)l
3
.:1eP Ntrial.:1p (12.81)

The algorithm is illustrated in Figs. 12.5a,b in lD and 3D, respectively.


318 12. Elasto-plasticity

O'trial
n+l

trial > O
f n+l
O'

In+! =0

Fig. 12.5. Geometric interpretat ion of return mapping algorithm: (a) 1D, (b) 3D

The elastic predictor/plastic corrector algorithm presented here is a par-


ticular case of an operator split methodology; see (Simo and Hughes, 1998)
and references therein.
The return mapping algorithm has to be modified in the case of plane
stress, where the out-of-plane strain is unknown and must be determined
from the condition that the out-of-plane stress is zero; see Appendix D.

12.10.3 Consistent tangent operator


As mentioned in Sec. 12.10.1, when the return mapping algorithm has con-
verged but the weak form of equilibrium -(12.58) or (12.59)- is not satisfied,
a new global iteration is needed in order to propose new approximations to
the nodal displacements (and therefore the strain fields) at t n +1! within the
same time interval [t n , tn+t]. When a Newton method is used to iterate on the
globallevel-Eq. (12.60)- a global Jacobian matrix is computed byassembling
local tangent operators.
The so-called consistent or algorithmic local tangent operator Halg is
determined by letting alI the variables, including the total strain tensor E,
vary slightly about the converged solution at tn+b so that:
(12.82)
IT the increment is entirely elastic (Le., if the elastic predictor is the solution of
the integration algorithm), then Halg = E; otherwise, we proceed as follows.
As in the previous subsection, the subscript (n + 1) is omitted for simplicity
and alI variables which do not contain this subscript are evaluated at t n +!.
Differentiation of the stress-elastic strain relation (12.63), the discretized
plastic flow rule (12.73), and the yield condition gives the following equations:
12.10 Numerical algorithms 319

ou = E: (oe - oeP )
âN
oeP Nop+ (Llp) âu : ou
dR
N: ou - dp op o (12.83)

Rewritiug the last equatiou usiug the first two leads to:

dp op = N : E: [oe - N op - (Llp) âN
dR âu : ou ] (12.84)

Which implies that:

op = h
1 N : E: [Oe - (Llp) â
âuN ]
: ou (12.85)

We substitute this expression into (12.83b) and then (12.83a); we find:

[ I+{,1p)E: âN Llp
âu -(-;;:)(E:N)® ( N:E: âN)]
âu :ou=H:Oe
(12.86)

where H is the "continuum" tangent operator defined by (12.37). It is easy


to eheck that:

(12.87)

From whieh it is dedueed that:

N: âN =0 aud âN: âN =_3_âN (12.88)


âu âu au 2aeq au
If the elasticity is isotropic, then:

E: âN =2G âN (12.89)
au âu
Therefore, Eq. (12.86) beeomes simply:

(12.90)

Using lemma (1.1), it ean be eheeked that:

( 1 + 2G âN
âu Llp
)-1 âN
= 1 - 2G€ âu

where the dimensionless factor € is given by:


320 12. Elasto-plasticity

Llp _ ~Ll (12.91)


~ =: 1 + 3G.:lp - triat P,
lI e • (J'eq

using Eq. (12.76). Finally, using (12.41c), the consistent or algorithmic tan-
gent operator is found to be:

(12.92)

This is a simple and explicit reIat ion which shows that w.r.t. the "contin-
uum" tangent H, the "consistent" tangent Halg has an extra term which
depends on two quantities: the plastic strain increment Llp and the curva ture
of the yield surface in stress space (aN jau = a2 f ja( 2 ). We see that if the
plasticity increment is very small (Llp -1 O) then Hatg -1 H, otherwise the
two operators are different. The importance of consistent linearization was
demonstrated by Simo and Taylor (1985) who showed that if we use New-
ton's method to solve the weak form of the equilibrium equations, then it is
essential to use a consistent tangent in order to preserve a quadratic rate of
convergence for the method.
Exercise: show that the inverse of Hatg has the simple expression:

(Halg)-l = H- 1 + aN Llp (12.93)


au
where H- 1 is given by (12.46). Hint: use lemma (1.1).
The reader can check that for a uniaxial stress state, (aN jaU)llll = O,
and therefore there is no difference between consistent and continuum tan-
gent operators for uniaxial tension-compression. For multiaxial stress states,
however, the two operators can be very different.

12.11 A general framework for material models

The model which was presented in Sec. 12.2 is perhaps the simplest -although
most widely used- elasto-plastic model. In this section, we present a frame-
work for more general constitutive models. We limit ourselves to the basic re-
sults; for more information, see (Coleman and Gurtin, 1967), (Germain et al.,
1983), (Halphen and Nguyen, 1975), (Kestin and Rice, 1970), (Lemaitre and
Chaboche, 1990), (Lubliner, 1990), (Maugin, 1992), and references therein.

12.11.1 State variables

In the method of local state, the thermodynamic state is completely defined


at given (material) point and instant by the values of certain state variables
at that location and time. Those variables only depend on the position in
space and time of the partide being considered.
12.11 A general framework for material models 321

The following additive decomposition of the total (observable) strain (E)


into thermo-elastic (Ee) and inelastic (EP) parts is assumed:

(12.94)

For the models studied in this book, the state variables are: E, the absolute
temperature T > O and a set of internal variables EP and V. Those variables
are macroscopic measures of irreversible phenomena such as plastic deforma-
tion, hardening or damage.

12.11.2 Equations of state

The equations of state associate thermodynamic forces to state variables


through a scalar-valued function called the thermodynamic potential. The
function chosen here is Helmholtz's free energy per unit mass, 'ljJ(Ee , T, V),
which is assumed to be convex w.r.t. Ee and V, and concave w.r.t. T, in order
to satisfy thermodynamic stability conditions.
It is shown in Sec. 14.9.3 that the Clausius-Duhem inequality reads:
"
u : e - p('ljJ + sT) - '\lT· Tq ~ O, (12.95)

where p[kg/m 3 ] is the mass density, q the heat flux vector and s the entropy
per unit mass. After differentiation w.r.t. time of (12.94) and the free energy
expression, and substitution in inequality (12.95), the latter becomes:

(u - p :~) : ee - p(s + :')1' -/}; •V + u : eP - '\lT· ~~O (12.96)

By imagining independent processes and requiring each one of them to be


thermodynamically admissible, Le. satisfy (12.96), we obtain:

â1jJ â1jJ
u=p- s=-- (12.97)
âE e ' âT
From (12.96), thermodynamic forces A associated with the internal variables
V are defined by equations similar to (12.97), namely:

(12.98)

Relations (12.97-98) are called equations of state. In this section as well as in


Sec. 12.12, the symbol (.) designates a sum of appropriate inner products.
For example, if V consists of a scalar s, a vector Vi and a second-order tensor
tij, and associated variables A are: S, Vi and 'Tij, respectively, then we have:
322 12. Elasto-plasticity

Example: In the isothermal case, and for many models, the free energy is
written under the following uncoupled form,
1 ~
p'lj;(ee, V) = _ee: E: ee + 'Ij;(V) (12.99)
2
Equations of state (12.97a) and (1298) are then:

A=â;j
u = E : ee , âV (12.100)

Using (12.94), it is clear that:


â'lj; â'lj; â'IjJ
p âe = -PâeP = Pâee = u
For the model of Sec. 12.2, we have: V = p and:

;j(P) = l P
R(r)dr

Equation of state (12.98) then gives: A = R(P).


12.11.3 Flow rules

Equations of state (12.98) do not allow to find the internal variables eP and
V because they only introduce new quantities- the thermodynamic forces
A associated with V. Similarly, even if the observable state variable e is
known, Eq. (12.97a) does not allow determinat ion of its elastic or inelastic
parts ee and eP • Complementary equations, known as ftow rules, are needed
in order to determine the internal variables eP and V. Using Eqs. (12.94) and
(12.97-98), inequality (12.96) becomes:
. q
V == U : i.P - A • V - \1T . T ~ O, (12.101)

thus requiring the dissipation to be non-negative. Inequality (12.101) also


shows that forces (u, A, \1T) are dual or (conjugate) with their respective
flux variables (i. P , - V, -q/T), since their inner products give a power per
unit volume. The total dissipation V is the sum of mechanical and thermal
contributions, V mec and V th , and each of them must be non-negative:

V mec == u : i.P - A • V ~ O, Vth == - \1T· ; ~ O (12.102)

The second inequality translates the experimental fact that heat flows from
hotter to colder regions of a body. It is verified by Fourier's heat conduction
law (Sec. 9.2).
We now introduce a new function: the pseudo-potential of dissipation,
cP*(u,A, \1T), which is a scalar-valued function that depends on the forces
12.11 A general framework for material models 323

and is supposed to be non-negative, convex, and equal to zero at the origin,


q,* (0,0, O) = o. The flow rules are derived from q,* according to:

i.P - 8q,* V 8q,* q 8q,*


(12.103)
- 8u' - = 8A' T 8(\1T)
Since q,* verifies by definit ion the properties stated above, then substitution
of Eqs. (12.103) into (12.102) shows that the Clausius-Duhem inequality is
automatically satisfied. We now introduce generalized "vectors" of internal
variables, X and thermodynamic forces, Y as follows:

X == (eP , -V); Y == (u,A) (12.104)

It appears that condition (12.102a) can be written as follows:


. 8q,*
Y • X ~ O or Y. 8Y ~ O (12.105)

Material models which are built within the framework described so far are
called generalized standard materials (Halphen and Nguyen, 1975). In sum-
mary, in order to construct such models, one needs to choose internal variables
V and write down expressions for two scalar-valued functions: the Helmholtz
free energy t/J(ee, T, V) and a pseudo-potential of dissipation q,*(u, A, \1T),
which satisfy mathematical conditions stated above. The equations of state
are (12.97-98) and the flow rules are (12.103).
The theory which was developed so far is valid for both elasto-plasticity
and elasto-visco-plasticity. In the former case however, the theory is usually
formulated in terms of a yield surface, and the flow rules are written differ-
ently, as explained in the next subsection.

12.11.4 Rate-independent plasticity

We study isothermal, rate-independent plasticity and consider, in the space


of forces Y = (u, A), a convex domain C which contains the origin (O, O)
and is defined by a scalar-valued and convex function f(Y) :c::; O. The pseudo-
potential of dissipation, q,*(Y), is defined as the indicator function of C:

q,*(Y) = O if Y E C,
q,*(Y) = +00 ifY ~ C (12.106)

Since q,* is not differentiable, normality rules (12.103a-b) are generalized as


follows:

X E 8q,*(Y), (12.107)

where X = (eP , - V) designates the internal variables and 8q,* the sub-
differential of q,*. It is possible to prove the following three results- e.g.
324 12. Elasto-plasticity

(Ladeveze, 1986) or (Maugin 1992),

(a) Dissipation inequality (12.105) is equivalent to:

(Y - Y*) • X ~ O, V'Y* E e (12.108)

(b) If Y is inside the domain e, then X = O.


(c) If Y is on the boundary of e, then X is normal to the boundary.
Results (b) and (c) can be re-written as follows using the yield junction f(Y),

X=O if f(Y) < O,


X = 'y 8f if f(Y) = O, (12.109)
8Y
with 'y > O a plastic multiplier. Equation (12.108) is Hill's maximum dis-
sipation principle, which was used in Sec. 12.6. Equations (12.109) are the
classical flow rules, e.g. (12.8-9) for the model of Sec. 12.2. Note that it is
possible to derive the flow rules directly from Hill's principle, as in Sec. 12.6.

12.11.5 Beat equation

It is shown in Sec. 14.9.1 that the first law of thermodynamics can be written
in the following local form:

pe = (T : i: + prezt - divq, (12.110)

where e [Jjkg] is an internal energy per unit mass, and Tezt [Wjkg] a mass
density of internal heat production due to external sources such as inductive
heating. Internal and free energies per unit mass, 1/J and e, are related by:

1/J(ee, T, V) =e - Ts, (12.111)

where s is the entropy per unit mass and T > O the absolute temperature.
Using equations of state (12.97-98), Eq. (12.110) can be rewritten as follows:

-divq + ,(prezt + V mec + pT1l), = pcr, (12.112)


..
r

where c [J jkgj K] is the specific heat, V mec the mechanical dissipation and
1l the structural heatingj they are given by the following identities:

c =
V mec = (T: i: P - A • V
1l ~(81/J) .·e 3...-(81/J). V (12.113)
- 8e e 8T . e + 8V 8T
12.12 A class of non-associative plasticity models 325

For thermo-elasticity, we have V mec = O. Moreover, the contribution of 1l


is negligible, therefore the heat equation can be written under the following
form which was used in Chap. 9:

-divq + prezt = pcr (12.114)

For metal plasticity, it is observed IA. VI is much smaller than 10- : EPI. The
1l term is also neglected. Moreover, in order to match experimental results,
the following expression is often used, with 'TI ~ 0.9,

r = pr ezt + 'TIO- : i=.P (12.115)

Because part of the plastic work is dissipated in heat, the thermo-mechanical


problem is coupled. Therefore, an uncoupled procedure such as the one which
was presented in Chap. 9 cannot be used. In practice, a staggered iterative
algorithm can be developedj see for example (Simo and Miehe, 1992) and
references therein.

12.12 A class of non-associative plasticity models

There are several material models which do not enter in the mould of gen-
eralized standard materials, for instance there are models for which the flow
rules are derived from a "plastic potential" which is diJJerent from the yield
function. This is called "non-associative plasticity" . Examples are the nonlin-
ear kinematic hardening model of Chap. 17 and the ductile damage model of
Chap. 18. A general class of non-associative models is studied in this section.
The internal variables V may be scalars, vectors or tensors. The total
strain t=. is assumed to be the sum of elastic and inelastic parts (t=.e and t=.P),
Eq. (12.94). The stress o- and the thermodynamic forces A are derived from a
specific free energy, 'lj;(t=.e, V) according to the equations of state (12.97a-98).
An elastic domain is defined by the yield function:

1(0-, Aj V) ::; O (12.116)

The evolution laws for t=.P and V are derived from a plastic potential,
F( 0-, Aj V), according to the generalized normality rules:

'p . oF V· . oF
t=. = "( 00-' = -"( oA' (12.117)

with l' obeying conditions (12.9). It is assumed that 'lj;, 1 and F are smooth
functions, so that their partial derivatives which appear in the equations are
well defined.
The inequality of Clausius-Duhem requiring the dissipation to be non-
negative reads:
326 12. Elasto-plasticity

V=u:i:P-A.V~O (12.118)

As in Sec. 12.11, the symbol (.) designates a sum of appropriate inner prod-
uets. When l' > O, the following consistency condition holds:

(12.119)

We introduce the following notation:

(12.120)

For classical models, E ia Hooke's operator. Differentiation of the equations


of state w.r.t. time gives:


= E"
: ~ - 'Y (E : âF A âF)
80' + • BA '

A· AT .. _ . (A T . âF âF)
= .E 'Y . alT + II • BA (12.121)

We assume that E ia invertible and introduce another set of notation:

(12.122)

We shall assume that h ia positive. Combining Eqs. (12.119, 121), we obtain


the plastic multiplier as:
. b: E: i:
'Y= (12.123)
h
Using the same procedure as in Sec. 12.7, one finds that the stress and strain
rates are related as follows:

ir = L : i, (12.124)

where L is a tangent operator defined by:

L=E iff<Oor f = O and b: E : i: < O,


h
L =E - i(E : a) ® (b : E) if f = O and b: E : i: >O
, ,.. , h
H
(12.125)
12.13 Further reading 327

From the equations found so far, a hardening modulus h* is defined:


b : iT
h* == h - b : E : a =- - (12.126)
t
Using lemma (1.1), it is found that the tangent operator H is invertible if
and only if h* =f O, and if so, then:

H-1 = E- 1 +a®b (12.127)


h*
Remarkably, the results found in this section are form-identical to those of
Sec. 12.7, although the constitutive equations here are much more general.

Exercise: Apply definitions (12.120, 122) to the model of Sec. 12.2, and
check that the expressions of t, H, hand h * found in this section correspond
to those of Sec. 12.7.

12.13 Further reading

There are some classical solutions and subjects (e.g., bending, torsion, neck-
ing, slip lines, limit loads, shakedown, etc.) which we did not studYj inter-
ested readers may consult (Hill, 1950), (Prager and Hodge, 1951), (John-
son and Mellor, 1962), (Nadai, 1963), (Thomsen et al., 1965), (Mandel,
1966), (Kachanov, 1971), (Lubliner, 1990), etc. For applications to structures
(beams, plates and shells), references include (Hodge, 1959), (Massonnet and
Save, 1967), (Calladine, 1969), (Save and Massonnet, 1972), (Massonnet et
al., 1979), (Stronge and Yu, 1993), etc. AIso, we are only concerned with the
macroscopic phenomenological theories of plasticity and viscoplasticitYj for
micro-mechanical aspects, see Chap. 20 and references therein.
13. Elasto-viscoplasticity

In the previous chapter, we studied rate-independent plasticity. In this chap-


ter, we present constitutive equations and computational algorithms for rate-
dependent plasticity or viscoplasticity. The theory allows the modeling of
material behavior presenting two features: a permanent (irreversible) defor-
mation and a sensitivity of the stress-strain response to the rate of straining
(Fig. 13.1). Visco-plasticity or creep in metals becomes important when the
absolute temperature exceeds one third of the absolute melting temperature.

13.1 One-dimensional model

t3 = const > t2
t2 = const > t1
_ _- - t1 = const

Fig. 13.1. Viscoplasticity: stress-strain curves for various strain rates

Let us illustrate an elasto-visco-plastic model in a simple uniaxial creep


experiment, where the imposed stress is held constant (after Owen and Hin-
ton(1980)):

a(t) = constant = (j (13.1)

We consider a linear viscoplastic model defined by:

(13.2)
330 13. Elasto-viscoplasticity

where U v is the viscous stress, 7J a viscosity coefficient, and i P the plastic


strain rate.
We assume that the total stress 7J is the sum of the viscous stress u v , the
initial yield stress Uy and a hardening stress R(tP) (which is a function of t P
because we only consider monotonie loading in this example),

(13.3)

Combining Eqs. (13.2-3) gives the following differential equation for tP

f
.p
= Ci - R(tP) , fP(O) = O (...
Uy -
lmtla1 cond··)
ltlOn (13.4)
7J
The total strain f is the sum of an elastic strain and the permanent strain fP,

(13.5)

where E is Young's modulus. We shall consider two models for isotropic hard-
ening. First we assume that there is no hardening: R = o. Time integration
of Eq. (13.4) gives the total strain t(t) as:
Ci Ci - Uy
t(t) = - + t (13.6)
E 7J
This is plotted in Fig. 13.2a which shows that t increases with time as one
expects in a creep experiment, but it does sa linearly, which is unrealistic.

t
Fig. 13.2. Simulation of a uniaxial creep experiment: (a) No hardening, (b) Linear
hardening

A better simulation is obtained if we consider linear isotropic harden-


ing: R(fP) = kf P, where k is a hardening modulus. In this case, differential
equation (13.4) becomes:
k Ci - Uy
i P + -t P = , tP(O) = O (initial condition) (13.7)
7J 7J
13.2 Three-dimensional model 331

Time integration of Eq. (13.7) gives the total strain f(t) as:

f(t) = -(j + (j-oy [ k ]


1 - exp( --t) (13.8)
E k 'f/
This is plotted in Fig. 13.2b which shows that 10 increases with time but -in
agreement with experimental observations- it saturates after a long period of
time. For linear isotropic hardening (and linear viscoplasticity) the saturation
strain is given by:
a (j - ay
t -+ 00 ==} f(t) -+ E + k (13.9)

Linear viscoplasticity -Eq. (13.2)- corresponds to the particular case (m = 1)


of the more general Norton's power law creep:
101> = ay (aV)m (13.10)
'f/ ay

13.2 Three-dimensional model


In this section, we shall see how the ideas which we discussed in the uniaxial
case are generalized to multi-axial stress states. The formulation that will be
presented is often named after Perzyna (1963).
The total strain € is assumed to be the sum of two parts: an elastic strain
€e and a plastic (inelastic, permanent or irreversible) strain €P:

€ = € e+ €,p 'l.e., fii = fiie+ fiip (13.11)


The Cauchy stress u and the elastic strain €e are related with a linear elastic
constitutive model:

(13.12)
A yield function f (u, R) is defined by:

f(u,R) == a eq - ay - R(P) (13.13)

where ay is the initial yield stress, a eq the von Mises equivalent stress, R(P)
the hardening stress and p the accumulated plastic strain (the definitions are
identical to those already given in Chap. 12). When the transformation is
elastic, then f(u, R) ::; O, but if there is plastic loading, then to the contrary
to rate-independent plasticity, f (u, R) may be positive. For a uniaxial stress
state, f(u,R) is simply the viscous stress av ofEq. (13.3).
The plastic ftow rule governs the evolution of the plastic strain:
of . of
'p
€ = "1-,
.

ou
.
l.e.,
.p
fi}' = "I~
vaii
(13.14)
332 13. Elasto-viscoplasticity

The scalar t is called the plastic multiplier, we have t ~ O and the precise
sign is determined by the following conditions:

t=O if f :::; O (elasticity),


t = 9,,(1) >O if f > O (plasticity), (13.15)

where 9,,(1) is a viscoplastic or creep functionj it is a monotonically increasing


function of f > O aud satisfies the condition:

9,,(1) = O {:=} f :::; O. (13.16)

An example of 9,,(1) -probably the most used one- is Norton's power law:

(13.17)

For a uniaxial stress state, Eqs. (13.15b, 17) simply become (13.10). Note that
to the contrary to rate-independent plasticity, the plastic multiplier i' is not
found from a consistency condition -which does not hold in viscoplasticity-
but is computed from Eq. (13.15b). The reader can easily ebeck that because
ofthe use of a von Mises criterion, we have p = t (as in elasto-plasticity). If
we combine the equations and write the plastic flow as:

(13.18)

we see that time appears explicitly in the constitutive equations, that is why
the material model is called rate-dependent. If we take the time derivative of
the stress-elastic strain relation (13.12) and replace the plastic strain rate by
its expression (13.18), we find that:

ir = E: (€ - 9,,(1) ~!) (13.19)

Therefore we see that when there is plastic loading, it is not possible to


find a tangent operator reIat ing the rates (or the infinitesimal increments)
of stress aud strain. In elasto-visco-plasticity, there is no counterpart of the
"continuum" tangent operator H of elasto-plasticity (Sec. 12.7).

13.3 Numerical algorithms

We shall see hereafter that the numerical algorithms for elasto-visco-plasticity


are very similar to those which we presented for elasto-plasticity in Sec. 12.10.
Therefore, we shall proceed at a quick pace aud insist only on those things
which are different from the rate-independent case. We will also use the same
notations unless otherwise indicated.
13.3 Numerical algorithms 333

13.3.1 Return mapping algorithm

The elastic predictor phase is identical to that for elasto-plasticity. IT f!r;.~' >
O, then we must enter the plastic corrector phase. As in elasto-plasticity, we
arrive to a radial return algorithm defined by Eqs. (12.75) which are:

N = Ntrial., 2G.1p + ~a
3 eq = ~atrial
3 eq (13.20)

Now the diJJerence with elasto-plasticity is that although we define a yield


function:

f == a eq - ay - R(P) (13.21)

we cannot replace a eq with [ay + R(P)] because f > O. We use however a


discretized form of Eq. (13.15b) -corresponding to a fully implicit integrat ion-
to write .1p as:

.1p = gv(f).1t (13.22)


where .1t is the time increment. Therefore, we have two scalar equations
which must be solved for the two unknowns p and a eq :

a eq + 3G.1p - atrial
eq = O (13.23)

The two equations are solved iteratively using Newton's method. At each
iteration (we omit the iteration index for simplicity of notation) one finds
the corrections cp and Cu from:

(13.24)

IT we substitute the expression of Cu from the second equation into the first,
we obtain:
_ 1 (k p ) • _ 1 dRdp
------
cp - - hv (dg v /df).1t + k u , h v = (dg v /df).1t + 3G + (13.25)

Note the relation between hv and h, the denominator of p, cp and op in elasto-


plasticity (Secs. 12.7, 12.10.2 and 12.10.3).

13.3.2 Consistent tangent operator

When there is plastic loading, we have seen in Sec. 13.2 that it is not possible
to find a tangent operator relating the rates (or the infinitesimal increments)
334 13. Elasto-viscoplasticity

of stress and strain. However, it is possible to compute a consistent or algo-


rithmic operator defined as in elasto-plasticity (Sec. 12.10.3) by:

(13.26)

Again, the subscript (n + 1) is omitted for simplicity and alI variables which
do not contain this subscript are evaluated at tnH' Differentiation of the
stress-elastic strain relation (12.63), the discretized plastic flow rule (12.73),
and the discretized creep law (13.22) gives the following three equations:

8u = E : (8f: - 8f:P)
âN
8f:P = N8p + (Llp) âu : 8u

dg v dR )
8p = dJ
(
N : 8u - dp 8p Llt (13.27)

Rewriting the last equation using the first two leads to:

(13.28)

where the denominator h v was defined in Eq. (13.25b). Expression (13.28)


is almost identical to that for elasto-plasticity -Eq. (12.85)- except that the
denominator here is hv instead of h.
Exercise: eheck that alI the remaining steps in the derivation of Halg
are exactly identical to that for elasto-plasticity, Eqs. (12.86) to (12.92). For
isotropic elasticity, we find at the end that:

(13.29)

The expression of Hv is almost identical to that of H (the "continuum"


tangent in elasto-plasticity) except that we have hv instead of h, but the two
tensors do not have the same meaning since a "continuum"tangent operator
does not exist in elasto-visco-plasticity. Limiting cases are an exception. For
example, for linear creep (m = 1), we have

hv =1t+ h (13.30)

and therefore (Llt being finite and positive):

(13.31)

i.e., in the limit of vanishing viscosity, we recover the elasto-plastic results.


13.4 Further reading 335

13.4 Further reading

This chapter is much shorter that the previous one because it uses many
concepts that were introduced there. Besides the correlations we pointed to
in Secs. 13.2 and 13.3, let us also recall that all the results of Sec. 12.11 and
its subsections -except 12.11.4- apply to visco-plasticity without any modifi-
cat ion.
For more informat ion about the physical aspects and the modeling of visco-
plasticity, see (Ashby and Jones, 1980), (Hertzberg, 1989), (Lemaitre and
Chaboche, 1990), (Fran~ois et al., 1993), and references therein.
In this chapter, we only dealt with Perzyna-type visco-plasticity models. An-
other formulation, without a yield function, called the endochronic theory
was proposed by Valanis (1971).
14. Nonlinear continuum mechanics

In alI previous chapters, we have worked within the small perturbation hy-
pothesis (SPH). As a consequence, we wrote (and solved for) equilibrium and
boundary condition equations on the initial, undeformed (thus known) con-
figurat ion of a body. The only exception was the study of possible buckling
modes in Chap. 10. In this chapter, we present basic continuum mechanics
in the presence of geometric non-linearities, namely finite strains, displace-
ments and rotations. Practical examples are metal forming problems or large
displacements of slender beams and thin shells.

14.1 Kinematics

We consider a body which occupies a configurat ion -called reference- Do C ]R3


at time t = O and a configuration -called current- Dt C ]R3 at time t > O (see
Fig. 14.1).

el
Fig. 14.1. Position vectors of a material partide are X in the reference configura-
tion (ilo) and x = rp(X, t) in the current configurat ion (ilt ), in a fixed Cartesian
orthonormal frame (O, el, e2, e3).

A given material partide is specified by its position vectors:

x = {XI}, 1 = 1,2,3 in the reference configuration,


X = {xd, i = 1,2,3 in the current configuration. (14.1)
338 14. Nonlinear continuum mechanics

For simplicity, we consider that all coordinates are w.r.t. a fixed Cartesian
system with origin O and orthonormal basis (el, e2, e3). We shall however
use upper case (e.g., I, J, K) and lower case (e.g., i,j, k) indices for the coor-
dinates of a partide in the reference and current configurations, respectively.
This will make it easier to identify referential -or Lagrangian or material-
objects on the one hand and current -or Eulerian or spatial- objects on the
other hand. For a more general presentation using curvilinear coordinates,
see (Marsden and Hughes, 1983), (Sedov, 1975), (Green and Zerna, 1968).
In the following sections, we will designate by 11:0 C Do an arbitrary
volume of material partides with boundary âll:o. That set of material partides
maps onto II:t C Dt, with boundary âll:t .

14.1.1 Description of mot ion

The motion of the body is described by a transformation c/J defined by:


(14.2)

The material velocity of a partide is defined by:

(14.3)

This means that we follow a given material partide (X fixed) and compute
the variat ion of its position over an infinitesimally small time increment. If
the velocity is viewed as a function of (x, t) then it is called spatial velocity:
(14.4)

14.1.2 Material time derivative


In general, if K(X, t) = k(x, t), Le. the same material quantity is expressed
as a function of (X, t) or (x, t), we have:
. â âk âk âx âk dk .
K := -K(X t) = - + - .- = - + (\7k)· v := - := k (14.5)
ât ' ât âx ât ât dt
By definition (dkjdt) is called the material time derivative of k, i.e. it is the
derivative of k w.r.t. time for X fixed (we follow a given material partide);
it will usually be denoted k for conciseness. On the other hand (âkjât) is
the derivative of k w.r.t. time for x fixed (we watch a given spatial -i.e. in
Dt - position). As an example, the material time derivative of the velocity is:
â âv dv .
âtV(X,t) = ât +(\7v)·v:= dt:=v (14.6)

This defines the acceleration:

r(X,t) = -y(x,t) = v (14.7)


14.2 Deformation 339

14.2 Deformation
In the following subsections and in Sec. 14.3, we give a mainly mathematical
description of the deformation. For a geometric picture and interpretation,
see (Novozhilov, 1953), for instance.

14.2.1 Defonnation gradient

The deformation gradient is defined by:

F - ârjJ . âXi
= âX' 1.e., F iJ = âXJ (14.8)

The component form shows that F has a "leg" in the current configurat ion
and a "leg" in the reference configuration. Since x = X + u where u is the
displacement, the deformation gradient is also given by:
âu
F= 1+ âX (14.9)

Consider two infinitesimally close material particles at positions (X) and


(X + dX) in the reference configuration. The positions of those particles are
(x) and (x + dx) in the current configuration. By definition we have:

(14.10)

This definit ion of F is equivalent to the previous one for Cartesian coor-
dinates, but it is more general because it applies to curvilinear coordinate
systems as well. It is very useful in practice. For computation, we can rep-
resent F with a 3 x 3 matrix (non-symmetric, in general). We introduce the
following notation:

J == detF (14.11)

We shall see that J represents the local compressibility.


The elementary volume in the reference configurat ion is given by the
mixed product formula:

(14.12)

Similarly, in the current configurat ion, we have (see Fig. 14.2):

dv (dxI, dX2, dX3)


(F ·dX 1 , F· dX 2, F· dX 3 )
(det F) (dXl,dX 2,dX3)
~' '" '
J dV

We then have:
340 14. Nonlinear continuum mechanics

dX 3

dX I
Fig. 14.2. Deformation of an elementary parallelepiped. Initial volume: dV
dXt . (dX2 x dX3). Volume after deformation: dv = dXt . (dX2 X dX3).

J- dv J>O (14.13)
- dV'
J must be positive in order to have positive volumes. Note that if a material
is incompressible, then J = 1. Since det F > O, F has an inverse, so that:

dX = F- I . dz, Le. dX/ = (F-I }Ijdxj (14.14)

14.2.2 Polar decomposition

It can be shown that F can be uniquely decomposed in the following way


(e.g., (Gurtin, 1981)):

(14.15)

This is called the polar decomposition. R is proper orthogonal and is called ro-
tation. U and V are symmetric and positive definite; they are called the right
and lefi stretch tensors, respectively (the justification of the word "stretch"
will appear later). U and V have the same eigenvalues, and those are real
and positive. The eigenvalues are called the principal stretches (Al, A2, A3)' A
deformat ion can be defined locally with a rotation followed by stretches along
three mutually orthogonal axes (principal directions), or vice-versa. The devi-
ation of AA from 1 measures a strain (see Sec. 14.3). Using the decomposition
of F, we obtain:

FT .F = U T . R T . R·U = U 2 == e, Le. F/kFkJ == C/J


'-v-"
1

F·F T = V·R·R T ·VT =V2 ==b,


'-v-"
Le.FiKFKj==bij (14.16)
1

where e and b are the right and lefi Cauchy-Green strain tensors, respec-
tively. They will allow the definition of strain measures (see Sec. 14.3). The
14.2 Deformation 341

component forms clearly show that C is defined w.r.t. the reference configu-
ration while b is w.r.t. the current configuration.
In general we try to use upper case characters (e.g., C) for Lagrangian
objects and lower case characters (e.g., b) for Eulerian objects. This will not
always be possible however, for instance the use of an upper case character
for F is an arbitrary choice.
In practice, we first compute the eigenvalues of C or b (see Appendix
B), we thus obtain the squares A~ (A = 1,2,3) of the principal stretches.
Let N(A) be the eigenvector (principal direction) associated with AA. By
definition:

C· N(A) = A~N(A) (no sum), (14.17)

The principal directions form an orthonormal basis:

N(A) . N(B) = 0AB (Kronecker's delta) (14.18)

14.2.3 Spectral decompositions

Decomposition of dX in the basis (N(I), N(2), N(3» gives:


3
dX= L(lAN(A) (14.19)
A=I

Therefore:
3
(N(I) ® N(I» . dX = N(I) L (lA~ = (lIN(I) (14.20)
A=I ,hA

Similarly, we have:
(N(2) ® N(2». dX = (l2 N (2) and (N(3) ® N(3». dX = (l3 N (3)
(14.21)

Note that, substituting these results in the expression of dX gives the fol-
lowing identity:
3
LN(A) ®N(A) = 1 (14.22)
A=I

We have:
3 3

dx = F·dX = L (lA~= L(lA~ (14.23)


A=l A=l
342 14. Nonlinear continuum mechanics

We shall show now that (n(1),n(2),n(3») that wejust defined is an orthonor-


mal hasis. The proof is easier if we switch to matrix notation (with n(A),s
and N(A),s stored as 3 x 1 arrays):

....!....{N(A)}T [F]T[F]{N(B)}....!....
AA '--v-" AB
[C]
_1_{N(A)}T A1{N(B)}
AAAB
= dAB (14.24)

Using (14.22), we can write:

dx = F· (t, N(A) ® N(A») . dX

(t,AAn(A)®N(A») ·dX (14.25)


, .... .,
F
Therefore, we have defined the spectral decomposition of F. In matrix nota-
tion, it reads:

=L
3
[F] AA{n(A)}{N(A)}T (14.26)
A=l
Since (n(l), n(2), n(3») is an orthonormal hasis, the right Cauchy-Green strain
C = F T . Fis found from (14.25) as follows:

L
3
C = A~N(A) ® N(A), (14.27)
A=l
which is expected since A~ (A = 1,2,3) are the eigenvalues of C and N(A)
its principal directions.
Since (N(l), N(2), N(3») is an orthonormal hasis, the left Cauchy-Green
strain b = F· F T is given hy (14.25) as follows:

L
3
b= A~n(A) ®n(A), (14.28)
A=l
which is expected since A~ are the eigenvalues of b and n(A) its principal
directions.
Since (N(l), N(2), N(3») is an orthonormal hasis, we deduce from the
polar decomposition F = R . U, the spectral decompositions of R and U:
14.2 Deformation 343

F ~ (ţ, n IA) ® NIA)) .


, ..... ", ,
Ct ),BNIB) ® NIB) )
'V' .,
(14.29)

R U
Similarly, from the polar decomposition F =V . R, we obtain:

(14.30)

It can easily be checked that Ris a rotation (RT . R = R· R T = 1).

14.2.4 Length variat ion

Let infinitesimal lengths be represented by:

dL = VdX . dX and d1 = Vda: . da: (14.31)

in the reference and current configurations, respectively. Using Eqs. (14.10,


17), and switching to matrix notation for convenience, we obtain:

(14.32)

Therefore, the length ratio is given by:

~=JdX'C'dX (14.33)
dL dX ·dX
Using (14.17, 19),

dX . C . dX = (ţ, aAN(A») . (t, aB)'~N(B)) = ţ, (aA),A)2

The length ratio can then be written as:

dl
(14.34)
dL

If we set a2 = a3 = O, we find that ),1 = d1 / dL (since),l > O). Therefore, ),1 is


the length ratio (Le., stretch) for a vector which was initially in the direction
of N(l). We obtain similar definitions for ),2 and ),3. Looking at (14.29-30),
the denomination of "stretch" tensors for U and V becomes clear.
344 14. Nonlinear continuum mechanics

14.3 Strain ~easures

14.3.1 One-dimensional case

We have just seen -Eq. (14.34)- that in 1D, the length ratio -or stretch- is given
by A = dl / dL. The unstrained state corresponds to A = 1. For injinitesimal
strains, the strain measure is given by:

dl-dL =A-1 (14.35)


dL
In this case, the unstrained state corresponds to zero strain. This is a property
that we would like to have in the finite strain case. Let us define a strain
measure by:

E = f(A) (14.36)

We require the following properties of f(A):

f( l) = O' df (1) = l' df > O VA > O (14.37)


, dA ' dA
We have already discussed the first property. The second property means
that for infinitesimal strains, we would like all strain measures to give the
same values -to the first order- as the infinitesimal strain measure (A - 1).
This is apparent if we write a Taylor expansion of f(A) around A = 1:
2f
f(A) = !2l +(A - 1) dAdf (1) + (A-1)2d
2! dA 2 (1) + ... (14.38)
o
The third property in (14.37) means that we would like the strain f(A) to
increase monotonically with the stretch A, for alI physically acceptable values
of A. With the above restrictions, we may define as many strain measures as
we wish. Examples are:
df
Nominal (or Biot's) strain: f(A) = A-l ~ - = 1
dA
1 2 df
Green's strain: f(A) = -(A - 1) ~ - = A
2 dA
df 1
Logarithmic strain: f(A) = In A ~ dA = ~ (14.39)

The three strain measures are plotted in Fig. 14.3

14.3.2 Three-dimensional case

In the 3D case, we may proceed as follows. By analogy with Eq. (14.27), we


may define a strain measure w.r.t. the reference configurat ion by:
14.3 Strain measures 345

12 ,
,,,
10
,
...
Q) 8 Green ----- ,/
;::1
6
Nominal - - "
gj Logarithmic ........... /
Q)

Ei 4
I'l ............ ,. ... ,.,.,.
•<il 2
..,... ":: .................. .
rn O
-2
-4~~~~~~~~~~--~

O 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5


Stretch (A)
Fig. 14.3. Three strain measures versus principal stretch.

(14.40)

where the function f(>.) was discussed in the 1D case.

It Green-Lagrange and Euler-Almansi strains


The Green-Lagrange strain is generalized from 1D to 3D as follows:
3
ea = ~ l)>'~ - l)N(A) ® N(A) (14.41)
A=l

Using Eqs. (14.22, 27), we find that:


1
ea = -(C
2
-1) (14.42)

Actually, this is how ea is computed, without having to compute the principal


stretches or the principal directions. Using Eq. (14.9), we have:

C =FT . F = [1 + (ax
au f] .(1 + ax
au) (14.43)

Therefore:

(14.44)

It appears then that ea is made up of two terms. The first one is the in-
jinitesimal strain tensor, which is valid for small strains and rotations. The
346 14. Nonlinear continuum mechanics

second term represents the nonlinearity due to finite strains. The presence
of Lagrange's name is due to the fact that EG measures a variat ion of length
w.r.t. the reference configuration. This can be seen from the following. We
have already established formulae for the squares of infinitesimallengths:

(14.45)

Therefore:
~ (dl)2 - (dL)2 = dX· EG . dX (14.46)
2 (dL)2 dX· dX
We may also define a variation of length w.r.t. the current configurat ion.
First, we switch to matrix notations to establish the following result:

(14.47)

where b = F . F T is the left Cauchy-Green strain. Therefore:


1 (dl)2 - (dL)2
(14.48)
2 (dl)2

Where EA is the so-called Almansi-Euler strain:

(14.49)

le Nominal (or Biot's) strain


The nominal (or Biot's) strain is generalized from 1D to 3D as follows:

=L
3
EB (AA - l)N(A) l8l N(A) (14.50)
A=l

Using the spectral decompositions of U (right stretch tensor) and 1, we find:

(14.51)

We have given a few examples of strain measures, for more strain tensors
see (Ogden, 1984). The choice of a particular strain measure depends on the
stress measure (see section 14.6).

14.4 Strain rates


Several constitutive models (e.g., plasticity models) use strain "rates" which
we need to define. Consider two material particles which are at positions (X)
and (X + dX) in the reference configurat ion, and (a:) and (a: + da:) in the
current configuration. The difference between their spatial velocities is:
14.4 Strain rates 347

âv
dv = âx . dx == l . dx = l . F . dX (14.52)

where l is the velocity gradient (it is a spatial object -components with lower
case indices):

âv . l âVi
l == âx' l.e. ij ==-â (14.53)
Xj

Computing the material time derivative of the deformation gradient, we find:

(14.54)

Therefore:

F= l . F, Le., l = F· F- 1 (14.55)

We can split l into symmetric and antisymmetric parts:

(14.56)

The symmetric part of l is the rate of deformation tensor d (dij = d ji ):

d= '12 [âV
âx
âv
+ (âx)
T]'
, l.e. d ij = '12 (âVi âVj)
âXj + âXi (14.57)

Note that dis identical in form to the infinitesimal strain tensor (the displace-
ment u is replaced with the spatial velocity v). We will see later that it is d
which appears in the weak form of the equations of motion (or virtual power
theorem). The anti-symmetric part of l is the spin tensor W (Wij = -Wji):

W = '12 [âV
âx -
âv
(âx)
T]'
, l.e. Wij = '12 (âVi
âXj
âVj)
- âXi (14.58)

Derivation W.r.t. time of the right Cauchy-Green strain gives:

C F T ·F+FT·F
F T . lT . F + F T . l . F
F T . (lT + l) . F,

Le.,

(14.59)

Note that the rate of Green-Lagrange strain is simply:


348 14. Nonlinear continuum mechanics

.0
E =-
C (14.60)
2
Using Eqs. (14.46, 59), and switching to matrix notations for convenience,
we have:

~{dX}T[C]{dX}
{dX} T[Ff[d][Fl{ dX}
= {dx}T[d]{dx}
Therefore:

{dxV [d]{dx} = ~~(dl) = ~ (ln~) (14.61)


dl dl dl dt dt dL
This shows that d measures the rate of change of a length (dl) or a stretch
(dl/dL). In general, it is not possible to write d as the time derivative of
some strain tensor, however, the last result shows that d is "linked" to the
time derivative of the logarithmic strain.

14.5 Balance laws


In this section, we will develop three balance laws, namely the conservation of
mass, linear momentum and of rotational momentum (these last two balance
laws generalize Newton's laws of dynamics to continuum media). We will find
it very useful to establish the so-called transport formula first.

14.5.1 Transport formula

We shall prove that the following result holds:

: [ pk(x, t) dv =[ [pk + k(p + pdiv v)] dv (14.62)


t J~t J~t
where p(x, t) is the current mass density [kg/m 3 ] and v the spatial velocity
field. Note that the integration domain K.t depends on time. Computing the
derivative will be easier if we transform the integral into an integral over the
reference -fixed- domain K.o. Using the result J = dv/dV, we have:

dl
d
t ~o
pkJ dV = 1 ~o
. .
(pkJ + pkJ + pkJ) dV (14.63)

Since J = det F (where F is the deformation gradient), we have:


. âJ·
J= âFiK FiK (14.64)
14.5 Balance laws 349

It can be shown (see for example (Ogden, 1984)) that:

aJ aJ
aFiK = J (-1)
F Ki,
.
l.e. aF = JF -T (14.65)

Therefore, using Eqs. (14.55-56), we obtain:

~(ln
dt
J) = :!.. = trl = tr d = div v
J
(14.66)

Using these results, the right-hand side of (14.63) can be rewritten as:

{ (pk + pk + pkdiv v)J dV (14.67)


}I<O
Rearranging terms and using J = dvjdV again, we find (14.62), Le. the de-
sired formula. The transport formula will be useful for deriving alternative
expressions for the balance laws and also for the laws of thermodynamics. Ac-
tually, we shall establish that the balan ce of mass implies that the transport
formula takes the very simple form:

:1 t 1<,
pk(x,t) dv = 1 1<,
pk dv (14.68)

For expressions of the transport formula in the presence of discontinuity


surfaces, see (Duvaut, 1990) for example.

14.5.2 Conservation of mass

IT the mass ofthe body is conserved throughout the motion, then the following
equality must hold:

ddt 1 1<,
p dv = o, "f/r;,t C fl t (14.69)

Using the transport formula (14.62) with k = 1, we obtain:

1 1<,
(p + pdiv v) dv = o, "f/r;,t C fl t

Since the equality must hold "f/r;,t C fl t , the localization theorem-e.g., (Gurtin,
1981)- gives the following local form of mass conservation:

p+ pdiv v =O (14.70)

We shall often use the localizat ion theorem, even if we do not mention it.
As a consequence of (14.70), transport formula (14.62) takes the simple
form (14.68).
350 14. Nonlinear continuum mechanics

Note that if we write the mass conservat ion condition as:

r pdv = r po dV, (14.71)


J Kt 1"'0
we find that the mass densities p and po in the current and reference config-
urations, respectively, are related by:

po = J p, Le., J = po = dv (14.72)
p dV

For an isochoric deformation, J = 1, and therefore p = po = constant, and


consequently: div v = O = tr l = tr d.

14.5.3 Conservation of linear momentum

=: r
The balance of linear momentum is written as:

r
lI<,
pb dv + r
laI<,
t da
t lI<,
pv dv, V"'t C [lt (14.73)

where b(x, t) is the field of body forces -Le. forces per unit mass, [Njkg]-
act ing in "'t (pb is a force per unit volume, [Njm 3 ]), t(x, t, n) is the field of
forces per unit area -[Njm 2 ]- applied on â"'t, and v is the spatial velocity
field; n designates the outward unit normal to â"'t. Using transport formula
(14.68) with k = v, it is found that:

:t lI<,r pv dv = r
lI<,
P'Y dv (14.74)

where'Y == il is the acceleration field.

14.5.4 Conservation of rotational momentum

The balance of rotational momentum written w.r.t. a fixed point taken here
as the origin of the coordinate system reads:

r
lI<,
a: x pb dv + r
laI<,
x x t da =: rt lI<,
px x v dv, V"'t C [lt (14.75)

Again, using transport formula (14.68) with k = x x v, it is found that:

:t jKtr pa: x v dv = r pa: x 'Y dv


J"'t
(14.76)
14.5 Balance laws 351

14.5.5 Cauchy stress tensor


Consider a point x on O"'t and let the unit vector n designate the outward
normal to O"'t at x. The Cauchy (or "true") stress u(x, t) is a second order
tensor which is defined with respect to the current configurat ion by:
t(x, t, n) = u T (x, t) . n, Le., ti = Ojinj (14.77)
For justification of Eq. (14.77) and detailed comments about it, see for in-
stance (Duvaut, 1990) and references therein. In a uniaxial tension test on a
cylindrical bar, the Cauchy stress at time t is equal to the current value of
the applied force divided by the current value of the cross section area. Using
the definit ion of the Cauchy stress, we have:

r
181<,
t da ru
181<,
T . n da

11<,
r div u T dv (by the divergence theorem) (14.78)

14.5.6 Eulerian strong formulat ion


Substituting result (14.78) into the balance of linear momentum, we obtain
the following local equation, called the first law of motion:
·
d IV T . OUji
U + pb = Pl, l.e., ~ + pbi = P'Yi (14.79)
uXj

It can be checked that the cross product of two vectors V and W can be
written as:
(V X W)i = fijkVjWk , (14.80)
where fijk is a circular permutation of (1,2,3) defined by:
f123 f231 = f312 = 1
f132 f213 = f321 = -1
fijk O in aU other cases
We now use Eqs. (14.77, 80) to rewrite a surface integral which appears in
Eq. (14.75):

r
181<,
fijkXj(Ulknt) da

r
11<,
fijk !la
UXI
(XjUlk) dv (divergence theorem)
352 14. Nonlinear continuum mechanics

The balance of rotational momentum can now be written as:

r EijkXjpbk dv + J~r Eijk [ajk + Xjp('Yk -


h.
bk)] dv = r PEijkXj'Yk dv
h.
(14.81)
This simplifies to:

(14.82)

Using the definition of the circular permutation, it is easy to check that Eq.
(14.82) implies the following local equations:

(14.83)

Le., the Cauchy stress is symmetric (aij = aji). In summary, we have the
following local -or strong- equations:

p + pdiv v = O in K.t (balance of mass)


(TT • n =t on âK.t (force boundary condition)
div (TT + pb = P'Y in K.t (equation of motion)
(T = (TT (symmetry of Cauchy stress) (14.84)

In component form in Cartesian coordinates, the previous equations become:

.
p+ p -
âVi
âXi
= O.In K.t

ajinj = ti on âK.t

âaji
-- + pbi = P'Yi in K.t
âXj
(14.85)

It is dear that alI the equations are written w.r.t. the current configuration
(recall that we use lower case indices for the coordinates of a partide in the
current configuration)j they are called Eulerian field equations. We shall see
in Sec. 14.5.11 how to write the Lagrangian field equations.

14.5.7 Eulerian weak formulation


As we shall see later, it is very useful to develop the so-called weak (or global)
forms of the equations of motion, w.r.t. the current and reference configura-
tions, respectively. The technique which will be used here is identical to the
one we have already presented while developing the virtual work theorem in
Chap. 2. Let w be a sufficiently smooth virtual velocity field. We take the dot
product of the equations of motion (14.84c) with w and integrate to obtain:
14.5 Balance laws 353

r (div u
J~
T + pb) . w dv = r
J~
P'Y' w dv (14.86)

It is easy to check the following derivations:


8 8aji 8Wi
-8 (ajiwi)
Xj
--Wi
8xj
+ aji-
8xj
~

8a"
8 Xj]'Wi + ____
aijdij(w) (u being symmetric)

where d( w) is the rate of deformation corresponding to w, Le.


_ 1 (8Wi 8wj)
dij (W ) - - -+-- (14.87)
2 8xj 8Xi
Now collecting the results and using the divergence theorem, we can rewrite
(14.86) as:

r (u
Jâ~t
T . n) . w da - r u: d(w) dv + jKtr
jKt
pb· w dv = r
Jltt
P'Y' w dv

(14.88)
Using (14.84b), Eq. (14.88) can be rewritten as:

J
r u: d(w) dv + Jr p,' w dv = Jr pb· w dv + leKt
Kt Kt
r t· w da
Kt
(14.89)

This is known as the weak form of the equations of motion, or the virtual
power theorem (V.P.T.)

14.5.8 Balance of work and energy rates


Consider the case w = v, where v is the actual spatial velocity field at time
t. Using transport formula (14.68), it is easy to show that:

dt J1<, 2
r
~pv. v dv =
!!:.-
JI<'
P'Y' v dv r (14.90)

Therefore, Eq. (14.89) written for w = v becomes:

r u:d(v)dv+!!..dt JI<'r ~pv.vdv=


JI<' 2
r pb·vdv+ J{)I<'
JI<'
r t·vda (14.91)

The right-hand side of Eq. (14.91) represents the power due to the external
loads, the left-hand side is comprised of two terms. Up to a sign, the term

r u: d(v) dv (14.92)
JI<'
represents the power of internat loads, or stress power. The term

!!.. r
~ pv . v dv
dt JI<' 2
(14.93)

is the rate of change of the kinetic energy.


354 14. Nonlinear continuum mechanics

14.5.9 Nominal stress

Consider a uniaxial tension test on a circular cylindrical bar. The current


value of the tension force is F, while the cross section area has the values Ao
and A in the initial and deformed configurations, respectively. The Cauchy
(o") and nominal (P) stresses are given by:

F F
u= Ai P= Aoi F=uA=PAo (14.94)

We now generalize these results to multi-axial stress states. Consider a surface


element of area da and outward unit normal n on 8K.t, and area dA and
outward unit normal N on 8K.o. According to Nanson's formula (e.g., see
(Ogden, 1984)), we have:

(14.95)

We use Eq. (14.95) to find a new expression for the traction force:

tda = (u T . n)da = (Ju T . F- T ) ·NdA,


'----"
=p
Where P is called the first Piola-Kirchhoff stress and

pT =JF- 1 . u, (14.96)

the nominal stress. The component form:

PiK =J(p-l )KjUji (14.97)

clearly shows that P is not symmetric. Actually, since PiK has a "leg" in
the current configuration (lower case index i) and the other in the reference
configuration (upper case index K), according to Marsden and Hughes (1983)
we should not even talk about symmetry or the absence of it.
Let the traction per unit area be designated by t(x, t, N) and T(X, t, N)
in the current and reference configurations, respectively. We have:

(u T . n) da = (p. N)dA, i.e. (Ujinj) da = (PiKNK) dA (14.98)


~ ~ ~' ... '
t T ~ n
Note that vectors (u T . n) and (p. N) have the same direction. Equation
(14.98) is the extension of (14.94) from 1D to 3D.

14.5.10 Lagrangian weak formulation

We shall see that it is the nominal stress which appears in the equations of
motion written w.r.t. the reference configurat ion. There are more than one
14.5 Balance laws 355

way to obtain those equations, the simplest is to write the balance of linear
momentum. We use another method which also gives the strong form of the
equations of motion. We have:

C7: d(w)dv =

where GRADw (written with upper case characters) is the gradient of w


w.r.t. the reference coordinates. Recalling that J = dvldV = Pol P and tda =
TdA, Eq. (14.89) is rewritten to give the Lagrangian weak formulation:

r pT: (GRADw)
Jlto
dV + r po(X)r· w
}lto
dV =
J
r Po(X)B· w
KO
dV

+ r
181<0
T· w dA (14.99)

where r(X, t) = ')'(x, t) is the accelerat ion field and B(X, t) = b(x, t) is the
body force (per unit mass) field.

14.5.11 Lagrangian strong formulation

In order to find the Lagrangian strong formulation, we need to rewrite the


following integral:

r
18K.o
T· w dA = r
18K.o
(p. N) . w dA

r
181<0
PiKNKWi dA

10 8~K (PiKWi) dV (by the divergence theorem)

10 (PiK : ; : + ~~; Wi) dV

10 (pT : GRADw + DIV p . w) dV

With this last result, the weak form of the equations of motion becomes:
356 14. Nonlinear continuum mechanics

r por. w dV = r poB· w dV + r DIVp· w dV (14.100)


J~ J~ J~
Since the equality is valid for any "sufficiently smooth" field w and V"'o C Do,
the "localization theorem" gives the following local form of the equations of
motion in "'o:

DIVP + PaB = por, . âPiK


l.e., âXK + Pa B i = po
ri
i.e., FiKPjK = PiLFjL (14.101)

Again, DIV P (written with upper case characters) is the divergence of P


W.Lt. the reference coordinates. Equation (14.101b) is a consequence of the
symmetry of the Cauchy stress (o- = o-T). The traction boundary condition
on â",o reads:

(14.102)

Equations (14.101-102) are Lagrangian field equations.

14.6 Conjugate stress and strain measures

It is known that the stress-strain response of metals in the finite strain regime
is better described by using logarithmic strain and Cauchy (or Kirchhoff)
stress. Also, the Green-Lagrange strain (in association with the second Piola-
Kirchhoff stress) is useful for "structural" elements (beams, plates and shells)
which may undergo large displacements and rotations but small strains. In
this section, we will see a systematic method for the determination of conju-
gate strain and stress measures.

14.6.1 Definition and examples

Let us examine the expression of the stress power again, Eq. (14.92),

ro-: d(v) dv = r (Jo-): d(v) dV (14.103)


JKt JKO
The stress:

T =Jo- (14.104)

is called Kirchhoff stress. The term [T : d(v)] clearly represents a power per
unit volume of the reference configuration. If we could -in general, we cannot-
write d( v) as the time derivative of the logarithmic strain or another strain
tensor, we would say that T is conjugate with that strain tensOL Nevertheless,
14.6 Conjugate stress and strain measures 357

this is the idea behind defining pairs of conjugate stress and strain measures
(Hill, 1968): they are such that the inner product ofthe stress and the strain
rate is equal to [T : d(v)]. Two examples follow .

.. Second Piola-Kirchhoff stress and Green-Lagrange strain


Using Eqs. (14.59-60), we find that:

T: d Ju: (F- T .€G. F- 1 )

= JUij(F-T)jKlj<L(F-l)Li
= J(F- 1 ) LiUij (F-T)jKll<L
(JF- 1 . u· F- T ) : iP (14.105)
",
... .,
=8
where 8 is the second Piola-Kirchhoff stress. It is conjugate with Green-
Lagrange strain eG ; it is also associated with (C/2). The component form:

(14.106)

clearly shows that 8 is defined w.r.t. the reference configurat ion (upper case
indices) and is symmetric (SLK = SKL) .

.. Nominal stress and deformation gradient


We use Eqs. (14.55-56) to write:

T:d Ju : l (u being symmetric)


Ju: CI'· F- 1 )
. 1
JuijFjK(F- )Ki
1 .
J(F- )KiuijFjK
(JF- 1 . u) : F (14.107)
'-....--'
pT

where P is the first Piola-Kirchhoff stress. The nominal stress pT -Eq.


(14.96)- is conjugate with the deformation gradient F .

.. Relations between various stress measures.


The first and second Piola-Kirchoff stresses are related by:

8 = pT . F- T = F- 1 • P (since 8 is symmetric) (14.108)

And the Kirchhoff stress is related to those stresses by:

T = F· 8 . FT = p. F T = F . pT (14.109)
358 14. Nonlinear continuum mechanics

14.6.2 Interpretation of the second Piola-Kirchhoff stress

The second Piola-Kirchhoff stress S can be interpreted easily in the case of


small strains but arbitrary rotations, which is typical for "structures" (Le.,
beams, plates and shells). IT we write the principal stretches AA in the form:

(14.110)

then Eqs. (14.29,22) give the deformation gradient F as:


3
F = R· (1 +e), with e = L fAN(A) ® N(A) (14.111)
A=l

The second Piola-Kirchhoff stress S can now be written as:

S JF- 1 ·(T·F- T
J(1 + e)-l . R- 1 • (T. R- T . (1 + e)-T
= J(l + e)-l . R T . (T. R· (1 + e)-T (since R· R T = RT . R = 1)
For small strains, we have: fA « 1, and therefore:
(14.112)

and

J det Rdet(1 + e) ~ 1 + tr e
= det F = -........-.- (14.113)
1

The stress S has now the following approximate expression:

S ~ (1 + tr e)(1 - e) . R T . (T. R· (1 - e) (14.114)

If we keep the dominant term only, we obtain:

S~ R T ·(T·R (14.115)

This shows that S is approximately the rotated Ca'ILchy stressj this is only
valid for small strains but arbitrary (possibly large) rotations. As an example,
consider a beam whose axis (mid-fiber) was initially along the fixed direction
el j 8 11 will always designate the stress along the fiber, regardless of the
current orientat ion of the beam. For these reasons, S is ca1led material or
co-rotational stress (again this is only true for small strains).

14.6.3 Uniaxial tension/compression

In order to better apprehend the stress measures defined previously, we con-


sider the simple case of a circular cylindrical bar subjected to uniaxial tension
14.7 Objectivity 359

or compression along the el axisj the axial force is F. In the undeformed


state, the bar length and cross section area are lo and Ao, respectively. The
corresponding values in the deformed configuration are 1 and A. The bar is
assumed to be incompressible: Aolo = Al (for large deformations, metals and
rubber are nearly incompressible).
The Cauchy (true) or Kirchoff stresses are: a = T = F / A. The associated
strain is the logarithmic strain tL = ln(l/lo). As reported in (Ashby and
Jones, 1980), if we plot experimental uniaxial stress-strain data for metals,
only the choice (stress = a) and (strain = tL) leads to a curve which is sym-
metric w.r.t. the origin, in agreement with physical expectation that metals
respond identically to tension or compression under such experiments.
The nominal stress is P = F/Aoj its associated strain is the deformation
gradient: âx/âX = l/lo. True and nominal stresses are related by:
Ao 1
a = P- = P- = Pexp(td
A lo
As a consequence, we have:

-da =exp(td [P+-


dP]
dtL dtL
Necking corresponds to dP /dt L = O, Le. the nominal stress reaches a maxi-
mum. However, the true stress continues to increase: da/dtL > O at necking.
This is due to the fact that a is inversely proportional to the current cross
section area, which is decreasing.
The second Piola-Kirchhoff stress S is given by Eq. (14.105) as:

S = âX a âX = (lO)2a = (...:!.)2a =...:!.P


âx âx 1 Ao Ao
The Green-Lagrange strain ta is associated with S,

14.7 Objectivity
Consider a circular cylindrical bar under uniaxial tension in a fixed Cartesian
frame (O, el, e2, e3). The current values of the force and the cross section area
are designated by F and A, respectively. First the bar is parallel to the x-axisj
the only non-zero Cauchy stress component is an = F / A. Next, the bar is
given a rigid rotation so that it becomes parallel to the y-axis. Now a22 = F / A
is the only non-zero Cauchy stress. Assume the material is (hypo )elastic. If
we write the constitutive equation as: aij = Cijkld'k, then since an f:. O
(and a22 f:. O) non-zero strains would occur. This is not physically acceptable
360 14. Nonlinear continuum mechanics

because we have a rigid rotation. The problem is that the material time
derivative of the Cauchy stress (iT) is not objective. We should write the
constitutive equation with an objective stress rate so that a rigid body motion
does not induce any deformation. In this section as well as in Sec. 14.8, we
define the notion of objectivity more precisely and give some examples.

14.7.1 Definition

Consider the following transformation which can be viewed either as a super-


posed spatial rigid body motion or as a change of observer:

x E {It ~ x+ = c(t) + Q(t) . x (14.116)

where c(t) is a "rigid" translation and Q(t) a "rigid" rotation:

(14.117)

We also assume that Q is proper orthogonal:

detQ =1 (14.118)

IT we apply transformation (14.116) to two points in {It, x ~ x+ and


y ~ y+, then:

II y+ - x+ 11 2 = (y - . Q(t) .(y -
X)T • Q(tf x) = II y - X 11 2
---..- 1

Therefore the transformation preserves the distances ("isometry"), this is


why it is called " rigid" . There is a property which we will find useful in some
derivations that we will carry on later, it is that

Q . QT is skew-symmetric (14.119)

The praof is easy:

(14.120)

It is well known that under transformation (14.116), a vector v = {Vi} and a


second-order tensor 11 = (Tlij) are objective if they obey the following trans-
formation rules:

vt = QijVj, i.e., v+ = Q·v,


l1t = Qill1lmQjm, Le., 11+ = Q '11 . QT (14.121)

For more general cases, we define "objectivity" as in (Hughes, 1984):


14.7 Objectivity 361

A tensor Aij ... kAB ... c transforms objectively if under the transformation
(14.116) it transforms according to the folIowing rule:

(14.122)

Recall that lower case indices (ij ... k) are for the current configuration ilt
and upper case indices (AB··· C) are for the reference -fixed- configurat ion
ilo. Note that indices AB· .. C do not change, e.g. FiJ is objective if:

(14.123)

A "material" or" Lagmngian" object, Le. defined w.r.t. the reference configu-
ration (e.g., a tensor with components A IJ , alI upper case indices) transforms
objectively and is unaffected by transformation (14.116) (AiJ = AIJ ). Ex-
amples are the right Cauchy-Green strain e, the right stretch tensor U, the
Green-Lagrange strain eG , etc.

14.7.2 Examples

le The deformat ion gradient F is objective:

F+ = âx+ = Q. âx =Q .F (14.124)
âX âX
We can use this result to double-check that the right Cauchy-Green strain e
is objective:

e+ = (F+)T . F+ = F T . QT . Q·F = e (14.125)


--....-...-
Note that J = det F transforms according to:

J+ = det(F+) = (det Q)(det F) = J (14.126)


---...-..-
1

which is consistent with the expectation that the volume is preserved in a


superposed rigid body mot ion or a change of observer.

le The velocity gradient 1 is not objective:


l+ = p+. (F+)-l
. . )
( Q·F+Q·F ·(F- 1 .Q-)
1

= Q. Q-l + Q .l. Q-l


= Q.QT +Q.I.QT
--....-...-
The presence of an extra term shows that 1 does not transform objectively.
362 14. Nonlinear continuum mechanics

.. The rate of deformat ion d is objective:

!(l+ + l+T)
2
~ (~+Q'l.QT +~ +Q'lT.QT)
1
"2 Q . (l + l T ). QT (using the skew-symmetry result)
Q·d·QT

Therefore, although the velocity gradient l is not objective, its symmetric


part, the rate of deformation tensor d does transform objectively.

.. The spin tensor w is not objective:

!(l+ _l+T)
2

~ (~+Q'l.QT _~ _Q'lT.QT)
. TIT T
Q . Q +"2 Q . (l -l ). Q (using the skew-symmetry result)
. T T
Q.Q +Q·w·Q

The extra term shows that the spin tensor w does not transform objectively.

.. The Cauchy (iT) and Kirchhoff (T) stresses are objective


It is assumed that both the outward unit normal n at O"'t in x and the
traction t(x, t, n) transform objectively, i.e.

n+ = Q . n; t+ = Q . t

Consequently, we have:

iT+ . n+ = t+ {::=:} iT+ . Q . n = Q . t


From which it is deduced that:

iT' n = t = (QT . iT+ . Q) . n


It appears then that

iT=QT·iT+·Q, or iT+=Q.iT·QT, (14.127)

which means that the Cauchy stress is objective. Consequently, the Kirchhoff
stress T is also objective:

(14.128)
14.8 Objective stress rates 363

using Eq. (14.126). A useful implicat ion is that the second Piola-Kirchhoff
stress 8 (a "material" object) is objective, that is:

8+ =8 (14.129)

This is easily established by using the identity 8 = F- 1 . T· F- T , and Eqs.


(14.124, 128) .

.. The material time derivatives of Kirchhoff or Cauchy stresses


are not objective
Assume that the material time derivative of 8 (a "material" object) is
objective, that is S+ = S. The Kirchhoff stress rate,
.. T . T .T
T=F·8·F +F·8·F +F·8·F

transforms according to:

T+ = (Q. F+ Q. p). 8· (F T . QT) + (Q. F). S. (F T . QT)


+(Q. F). 8. (F T . QT + pT. QT)

Using previously found results, the following final expres sion is found after
some algebraic manipulation:

(14.130)

Because of the presence of two extra terms, T does not transform objectively
(and neither does u). A practical consequence is that for constitutive models
which are defined in rate form (e.g., elasto-plasticity) we cannot use T or u
because a rigid body motion would generate strains.

14.8 Objective stress rates

We have seen in the previous section that the time derivatives of Kirchhoff
or Cauchy stresses are not objective. In this section, we shall examine a few
objective rates. For many more examples, see (Truesdell and Toupin, 1960).

14.8.1 Examples

Let 1] be a second order Eulerian tensor (e.g., the Cauchy stress u, the Kirch-
hoffstress T, etc.) which transforms objectively (Le., satisfies Eq. (14.121b)) .

.. J aumann rate
The Jaumann rate of 1] is defined by:
364 14. Nonlinear continuum mechanics

J .
11= 11 + 11 . w - w . 11 (14.131)

where w is the spin tensor. Let us show that the Jaumann rate is objective:
J+
11 17+ + 11+ . w+ -
w+ '11+
T . T· T
Q·17·Q +Q·11·Q +Q·11·Q
+(Q '11' QT). (Q. QT + Q. w. QT)
_(Q. QT + Q. W. QT). (Q '11' QT)
Q . 17 . QT + Q . 11 . QT + Q
. 11 . QT + Q . 11 . (_QT . Q) . QT
+Q . 11 . w . Q - Q . 11 . Q - Q . w . 11 . Q T
T ' T
Q '17' QT + Q . 11 . w . QT - Q . w . 11 . QT
= Q'~'QT
The Jaumann stress rate is perhaps the one that is implemented the most
in finite element codes in conjunction with finite-strain elasto-plasticity, al-
though it has a number of shortcomings. For example, in metal plasticity with
kinematic hardening, Nagtegaal and de Jong (1981) have shown that the use
of Jaumann's rate leads to unacceptable oscillatory response in simple shear.

It Green-Naghdi-McInnis rate
This objective rate is defined by:
G. . T . T
11= 11 + 11' (R· R ) - (R· R )'11 (14.132)

where R is the rotation coming from the polar decomposition of the defor-
mation gradient.
Exercise: show that the rate thus defined is objective (hint: R+ = Q·R).

It Lie derivative or Truesdell rate


The following rate of the Kirchhoff stress 7" is known as the Lie derivative
of 7", or up to a factor of J, the Truesdell rate:

L .
7"=7"-
1 '7" -7"'
IT (14.133)

Exercise: show that this stress rate is objective.


Actually, objective rates (14.131-133) and any other possible objective rate
are particular instances of a general geometric concept called the Lie deriva-
tive (Marsden and Hughes, 1983).
Exercise: Prove the following identities

(14.134)
14.8 Objective stress rates 365

14.8.2 A family of objective rates


Several objective rates (not aH of them) can be written as:
*
rl'==iJ+."·W-W·.,,
.. (14.135)

where W is a skew-symmetric tensor chosen so that ~ transforms objectively,

."*+ = Q. ."* .QT (14.136)


As noted by Hughes (1984), given W, we may generate a group of rotations
'R,such that:
'k == W . 'R" and 'R,(t=O) =1 (14.137)
The Jaumann stress rate corresponds to the case:
W=w (14.138)
where w is the spin tensor. It generates a group of rotations 'R, defined by:
'k, == w . 'R" and 'R,(t=O) = 1 (14.139)
The Green-Naghdi-McInnis stress rate corresponds to the case:
W = il·RT (14.140)
where R is the rotation coming from the polar decomposition of the deforma-
tion gradient. Using a procedure presented in Hughes (1984), we now arrive
to the nice thing about the family of stress rates described in this sub-section.
Rotate ." with 'R, to find "''R, as:
"''R, == 'R,T . ." . 'R, (14.141)
The time derivative of "''R, is given by:
iJ'R, 'kT . ." . 'R, + 'R,T . it . 'R, + 'R,T . ." . 'k,
= 'R,T .'R,.'k,T . .".'R,+'R,T .iJ.'R,+'R,T . .".'k,.'R,T.'R,
~ ~

= T· T
'R, . (- W) . ." . 'R, + 'R, . iJ . 'R, + 'R,T ·
. .". (W) . 'R,

'R,T . ( iJ+."·W-W·.,,
. . ) .'R,

This is equivalent to:


.
"''R, = -nT * -n"-,
"- . .". or .,,=
* "- ."''R, . "-
-n. -nT (14.142)
This means that a rather complicated objective stress rate can be computed
as the simple time derivative of a rotated stress (the rotation being 'R, as
defined above). As we shall see in Chap. 16, for the rate formulation of elasto-
plasticity, this result allows the development of algorithms for finite strains
which are identical to those for infinitesimal strains, provided that objects
are rotated properly.
366 14. Nonlinear continuum mechanics

14.9 Laws of thermodynamics


14.9.1 First law

The first law of thermodynamics (or balance of energy law) states that:
(the rate of kinetic energy) + (the rate of internal energy) =
(the power due to externalloads) + (the rate of heat)
This statement is translated as follows:

r !2 pv . v dv +.!!:..dt JK.tr pe dv = JK.tr


.!!:..
dt JK.t
pb· v dv + r t· v da
J8K.t
, .... ' "-".-.' , .... '

+ r pr dv - r q. n da, V/'i,t Dt C (14.143)


JK.t , J8K.t ,
....

where p [kgjm 3] is the current mass density, {Vi} the spatial velocity field,
e [Jjkg] an internal energy per unit mass, {bi} [Njkg] the body force
field, {ti} [Njm 2 ] the field of surface (or traction) forces applied on 8/'i,t,
r [W jkg] a mass density of internal heat production due to external sources
such as inductive heating, and {qi} the heat flux vector through 8/'i,t. Using
(14.91), transport formula (14.68) with k = e and the divergence theorem,
Eq. (14.143) becomes:

r pe dv = JK.tfu: d(v) dv + Jlttr pr dv - }"'tr div q dv,


int
V/'i,t c Dt (14.144)

Therefore, we obtain the following local form of the first law:

pe=u:d(v)+pr-divq, in Dt (14.145)
Equation (14.145) enables the derivation of the heat equation (Sec. 12.11.5).

14.9.2 Second law

The second law of thermodynamics states that:


(the rate of entropy) ~ (the rate of heat divided by the absolute
temperature)
This statement is translated as:

.!!:..
dt JK.t
r (!!..)T . n da,
r ps dv ~ JK.tr p!...T dv - J8K.t V/'i,t C Dt (14.146)

where s is the entropy per unit mass and T > O the absolute temperature.
Using transport formula (14.68) with k = s and the divergence theorem,
(14.146) becomes:
14.10 Further reading 367

(14.147)

Therefore, we obtain the following local form of the second law:

(14.148)

14.9.3 Clausius-Duhem inequality

When we substitute pr into (14.148) by its expression from (14.145), we find:

u:d(v)+p(Ts-e)-q· \lTT ~O, innt (14.149)

This is called the Clausius-Duhem inequality. It can be re-written by intro-


ducing another quantity, the free energy par unit mass, 'I/J,

'I/J = e - Ts (14.150)

Consequently, Eq. (14.149) becomes:


. . \lT
u : d( v) - p( 'I/J + sT) - q . T ~ O, in nt (14.151)

This inequality plays a fundamental role in the formulat ion of constitutive


equations; see chapters 12, 16 and references therein.

14.10 Further reading

In this chapter, we presented results which are independent of any constitu-


tive model. Other results, although material-independent, are given in other
chapters on a per-needed basis. Examples are:
- computation of the strain increment (Sec. 16.2.2),
- a polar decomposition algorithm (Sec. 16.2.3),
- a time-integration algorithm for the rotation matrix (Sec. 16.2.4),
- an algorithm for the mid-point spin (Sec. 16.2.5),
- the Piola identity (Sec. 15.8.1),
- linearization of a pressure B.C. (Sec. 15.8.2),
- differentiation of an isotropic function of a second-order symmetric tensor
(Sec. 15.8.3).
15. Nonlinear elasticity

When the stress-strain response is reversible, the material is called elastic,


the simplest model being linear elasticity, which is valid under the small
perturbation hypothesis (SPH). In this chapter, we study nonlinear elasticity
in the finite-strain regime.

15.1 Hyperelasticityand hypoelasticity


15.1.1 Definitions

In linear elasticity the stress (T and strain ~ are related by:


â 1
(T = -â~ (-~ : c : ~)
2
'--v-"
W(~)

where c is Hooke's operator and W the stored (or elastic or strain) energy
per unit volume. Similarly, in the nonlinear regime, a hyperelastic constitutive
model is derived from a function W(X, F(X, t)), which is a stored energy per
unit reference volume. It is a function of X, the position of a material partide
in the reference configuration, and F(X, t), the deformation gradient. When
there is no ambiguity, we will simply write W(F).
In linear elasticity, the stress and strain rates are related by iT = c : €.
Similarly, in the nonlinear regime, when an objective stress rate; is related
* a : d which cannot be derived
to the rate of deformat ion d via a relation T=
from a stored energy function, the material model is called hypoelastic.

15.1.2 Hyperelasticity and material objectivity

Assuming the existence of a strain energy function for a hyperelastic material


implies that W equals the stress power per unit reference volume, which was
found in Sec. 14.6 to take the following expressions:

. CJI.
W = Tijdji = SIJ-- = PiKFiK (15.1)
2
370 15. Non1inear elasticity

where T is the Kirchhoff stress, S the second Piola-Kirchhoff stress, pT the


nominal stress, d the rate of deformation and C = F T . F the right Cauchy-
Green strain. Now, time derivation of W(F) gives:
. 8W·
W= 8FiK F iK (15.2)

Comparison with (15.1) yields the following equalities:

8W. T 8W
P iK = 8 F iK' l.e. P = 8F (15.3)

As in Sec. 14.7, consider a rigid body mot ion which is superposed on the
deformed configuration:

x E Dt t---+ x+ = c(t) + Q(t) . x (15.4)

We have seen in Sec. 14.7 that F transforms according to:

F+=Q.F (15.5)

The principle of material objectivity requires that:

W(X,Q ·F) = W(X,F) (15.6)

Since F = R· U from the polar decomposition, choosing Q = R T shows that


W is a function of U = C 1 / 2 , and therefore W is a function of C:

W(X,F) = W(X,C) (15.7)

When there is no ambiguity, we will simply write W(C). Time derivat ion of
W(C) and comparison with (15.1) implies that:

S=2 8W (15.8)
8C
We can retrieve (15.8) from (15.3). FirstIy, we compute (8C j8F),

8CKL 8
8FiJ = 8FiJ (FjKFjL ) = ~ji~KJFjL + FjK~ji~LJ = ~KJFiL + FiK~LJ
(15.9)

SecondIy, using the chain ruIe, Eq. (15.9) and the symmetry of C, we obtain:

(15.10)

Combining Eqs. (15.3, 10), it is found that:


15.1 Hyperelasticity and hypoelasticity 371

aw.1.e. P
PiJ = 2FiK acKJ ' = 2F·
aw
ac (15.11)

and since P = F . S from Sec. 14.6, Eq. (15.8) is deduced. Also from Sec.
14.6, we have T = F· S . F T ; therefore the following stress-strain relations
can be derived from W(C) (O' being the Cauchy stress):

T = 2F . aw . F T . O' = ~ F . aw . F T (15.12)
ac ' J ac
A strain energy function satisfying material objectivity as explained and
from which the stress-strain relations are derived according to (15.3), (15.8),
(15.11) or (15.12) defines a so-called hyperelastic constitutive model.

15.1.3 Elasticity tensors

We shall now establish relations between stress and strain rates. Time deriva-
tion of (15.8) gives:

(15.13)

where C is the material elasticity tensor and e G the Green-Lagrange strain.


Unlike Hooke's operator in linear elasticity, C is not constant in general
but depends on the strain; it has however the same symmetries as Hooke's
operator:

CIJKL = CJIKL = CIJLK = CKLlJ (15.14)

Computing the material time derivative of T, we obtain:

T = (l . F) . S . F T + F . (C : C) . F T + F . S . (F T ·lT) (15.15)
2
where we have used a result from Sec. 14.4, il' = l . F, with l the velocity
gradient. Another result from Sec. 14.4, C = 2FT . d· F, allows to rewrite
(15.15) as:

T - l .T - T . lT = F . (C : F T . d . F) . F T (15.16)

The left-hand side of Eq. (15.16) is the Lie derivative of T, ~, which was
shown in Sec. 14.8.1 to be objective. The right-hand side can be rearranged
so that Eq. (15.16) is rewritten as:
L
T= c: d (15.17)
372 15. Nonlinear elasticity

where c is called the spatial elasticity tensor. In component form, we have:

where c is given by the "push-forward" relation:

(15.18)

The tensor c has the same symmetries as C:

Cijkl = Cjikl = Cijlk = Cklij (15.19)

We shall derive a reIat ion similar to (15.17) for the (objective) Jaumann stress
rate. We have:

~= T - T . wT - W .T =T- T . (IT - d) - (1 - d) . T =~ +T . d +d .T
Using Eq. (15.17), this can be written as:
J
T= a: d, where aijkl = Cijkl + OikTjl + OjlTik (15.20)

Actually, any objective stress rate can be written under the form (15.20a).
Now, if we consider a rate relation of the form (15.20a), it is not always
possible to find a stored energy function such that T is given by (15.12),
unless some compatibility conditions are satisfied. A constitutive model of
the form (15.20a) which cannot be derived from a strain energy function is
called hypoelastic. Hypoelasticity by itself is not a good constitutive theory,
for example hypoelastic models can generate non-zero dissipation in closed
cycles. The main use of hypoelastic theory is in conjunction with finite-strain
elasto-plasticity, as we shall see in Chap. 16.
Another useful elasticity tensor can be defined by taking the time deriva-
tive of Eq. (15.3):
. â âW .
PiK = âFjL (âFiK) FjL
(15.21)
, .
AK;L;
.,

It is seen that the elasticity tensor A has the following symmetries:

(15.22)

Using Eqs. (15.8-10) and (15.13), and after some algebraic manipulation, the
following useful relation is found:

(15.23)

Elasticity tensors play a key rale in the theory of linearization (Sec. 15.6).
15.1 Hyperelasticity and hypoelasticity 373

15.1.4 Incompressibility constraint

When the material is constrained, the stress-strain relations presented in


Sec. 15.1.2 have to be modified. We first present the ideas in a heuristic
wayas in (Ogden, 1984). A justification within a variational Iramework will
be presented in Sec. 15.7. Consider a single scalar constraint of the form
h(F) = o. Time derivation of this constraint gives:
. âh.
h = --FiK =0
âFiK

Since this term is zero, we can add it to the expression of IV, Eq. (15.2):

. âW. âh·
W = âFiK FiK - P âFiK FiK

where p is an arbitrary scalar. Now since IV = PiKFiK , it is deduced that:


âW âh
PiK = - - - p - - (15.24)
âFiK âFiK
The most commonly used constraint for hyperelastic rubber-like materials
is incompressibility, for which the constraint is h(F) = J - 1 = O, with
J == det F. Using (14.65), Eq. (15.24) written for this constraint becomes:

âW (1
PiK = âFiK - pJ F- )Ki

Since Jaij = PiKFjK Irom Eq. (14.109), we find the Cauchy stress as:

1 âW
aij = J âFiK FjK - Pdij

Using Eqs. (15.3, 11), the expression of (T can be rewritten as:

(T
2
= - F .-
âW . F T - pl (15.25)
J âC
It is seen that (T is made up of two terms. The first term is the Cauchy stress
that we would find if the material were compressible, the second term is an
additional hydrostatic pressure. The Lagrange multiplier p is not given by the
constitutive model but can be determined from solving the boundary value
problem together with the incompressibility constraint.
In the remainder of this chapter, we shall always consider compressible,
unconstrained materials, unless otherwise indicated. We shall return to the
issue of incompressibility in Sec. 15.7.
374 15. Nonlinear elasticity

15.2 Principal invariants and principal stretches

For isotropic hyperelasticity, W(C) is either a function of the principal in-


variants of C = F T . F or a symmetric function of the principal stretches.
The principal invariants of Care given by:

1 2
Il = tr C, 12 = "2(11 -
2
tr C ), h = det C = J 2 (15.26)

The principal stretches (Al, A2' A3) are related to the principal invariants of
C by the following relations:

h = Ai + A~ + A~, 12 = (AlA2)2 + (A2A3)2 + (A3Ad2, h = (AlA2A3)2


(15.27)

For the constitutive models which will be presented in this chapter, we will
find it useful to split the deformat ion gradient F according to the following
multiplicative decomposition:

(15.28)

It is seen that

det F vol = J = det F, and det F = 1 (15.29)

Therefore F vol rep~sents the volumetric or dilatational part of the deforma-


tion gradient, and F the isochoric or "deviatoric" part of F.
- - - - -T-
The "deviatoric" principal invariants 11 ,12 and 13 of C == F . F =
J-2/3C are related to the principal invariants 11 ,12 and 13 = J2 of C by:

(15.30)
- - -
Similarly, the "deviatoric" principal stretches Al, A2 and A3 are related to the
principal stretches Al, A2 and A3 by:

(15.31)

Therefore:

(15.32)

The "deviatoric" principal invariants are related to the "deviatoric" principal


stretches by:
15.3 Isotropic hyperelasticity in principal invariants 375

15.3 Isotropic hyperelasticity in principal invariants


15.3.1 Formulation

In this formulation, the stored energy function depends on the principal in-
variants II, 12 , 13 of c:
W(X,C) = W(X,lt,hI3) (15.34)
Using Eq. (15.8) and the chain rule, the second Piola-Kirchhoff stress 8 is
given by:
3 -
8=22: aWaIA
A=l aIA ac
The derivatives of IA w.r.t. Care easily found:
aII =1 aI2 aI3 _l
(15.35)
ac 'ac =I1 1-C, ac =I3C
Therefore, 8 is given by:
-
8 = 2(W,l + Il -W,2)1- 2W,2C
- --1
+ 2I3W,3C (15.36)

where we have used the short-hand notation

The "push-forward" relation T = F· 8· F T gives the stresses T and u as:


- - - 2 -
T = Ju = 2(W,l + Il W,2)b - 2W,2b + 2l3W,31 (15.37)

where b = F . pT is the left Cauchy-Green strain.


The material elasticity tensor is given by:

C
a2
= 4acac = 2ac
w a8
(15.38)

In order to compute the derivative of 8 -Eq. (15.36)- w.r.t. C, the following


result is useful:

(15.39)

Using Eqs. (15.35-36) and (15.38), the following expression for the material
elasticity tensor is found:
C nC- l ® C- l + r2 (1 ® C- l + C- l ® 1)
+ r 3(C ® C- l + C- l ® C) + r41 ® 1 + r s (1 ® C + C ® 1)
ac- l
+ r 6c ®C + r 7 1 - r 8 ac (15.40)
376 15. Nonlinear elasticity

where I ijkl = ~(dikdjl + di/djk) is the fourth-order identity tensor and


o~;~~/J = _~ [(C- 1hK(C- 1)JL + (C- 1hdC- 1)JK] (15.41)

(This result is found by differentiating the relation C . C- 1 = 1 as in Sec.


15.8.2 and using the symmetries of C and C- 1 ). The coeflicients r b . .. , rs
are given by:
r1 = 413(W,3+13W,33); r2=413(W,13+I1W,23); Fa=-413W,23;
n =
-
4(W,2+ W
- - 2-
+2hW,12 +I1W,22); Fs=-4(W,12+hW,22);
,11
- -

rs = 4W,22; Fr = -4W,2; = -413W,3rs (15.42)


We have used the short-hand notation:
- Q2w
W,AB == oIAoIB
The spatial elasticity tensor c is obtained by the push-forward of e -Eq.
(15.18)- using Eqs. (15.40-41). In order to carry out the push-forward oper-
ations, the following result is helpful:
FiJFjK FiLFmM'T]JKeLM = [(F· TI· F T ) ® (F . e.FT)]ijlm (15.43)
After some algebraic manipulation, one finally finds the following final result:
c = r 11 ® 1 + r 2(b ® 1 + 1 ® b) + r 3(b 2 ® 1 + 1 ® b2)
+r4 b ® b + Fs(b ® b2 + b2 ® b) + r 6 b2 ® b2 + r 7 I b + r s1(15.44)
where:
(15.45)

Note that the expression of the stresses and elasticity operators which were
obtained are valid for any isotropic hyperelastic model formulated in principal
invariants. For a given model, one only needs to provide the first derivatives
of the stored energy function w.r.t. the principal invariants in order to com-
pute the stresses, and the second derivatives for the computation of elasticity
operators. This allows a neat design of computer software for this kind of
models.
Constitutive models must observe some mathematical restrictions. For
instance, an undeformed state corresponds to F = 1, and therefore Il =
h = 3 and 13 = 1; the stresses computed for this state must be zero. For
other restrictions, see (Marsden and Hughes, 1983) and (Ogden, 1984).
Note that if one is only interested in the second Piola-Kirchhoff stress
S or the material elasticity tensor e, then the inverse of the right Cauchy-
Green strain C can be computed without having to invert C. Indeed, the
Cayley-Hamilton theorem gives:

C 3 - I 1C 2 + 12C - 131 = O ~ C- 1 = :3 (C 2 - I1C + 121) (15.46)


15.3 Isotropic hyperelasticity in principal invariants 377

15.3.2 Modi6.ed neo-Hookean model

The neo-Hookean model is able to describe the stress-strain response of a


rubber-like material for moderate strains. The modified stored energy func-
tion for this model in the compressible case is written as:

(15.47)

where it is the first "deviatoric" principal invariant, G the shear modulus and
K a bulk modulus whose value is large compared to G. Other formulations of
a compressible neo-Hookean model exist in the literature, e.g. (Ogden, 1984):
- G K 2
W(Il,h13 ) = 2"(11 - 3 - 21nJ) + "2(J -1)
In order to compute the stresses, we need to compute the first derivatives of
W w.r.t. the principal invariants, and these are given by:

W ,1 = G2 r3 1 / 3 •, W ,2 = o·, W ,3 = _ G6 11 r3 4 / 3 + K2 (1 _ r3 1/ 2 ) (15.48)

The Cauchy stress u = T / J is then given by Eq. (15.37). After rearranging


terms, we find:

u = GJ- 5 / 3 dev b+ K(J -1)1; dev b == b - i l (15.49)

From F = 1 + au / ax, one easily finds that for infinitesimal strains (€ being
the infinitesimal strain tensor),
J ~ 1 + tr €, b ~ 1 + 2€ (15.50)

The Cauchy stress is then:


5
u ~ 2G(1 - '3tr €)dev e + K(tr €)1 (15.51)

When we neglect the product (tr €)dev e, we retrieve the constitutive stress-
strain relation of isotropic linear elasticity, which explains the name "neo-
Hookean" for the model studied here.
We now return to the finite strain case. In order to compute the spatial
elasticity operator, we need the second derivatives of W W.r.t. the principal
invariants, and these are given by:

W,33
= ~G1 r 7 / 3 + K4 r3
9 1 3
3/ 2 .
,
W __ G r
,13 - 6 3
4/ 3 .
,

W,ll = W ,22 = W ,12 = W ,23 = O


In the incompressible case, the stored energy function of a nea-Hookean model
is given by:
378 15. Nonlinear elasticity

- G
W = 2"(h - 3); J = 1 (15.52)

Equations (15.25, 37) show then that the Cauchy stress is given by:

(7 = Gb - p1, Le. (Tij = Gbij - pc5ij (15.53)

15.3.3 Modified Mooney-Rivlin model

The neo-Hookean model is able to describe the stress-strain response of


rubber-like materials under moderate strains; for larger strains a Mooney-
Rivlin model is used. The modified stored energy function for this model in
the compressible case is written as:

(15.54)

where it and h are the first and second "deviatoric" principal invariants,
respectively, al and a2 are material parameters, and a is a penalty term
whose value is large compared to laII or la21.
The neo-Hookean model is a particular case of a Mooney-Rivlin model; it
corresponds to the following parameter values:

G
al = 2"; a2 = O; a =K
In order to compute the stresses, we need to compute the first derivatives of
W W.r.t. the principal invariants, and these are given by:

W,l al 13
-1/3
;
W ,2 = a2 3
1- 2 / 3 ;

a l I 1- 4 / 3 2 1 1- 5 / 3 a( 1- 1 / 2 )
W,3 -3 l 3 - '3 a2 2 3 + '2 1- 3 (15.55)

The second derivatives of W w.r.t. the principal invariants are needed for the
spatial elasticity tensor; they are found to be:

W,33 4 1
gal 1 3
r 7/ 3 10 1 r
+ ga 2 2 3
8/ 3 ar
+ '4 3
3/ 2 .
,
W _
13 -
al r 4 / 3 .
-3 3 ,

2 -5/3 - - -
W,23 -'3a213 ; W,ll = W,22 = W,12 = O
In the incompressible case, the stored energy function of a Mooney-Rivlin
model is given by:

(15.56)
15.4 Isotropic hyperelasticity in principal stretches 379

15.4 Isotropic hyperelasticity in principal stretches

15.4.1 Formulation

In this formulation, the stored energy function is a symmetric function of the


princi pal stretches Al, A2' A3:

(15.57)

When there is no ambiguity, we will simply write W(AI, A2' A3)' The following
symmetry relations must hold:

(15.58)

For simplicity, we only consider the case of three distinct principal stretches
(Al i- A2 i- A3), for the other two cases (Al = A2 i- A3, and Al = A2 = A3)
we refer to (Simo and Taylor, 1991).
As shown in Sec. 15.8.3, if '11 is a second-order symmetric tensor with
eigenvalues (1]1,1]2,1]3) and -orthonormal- eigenvectors (e(1), e(2), e(3)), and h
is an isotropic function of '11, i.e. h(1]I, 1]2,1]3), then:

(15.59)

Two important implications are that '11 and 8h/8'11 are coaxial and commute:

8h 8h
'11'-=-''11 (15.60)
8'11 8'11
An application of Eq. (15.59) is:

8TJB = e(B) 181 e(B) (no sum over B) (15.61)


8'11

Consider the right Cauchy-Green strain C = F T . F with eigenvalues A~ and


eigenvectors N(A), A = 1,2,3. Applying Eq. (15.61) gives:

8AB = _1_ N (B) 181 N(B) (no sum over B) (15.62)


8C 2AB

Using Eq. (15.59), the second Piola-Kirchhoff stress is found to be:

(15.63)

The eigenvectors of Sare N(A) and the eigenvalues are:


380 15. Nonlinear elasticity

1 8W
SA = - - - (no sum over A) (15.64)
AA 8A A

We recall from Chap. 14 that the vectors n(A) == (I/AA)F· N(A) (no sum
over A) are orthonormal and are the eigenvectors of the left Cauchy-Green
strain b = F . F T . The push-forward reIat ion T = F . S . F T gives then the
following expression for the stresses T and u:

(15.65)

The eigenvectors of T and u are n(A) and their eigenvalues are:

(15.66)

A result which is computationally interesting is that in order to compute the


stresses, we do not need to compute the eigenvectors N(A) or n(A), only the
tensor products N(A) ® N(A) or n(A) ® n(A) are needed, and these have
simple expressions. lndeed, for a symmetric tensor 1] as defined earlier, using
the spectral decompositions of 1,1] and 1]2 we easily find the following result:

(15.67)

Where (A,B,e) is an even permutation of (1,2,3). Applying Eq. (15.67)


allows to compute the stresses T and u as follows:

(15.68)

Computation of elasticity operators is rather involved and reported to Sec.


15.8.4.

15.4.2 Modified Ogden's model

Ogden's model is useful for simulating the stress-strain response of rubber-


like materials under very large strains. The modified stored energy function
of this model in the compressible case is given by:

t; ~i~ (~~; + ~~; + ~;'; -


N
W(X, Al, A2' A3) = 3) + ~(J _1)2 (15.69)

where ~1' ~2 and ~3 are the "deviatoric" principal stretches, Ci, mi, (i =
1, ... , N) are material-dependent parameters and a is a penalty parameter.
15.5 Examples of homogeneous deformations 381

For more informat ion about the model and the identification of its material
parameters, see (Ogden, 1984).
Since >:1>:2>:3
= 1, Eq. (15.69) can be rewritten as

N
W(X, AI, A2, A3) = L ~~ [>:~. + >:;n' + (>:1>:2)-m. - 3] + ~(J _1)2
i=l •
(15.70)

Using relations (15.33) between "deviatoric" principal invariants and stretches


and comparing Eqs. (15.70,54), it is seen that Mooney-Rivlin model is a par-
ticular case of Ogden's model corresponding to the following parameters:

We now return to the general case. In order to compute the stresses, the first
derivatives of W w.r.t. the principal stretches are needed, and these are:

(15.71)

where (A, B, C) is an even permutation of (1,2,3).

15.5 Examples of homogeneous deformations

15.5.1 Homogeneous simple shear

arctan 'Y

el
Fig. 15.1. Simple shear of a rectangular block

A homogeneous simple shear deformation of a rect angular block is de-


picted in Fig. 15.1. The deformation is defined by the following equations
W.r.t. a fixed Cartesian coordinate system:

(15.72)
382 15. Nonlinear elasticity

The following matrix representations for the deformation gradient F, the left
Cauchy-Green strain b and b 2 easily follow:

[F)

(15.73)

Note that the deformat ion is isochoric (J = 1) and that the deformation
gradient can be written as:

(15.74)

where e2 is the normal to the slip plane and el is the slip direction in the slip
plane. The principal invariants 11 ,12 and 13 of the right or left Cauchy-Green
strain tensors are given by:

(15.75)

For any isotropic hyperelastic model, we have seen in the previous sections
that the Cauchy stress tensor is given by:

(15.76)

where /30, /31 and /32 are scalar functions of the principal invariants 11 ,12 and
13' For simple shear, the Cauchy stress matrix is:

0"22 + ''(0"12 ')'[/31 + /32 (2 + ')'2)]


[0"] = [ ')'[/31 + /32(2 + ')'2)] /30 + /31 + /32(1 + ')'2)
O O

It is seen that unlike the infinitesimal strain case, 0"11,0"22 and 0"33 do not
vanish in general. AIso, the following relation holds:

(15.78)

This is called a universal relation because it is valid for any isotropic hyper-
elastic model in simple shear. The stress vector on the top horizontal face is
given by:

(15.79)

So it is seen that in general we cannot maintain a state of simple shear by


shear forces alone, we also have to apply normal pressure (this is known as the
15.5 Examples of homogeneous deformations 383

Poynting effect). The same conclusion applies to the inclined facets. Consider
the right facet for instancej the normal and tangential directions are:

After some algebra, the stress vector on the facet is found to be:

(T T • n = ( 0"22 - "(
--20"12 ) n + 1
--20"128 (15.80)
1+"( 1+"(
We now apply apply the results to the neo-Hookean model. In the compress-
ible case, Eq. (15.49) gives:

(15.81)

On each horizontal facet, we need to apply a shear force (G"() and a pressure
(G"(2/3). We also need to apply shear and pressure forces on the inclined
faces according to Eq. (15.80).
For the incompressible neo-Hookean model, Eq. (15.53) gives:

(15.82)

The shear and pressure forces on the horizontal facets are (G"() and (G - p) j
on the inclined facets these forces are given by Eq. (15.80) as (G"( / (1 + "(2))
and (G/(l + "(2) - p), respectively.

15.5.2 Uniform extension

Consider a circular cylindrical bar whose axis is aligned with el. A uniform
axial extension is defined by the following equations in a fixed Cartesian
system:

(15.83)

The following matrix representations for the deformation gradient F, the left
Cauchy-Green strain b and b2 easily follow:

~2 ~ 1 [b] = [ ~~ ~~ ~ 1 [b]2 = [~t ~~ ~ 1


j j
O A2 O O A~ O O A~
(15.84)
384 15. Nonlinear elasticity

The principal invariants 11 ,12 and 13 are given by:

J = VIT 2
13 = AlA2i Il
2J
= Al2 + ~i (2
12 = 2Al
J )J
+~ Al (15.85)

For the compressible neo-Hookean model, Eq. (15.49) gives the non-zero com-
ponents of the Cauchy stress as (alI shear stresses vanish):

2 -5/3 2 J
au = K(J - 1) + "3 GJ (Al - AI)i
1
a22 = a33 = K(J - 1) - -GJ (Al - -J)
-5/3 2
(15.86)
3 Al
IT a22 = a33 = O (simple tension), then:
(15.87)

One expects that in simple tension: au > O (tensile stress), Al > 1 and
increases, A2 < 1 and decreases. Equation (15.87b) then implies that G > O
and Eq. (15.87a) gives that J > 1 (increase in volume), since it is assumed
that K > O.
For an incompressible neo-Hookean model, it is found that:

G
au = GAl2 - Pi a22 = a33 = Al - Pi J = AlA22 = 1 (15.88)

IT a22 = a33 = O (simple tension), then:

G 2 1
P = Al i au = G(AI - Al) (15.89)

Note that the pressure P is now determined. The nominal stress Pu is re-
lated to the Cauchy stress by: auA = PuAo, and incompressibility implies
that Aolo = Al, where (A o, lo) and (A, l) designate the cross section area
and the axiallength in the initial and deformed configurations, respectively.
Therefore, the nominal stress is given by:
au 1
Pu = - = G(AI - 2")
Al Al
The stress-stretch relations au (Al)/G and Pu (Ad/G are shown in Fig. 15.2.

15.5.3 Pure dilatation

Consider a cube equalIy stretched in the three directions: Al = A2 = A3. The


three invariants 11 ,12 and 13 are given by:

(15.90)
15.6 Linearization 385

3.5
3
\!l
>. 2.5
,t:l
'"O
Il:>
'"O 2
'>
;.a 1.5
<Il

...,...Il:>
<Il
1
(J}
0.5

Axial stretch (.xI)


Fig. 15.2. Cauchy and nominal stresses vs. axial stretch for an incompressible
neo-Hookean material in simple tension

For the compressible neo-Hookean model, Eq. (15.49) gives the Cauchy stress:
a = K(J - 1)1; Le. aij = K(J - l)Oij (15.91)
This is a state of pure dilatation where the only material parameter which is
present is the bulk modulus K.

15.6 Linearization
Linearization plays an important role in the formulat ion and analysis of prob-
lems of solid and structural mechanics, where it is often viewed as "small
deformations superposed on finite deformation" (Green and Zerna, 1968).
Linearization is also important in computational mechanics, since a nonlin-
ear problem is replaced by a cascade of linear problems which are solved at
each iteration. In both instances, linearization needs to be properly defined
and carried out. For an intuitive introduction to the subject, see (Hughes
and Pister, 1978), for a more mathematical presentation, see (Marsden and
Hughes, 1983). Linearization may be defined as follows.
Consider infinite-dimensional spaces e and F, a continuously differen-
tiable ("Cl,,) mapping G : e --t F and elements x, u in e. The linear part
of Gat x is:

L[G]x = G(x) + DG· u (15.92)


where (DG . u) is the direction al derivative of G at x in the direction u.
Equation (15.92) is a Taylor expansion truncated at the first order. In prac-
tice, the directional derivative is computed according to the following formula
(often known as the Gâteaux derivative):
386 15. Nonlinear elasticity

d
DG . u = L [G(a: + fU)] (15.93)
ut (f=O)

In classical terminology, u is called a variation of a: and is denoted (8a:) aud


(DG . u) is called the first variat ion of G and is denoted (8G).

15.6.1 Linearization of the deformation

Consider a deformed state defined by: cp : X E Do I----t a: = cp (X, t). The


derivative of Fat cp in the direction (8cp) is:
d 8 d 8(8cp) 8(8cp)
df(f=O} 8X (cp(X, t) + E8cp(X, t)) = -df(f=O} (F + E 8X ) = 8X
(15.94)
Therefore, the linear part of Fis:

L[F] = F + GRAD (8cp) (15.95)


We have seen in Sec. 14.5.1 that 8J/8FiK = J(F-l)Ki, therefore the deriva-
tive of J at cp in the direction 8cp is:

d det (F + E8(8cp)) = J 8XK 8(8<pi) = J 8 (8<pi) (15.96)


dE (f=O) 8X 8Xi 8XK 8Xi
Consequently, the linear part of J is:

L[J] = J[l + div (8cp)] = J [1 + tr (F- 1 • GRAD (8cp))] (15.97)


The derivative of the right Cauchy-Green strain C at cp in the direction 8cp
is given by (using the chain rule and Eq. (15.94)):
8(8<pi) 8(8<pi)
D(FiKFiL)' 8cp = aXK FiL + FiK aXL (15.98)

Therefore the linear part of C is:

L[C] =C + FT . GRAD (8cp) + GRAD T (8cp)· F (15.99)

The linear part of Green-Lagrange strain ~.a is then given by:

(15.100)

It is interesting to consider as au application the particular case where the


current state corresponds to an undeformed reference state:

Xi = 8iL XL; FiK = 8iK; €


G = O; L[€ G ] = '1[
2 GRAD (8cp) + GRAD T (8cp) ]
It is seen that the linear part of €G is the usual injinitesimal strain tensor.
15.6 Linearization 387

15.6.2 Linearization of constitutive equations


The constitutive relation for a hyperelastic material can be written as: PiK =
PiK(F). Using the chain rule, the derivative of PiK at ep in the direction dep
is found to be:

(15.101)

where A is the elasticity tensor defined in Eq. (15.21). The linear part ofthe
nominal stress pT is then:
L[pT] = pT + A: GRAD (dep) (15.102)

15.6.3 Linearization of the equations of elasto-statics


The linear part of DIV P is obtained from Eq. (15.102)
L[DIV P] = DIV P + DIV (A: GRAD (dep)) (15.103)
Considering dead loads for simplicity, the linearized equations of elasto-statics
are:
DIV P + DIV (A: GRAD (dep)) + poB = O (15.104)
We shall now rewrite Eq. (15.104) in the current configuration. Let Y =
(YiK), the Piola identity (which is demonstrated in Sec. 15.8.1) gives:
. (1 T)' aYiK a (1 )
DIV Y = JdlV JY' F , 1.e. aXK = J aXl JYiKFlK
Taking Y =P and using Eq. (14.96), we obtain:
DIV p = Jdiv (O"T) (15.105)
Now, applying the Piola identity to Y = A : GRAD (dep) and using Eqs.
(15.23) and (15.18), we obtain:
~ (
aXK
. . a(o<Pj)) _ ~
A K' L3 aXL - J aXI
(.!.JA K'.L3.a(o<pj)
aXL FlK
)

= J aXI
a (1JA KiLj a(d<pj)
aX m FmLFlK
)

_J~(a(d<Pi)
- a Xl a Xm Uml
.!. . .a(d<Pj))
+ JC hm3 a Xm (15.106)

where c is the spatial elasticity tensor defined in Sec. 15.1.3. Using Eqs.
(15.104-106) and recalling the following relations from Chap. 14 between
densities and body forces in the reference and current configurations: Pa = pJ,
B(X, t) = b(x, t), the linearized equations of elasto-statics written w.r.t. the
current configurat ion are found to be:

div [O" + \7(oep) . O" + -Jc: \7(dep)] + pb = O (15.107)


388 15. Nonlinear elasticity

15.6.4 Variational formulations

As in Chap. 2, we consider the folIowing mixed boundary conditions (B.Cs.):

- Displacements are prescribed on a portion an~U) of the boundary ano,

cp(X,t) = cp(X)j for X E an~U) (15.108)

- Traction is prescribed on the remaining port ion an~F) of the boundary


(we assume that an~U) n an~F) = 0 and an~U) U an~F) = ano),
P(X, t) . N(X) = T(X), for X E an~F) (15.109)

where N is the unit outward normal to the reference surface an~F) . Equation
(15.109) corresponds to a dead load B.C. since the prescribed traction T
depends on the reference position X and not on the current position z. For
more general B.Cs., see (Ogden, 1984). An example of a B.C. which does
change with the deformation is pressure:

(TT(Z, t) . nea:) = -p(t)n(a:) == t, for a: E anjF) = cp(an~F») (15.110)

where n is the unit outward normal to the deformed surface anjF). Lin-
earization of a pressure B.C. is given in Sec. 15.8.2.
As in Chap. 2, we introduce the folIowing two sets:

U = {ifJ: n o -+ ~ I ifJ "sufficiently smooth" and ifJ = cp an an~U)}

U* { ifJ* : n o -+ ~ I ifJ* "sufficiently smooth" and ifJ* = O an an~U)}


(15.111)

We now define:

_ [ pT : (GRAD ifJ*) dV - IIext(ifJ*)j with


lao
_ [ poB· ifJ* dV + [ T· ifJ* dA (15.112)
lao llw~F)
We have seen in Sec. 14.5.10 that the Lagrangian weak formulation of the
boundary value problem in statics is given by:

Find ifJ E U such that G( ifJ, ifJ*) = O for alI ifJ* E U* (15.113)

This is also known as the virtual work theorem (V.W.T.)j it is valid for any
constitutive model. Equivalence between strong and weak formulations can
be established by a procedure identical to that of Sec. 2.2. For a hyperelastic
15.6 Linearization 389

material, the V.W.T. can be retrieved by writing a stationarity condition for


a potential energy functional II(c/» which is defined for c/> E U by:

II(c/» == ( W(F(c/») dV - IIext(c/» (15.114)


10 0

Assuming dead loads, the derivative of IIext(c/» at c/> in the direction oc/> is:

Dilext . oc/> = { PoB· (oc/» dV + ( T· (oc/» dA = IIexd0c/»


10 0 1aoăF)
(15.115)

Consequently, the derivative of II(c/» at c/> in the direction oc/> is given by:

(15.116)

The stationarity condition DII . oc/> = O gives the weak formulat ion (15.114).
The question: "can II be minimized?" turns out to be a difficult onej for a
discussion see (Marsden and Hughes, 1983).

15.6.5 Linearization of the weak formulat ion

The derivative of G(ep,oep) in the direction dep is given by (using Eq.


(15.106)):

This can be rewritten as:

DG(ep,oep) . dep = i, []V(oep) : c: V (dep).. + y(oep)· ~. V(dep)) dv


'" (15.118)

The first term in the right-hand-side of (15.118) is a material stiffness con-


tribution, while the second is a geometric stiffness contribution.
390 15. Nonlinear elasticity

15.7 Mixed variational formulation

In this section, we present a three-field variational formulation which was


proposed by Simo and Taylor (1991). Its main advantage is that it can handle
quasi-incompressible, compressible and incompressible cases within the same
framework.

15.7.1 Fonnulation

Let F( ep) be the deformation gradient corresponding to ep, Le. F( ep) =


GRAD ep and J(ep) the associated change in volume, Le. J(ep) = detF(ep).
The basic idea behind the three-field variational formulation is to treat the
change in volume as an independent variable O and to enforce the condition
J(ep) = O via a Lagrange multiplier p. A modified deformat ion gradient F
such that det F = O is defined by:

(15.119)

A potential energy functional is defined for ep E U, O > O and p by:

where IIext(ep) is defined in Eq. (15.1l2b). The three-field variational formu-


lation is obtained by computing the first variat ion of II at (cp,O,p) defined
as above, for arbitrary variations (oep, 08, op). We use the following notation
for the derivative of II in the direction dep (and similar notations for the
derivatives in the directions dO and dp):

(15.121)

Using Eqs. (15.94) and (15.96), the derivative of Fin the direction dep is:

1 âd<pj ) - -
-
DFiK . dep = (âdCPi
âXl -"3 âXj du FlKi i.e. DF· dep = dev (Vdep) . -F
(15.122)

A modified Cauchy-Green strain is defined by:


- -T-
C=.F ·F (15.123)

Modified nominal and Cauchy stresses are defined by:

p T _= âW __ 2âW .FT. _ 2- âW -T
(15.124)
âF âC ' (J' =. OF. âC . F
15.7 Mixed variational formulat ion 391

Using the chain rule and Eqs. (15.122, 124), the derivative of W in the di-
rection oI.{) is given by:

DW . oI.{) = -T -
P : (dev ('VOI.{)) . F) = (J(j: dev ('VOI.{)) (15.125)

Using Eqs. (15.125) and (15.96) the derivative of II in the direction oI.{) (writ-
ten w.r.t. the current configuration) is given by:

DII·ol.{) = L. [JU:dev ('Vol.{))+pdiv (oCP)] dv-DIIext·ocp

=L. [JO'+(p-
,
30JtrO')1] : 'V (ocp)dv-DIIext'ocp
'
(15.126)
'"
ii'
A very interesting feature of the mixed variational formulat ion is that in a
computational framework, the constitutive routine does not change: F is sent
instead of F and the material model returns a "modified" Cauchy stress 0'.
The tensor II defined in Eq. (15.126) is often called "mixed" stress.
The derivative of F in the direction 00 is:
-
DP'K '00 = -F-K-
00
(15.127)
• • 30
Therefore, the derivative of W in the direction 00 is (using the chain rule and
Eqs. (15.124)):
-T -00 00
DW . 00 = P : F 30 = (tr 0')'3 (15.128)

Using Eq. (15.128), the derivative of II in the direction 00 (written w.r.t. the
current configuration) is given by:

DII· 00 = r
lOt
_Jl (~tr
3
O' - p)(oO) dv (15.129)

Finally, the derivative of II in the direction op (written w.r.t. the current


configuration) is:

DII . op = { ~(J - O) (op) dv (15.130)


lOt
We now check that the solution of the three-field mixed formulat ion is the
same as that of the original formulation. The stationarity conditions for II
for arbitrary variations op and 00 (Le., setting Eqs. (15.130) and (15.129)
to zero) give O = J and p = (tr 0')/3, respectively. Using these results, the
stationarity condition of II for arbitrary variations oI.{) (Le., setting (15.126)
to zero and using (15.115)), gives the Eulerian weak form of the equilibrium
equations.
392 15. Nonlinear elasticity

15.7.2 Incompressibility constraint

For the material models we have presented (neo-Hookean, Mooney-Rivlin and


Ogden's), the stored energy function has the following form:

W(F) = W(F) + o')'(J) (15.131)

The first term involves the isochoric part F of the deformation gradient F
and the second term the volumetric part. The function ')'(x) (x > O) satisfies
the following conditions:

')' is convex; ')'(x) ~ O; ')'(x) = O{=:} x = 1 (15.132)

A typical expression of ')'(x) -which was used in this chapter- is:

(15.133)

In the mixed variational formulaţ.ion, F is sent to the constitution instead of


F; its isochoric part is equal to F however:

(15.134)

Therefore, the energy functional II -Eq. (15.120)- is written as:

II(cp,O,p) =( [W(F) + 0')'(0) + p(J(cp) - O)] dV - IIext(cp) (15.135)


Jao
Incompressibility may be enforced via a penalty method: o > O is viewed as
a penalty parameter and as o -+ 00, () -+ 1; see (Luenberger, 1989). The
penalty method however becomes ill-conditioned as o is increased. Simo and
Taylor (1991) report numerical problems with the neo-Hookean model for a
ratio o/G ~ 103 . They recommend using an augmented Lagrangian method
which we now describe (see also (Luenberger, 1989)).
A Lagrangian functional is defined by:

C(cp,(),p;>.) == II(cp,O,p) + ( >'h(O) dV (15.136)


Jao
where >. is a Lagrange multiplier, and h(x) (x > O) is a continuously differ-
entiable function which satisfies:

h(x) = O{=:} x = 1 (15.137)

A typical example for h(x) is:

h(x) = x-1 (15.138)


15.8 Appendices 393

The stationarity conditions of C in the directions 81.{) and 8p are identical ta


those for II; the stationarity conditions in the directions 80 and 8A are:

DC·80 1[~tr(lT)
ilo 3
- p + Ah/(O)](80) dV =O
DC·8A = ( h(O)(8A) dV =O (15.139)
}ilo

For a pure penalty method, A = O. For an augmented Lagrangian method, A


is updated according ta the Uzawa algorithm (see (Luenberger, 1989)).

15.8 Appendices

15.8.1 The Piola identity

As in (Marsden and Hughes, 1983), we integrate (DIV Y) over K.o ta obtain:

{ âYiK dV = { YiKNK dA
}I<.O âXK }81<.0
{ 1
YiK- (Eq. (14.95))
J FlKn, da
}81<.'
= { ~(.!.YiKFlK) dv
}I<., âXI J
{ ~(.!.YiKFlK)J dV
}I<.O âXI J
Since the equality holds VKO E Do, then:
âYiK â 1 1
âXK = J âXl C:J YiKFlK ), Le. DIV Y = Jdiv (JY. F T ) (15.140)

which is the Piola identity.

15.8.2 Linearization of a pressure B.C.

A pressure B.C. is such that an the deformed surface âDt, we have t = -pn.
We have seen in Sec. 14.5.9 that tda = TdA and nda = JF- T . NdA,
therefore:

(15.141)

We note that N is based an X E âDo, the surface an the reference configu-


ration. We use the short hand notation 8F for (DF . 81.{)), the derivative of
F in the direction 81.{). Similarly, we use the notations 8F- 1 and 8J.
394 15. Nouliuear elasticity

We have already computed ~F aud ~J (Eqs. (15.94, 96)), we only need


to compute ~F-l. Noting that:

we multiply by (F-l )Lj to find:

8(F-l)Kj = _(F-l)Ki(~Fid(F-l)Lj, i.e. ~(F-l) = _F- 1 . (8F)· F- 1


(15.142)

Finally, 8T is given by:

8T = -(8p)JF- T . N - p(8J)F- T . N + pJF- T . (8F T ) . F- T . N


(15.143)

If the pressure is uniform then: 8p = O.

15.8.3 Differentiation of an isotropic function of a second-order


symmetric tensor

Consider a second-order symmetric tensor "1 with eigenvalues 7]A and -


orthonormal- eigenvectors ee A), A = 1,2,3. We only consider the case when
the three eigenvalues are distinct. The spectral decomposition gives (we use
matrix notation for convenience):

L
3
[7]] = 7]A{e(A)}{e(A)}T
A=l
After differentiation, multiplication with {ee B)V at the left, {e(C)} at the
right and use of the orthonormality conditions, the following result is ob-
tained:

{ee B) }T[d7]]{ ee C )} = (d7]B )8 Bc + 7]0{ eeB)}T {de(C)} + 7]B{ de(B)V {e(C)}


(no sum over B or C)

In the basis (e(1), e( 2 ), e( 3 )) the left-hand side of the equation is simply [d7])BC
and the right-hand side can be simplified by using the orthonormality condi-
tion:

Therefore, the following result is found:

[d7])BC = (d7]B)8 Bc + (7]C -7]B){ee B )}T{deeC )} (no sum over B or C)


(15.144)
15.8 Appendices 395

In the basis (e(l) , e(2) ,e(3»), the diagonal terms of (d"1) are (d1]l' d"72, d1]3). In
(81]A/8"1) the A-th diagonal term is equal to 1 and alI other terms are zero;
therefore:

81]A = etAJ (8) etAJ (no sum over A) (15.145)


8"1
Now, if h is an isotropic function of "1, i.e. h(1]I' "72,1]3), then using the chain
rule and Eq. (15.145), it is found that:

(15.146)

It is useful to notice that "1 and (8h/8"1) have the same eigenvectors and
commute:

(15.147)

15.8.4 Elasticity tensors for principal stretch formulat ion

The objective of this subsection is to compute the spatial tangent operator c


defined by:
8S/J
Cijkl = 2Pi/Pp PkK FlL 8CKL '
for the principal stretch formulation of Sec. 15.4. Using Eqs. (14.106), (15.65),
and a definit ion from Sec. 14.2.3, F· N(A) = AAn(A) (no sum), we obtain:

S/J (P- 1) IiTij (F-l)Jj


3

L....t (3A (P- 1 ) /tn


"" . (A)(p- 1 ) . (A)
i JJn j
A=l
3
L(3AM;1) (15.148)
A=l
where we used the following notation:

(15.149)

Differentiation of S/J w.r.t. CKL gives:

8S/J =
8CKL
t[
8(3A M}1)
A=l 8CKL
+ (3A 8M}1)]
8CKL
(15.150)

From Eq. (15.146), we obtain (8(3A/8CKL) ,


396 15. Nonlinear elasticity

Exercise: Compute (âMiJ) jâCKL). Hint: use results of Sec. 15.8.3. Answer:
see (Simo and Taylor, 1991).
Using the previous results and the following notation:

(15.151)

it is found that Cijkl has the following expression:

(15.152)

In tensor notation, this can be rewritten as follows:


3 3 3
C =L L 1'ABm(A) ® m(B) +2L PAC(II)(b) (15.153)
A=l B=l A=l

where the expression of the fourth-order tensor c(II) (b) is given in (Simo and
Taylor, 1991)
Note that the expressions of PA and 1'AB become remarkably simple if we
introduce the principal logarithmic stretches defined as follows:

(15.154)

In this case, it can be easily checked that we have:

(15.155)

For alternative derivations of the spatial elasticity tensor in principal stretch


formulations, see (Ogden, 1984) and (Bonet aud Wood, 1997).
16. Finite-strain elasto-plasticity

In this chapter, the infinitesimal theory of elasto-plasticity of Chap. 12 is


generalized to finite strains. Two formulations are presented. The first one is
based on a multiplicative decomposition of the deformat ion gradient into elas-
tic and plastic parts and on hyperelastic stress-strain relations. The second
formulation assumes an additive decomposition of the rate of deformat ion
and hypoelastic constitutive relations. Numerical algorithms are presented
for both theories. Although the hyperelastic-based plasticity formulation is
more satisfying from a theoretical point of view, the rate models are stiH
widely used in numerical analysis.

16.1 First theory

In this section, we present a formulation of finite strain elasto-plasticity which


is based on a multiplicative decomposition of the deformation gradient and
hyperelastic constitutive equations. Our presentation of the theory mainly
follows (Simo, 1992); see also (Halphen and Nguyen, 1975), (Lubliner, 1990),
(Maugin, 1992), and references therein. The numerical algorithms were pro-
posed by Simo (1992). 1

16.1.1 Multiplicative decomposition of the deformat ion gradient

Following Lee (1969), it is assumed that the deformation gradient F(X, t) is


split into elastic and plastic parts as follows:

(16.1)

First, we motivate this basic assumption in lD. Consider the uniaxial tension
test of Chaps. 1 and 12 again, the stress-strain response is plotted in Fig.
16.1. In this test, a bar of uniform cross section is subjected to the uniaxial
1 Juan Carlos Simo (1952-1994), a native of Spain, made seminal contributions to
computational mechanics during his research career at University of California,
Berkeley and Stanford University, unt il his untimely death at the age of 42. 1
had the good fortune of working with him when he was a consultant for Centric
Engineering Systems, Inc.
398 16. Finite-strain elasto-plasticity

loading history: (O) --? (B) --? (C), during which the length of the bar
takes the following values: (lo) --? (l) --? (lp). Stage (O) --? (B) corresponds
to a monotonie loading beyond the elasticity domain, and (B) --? (C) to
elastic unloading. RecaB that (C) --? (B) is an elastic loading process. State
(C) corresponds to a stress-free, unloaded configuration. We can write the
following trivial identity:

(16.2)

where A = l/lo is the axial stretch at the end of (O) --? (B), Ae = l/lp can
be viewed as the stretch at the end of the elastic transformat ion (C) --? (B)
and Ap = lp/lo corresponds to the length ratio between states (O) and (C).
Equation (16.2) is the particularization of (16.1) to lD.

B
a(B) ~-----:::=-----

of----+--+-----

f(B)

Fig. 16.1. Uniaxial stress-strain response of a metallic specimen

The illustration of (16.1) in 3D is similar to that of the lD case. If (O)


designates the initial state, (B) the current state and (C) a local intermediate,
stress-free, unloaded state, then the deformation gradients are: F for (O) --?
(B), (Fe)-l for (B) --? (C) or F e for (C) -+ (B), and FP between states (O)
and (C); see Fig. 16.2.
Decomposition (16.1) is not unique, it is defined up to a rigid rotation.
lndeed for any Q such that Q. QT = QT . Q = 1, we have:

However, the formulation which will be presented uses be = F e • F eT and


C P = FpT. FP whose expressions are unaffected by rigid rotations Q, unlike
F e and FP.
19 stale (B)
16.1 First theory 399

d: 'y'
stare (o()"Z-L) state (C)
Fig. 16.2. Illustration of the multiplicative decomposition of the deformat ion gra-
dient

16.1.2 Hyperelastic-plastic constitutive equations

Similarly to the infinitesimal strain case (Chap. 12), we make hereafter some
basic assumptions in order to derive the constitutive equations of finite-strain
elasto-plasticity.
We assume the existence of a specific free energy per unit reference volume
'IjJ(b e , ~), where be is the elastic left Cauchy-Green strain:

(16.3)

and ~ an internal variable (only a single scalar variable is assumed for sim-
plicity). An elastic domain is defined by the following yield criterion:

f(T,q) ~ O, (16.4)

where T is the Kirchhoff stress and q a hardening stress. The material is


assumed to be isotropic, hence 'IjJ( b e ,~) and f (T, q) are isotropic functions of
b e and T, respectively.
With (T = TI J and J = Pol P from mass conservation, the Clausius-
Duhem inequality of Sec. 14.9.3 requiring the mechanical dissipation (per
unit reference volume) V to be non-negative can be written as follows:

(16.5)

where d is the rate of deformation tensor (i.e. the symmetric part of the
velocity gradient 1). We recall the following relations from Sec. 14.4:

(16.6)

The following result immediately follows from the chain rule:


400 16. Finite-strain elasto-plasticity

The right-hand side will be rearranged hereafter. Designating by CP the


plastic right Cauehy-Green strain,
CP == FpT. FP,

the folIowing identity folIows from the multiplieative decomposition (16.1):

(16.7)

Time differentiation of this relation and use of Eq. (16.6b) gives:


. L L d
be=l·be+be·lT +be, with be==F· dt(CP-l).F T (16.8)

L
where b e is the Lie derivative of be (see Sect. 14.8.1).
An important result which was proven in Sect. 15.8.3 is that sinee 'Ij; is
an isotropic function of be , then (â'lj; / âb e) and be commute:

â'lj; . be = be . â'lj;
âbe âb e
Consequently, and after some manipulation, the folIowing identities are found:

= 2tr (:t . be • d);

tr (â'lj; . be .
âb e
fi .be- 1)

Substitution of the last two identities into the ehain rule expression of (d'lj; / dt)
gives:

d'lj; = tr [â'lj; . be . (2d + (;e . be- 1 )] + â'lj; ~ (16.9)


dt âb e â~

Plugging this expression into (16.5), it appears that the dissipation V ean be
written as folIows:

V = (r - 2 â'lj; . be) . d + (2 â'lj;e . be) . (_~{;e . be- 1 ) _ â'lj; i > O (1610)


âb e . âb . 2 â~'" - .

Sinee the inequality must hold for alI admissible proeesses, a standard ar-
gument which was mentioned in Sec. 12.11.2 gives the folIowing equation of
state:

(16.11)
16.1 First theory 401

The dissipation then takes the following reduced form (with q == â'ljJ / â~):

(16.12)

Hyperelastic stress-strain relation (16.11) will be studied in Sect. 16.1.3, and


it will be shown in Sect. 16.1.4 that the flow rules can be derived by applying
the maximum dissipation principle to Eq. (16.12).

16.1.3 Stress-strain relations

Designating the principal directions of the elastic left Cauchy-Green strain


be by (n(l),n(2),n(3») -and these were shown in Sec. 14.2.3 to form an or-
thonormal basis- and the principal elastic stretches by AA' the spectral de-
composition of be is the following:

(16.13)

It was shown in Sect. 15.8.3 that since 'ljJ is an isotropic function of


be , then be and (â1jJ / âb e ) have the same eigenvectors. Writing 'ljJ( be ,~) =
~((AIY,(A2)2,(A~)2,~), and using Eqs. (15.146) and (16.13), and the or-
thonormality properties, the stress-strain relation (16.11) becomes:

(16.14)

where

(3A == 2(AA)2 â(~~)2' (A = 1,2,3, no sum over A), (16.15)

designate the principal Kirchhoff stresses.


Define the principal elastic logarithmic stretches by:

(16.16)

and introduce the following array notation:

(16.17)
402 16. Finite-strain elasto-plasticity

Since 'IjJ is an isotropic function of be, it can be expressed in terms of A~, or


equivalently, of E~:

'IjJ(be,~) = ~((An2, (A~)2, (A;)2,~) = ~(E~, E~, E;,~)


Equation (16.15) gives:

e 2 â~ dE~ )
f3A = 2(AA) âE~ d(A~)2 (no sum

Using the foUowing result:

l [(Ae )2] dE~ 1


EA
e
= "2 ln A '* d(A~)2 = 2(A~)2'
it appears that the principal Kirchhoff stresses f3A are related to the principal
elastic logarithmic stretches E~ by the foUowing simple relations which are
form-identical to those of the infinitesimal strain case:

Le. (16.18)

16.1.4 Flow rules

Similarly to the infinitesimal strain theory (Sect. 12.6), the flow rules can
be derived from Hill's maximum dissipation principle which can be stated as
follows.
Let 5 designate the elasticity domain, a set of admissible states of Kirch-
hoff stress T and hardening stress q:

5 = {(T, q), Tij = Tji I f(T, 71) ~ O} (16.19)

For a (plastically deformed) body at given current and intermediate config-


L .
urations, with prescribed rates (b e , ~), Hill's principle states that among aU
possible admissible states (T,q), the solution (T,q)
is the one for which the
mechanical dissipation is maximum:

v(T,qJ;e,~) = ~ax {v (T,qJ;e,~)} (T,q}E-


'" (16.20)

Using Eq. (16.12) and following the same procedure as in Sect. 12.6, it is
found that the solution to this optimization problem is given by:

1 bIe bc - _ 1 â f( )
-"2' - .
'Y âT T,q
. â
~ = -"y-f(T,q)
âq
"y~0, f(T,q)~O, "yf(T,q) =0 (16.21)
16.1 First theory 403

Equation (16.21a) is the plastic ftow rule, Le. the finite strain counterpart of
the well-known infinitesimal rule: €'1' = "raf/ar. In order to illustrate this
point, we shall prove the following result:

~ (In J1') = tr ("r ~~); with J1' =:: det F1' (16.22)

An important consequence is that if the yield criterion is pressure insensitive,


(Le. tr ("raf Jar) = O), then plastic deformat ion is isochoric (J1' = 1).
We shall now prove Eq. (16.22). First, setting:

r =:: det F e = (det be )1/2


and using Eq. (16.9) with Je instead of 1/;, we obtain:
. = tr
Je [aJe
ab e ' be . (2d + bLe . be-l)]

A well-known result used several times in Chaps. 14 and 15 gives:

Consequently, we obtain:

Now since J =:: det F = Je J1' and j = Jtr d (Eq. (14.66)), we have:

d
-(lnJP) d-.1af)
= -j - -je = tr (d) -tr ( -
& J P ~

Therefore, we have proven Eq. (16.22). Finally, using Eqs. (16.7-8), the flow
rule (16.21a) can be written in the material description as follows:

(16.23)

with:
T af
N=::F . ar ·F,
being interpreted as the pull-back to the reference configurat ion of the normal
(af Jar) to the yield surface.
404 16. Finite-strain elasto-plasticity

16.1.5 Elastic predictor

Similarly to the infinitesimal strain case, a two-step algorithm is used in


the finite strain regime. First, a trial state (elastic predictor) is computed
by assuming that no plastic deformation develops during the time interval
[tn, t n+1l. This amounts to assuming -using Eq. (16.23)- that:

Cf = constant = C~; t E [tn, tn+1l


The corresponding trial elastic left Cauchy-Green strain is given by (16.7) as:

b~ tr = F t . C~-1 . Fi (16.24)

The trial Kirchhoff stress T~r is then computed from the constitutive equa-
tions (16.11) or (16.14), (16.18). If the trial state verifies the yield condition
f( T~r, qn) ::; O, then the computation for this time step is completed, other-
wise plastic flow has occurred and the elastic predictor has to be corrected
as will be explained later.
Before dos ing this subsection, we report an alternative expression of b~ tr
which will be useful later. Equations (16.24, 1) imply that:

(16.25)

It is easily checked that ft is the relative deformat ion gradient between tn


and t:

f = aXt = aXt . ax = F t . F- 1 (16.26)


t aX n ax aX n n

where the notations are those of Chap. 14, i.e. X is the position vector of a
partide in the reference configurat ion and Xt and Xn those in the deformed
configurations at times t and tn, respectively.

16.1.6 Time discretization of the plastic flow rule

For infinitesimal strains (Chap. 12), we used a backward Euler time dis-
cretization of the flow rule. For finite strains, Simo uses an exponential ap-
proximation. RecaB that the solution of the first order linear problem:

is given by:
z(t) = exp[(t - tn)Al . zn
Based on this result, Simo's exponential approximation to the flow mIe
(16.23) within the time interval [tn, tl c [tn, tn+ll is the following:
- p-l
Ct = exp( -2..11'tCi 1 . Nt}
- -1
. CP n (16.27)
16.1 First theory 405

where a superposed - refers to the algorithmic approximation, Ll'Yt == 'Yt - 'Yn


and the subscript t refers to quantities evaluated at time t E [t n , t n+1]'
Using Eqs. (16.7, 24), result (16.27) can be transformed to the current
configurat ion as follows:

(16.28)

Using the following property:

F· exp(A) . F- 1 = exp(F· A· F- 1 )

together with the expression of N, the discretized flow rule (16.28) can be
written under the following simple form:

(16.29)

Following Simo, we now show that the algorithm has two nice properties,
which are the exact preservation of the constraint det(bD > O and of the
plastic volume. The proof for the first property proceeds as follows.
Applying the well-known result (e.g., (Gurtin, 1981))

det[exp(A)] = exp[tr(A)]

to Eq. (16.29), it is seen that:

det(b;) = exp { -2Ll'Yttr [:r!(Tt,iit)]} det(b~tr) (16.30)

Taking the determinant on both sides of Eq. (16.25a), we find that:

Assuming that det(b~) > 0, we have det(b~ tr) > O. Consequently, it is seen
that det(b;) > 0, Vt E [t n , tn+l]'
We shall now prove that the algorithm preserves plastic volume for pres-
sure insensitive yield criteria. Using the following identities

together with (16.30), it is seen that for pressure-insensitive yield criteria

i.e. ( Jt)2 = (~)2


lf JJ;'

and therefore li = JJ;.


406 16. Finite-strain elasto-plasticity

16.1. 7 Return mapping algorithm in principal stresses and strains

In the upcoming developments, one of the better features of Simo's algorithm


will be proven: it is form-identical to its well-known infinitesimal strain coun-
terpart, when expressed in terms of principal Kirchhoff stresses and elastic
logarithmic stretches.
First, the yield function being isotropic, it can be expressed in terms of
principal stresses: f(T, q) = Î(f31, f32,f33,q), and Sect. 15.8.3 shows that the
following result holds:

(16.31)

Equation (16.29) then gives the spectral decomposition of b e tr as:

b~ tr = exp [2Ll'Yt! f(-ft , qt)] . b:


~ e 2 exp [ 2Ll'Yt âf3A
~(AA) âÎ n t(A) ® n t(A) , 1
were we used Eqs. (16.13,31) and the orthonormality properties.
On the other hand, designating the trial principal elastic stretches by A~tr
and the principal directions by n: r (A), the spectral decomposition of betr can
be written as:
3
bte tr -_ ~ (\ e tr)2 tr
L..J A A nt
(A) 10.
'OI
tr
nt
(A)

A=l
From the uniqueness of the spectral decomposition, it is concluded that:

(16.32)

The important implicat ion is that the eigenvectors n~A) of the elastic left
Cauchy-Green strain b~ coincide with those (n!r (A») of the trial guess b~ tr.
Taking the logarithm on both sides of Eq. (16.32b) and using array nota-
tion (16.17), we obtain:

e
e t = et
e tr A âÎ(a
- Ll'Yt â{3 tit' qt
) (16.33)

This reiat ion shows that the finite strain algorithm is form-identical to that
of the infinitesimal strain case (Chap. 12) with ee designating the principal
elastic logarithmic stretches in the former case and the infinitesimal elastic
strains in the latter.
16.1 First theory 407

In summary, the return mapping algorithm can be stated as follows: Given


the values of the variables at tn and that of ee tr, compute {3n+ l' e~+ l' ~n+ 1,
qn+1 and ')'n+1 such that the following system of equations is satisfied:
8 h

8e e 'l/J(e~+1' ~n+1),
8 h

8~ 'l/J(e~+1' ~n+1),

= ee tr - .::1')' :{3 Î ({3n+1 , qn+ d


8 h

~n -.::1')' 8q!({3n+1,qn+1)

= O (16.34)
Formally, there are 9 scalar unknowns and 9 scalar equations. In practice
however, the problem can be greatly simplified. For instance, it will be shown
in Sect. 16.1.10 that for h plasticity and a free energy function which is
quadratic in terms of the principallogarithmic elastic stretches, the problem
is reduced to solving one scalar equation for a single scalar unknown.

16.1.8 Algorithmic tangent moduli

The objective of this subsection is to compute the spatial tangent operator c


which is defined by (see Sect. 15.1.3):
8S/J
Cijkl = 2FiIFjJFkKFlL 8CKL ' (16.35)

where S is the second Piola-Kirchhoffstress and C the (total) right Cauchy-


Green strain. Introducing the following notation:

M(A) = _l_ N
- (AA)2
(A) ® N(A)
,-
meA) = neA) iO.
'0/
neA)
,
(16.36)

we have seen in Sect. 15.8.4 that for the principal stretch formulation of
isotropic hyperelasticity, the spatial elasticity tensor is given in component
form as follows:

(16.37)

where the term 8Mi1 l /8CK L is computed in (Simo and Taylor, 1991). In
tensor notation, Eq. (16.37) can be rewritten as follows:

(16.38)
408 16. Finite-strain elasto-plasticity

where the expression of the fourth-order tensor c(II)(b) is given in (Simo and
Taylor, 1991).
It is important to keep in mind that expressions (16.37-38) were obtained
for an elastic material, thus n(A) and N(A) are the principal directions of
the totalleft and right Cauchy-Green strains b and C, respectively, and .hA
are the total principal stretches. The derivation of an expression for c for an
elasto-plastic material relies on the following observations:
- (i) In general, the directions of be and b or CP and C do not coincide,
but incrementally, time-discretization (16.27) implicitly assumes that the
incremental permanent strain is collinear with the current stresses and
therefore with the current elastic strains (and recall that b~+1 is collinear
with be tr).
- (ii) The algorithm defines the principal Kirchhoffstresses (lA as an implicit
function of the trial elastic logarithmic strains fir.
- (iii) In the trial state, the intermediate (unloaded) configurat ion is held
fixed (modulo a rigid rotation) and no plastic flow is assumed to occur
during the time interval [t n , t n+1], thus fÎJtr are function of the total de-
formation.
With these observations, the derivation of the tangent operator c proceeds
along the same lines as for elastic materials, and an algebra similar to that of
Sec. 15.8.4 leads to the following generalizat ion of Eq. (16.38) to the elasto-
plastic case:

L::: L::: a'1B mtr (A) ® m tr (B) + 2 L::: (lA (n+1)c(Il)(b e tr),
3 3 3
C = (16.39)
A=lB=l A=l

where:
_ ( â(lA )
ep
a AB = âf e tr n+ 1
B
(16.40)

is the 3 x 3 matrix of elasto-plastic moduli obtained by exact linearization of


algorithm (16.34).

16.1.9 Summary of the algorithm

We choose to keep as a history variable -in addition to e- the Green-Lagrange


plastic strain eP defined by:
1
e P ::-(CP-l) (16.41)
2
This choice is liseful for output purposes and for comparison with the in-
finitesimal strain case. Consider a time interval [t n , tn +11. A converged so-
lut ion (at local and structurallevels) at tn is known and the problem is to
16.1 First theory 409

find a solution at t n +1' At a given quadrature point, the data is: F n, F n+b
e~, ~n and the material parameters. The problem is to compute 7"n+1, e~+1'
~n+1 and e n +1' The algorithm proceeds as follows.

- (1) Compute the plastic right Cauchy-Green strain at tn:

and its inverse.


- (2) Compute the trial elastic left Cauchy-Green strain at tn+I:

be tr = F n+1 . (C~)-l . F~+1


- (3) Compute the principal values of be tr by solving a cubic equation in
closed form using Cardan's formulae (see Appendix B). Perturb the values
if they are equal or tooc1ose:-€omput~-their square-roots to obtâin the
trial principal elastic stretches at t n +1:

AA tr , (A = 1,2,3)
- (4) Compute the trial principal elastic logarithmic stretches at t n +1:

tA tr = In(AA tr ), (A = 1,2,3)
- (5) Compute the trial principal Kirchhoff stresses at t n +1:

f3Atr -- ~.Î.(te
8 e 'P 1 tr , te2 tr , te3 tr , <"n
C )
tA

- (6) Test yield condition at tn+I:

Î (f3fr, f3~r, f3~r , qn) ::; O ?

If the answer is yes, skip step (7) and proceed directly to step (8). Other-
wise, go to step (7).
- (7) Return mapping algorithm to compute: /3n+1' €~+1 and ~n+1' See Eqs.
(16.34) for the general case and Sec. 16.1.10 for the particular case of a
quadratic logarithmic free energy model and h flow theory.
- (8) Compute (3 x 3) elasto-plastic tangent matrix:

ep
a n+ 1 = (8/3)
8e e tr n+l

- (9) Compute tensor product of principal directions of be tr:


410 16. Finite-strain elasto-plasticity

- (10) Corn pute Kirchhoff stress at t n +1:


3
"f3 tr (A)
T n+l = ~ A (n+l)m
A=l

- (11) Cornpute (6 x 6) stiffness rnatrix at t n +1:

C -
n+1 - "
3
"
~ ~
3
ep
a AB (n+1) m
tr (A) /O>,
'0/
m tr (B) + 2"
~
3
f3 A (n+l) c(1I)(b e tr)
A=lB=l A=l

- (12) Cornpute (b~+1)-l, the inverse of elastic left Cauchy-Green strain at


t n +1:
3
(b~+1)-l = L exp[-2f (n+1)]m tr (A)
A=l
- (13) Cornpute plastic right Cauchy-Green strain at t n +1:

C:+1 = F~+1 . (b~+1)-l . F n+1


- (14) Update Green-Lagrange plastic strain at t n +1

p
€n+1 = '12 (P
C n+1 - 1)

Note that if a function for hyperelasticity in principal stretches already


exists, steps (9) to (11) becorne trivial: it suffices to call that function with the
following argurnents: be tr (instead of b), f3A = f3A (n+1) and "(AB = a1B (n+1)
(see Chap. 15).
AIso, if a function for infinitesimal elasto-plasticity exists, steps (6) to (8)
are reduced to calling that function with minor changes due to the fact that
stresses and strains are stored here in 3 x 1 arrays instead of 6 x 1, and only
a 3 x 3 rnatrix for a ep is needed instead of a 6 x 6; see Sec. 16.1.10 for an
example.

16.1.10 Application: Quadratic logarithmic free energy and J 2


flow

Consider the following uncoupled expression of the free energy function,


quadratic in the principal elastic stretches (= ln(Ă»: <4
1ÎJ(€e,~) = ~(I\: - ~G)(f~ + f~ + f~)2 + G[(fD 2 + (f~)2 + (f~)2] + h(~),
(16.42)
were 1\: and G are the elastic bulk and shear moduli, respectively. Following
Simo, it will be shown that for this particular choice of the free energy func-
tion and for h flow theory, the return mapping algorithm of Sec. 16.1.7 boils
16.1 First theory 411

down to the simple radial return algorithm 01 the injinitesimal strain case
(Chap. 12). The stress-strain relations (16.18) give:
2
(3A = (~ - 3G)(tr ee) + 2Gt:Â (16.43)

Introducing the following matrix notation:

(16.44)

relations (16.43) can be written under the following matrix forms which are
lorm-identical to linear elasticity in principal stresses and strains,

f3 = ae e , where a = 2GI- + (~- 3G)11


2 --T
(16.45)

The trial principal Kirchhoff stresses at tnH are simply:


{3tr = aee tr
Multiplying both sides of the discretized flow rule relation (16.34c) by the
constant matrix a, we obtain:
â
f3 n+1 = {3tr - A

L1"(a â{3f({3n+ll qn+d (16.46)

For h elasto-plasticity, the yield criterion is the following (Chap. 12):

f(r,q) == J2(r) - (Jy - q(€) ~ 0, (16.47)

where (Jy is the initial yield stress (a material parameter) and J2 (r) the von
Mises equivalent stress defined by ({3 A being the principal values of r):

J 2(r) [~dev (r) : dev (r)]1/2

J2 [((31 -
""2 (32) 2 + ((32 - (33) 2 + ((33 - (3d 2]1/2
= Ildev ({3)11 (16.48)

where dev (.) designates the deviatoric part of (.). It was already established
in Chap. 12 that:
âi (r,~) = ~ dev (r)
âr 2 J2(r)
This result carries on as is to the principal values {3A of r:

_ âÎ 3 dev ({3)
v({3) = â{3 ({3, €) = 211dev ({3) II (16.49)
412 16. Finite-strain elasto-plasticity

Multiplication with matrix a of Eq. (16.45b) gives:

(16.50)

Therefore, the fundamental algorithmic equation (16.46) becomes:

(16.51)

By comparison with Sect. 12.10.2, it is clear that the return mapping algo-
rithm in finite strains is identical to the classical radial return algorithm of
the infinitesimal strain case. The results of Sect. 12.10.2 can be used as they
are, with one further simplification: 3 x 1 arrays for {3, ee, etc., can be used
instead of 6 x 1 arrays, and 3 x 3 tangent matrices are computed instead of
6 x 6. Application of the algorithm of Sec. 12.10.2 shows for instance that
the normals to the yield surface at trial and final states are identical:

(16.52)

and the plastic correction phase is reduced to solving one scalar equation
with the sole unknown ~n+1:

(16.53)

This equation can be solved iteratively as explained in Sect 12.10.2, and once
the iterations have converged, the final solution at t n +1 can be updated as
follows:

IIdev ({3n+1)11 O"y+ q(~n+d (16.54)

dev ({3n+1) ~lldev ({3n+1)lIv({3tr) (16.55)


1 tr -
{3n+1 = dev ({3n+1) + 3(tr {3 )1 (16.56)

e~+1 = ee tr - (~n+1 - ~n)v({3tr) (16.57)

The elasto-plastic tangent matrix a ep is also taken directly from Sec.


12.10.3 as the first (3 x 3) block of the (6 x 6) consistent tangent matrix.
Consequently, the expression of a ep is as follows:

(2G)2 T (2G)2 Ll~


a - 3G + ~ vv - 3GLl~ + IIdev ({3)1I
3- - 211
x ( 21 1-T - vv T) (16.58)
16.2 Second theory 413

16.2 Second theory

16.2.1 Additive decomposition of the rate of deformation and


hypoelasticity

We have seen in Chap. 12 that for infinitesimal elasto-plasticity, two key


assumptions are the following:

where €, Ee and €p are the total, elastic and plastic strain tensors, respectively,
(Tis the stress tensor and c Hooke's operator. In rate form, we have:

The second formulat ion of finite strain elasto-plasticity we shall examine in


the upcoming developments is based on a straightforward extension of the
above relations as follows:

(16.59)

where d, de and ~ are the total, elastic and plastic rate of deformation ten-
sors, respectively and ;. is an objective rate of the Kirchhoff stress tensor r
(see Sec. 14.8). It appears that this formulat ion is based on an additive split
of the rate of deformation into elastic and plastic parts, Eq. (16.59a), and
on hypoelasticity, Eq. (16.59b). This formulation is widely used in numerical
analysis, although -as alluded to in Sect. 15.1.3- it is objectionable on fun-
damental grounds (and to make matters worse, the elasticity operator c in
Eq. (16.59b) is usually taken to be the constant Hooke's operator of linear
elasticity). From a practical point of view however, experience shows that for
monotonie loadings, hyperelastic- and hypoelastic-based formulations basi-
cally lead to the same results.
We recall hereafter some results from Sec. 14.8.2. We designate by TI a
symmetric second order Eulerian tensor (e.g., r) and consider objective rates
ofthe form:

(16.60)

where W is a skew-symmetrie tensor chosen so that ;, transforms objectively.


Given W, we may generate a group of rotations 'R. such that:

'k == W . 'R., and 'R.(t=O) =1 (16.61)

Now rotate TI with 'R. to find TI'R. as:

(16.62)
414 16. Finite-strain elasto-plasticity

The following key results were established in Sec. 14.8.2:


.
TI'R, = 'DT
"- . TI* . "-,
'D
or *..... 'DT
TI= "-. TI'R, . "- (16.63)
This meaus that a rather complicated objective stress rate can be computed
as the simple tirne derivative of a rotated stress (the rotation being 'R, as
defined above). Using (16.63a), Eq. (16.59b) becomes:
(16.64)

Repeated use is made ofthe identities: 'R,.'R,T = 'R,T.'R, = 1, in the following


manner:

---------......-....
d~p = (RqrRmr) (RptRst)d=m
ISmq ISop

Consequently, Eq. (16.64) becomes:


(rn)ij = (RkiR/jRptRqrCklpq)(RstRmrd:m) (16.65)
And this can be rewritten as:
(16.66)
where c'R, is the rotated Hooke's operator and d'R, aud d P are the rotated
:lf
total aud plastic rate of deformation tensors, respectively. tensor notation,
Eq. (16.66) takes the simple form:
(16.67)
This equation is jorm-identical to its infinitesimal strain counterpart except
that alI tensors here have to be properly rotated with 'R,. In order to clarify
the ideas, we give a summary of the constitutive equations in the rotated con-
figuration for an elasto-plastic model with a single scalar hardening variable
(Chap.12):
T'R, = c'R, : (~ - d~); q = q({)
f(T'R" q) ~O

dR = 18T'R,
'8 - f(T'R"q); ~ = -1' 88 f(T'R"q)
q
l' > O if f = O aud j = O;
l' = O if f < O or f = O and j < O (16.68)
Note that for isotropic materials: C'R, = c, and for anisotropic materials, the
rotation of C -Eq. (16.65)- can be be written under the following 6 x 6 matrix
format (see Appendix C):
(16.69)
Finally, note that for J2 elastoplasticity, f(T'R"q) = f(T,q).
16.2 Second theory 415

16.2.2 Computation of the strain increment

Time discretization of the basic constitutive equation (16.67) over a time


interval [t n , t n +1] with a generalized mid-point rule gives:

(16.70)

where the following notation was used:

,1(e) == (e)n+1 - (e)n;


(e)n+a = (e)(t=t,,+a);
t n +a = (1 - a)t n + at n +1, a E [0,1] (16.71)

Incremental stress-strain relation (16.70) can be rewritten in a form similar


to its infinitesimai strain counterpart (Sec. 12.10.2) as follows:

(16.72)

where ,1€'R., (n+a) is a rotated strain increment defined as:

(16.73)

We now need to compute the strain increment

(16.74)

We have established in Sec. 14.4 the following reIat ion (C = FT . F being


the right Cauchy-Green strain):
. T
C= 2F ·d·F

Time discretization of this reIat ion gives:


. T
,1C = (C)n+a,1t = 2F n+a . d n+a . F n+a,1t,

or, extracting (dn+a,1t),

,1€ = 21 F;;:ra . (,1C) . F;;:!a (16.75)

In order to gain further insight into the "strain increment" given by


(16.75), we shall prove the following identity (Hughes, 1984):

(16.76)
416 16. Finite-strain elasto-plasticity

First, we explain the notation. Consider two infinitesimally close material


particles at positions (X) and (X +dX) in the reference configuration. Vector
dX transforms onto dz n , dZ n+a and dZ'{;H in the deformed configurations
at times tn, t n+a and tn+b respectively. The deformation gradients at those
times are defined as follows (see Sec. 14.2):

dZn = F n . dX, dZn+a = F n+a . dX, dZ n+1 = F n+l . dX, (16.77)

from which the following equalities are deduced:

IldZnH 11 2 - IIdzn ll 2 = (F nH . F;;!a . dzn+a)T . (F nH . F;;!a . dzn+a )


-(Fn' F;;!a . dzn+a)T . (F n' F;;!a . dzn+a )
dz~+a . F;;Ja . (CnH - Cn) . F;;!a . dZ n+a
= dz~+a' (2dn+aL1t) . dzn+a,
using (16.74-75). We have thus proven (16.76). That equatiOn gives the fol-
lowing interpretat ion of the strain increment (d n+a L1t): it coincides with the
incremental Green-Lagrange strain if a = O and the incremental Almansi-
Euler strain if a = 1, see Sec. 14.3.2. Mid-point strain corresponds to the
case a = 1/2.
Following Hughes (1984), we shall now prove a third equivalent expression
for the strain increment:

A basic kinematic assumption is that of a linear interpolation of the current


positions over each time interval [t n , tnH]:

Zn+a = (1 - a)zn + aZnH, a E [0,1] (16.79)

Differentiation w.r.t. X gives the following relation between the deformation


gradients:

F n+a = (l-a)Fn +aFnH (16.80)

On the other hand, we have the following equality:

F n+1 - Fn = G a · F n+a, (16.81)

which is easily proven using the definit ion of G a :

G - â(ZnH - Zn). âX _ (F _ F ). F- 1
a - âX â - n+l n n+a
Zn+a
From Eqs. (16.80-81), the following relations are deduced:
16.2 Second theory 417

F n+l = F n+a + (1 - a)Ga . F n+a; F n = F n+a - aGa' F n+a, (16.82)

which lead to the following equalities:

C n+l - C n = F~+l' F n+l - F~ . F n


= F~+a' [Ga +G~ + (1- 2a)G~. GaJ· F n+a
Using Eq. (16.75), the desired result (16.78) immediately follows.
It is shown in (Hughes, 1984) that second-oroer accuracy is achieved for
a = 1/2; this corresponds to the mid-point strain:

1 T _ 8(Zn+l - zn)
dn+l/2Llt = -2 (G1/ 2 + G 1 / 2); G 1/ 2 = ---'-8~"::""-----'-:":" (16.83)
Zn+l/2

To summarize, the strain increment can be computed as follows:


Given F n aud F n+l> compute C n = F~ . F n, C n+ 1 = F~+l . F n+l,
F n+a = (1 - a)F n + aF n+l, F~!a, aud finally Lle from Eq. (16.75).
Now, it remains to compute the rotations 'R,n, 'R,n+l aud 'R,n+l/2' in order
to be able to compute the rotated variables which appear in the incremental
constitutive relation (16.70). There are two possible cases, depending on the
objective stress rate chosen:
(i) 'R, = R, where R is the rotation corresponding to the polar decompo-
sition F = R· U. This case is treated in Sec. 16.2.3.
(ii) Given a skew-symmetric matrix W, 'R, needs to be computed by
integrat ion of the differential equation (16.61). This problem is examined in
Sec. 16.2.4.

16.2.3 Polar decomposition algorithm

For the Green-Naghdi-Mclnnis objective stress rate (Sec. 14.8.1), the rotation
'R,coincides with R, the polar decomposition rotation defined by: F = R·U,
where U is the right stretch tensor. In the forthcoming developments, au
algorithm for the computation of R is presented. The algorithm is also useful
when R is needed for other purposes or when U is needed in order to define
various strain measures as in (Ogden, 1984).
The principal values (A~, A~, A~) of C = F T . F can be computed in closed
form using Cardau's formulae (see Appendix B). The principal directions
(N(l), N(2), N(3)) of C form an orthonormal basis, aud as noted in Sect.
15.4.1, their tensor products can be computed in closed form as follows:

where (A,B,e) is an even permutation of (1,2,3). Since C = U 2 , then


as seen in Sec. 14.2.3, the principal directions of U are those of C and its
418 16. Finite-strain elasto-plasticity

eigenvalues are (Al, A2' A3) (the principal stretches). Consequently, U can be
computed from its spectral decomposition as folIows:

~ (C - Ai:,t) . (C - A~t)
U = L..J AA (A2 _ A2 )(A2 _ A2) , Al f; A2 f; A3, (16.84)
A=l A BAC

This result is valid when alI three s"retches are different. When two of them
coincide (Al = A2 f; A3), the spectral decomposition of U becomes (e.g.,
(Gurtin, 1981)):
U = Alt + (A3 - Al )N(3) ® N(3)
A similar expression for C allows to compute N(3) ® N(3):

N(3) N(3) = C - A~t


® A32 - A21

Since N(3) is a unit vector, we have:

N(3) ® N(3) = [N(3) ® N(3)j2 = (CA3 - Ai:)2


2-
Al
This allows to put U under a form similar to (16.84):
_ (C - A~t)2
U - Alt + (A3 - Al) (A~ _ AV2' Al = A2 f; A3 (16.85)

When alI three principal stretches are equal, U is simply given by:
U = Alt, Al = A2 = A3 (16.86)
Remarkably, it is shown in (Hoger and Carlson, 1984) and (Ting, 1985)
that alI three cases (16.84-86) can be combined into a unique, singularity-free
formula:

(16.87)

where (it, i 2 , i 3 ) are the principal invariants of U which are computed from
the principal stretches as folIows (see Sec. 15.2):
(16.88)
The folIowing result, which is easily obtained after some algebraic manipu-
lation, allows to check that the expression of U given by (16.87) is indeed
singularity-free.
i l i 2 - i 3 = (Al + A2)(A2 + A3)(A3 + At) > O (16.89)
Using the Cayley-Hamilton theorem, the inverse of U is given by:

U- l = ~(C
Z3
- ilU + i 2 t) (16.90)

In summary, the polar decomposition algorithm has the folIowing steps:


16.2 Second theory 419

- (1) Given F, compute C = F T . F and C 2 •


- (2) Using Cardan's closed form formulae (Appendix B), compute the eigen-
values (Ai, A~, A~) of C. Compute their square roots: the principal stretches
(Al, A2, A3)'
- (3) Compute U from Eq. (16.87), using also (16.88-89).
- (4) Compute U- l from Eq. (16.90).
- (5) Compute R = F· u- l .

16.2.4 A time-integration algorithm for the rotation matrix

For objective stress rates such as Jaumann's, given a skew-symmetric matrix


W(t), we need to find a rotation 'R.(t) solution of the differential equation
(16.61). With a time-stepping scheme, the problem becomes: for each time
interval [tn, t n+1], given skew-symmetric matrices Wn+"" a E [0,1], and a
proper-orthogonal matrix 'R. n , find a proper-orthogonal matrix 'R. n +1 such
that:

(16.91)

Time-discretization of Eq. (16.91) using an exponential scheme as in Sect.


16.1.6 leads to:

'R.n+1 = (exp W) . 'R. n; W == W n+",Llt; a E [0,1] (16.92)

It can be shown (e.g., (Gurtin, 1981)) that if A is skew-symmetric, then


(expA) is a rotation. Also, for any 3 x 3 matrix, det(expA) = exp(tr A).
So if (tr A = O), then det (exp A) = 1. Therefore, algorithm (16.92) gives a
matrix 'R. n +1 which is indeed proper-orthogonal.
Matrix (exp W) can be evaluated with the following formula
00 1
expW = "" -W n, (16.93)
L..J n!
n=O

which is similar to the well-known scalar function expansion. However, better


representations of (exp W) do exist as explained hereafter.
Introduce the following notation:

(16.94)

where w is called the axial vector of W. It can be checked that the following
property is satisfied for any vector u:

W·u=wxu (16.95)
420 16. Finite-strain elasto-plasticity

Since (exp W) is proper-orthogonal, it can be represented by way of the


Rodrigues formula. Several alternative expressions of the formula exist, for
instance the following (see, e.g., (Whittaker, 1937), (Goldstein, 1981) and
(Guo, 1981)):
_ 2 - -2. __ W IIwll
expW -1 + 1 + IIwIl 2 (W + W), w = IIwll tan-2- , (16.96)

where w is the so-called Gibbs vector, and W is the skew-symmetric matrix


associated with w by Eqs. (16.94). Finally, it is shown in (Simo and Hughes,
1998) that the well-known algorithm of Hughes and Winget (1980) can be
retrieved from the exponential algorithm which was presented here.

16.2.5 Application: the Jaumann ohjective stress rate


The Jaumann objective stress rate is widely used in practice despite its short-
comings (see Sect. 14.8.1). For this rate, the skew-symmetric matrix W is
equal to the spin tensoT W (see Sec. 14.4):
. 1 T
W=Wj w=-(l-l ), (16.97)
2
where 1 is the velocity gradient. As shown in Sect. 16.2.4, once the following
skew-symmetric tensor:
W = wn+aLlt (16.98)
is known, the rotation at t n +1 is given by:
'R. n +1 = (exp W) . 'R. n , (16.99)
where (exp W) is computed with the algorithm of Sect. 16.2.4. We consider
a = 1/2 and we shall show that a good approximation to the mid-point spin
is given by:
1 T â(x n +1 - xn )
W n+1/2 Llt = -2 (G1 / 2 - G 1/ 2 )j G 1/ 2 == â (16.100)
Zn+1/2
We recall the following reIat ion from Sec. 14.4:
F=I·F
Time discretization with the mid-point rule gives:
F n+1 - Fn = In+1/2 . F n+1/2Llt
This can be rewritten as:
In+1/2 Llt = (F n+1 - F n) • F~!1/2 = G 1/ 2 ,
using Eq. (16.81). Now, definition (16.97b) leads to the desired result (16.100).
Finally, as shown in (Simo and Hughes, 1998), the mid-point Totation is:

'R.n+1/2 = exp (~) . 'R. n (16.101)


16.2 Second theory 421

16.2.6 Summary
We summarize hereafter the different algorithmic steps for the numerical
implementat ion of finite strain elasto-plasticity formulated in rate form. For
a time interval [tn, t n+11 and a given quadrature point, the data is: F n, F n+1,
'R. n , T n , constitutive history variables at tn (e.g. ~n) and material parameters.
The problem is to compute T n+1, 'R. n+1, history variables such as ~nH and
the tangent operator C~~l' The step-by-step procedure is given below for the
integration parameter a = 1/2.
- (1) Compute the rotations 'R.n+1 and 'R.nH / 2 ; see Sect. 16.2.3 if 'R. = R
and 16.2.4 if not.
- (2) Compute the strain increment dnH / 2Llt; see summary after Eq.
(16.83).
- (3) Compute the rotated strain increment:

Ll€'R. (n+1/2) = 'R.~+1/2 . (d n+1/2 Llt ) . 'R.nH / 2


- (4) Compute the rotated Kirchhoffstresses at tn and t n+1:

T'R. (n) = 'R.~ . T n ' 'R.n ; T'R. (n+1) = 'R.~+1 . T n+1 . 'R.n+1

- (5) Compute C'R, (n+1/2)' the rotated Hooke's operator at t n+1/2' If the
material is isotropic, then C'R, = c, otherwise, the transformation is given
by Eq. (16.65), which can be written under the following 6 x 6 matrix
format (see Appendix C):

[c'R. (n+1/2)1 = [Q('R.n+1/2)lT[c][Q('R.n+1/2)1


- (6) Return mapping algorithm in the rotated configuration: salve the in-
cremental stress-stress relation:

T'R. (n+1) = T'R. (n) + c'R. (n+1/2) : (Ll€'R. (n+1/2) - Ll€!R),

and the other discretized relations in the rotated configurat ion (Sect.
16.2.1). In this rotated format, the equations and the algorithm are iden-
tical ta those for infinitesimal elasto-plasticity. The trial (rotated) stress is
defined by:

Tn (nH) = T'R. (n) + C'R, (n+1/2) : Ll€'R. (nH/2)

For classical isotropic J2 elasto-plasticity, for instance, the plastic correc-


tion phase is reduced ta solving one scalar equation with the sale unknown
~nH:

(16.102)

The rotated algorithmic elasto-plastic tangent c':Jl


(nH) is also found di-
rectly from the infinitesimal strain expression of Sec. 12.10.3.
422 16. Finite-strain elasto-plasticity

- (7) Rotate back the stresses and elasto-plastic tangent at t n+1 (also non-
scalar history variables if they exist):

= 7ln + 1 · T'R, (n+l) . 'R,~+1;


= [Q('R.n+1 )][c~ (n+1)][Q('R.n+1)lT

16.2.7 Incremental objectivity

We shall show hereafter that the algorithms for the computation of the strain
and stress increments obey the notion of incremental objectivity, which was
formalized by Hughes and Winget (1980) and which means that under a
rigid body motion, the strain increment must vanish and spurious stresses
are preduded. Consider a rigid body mot ion defined over a time interval
[t n , t n +11 by:
X n +l = c + Q . X n ,

where c is a constant translation and Q is a constant rotation which verifies:

Q. QT = QT . Q = 1; det Q =1
The deformation gradient at t n +1 is:

F n +1 =Q·Fn ,

and the right Cauchy-Green strain at t n +1 is given by:

C n+1 = F~ .QT ·Q·Fn = F~ ·Fn = C n


'"--"
1
Equation (16.75) then gives LlE = O and Eq. (16.73), LlE'R. = o. From Eq.
(16.72), with LlE'R. = LlE!R. = O, we obtain:

T'R. (n+1) = T'R. (n) (16.103)

This can be rewritten as follows:

'R.~+1 . T n+1 . 'R. n+1 = 'R.~ . T n . 'R.n, Le.


T n+1 = ('R.n · 'R.;;~l)T . T n · ('R.n · 'R.;;~1) (16.104)

It can be shown that (16.103) or (16.104) are equivalent to:

(16.105)

Le., T transforms objectively under a rigid body motion. Consider the case
'R. = R. Since F n+1 = Q . F n and U n+1 = U, we have: Rn+1 = Q . Rn.
Substituting in (16.104) and using the fact that Rn is proper orthogonal,
(16.105) immediately follows.
17. Cyclic plasticity

As explained in Chap. 12, isotropic hardening alone is unable to predict


the experimentally-observed behavior of metals under cyclic loadings (e.g.,
Baushinger effect and stabilized stress-strain loops). In this chapter, we
present the constitutive equations and the algorithmic treatment of a non-
linear kinematic hardening model which is useful for the description of cyclic
plasticity of metals.

17.1 One-dimensional model

Consider a metallic cylindrical specimen undergoing a cyclic uniaxial ten-


sion/ compression test under imposed total strain. The strain history is shown
in Fig. 17.1. The corresponding stress-strain response is shown in Fig. 17.2.
The yield "surface" is actually a segment collinear with the ull-axis, its size

,
" time

Fig. 17.1. Cyclic strain history

is 2(uy + R), where Uy is the initial yield stress and R the hardening stress.
Figure 17.2 shows that the center Oi of the yield segment is not situated on
the fll-axis; in other words the yield segment is not symmetrical w.r.t. the
fl1-axis. This cannot be predicted with an isotropic hardening model alone;
that is why a new variable X l1 -the kinematic hardening stress- which mea-
sures the translation of the center Oi of the yield segment along the Ul1-axis
is introduced; see Fig. 17.2. From that figure, the values of X l l and the ac-
424 17. eyctic plasticity

R+ay

R+ay

Fig. 11.2. Idealized stress-strain response to a eyclic strain history. Identification


of isotropic and kinematic hardening stresses, R and Xu.
17.2 Three-dimensional model 425

cumulated plastic strain p can be measured for each cyclej thus experimental
data points (p, X n ) are available. The following rate model was proposed:

Xu = (a - bXu)p, (17.1)

where a (> O) and b (~ O) are material parameters. The differential equation


can be integrated in time to give:
a
X u =1; [1 - exp( -bp)] if b # O and a # O
X l l =ap if b = O and a #O (17.2)

Kinematic hardening is linear if b = O and nonlinear if b # O. In the latter


case, parameters a and b can be identified from experimental data as shown
in Fig. 17.3. Equation (17.2a) shows that the nonlinear kinematic hardening
stress saturates for large values of plastic strain (Xn --t alb when p --t 00).
Finally, Fig. 17.2 shows that the following yield criterion holds:

p
Fig. 11.3. uniaxial kinematic hardening stress, Xll, versus accumulated plastic
strain, p. Identification of material parameters a and b.

(17.3)

17.2 Three-dimensional model

In this section, we present a generalization of the ideas introduced in Sec.


17.1 rrom uniaxial to multiaxialloadings. The model was initially proposed
by Armstrong and Frederick (1966) and further developed by Chaboche and
Marquis; see (Lemaitre and Chaboche, 1990).
We follow the procedure of Sec. 12.12 for non-associative plasticity. The
total strain is assumed to be the sum of elastic and plastic parts:
426 17. Cyclic plasticity

(17.4)

Internal variables are defined:

to which thermodynamic forces are associated:

(O",R,X)
'-v-'
A
The scalar variable r models isotropic hardening and the strain-like tensor
n models kinematic hardening. The scalar variable R measures the radius of
the yield surface in the space of deviatoric stresses while the so-called back
stress X measures the translation of the center of that surface in the same
space. A specific free energy is defined by:

(17.5)

where p denotes the material density and is assumed to be constant, and E


is Hooke's operator. The equations of state are derived according to:

a'ljJ â'ljJ a'ljJ


O" = p-;
alOe
R = p-;
ar
X = pan- (17.6)

Using (17.4-5), the following expressions are obtained:

O" =E : (€ - lOP); R = R(r); X = an (17.7)

An elastic domain is defined by the yield function:

f(O", R, X) = h(O" - X) - ay - R(r) :::; O, (17.8)

where J2 (0" - X) is the von Mises measure of (O" - X):

3 ] 1/2
J2 (0"-X)= [ 2(S-X):(s-X) , (17.9)

with s being the deviatoric part of the stress O" (the tensor X is deviatoric).
It can be checked that for a uniaxial stress state,

[a] = (
a11

~
O O)
O O , [X] =
( X 11
O
O
O (17.10)
O O O _.Ku.
2
17.2 Three-dimensional model 427

equation (17.8) -combined with (17.9)- simplifies to the 1D inequality (17.3).


The evolution laws for the plastic strain and the internal variables are derived
from a plastic potential, F(u, R, X), using the generalized normality rules:

.P _ . âF ._ . âF. . âF
€ - "( âu' r - -"( âR' o: = -"( âX' (17.11)

with the Kuhn-Tucker conditions:

i' 2': 0, i'I = 0, i' j = ° (17.12)

The plastic potential is chosen as follows:

b
F(u,R,X) = l(u,R,X) + -X: X (17.13)
2a
The fiow rules (17.11) now give:

. âl .3 (8 - X)
"( âu = "(2 Jz(u - X)'
. âl .
-"( âR = "(,

_i'(â l +!?.X)=i'(â l _!?X)=i:P_!?..Xi' (17.14)


âX a âu a a

Now, using (17.7 c) and the definit ion of the accumulated plastic strain rate

equations (17.14b-c) can be rewritten as follows:

r = i' = Pi X = ai: P - bX i' (17.15)

So, it appears that the internal variable r which models isotropic hardening
is the accumulated plastic strain p. Also, (17.15b) is the generalizat ion of
(17.1) from 1D to 3D.
In summary, the model is described by the following set of equations:

u E: (€ - €P),
I J2 (u - X) - R(p) - ay :S 0,
i:P pN,
X ai: P - bXp,
P > 0, p! = 0, pj = 0, (17.16)

where we introduced the following notation for convenience:


428 17. Cyclic plasticity

(17.17)

In practice, for isotropic hardening, a power law or an exponentiallaw (with


saturation) are often used:

R(P) = kpm, or R(P) = Roo[l- exp(-mp)) (17.18)

where k (2 O), m (2 O) and Roo (2 O) are material parameters.

17.3 Dissipation inequality

The inequaIity of Clausius-Duhem requires the dissipation to be non-negative


(see Sec. 14.9.3):
u : i: - P:t 1jJ( Ee, r, 0:) 2 O

Using the chain rule and equations of state (17.6), the inequality becomes:

u : i'P - Rf - X : it 2 O,
or, using flow rules (17.14),

b
u: N - R - X: (N - -X) 2 O
a

After some algebraic manipulation, this requirement takes the simple form:

b
ay + -X: X 2 O, (17.19)
a

and is always satisfied.

17.4 Plastic multiplier

When f = O, the consistency condition j = O holds. Using the same pro-


cedure as in Sec. 12.7, the following results are found, after some algebraic
manipulat ion,

(17.20)

where < x > designates the positive part of (x) and:


dR 3
h== N :E: N+- +-a-bN:X (17.21)
dp 2
17.5 Tangent operator 429

In this chapter, we assume that linear elasticity is isotropic, with G and K,


being the elastic shear and bulk moduli, respectively. Since E : N = 2GN,
the expression for p becomes:

(17.22)

with
3
h = 3G + -dR
dp
+ -a -
2
bN : X (17.23)

As in ehap. 12, we assume that h is positive (if b O, this is trivially


satisfied) .

17.5 Tangent operator

The tangent modulus, L, is defined by the rate relation: iT = L : i=.. It is


found that L has two possible expressions, E and H:

- (i) If f < O (elasticity) or f = O and j < O (elastic unloading), then p=O


and L = E (Hooke's operator).
- (ii) If f = O and j = O (plastic loading), then p > O and L = H.
Using the results of Sec. 17.4, it is found that elastic unloading corresponds
to the quantity (N : i=./h) being negative, while plastic loading corresponds
to that same quantity being positive. The elasto-plastic 'continuum' tangent
operator H is found to be:
1
H=E--(E:N)I8I(E:N) (17.24)
h
Since linear elasticity is isotropic, the expression simplifies to:

H = E - (2G)2 N 181 N (17.25)


h
Note that the expressions of p and H are form-identical to those of Sec. 12.7.

17.6 Hardening modulus

We consider the case of plastic loading. With the previous results and nota-
tions, we have:
N:E:i=.
p= iT = E : (i=. - pN) (17.26)
h
430 17. Cyclic plasticity

From the above relations, it is obtained that:

N : iT = P,(h - N : E : N),, (17.27)


..
h*

where h* is the so-called hardening modulus. It has the following expressions:


N · . .p
h* =~ = (1' :e (17.28)
p (P)2
The definit ion of h* together with the expressions of hand N leads to:

h* = ~: + ~a - bN : X (17.29)

Using lemma (1.1) and Eqs. (17.24,27), it appears that the tangent operator
H is invertible if and only if h* =f. O (if b = O, this is always the case). The
inverse of H is given by:

(17.30)

Exercise: Retrieve the results of Secs. 17.4 to 17.6 by directly applying the
notations and results of Sec. 12.12.

17.7 Return mapping algorithm


In this section, we present the return mapping algorithm which was proposed
by Doghri (1993). The method used is a strain-driven algorithm as described
in Chap. 12. AIso, as in that chapter, Ll designates an increment and alI
variables which do not contain the subscript (n + 1) are computed at t n +1.
The elastic predictor is defined by:

(1' = (1'tr = E : (e - e~) (17.31)

AlI other 'plastic' variables are equal to their values at tn. IT this trial state
does not satisfy the yield condition f ::; O, then the solution at t n +1 has to
satisfy the following system where the backward Euler scheme was used to
integrate the rate constitutive relations:

(1' = (1'tr - 2G Lle P


f J2 ((1' - X) - R(P) - O"y = O
LleP N Llp
LlX = aLleP - bX Llp (17.32)

Recall that eP , N and X are deviatoric tensorsj also:


17.7 Return mapping algorithm 431

3 {3
N = 2 J2 ({3) , with {3 == 8 - X

As in Sec. 12.10.2, the following relations hold:


tr (T = tr (Ttr = 311:tr €
They imply that only the deviatoric part of (17.32a) needs to be considered.
It is easily shown that problem (17.32) is reduced to finding two unknowns
{3 and p from the two equations:

k a_ tr 2GN d. (X n + aN d.p) = O
fJ 8 + P+ 1 + bd.p ,
f J2 ({3) - R(P) - ay = O (17.33)

When b = O, then N = N tr and it is well known that the problem can


be reduced to solving a scalar equation. However, if b =1- O, then N =1- N tr •
Solving the nonlinear system (17.33) with a Newton scheme, we obtain at
each iteration:

k + [[ + (2G + 1 +~d.p) (d.p) ~~] : c{3


+ [2GN +
aN - bXn]
(1 + bd.p)2 cp =O
dR
f +N : c{3 - ( dp )cp = O (17.34)

Using the expressions of N and J2 ({3), it is found that:

aN N N)
â{3 = J2 1({3) (3[deV
2 - ®
(17.35)

where:
1
I dev == 1 - -1 ® 1,
3
with 1 and 1 the fourth and second-order identity tensors, respectively. Since
3 aN
N: N = 2' and N: a{3 = O, (17.36)

the iterative system (17.34) gives the correction for p explicitly:


f-N:k
Cp = (17.37)

where the denominator halg is given by:

h alg - 3G dR 1 (3 bN . X ) (17.38)
= + dp + (1 + bd.p)2 2a - . n
432 17. Cyclic plasticity

We shall now prove that the correction for {3 can also be found explicitly.
Using the fact that c{3 is deviatoric, together with (17.35), Eq. (17.34a) can
be rewritten as folIows:
aN -bXn ]
B : c{3 = -k - [2GN + (1 + bLlp)2 cp , (17.39)

with:
3
- (1
B = + -g)1
2 - gN ® N (17.40)

where

a ) Llp (17.41)
9 == ( 2G + 1 + bLlp J2({3)
Using lemma (1.1), it is found that B is invertible if and only if 1 + (3/2)g i:-
O (which is always verified by an acceptable iterative approximation). The
inverse of B is given by:

-1 1
B = 1 + (3/2)g (1 + gN ® N) (17.42)

Using this expression, the correction for {3 is found explicitly:

We remark that the corrections over the elastic predictor have very simple ex-
pressions. Since for this trial state we have Llp = O and k = O, the corrections
are:
ftr
cp = h tr ; c{3 = -(2G + a)Ntrc p + bXncp

For linear kinematic hardening (b = O), expressions (17.37, 43) become very
simple:
f-N:k
h
1
c{3 = -1+(3/2)g[k+ g (N:k)Nj-(2G+a)Nc p (17.44)

with:
2G+a
9 = J2({3) (Llp)

If there is no kinematic hardening at alI (classical h flow theory of Sec.


12.2) we have a = O and {3 = s.
17.8 Consistent tangent operator 433

The iterative approximations to p and ţj are updated according to:


p(it+1) = p(it) + cp ; ţj(it+1) = ţj(it) + eţj
Once tolerance criteria on the residuals k and f are satisfied, a converged
solution is found. AlI variables can then be easily updated according to the
folIowing formulae:

Llt:P = NLlp
X = 1+
~Llp (X n + aLlt: P )
1
u = ţj + X + '3tr(utr)l (17.45)

Despite the complexity of the constitutive equations, closed-form expressions


were derived, without any approximations.
FinalIy, it is shown in (Hartmann and Haupt, 1993) and (Auricchio and
Taylor, 1995) that the return mapping algorithm can be reduced to solving
one scalar nonlinear equation. It remains to compare the computational cost
and the numerical behavior of the different algorithms.

17.8 Consistent tangent operator


When the local integration algorithm described in Sec. 17.7 has converged,
the corresponding consistent tangent operator Halg is computed as in Chap.
12 according to 8u = Halg : 8t:. IT the increment is entirely elastic (Le.,
if the elastic predictor is the solution), then Halg = E, otherwise, Halg is
computed as in (Doghri, 1993).
Differentiation of the incremental relations (17.34) w.r.t. ţj, p and e gives:

[1 + (Llp) ( 2G + 1 +a ) âN] [ aN - bXn]


bLlp âţj : 8f3 + 2GN + (1 + bLlp)2 8p
_2Gldev : 8t: = O
(17.46)

This system gives the expression for (8p) explicitly:


8 _ 2GN: eSt:
p- halg (17.47)

where h alg is given by (17.38). IT Llp ~ O then h alg ~ hand we retrieve the
expres sion ofthe plastic multiplier p given by (17.22). From Eq. (17.46a), the
following relation is derived:

tr(8f3) = O (17.48)
434 17. Cyclic plasticity

Equation (17.46a) can now be rewritten as:

r = - [2G N
B : uf3 + aN - bX n] r G de" : ue,
(1 + bL1p)2 up + 2 1
r (17.49)

where B is given by (17.40). Using the expression of B- 1 given in (17.42), it


is found that:

8f3 1+ ~~2)g[Ide" + gN ® N] : 8e - [2G + (1 + ~L1p)2] N8p


1 b
+ 1 + (3/2)g (1 + bL1p)2 [X n + g(N : X n)N]8p (17.50)

After differentiation of the following identity

1
X= l+bL1p(Xn +aNL1p),

and use of the expression of (8f3), it is found that:

Differentiation of (17.45e) gives:

8u = 8f3 + 8X + 1b(1 ® 1) : 8€ (17.52)

Using the expressions of 8X, 8f3 and 8p, an explicit expression of the consis-
tent tangent operator is obtained:

Halg = H mod _ (L1p)(2G)2 âN _ (L1p)(2G)2 x


[1 + (3/2)g] âf3 [1 + (3/2)g]
b 1 1 [3'2
(1 + bL1p)2 J2(f3) halg Xn - (N : Xn)N ® N
] (17.53)

where H mod is the modified eontinuum tangent operator defined by:

(17.54)

It is form-identical to the eontinuum operator H -Eq. (17.25)- with the de-


nominator being halg instead of h. If L1p ~ O, then Halg ~ H.
If kinematie hardening is linear (b = O), then the expression of Halg
simplifies to:
17.9 Numerical simulation 435

Halg = H _ (Llp)(2G)2 aN
(17.55)
[1 + (3/2)g] a/3
This simple and explicit reiat ion clearly shows that w.r.t. the continuum case,
there is an extra-term which depends on the finite plasticity increment (Llp)
and the curvature of the yield surface:

17.9 Numerical simulat ion


The constitutive equations presented in Sec. 17.2 have been implemented in
the general-purpose finite element programs ABAQUS (via a user subroutine)
and SPECTRUM using the algorithms developed in Secs. 17.7 and 17.8.
We consider hereafter the example of a notched bar subjected to a cyclic
4-point bending load. A state of plane strain is assumed. Figures 17.4 and
17.5 show the mesh used and the B.Cs. The loading history is depicted in Fig.
17.6. The material considered is a low-carbon (AISI 1010) steel in a laminated

A:(
11
4 -
dVi
HI-(1
-
-c:--.-
t-
1r

I'
p p

21.
,1;; • .1

Fig. 17.4. Notched bar under cyclic loading: mesh and boundary conditions.

state. The following values of material parameters have been identified: E =


210GPa, v = 0.3, O"y = 200MPa, Roo = 2 GPa, m = 0.26, a = 17GPa
and b = 21. The non-annealed initial state of the material was modeled by
introducing initial values of p, R and X:

po = 0.43; Ro = 211 MPa; Xo= [


-181
~
o
128
O
II (x,y,z)
MPa

Numbers (8,12, ... ,32) indicated in Fig. 17.6 correspond to the cumulative
numbers of time steps used in the simulation. Figure 17.7 illustrates the
time evolution of stress, strain and kinematic stress components at the point
436 17. Cyclic plasticity

.I
.-t._-_._.
I
2.

Fig. 11.5. Notched bar under cyclic loading: zoom showing the mesh and boundary
conditions at the root of the notch.

P N

6~+-----~----------~----------~r-----------~

O.L-..............................~..............................JL..............................~.........................~t. .

Fig. 11.6. Notched bar under cyclic loading: bending load VS. time.
17.9 Numerical simulation 437

marked M in Fig. 17.5, in the vicinity of the root of the notch. Because of
the nonlinear (the b-) term in the evolution law (17.16d) of the kinematic
hardening, the mean value of X tends to vanish cyclically (Fig. 17.7b), so
that the stress-strain loops tend to stabilize with a zero mean stress value
(Fig. 17.7a).
438 17. Cyclic plasticity

Fig. 17.7. Notched bar under cyclic loading: (a) U yy vs. Eyy stress-strain loops at
the point M in Fig. 17.5 and (b) time evolution of kinematic hardening components
Xxx, X yy and Xzz at point M.
18. Damage mechanics

In this chapter, we present an introduction to 'Damage Mechanics' which


is a local approach to fracture (whereas 'Fracture Mechanics' is a global
approach). A ductile damage model for metal plasticity is studied in detail.
Closed-form solutions, a computational algorithm and numerical simulations
are presented and discussed. Although by comparison to the constitutive
model of Chap. 17, only one scalar variable is added in order to describe
damage, we shall see that this adds a great deal of complexity to both the
constitutive equations and the computational algorithm.

18.1 Damage variable

In damage mechanics, the constitutive equations take into account the pro-
gressive deterioration of a material via an internal variable called damage.
The latter is a macroscopic measure of the microscopic degradat ion of a repre-
sentative volume element (RVE). The ultimate phase of the damage evolution
is detected by a local criterion and corresponds to the failure of the RVE,
and hence to a macro-crack initiation. The spatial evolution of a completely
damaged zone corresponds to to macro-crack propagation.
In this chapter, we use a definit ion of 'damage' which was initially pro-
posed by Kachanov in 1958 and Rabotnov in 1968, and further developed and
popularized by Chaboche, Hult, Leckie, Lemaitre, Kraczinovic and others in
the 1970's, and many more researchers since. See Sec. 18.11 for references.
Suppose that a RVE is an elementary parallelepiped depicted in Fig. 18.1,
and consider a facet of outward unit normal (n). A measure of 'damage' on
the facet is given by a scalar D(n) defined as follows:

AD
D(n) == Ao' (18.1)

where Ao is the total area of the facet and AD the flaws (or damaged) area.
We shall assume that damage is 'isotropic', i.e. the damaged state of a RVE
is measured by a single scalar variable (D),

D(n) = D, '<In (18.2)


440 18. Damage mechanics

Equation (18.1) shows that D E [0,1] and that the extreme values of the
damage variable are D = O for the sound material (A v = O) and D = 1 for
a RVE with null stress-carrying capability (A v = Ao). One method which is

Ao n

Av

D(n) = -t-

Fig. 18.1. Definition of damage, D(n) = AD/Ao, on a facet with outward unit
normal nj Ao: total area, AD: flaws area.

used to identify damage is the decrease in stiJjness:

D = 1- E v , (18.3)
Eo
where Eo and Ev are Young's moduli for the sane and damaged material,
respectively. Equation (18.3) shows that damage can be measured from load-
ing/unloading tests in a uniaxial tension experiment, as illustrated in Fig.
18.2. The figure also shows that the stress Uu is related to the elastic strain
lOII by:

(18.4)
Note that because (D) is not constant, the relation between stress and elastic
strain is not linear. Designating the tensile force by (F), a so-ca1led eJjective
stress, uu, is defined as the force divided by the resisting area of the cross
section:
_ F Uu
(18.5)
au == Ao - Av = 1- D'
where au = F/Ao is the true (Cauchy) stress. Equation (18.4) shows that
the effective stress is linearly related to the elastic strain:
(18.6)
The following evolution law for ductile damage was proposed by Lemaitre
18.1 Damage variable 441

fn
Fig. 18.2. Identification of damage, D = 1- (ED / Eo), by the decrease in stiffness;
Eo and ED: initial and damaged Young's moduli, respectively.

and Chaboche:
. y
D = - <jJ > if p ~ PD and D ~ Dc,
So
iJ =0 otherwise, (18.7)

where < • > designates the positi ve part of (• ), (P) is the accumulated plastic
strain, (PD) a damage threshold which can be related to the energy stored
in the material (Lemaitre and Doghri, 1994), (Dc) a critical damage value,
(So) a material parameter and (Y) designates the strain energy release rate.
It will be shown in Sec. 18.2 that (Y) has the following expression:

Y= _1_ [h(U)] R
2Eo 1- D
2
v,
2
with R V =3(1+v)+3(1-2v) [ UH ]
J2 (u) ,
2

(18.8)

where UH == u mm /3 is the hydrostatic stress. The variable Y and conse-


quently the damage rate iJ depend strongly on the triaxiality ratio U H / h (u),
which is known to be a main feature in failure. Indeed, as the stress triaxial-
ity increases, the ductility at fracture decreases (the material becomes more
brittle)j see Fig. 7 in (Lemaitre and Doghri, 1994).
In a uniaxial tension test, h(u) = Un, uH/h(u) = 1/3, Rv = 1 and the
damage evolution law (18.7) becomes:
2
. 1 ( Un ) .p
D = 2EoSo 1_ D f 11
(18.9)

The variation of the damage (D) versus the plastic strain (fiI) is plotted in
Fig. 18.3.
442 18. Damage mechanics

D
failure
De

PD
Fig. 18.3. Variat ion of the damage (D) versus the plastic strain (Eid in uniaxial
tensionj PD: damage threshold, De: critical damage.

18.2 Three-dimensional constitutive model

In this section, the damage model of the previous section is generalized


from 1D to 3D, and the equations are coupled with those of Chap. 17. In
other words, we consider a family of J2 plasticity models exhibiting non-
linear isotropic hardening, nonlinear kinematic hardening (Chap. 17), and
ductile damage (Lemaitre-Chaboche model).
We follow the procedure of Sec. 12.12 for non-associative plasticity. An
additive strain decomposition into elastic and plastic parts is assumed:

(18.10)

Internal variables are defined:

to which thermodynamic forces are associated:

(u,R, X, -Y)
~
A
As in Chap. 17, the scalar variable r models isotropic hardening and the
strain-like tensor a models kinematic hardening. The scalar variable R mea-
sures the radius of the yield surface in the space of deviatoric stresses while
the back stress X measures the translation of the center of that surface in the
same space. The internal variable D models the damage, which is assumed
to be 'isotropic'; its associated variable is denoted by Y.
A specific free energy, 1{;(ee, V), is defined by:

1
p1{;(ee, r, a, D) = -ee:
2
(1 - D)Eo : ee + -a
2 o
l
a : a + T R(~)d~ (18.11)
18.2 Three-dimensional constitutive model 443

where (Eo) is Hooke's operator of the sound material and (a) a material
parameter. The equations of state read:
81/1 81/1
O"=p-' A=p- (18.12)
8e e ' 8V
Using (18.10,11), the following expressions are obtained:

O" = (1- D)Eo : (e - eP ); R = R(r); X = au; Y = 2"1 ee : Eo : ee


(18.13)
It appears that the variable Y associated with the damage D is the strain
energy release rate. An elastic domain is den.ned by the yield function
i(O",A; V):
i(O", R, X; D) = J2 (0- - X) - ay - R(r) ~ 0, (18.14)
where ay is the initial yield stress, J2(0- - X) the von Mises measure of
(O- - X) and O- the so-called effective stress:

(18.15)

The evolution laws for the plastic strain and the internal variables are derived
from a plastic potential, F(O",A; V), using the generalized normality rules:
'p .8F· 8F
e = 'Y 80"'
V = -t 8A ' (18.16)

together with the Kuhn-Tucker conditions

t ~ 0, ti = 0, tj = ° (18.17)
The plastic potential is chosen as:
b
F(O",R,X, Y;D) = i(O",R,X;D) + 2aX : X + FD(Y;D) (18.18)

We introduce the following notation:


-_8i 8i 3 (S-X)
N = 80- = - 8X = 2 h(o- - X) , (18.19)

where s is the deviatoric part of 0-. If damage is inactive, N is the normal


to the yield surface in the stress space. The evolution laws (18.16) give:

i:P
t -
1_D N ,
i- t,
a = t(N- - -X),
b
a
.8FD
b 'Y 8Y
(18.20)
444 18. Damage mechanics

- --
Note that eP , N and X are deviatoric. Since N : N = 3/2, it is easily shown
that the accumulated plastic strain rate p is related to l' and r by:

l' = r = (1 - D)p (18.21)


So it appears that when damage is present, the isotropic hardening variable
r is not equal to the accumulated plastic strain p.
In summary, the model can be described by the following set of equations:
u = (1 - D)Eo : (e - eP ),
f J2 (u - X) - R(r) - O"y ~ O,
i=.P pN,
X = (1 - D)(ai=.P - bXp),
D· .8FD
= r 8Y'
r = (1 - D)p > O, r f = O, r j = O (18.22)
In practice, a power law or an exponentiallaw (with saturation) are often
used for isotropic hardening:
R(r)=krm , or R(r) = Roo[l-exp(-mr)] (18.23)
where k (~ O), m (~ O) and Roo (~ O) are material parameters. In practice
also, the damage potential FD(Yj D) is chosen as:

(18.24)

which gives the damage evolution as:


. Y
D= 8l~0, (18.25)

where 8 0 is a material parameter. A damage threshold (PD) and a critical


damage (De) can be used as in Eqs. (18.7).
In the remainder of this chapter, we will assume that linear elasticity is
isotropic. We then have, with Go and 11:0 being the elastic shear and bulk
moduli for the sound material, respectively,

= 2Go(! - 31 ® 1} + 11:01 ® 1, Eo
1 Eo
Eo G o = 2(1 + v)' 11:0 = 3(1 - 2/1)
..
Id.~

(18.26)
where (/, 1) are the fourth and second-order identity tensors, respectively,
and (Eo, /1) Young's modulus and Poisson's ratio ofthe undamaged material,
respectively. Using Eqs. (18.13d, 26) it can be shown that the expression of
(Y) is that given by (18.8). c
18.3 Dissipation inequality 445

18.3 Dissipation inequality


The inequality of Clausius-Duhem requires the dissipation to be non-negative

!
(see Sec. 14.9.3):
(T : i: - P 'I/J(ee, V) ~O
Using the chain rule and equations of state (18.12), the inequality becomes:

(T : i:P - A. 11 ~ O, Le
(T : i:P - Rf - X : it + Y D ~ O

Using flow rules (18.20), and after some algebraic manipulation, this require-
ment takes the simple form:
b 8FD
ay + -;;; X :X +Y 8Y ~ O, (18.27)

and is always satisfied.

18.4 Plastic multiplier


Using the same procedure as in Sec. 12.7, the following results are found,
after some algebraic manipulation,

j = 2GoN : i: - hp, (18.28)

where:
dR 3 - )
h == 3Go + (1 - D) ( dr +"2a - bN : X (18.29)

When f > O, the consistency condition j = O holds. This gives:

. r 2GoN: i:
P=l-D= h (18.30)

where we assumed that modulus h is positive. Note that this condition is


always satisfied if b = O.

18.5 Tangent operator


The tangent modulus, L, is defined by the rate relation: ir = L : i:. It is
found that L has two possible expressions, E and H.
(i) If f < O (elasticity) or f = O and j < O (elastic unloading), then p = O
and L = E, where:
446 18. Damage mechanics

E == (1- D)E o (18.31)

For classical models, E is Hooke's operator, but for the model considered
here, E is the elasticity tensor of the damaged material. Equation (18.31)
also shows that in uniaxial tension, we have: D = 1 - (EIEo), Le. the dam-
age variable can be measured by the decay of Young's modulus.

(ii) If f = O and j = O (plastic loading), then p > O and L = H. Using


the results of Sec. 18.4, it is found that plastic loading corresponds to the
quantity (2G oN : i:lh) being positive, while elastic unloading corresponds to
a negative sign of the same expression.
Using the same procedure as in Sec. 12.7, the elasto-plastic 'continuum'
tangent operator H is found to be:

1 ( 2GoN
H = (I-D) [ Eo - h - 8Y ~)
+ 8FD (J'
- ]
® (2GoN) (18.32)

Due to the coupling with damage, H is not symmetric. Note that H can be
written in the following form:
1
H = E - - (E : a) ® (b : E) (18.33)
h
where:

and (18.34)

18.6 Hardening modulus

With the notations and results of Secs. 18.4 and 18.5, we obtain:

b : ir = P(h - b : E : a), (18.35)


'--'"
h*

where h* is the so-called hardening modulus. Using lemma (1.1), it is found


that the tangent operator, H, is invertible if and only if the hardening mod-
ulus, h* is non zero, and in this case:

(18.36)

Relations (18.34-35) give the hardening modulus as:

h * =b-
: ir ir : i: P
-=-- (18.37)
P pr
18.7 Closed-form solutions for loadings with constant triaxiality 447

Since b = âllâu, it is seen that when h* is positive or negative, we have


hardening or softening, respectively. Definition (18.35) together with the ex-
pressions of h, a and b leads to:

dR
h* = (1 - D) ( -
3 -
+ -a - ) âFv-
bN . X - - N . ii (18.38)
dr 2 . âY·

18.7 Closed-form solutions for loadings with constant


triaxiality

In this section, we present analytical solutions developed by Doghri (1995)


for loadings such that the triaxiality ratio, and therefore the triaxiality term,
RII' are constant. Examples are: uniaxial tension (R II = 1) and simple shear
(R II = 2(1 + /1)/3). We consider the version of the model without kinematic
hardening, Le. with only isotropic hardening R(r) and ductile damage D.
Several curves will be plotted in Figs. 18.4-9; large dots in those figures
correspond to the transition from strain hardening to strain softening as
detected by the change in sign of the hardening modulus h* (this will be
explained later).
Using Eqs. (18.8a), (18.21) and the yield condition
__ J2 (u)
h(u) = l-D =R+ay,

the damage evolution Eq. (18.25) can be rewritten as:

(1 - D)D =~ 2 (
1+ -R)2 Rllr (18.39)
2&Bo ay
As proved in Sec. 18.6, the plastic tangent operator, H, becomes singular
when the hardening modulus h* is zero. This was also found to be a condition
for softening for the model under consideration. Using the expressions of Fv,
N and h, reIat ion (18.38) becomes:

h* = (I-D)~~ - (1 ~ D) ~ J2 (u)
Using Eq. (18.8a) and the yield condition again, the expression of h* becomes:

h* = (1 _ D) dR _ 1 (R + uy )3 RII (18.40)
dr (1 - D) 2EoBo
This can be rewritten, using the damage evolution equation (18.39) as follows:

h* dR
= (1- D)- - dD
(R+ay)- (18.41)
dr dr
448 18. Damage mechanics

Taking the derivative of J2 (0") w.r.t. r in the yield condition and substituting
in Eq. (18.41), one finds:

h* = dJ2 (0") (18.42)


dr
This equality shows that h*(r) is the slope of the curve J2 (0") versus r, and
the condition (h* = O) corresponds to the peak of that curve.
From now on, we consider linear isotropic hardening (R = kr). Time
integration of Eq. (18.39) with the initial conditions ro = Do = O gives:

(1 - D)
2
=1 - CTy
- 2- -
3EoSo k
CTy
[ (1 + -r k)3 - 1] Rv
CTy
(18.43)

Figure 18.4 shows the evolution of (D), as a function of (r). From Eq. (18.43),
it is deduced that for a given (r), the value of (D) is higher for a Iar ger value
of (Rv ). This is illustrated in Fig. 18.4 which shows that for a given (r), a
tension test is more damaging than a shear test. Taking the derivative of (D)
w.r.t. (r), one finds:

dD _
dr -
1 CT}
(1 _ D) 2EoSo
(
1+
k)2
r Rv > O
CTy
(18.44)

The slope of (D) at the origin is given by:

dD a2
a;:(O) = 2E;So Rv

The slope of (D) is infinite for (D = 1), and corresponds to the following
failure value of (r):

r1ail = CTy
k
[(1 + ~ Eo;o ~)1/3
Rv CTy CTy
-1] (18.45)

A numerical application gives r 1ail = 12.43 % in tension and 13.14 % in shear.


These failure values correspond to vertical slopes in Fig. 18.4.
Using the yield condition, the von Mises equivalent stress is found to be:

J2 (0") = (1- D)(ay + R) = (1- D)ay (1 + :y r) (18.46)

Using Eq. (18.43), an analytical expression of J2 (0") as a function of (r) is


obtainedj it is plotted in Fig. 18.5. The softening behavior due to damage is
illustrated in that figure. Also in Fig. 18.5, the 'effective' von Mises equivalent
stress J2 (u) is plotted versus (r). That curve is a straight line with a positive
slope, which is explained by the fact that as damage increases, the resisting
18.7 Closed-form solutions for loadings with constant triaxiality 449

areas decrease, and therefore the effective stresses increase. The straight line
in Fig. 18.5 would also represent the true von Mises equivalent stress, if
damage were inactive.
From Eqs. (18.41) and (18.44), the expres sion of the hardening modulus
h* as a function of (r) is foundj it is plotted in Fig. 18.6. Using Eqs. (18.41)
and (18.44), it is found that (h* = O) is equivalent to (D = DBOlt) where:

DBOl t = 1- (18.47)

A numerical application gives DBOl t = 0.22 for both tension and shear. Note
that this critical damage value is much smaller than the ultimate failure
value of l.
Using (18.47) and (18.43) it is found that softening occurs for the following
value of the isotropic hardening strain:

r Bolt = ay [(~ + ~ Eo~o ~) 1/3 -


k 5 5RII a y ay
1] (18.48)

A numerical application gives r Bolt = 8.63 % in tension and 9.15 % in shear.


Combining Eqs. (18.48) and (18.45) the following simple relations between
softening and failure values of (r) are found:

rBolt = (~)
5
1/3 r'ail _ [1 _ (~) 1/3]
5
ay '" O74r'ail _ O.26 ay
k '" . k
(18.49)

We now compute the value of the von Mises equivalent stress at softening.
Using the yield condition (18.46b), we obtain:

h(u BOlt ) =
ay
(1 _D 801t ) (1 + ~rBolt)
ay
(18.50)

A numerical application gives J2(u8olt)/ay = 4.12 in tension and 4.35 in


shear.
The increment of plastic work is defined by dw P = u : deP. After some
algebraic manipulation, this is found to be:

dw P = J2 (u)dr = ay (1 + ~ r) dr (18.51)

Time integrat ion ofEq. (18.51) with zero initial conditions gives the following
expression for the plastic work, which is independent of the triaxiality ratio

(18.52)
450 18. Damage mechanics

Figure 18.7 shows w P versus r. Using Eq. (18.49), the following constant ratio
is found:
(dw Pjdr)Bo~t = (~) 1/3 ~ 0.74 (18.53)
(dwPjdr)/ad 5
The denominator of the plastic multiplier, h, is given by the following ex-
pression:
h = 3Go + (1 - D)k >O (18.54)
The curve h versus r is plotted in Fig. 18.8; it is seen that after the occurrence
of strain softening, h decreases steeply, with a vertical slope at failure.
Figure 18.9 shows the variation of the accumulated plastic strain (P) as
a function of the isotropic hardening strain (r). Before softening, p and r
are almost identical, but in the softening regime, p increases rapidly, with a
vertical slope near failure. We recall that (p = r) when there is no damage.
We now consider nonlinear isotropic hardening and give more results
about strain-softening from (Doghri and Billardon, 1995). It is found -using
(18.40)- that h* is zero if and only if:

1-D[2;:8 (1+ :J
= 0 (dRldr) R·r' (18.55)

It appears that for a given triaxiality term Rv, the critical state h* = O
corresponds to critical values of D and r which simultaneously satisfy Eqs.
(18.55) and (18.39).
As an application, consider power law isotropic hardening (R = kr n )
and three materials, each corresponding to a different value of the hardening
exponent (n). For each material, Fig. 18.10 shows the curve D versus r -
from Eq. (18.39)- and the locus of points (r, D) obeying Eq. (18.55). The
intersection of the two curves gives the critical values (r BO / t , DBO/t). From
Fig. 18.10, it is seen that: r BO / t = 0.007 to 0.086, while DBO/t = 0.14 to 0.22,
which implies that r BO / t strongly depends on the hardening exponent (n),
but DBO/t is less dependent on n. Figure 18.10 also shows the influence ofthe
triaxiality term Rv. It appears that for a given material, DBO/t is independent
of Rv, while r BO / t decreases with increasing Rv. In other words, for a given
material, and regardless of the stress state, strain-softening always occurs for
the same value of damage. This critical value can thus be identified from
a uniaxial tension test. Remarkably, values of damage at softening close to
0.2 have been found using the Lemaitre-Chaboche model with completely
different sets of material parameters and stress states (see Sec. 19.5).

18.8 Return mapping algorithm


In this section, we present the return mapping algorithm which was proposed
by Doghri (1995). It is based on the strain-driven procedure described in Sec.
18.8 Return mapping algorithm 451

tenelon -

·
···
0.9

,,
0.8
,,
,
0.7 ,,

0.6
• ...:
1 0.5

0.4
,//
,~",
"
", ..,

,.",..-
0.3

0.2 ..-
0.1 ................" ....,"',.,'

0L-----====~~--~----~----~----~----~
O 0.02 0.04 0.08 0.08 0.1 0.12 0.14
lsoIroplc lIInI8nIng strain

Fig. 18.4. Damage (D) vs. isotropic hardening strain (r). Material parameters:
Eo = 200 GPa, 1/ = 0.3, Uy = 200 MPa, k = 10 GPa, m = 1, So = 0.5 MPa.

8~----~-------r------~------~----~-------r------,

I
8

i
.~
5

II"
I
~ .....//......-..,.,""'''''"'''*.!:::::~~<~\:.
I
3

\
\
O~----~----~----~~----~-----L----~----~
O 0.02 0.04 0.08 0.08 0.1 0.12 0.14
iIoIropIc hMIer*Ig strain

Fig. 18.5. True and effective von Mises equivalent stresses, divided by uy, vs.
isotropic hardening strain (r).
452 18. Damage mechanics

0.1
Wnsion
lhear ••••.
O
-~.""-""
..................
",
-0.1 ..........

'\
\

3
-0.2

I -0.3

i•
\

-0.4
J:

1 -0.5

1 -0.6

-0.7

-0.8

-0.9
O 0.02 0.04 0.08 0.08 0.1 0.12 0.14
lIoIropic hardenlng strain

Fig. 18.6. Hardening modulus (h*), divided by Eo, vs. isotropic hardening strain
(r).

0.6
tenllian anei sllear -

0.5

0.4
i
!
t

I
0.3

0.2

0.1

o
o 0.02 0.04 0.08 0.08 0.1 b..12 0.14
lIoIropIc hardIning lIr8in

Fig. 18.1. Plastic work (w P), divided by uY, vs. isotropic hardening strain (r).
18.8 Return mapping algorithm 453

1.205
tension -
shear ••••.
1.2

} 1.195
i
"E 1.19

t. 1.185
=
o
t'g
1.18

c 1.115
-8

I
1.17

1.185

1.16

1.155 L -_ _ _L -_ _--J......._ _--L_ _ _.....J.._ _ _- ' -_ _ _-'-....!.._ _......

O 0.02 0.04 0.06 0.06 0.1 0.12 0.1.


Isotroplc hardenino strain

Fig. 18.8. Denominator of the plastic multiplier (h), divided by Eo, vs. isotropic
hardening strain (r).

0.2
tension -
0.18
~,t···
0.16 ,',l
,/ ..,
0.14
"
·i
i
,j
0.12
..//~/
i 0.1
i
'5
§
OI
0.08

0.08

0.04

0.02

O
O 0.02 0.04 0.06 0.08 0.1 0.12 0.1.
IsoIfOpic hardenlng strain

Fig. 18.9. Accumulated plastic strain (P) vs. isotropic hardening strain (r).
454 18. Damage mechanics

1.0

0.8

0.6
I
D
I
I
0.4 I

,,
I
, 0.3

0.2

O
O 0.02 0.04 0.06 0.12

Fig. 18.10. Strain-softening condition h* = O. Critical values of damage D and


isotropic hardening strain r. Power-law isotropic hardening (R = kr n ), with n =
0.3, 0.6, 1. Other material parameters as in Fig. 18.4.

12.10. AIso, as in that section, Ll designates an increment and all variables


which do not contain the subscript (n) are computed at tn+!. The elastic
predictor is defined by:

(18.56)

All other 'plastic' variables are equal to their values at tn. If this trial state
does not satisfy the yield cond it ion ftr :'S 0, then the solution at tn +! is found
as follows.
Firstly, we note the following relation between iT and iT tr :

(T
-tr
- = (T - 2GOL..l€
A P
(18.57)

If the true stresses, (T, are used instead of the effective stresses, iT, such a
simple relation cannot be found. Secondly, noting that:
- -tr
tr (T = tr (T = 31\;otr €, (18.58)

only the deviatoric part of (18.57) needs to be considered. Time discretization


of constitutive equations (18.22) using the (fully implicit) backward Euler
scheme gives the following set of equations:
18.8 Return mapping algorithm 455

-tr
8 8 - 2Go.1eP
f J2(Oo-X) -R(r) -ay =0
.1eP = N.1p
.1X = (1- D)(a.1e P - bX.1p)
.1D y(s).1p
.1r (1 - D).1p (18.59)

where the notation y == Y/So was introduced for convenience. Noting that
aH / J2(u) = aH / h(Oo), and using (18.8, 58), it appears that (y) only de-
pends on (s), through J2(Oo). After substitution of (.1eP ) and (.1p) by their
expressions, it becomes clear that the problem is reduced to finding the four
unknowns s, r, X and D which satisfy the following system offour equations:

ks = -
8-8
-tr
+ 2GoN .1r -
1-D-
o
f = J2 (Oo - X) - R(r) - ay =o
- .1r
kX .1X - (aN - bX n )
1+ r
b.1 =O
kD .1D - Y(8)~
1-D
=O (18.60)

Because of the damage coupling, it is not possible to reduce the problem to


finding two unknowns, (r) and (s - X), as in Chap. 17. Newton's method is
used to solve the nonlinear system of equations (18.60). For each iteration,
the corrections obey the following linear system:

- aN
k s + 2GoNc p + C s + 2Go.1p as : (c s - cX) = O
- dR
f + N: (c s - cX) - (-)c r
dr
=O
- Cr a.1r aN
kx + Cx - (aN - bX n ) (1 + b.1r)2 - (1 + b.1r) Os : (c s - cX) = O
ay
k D + CD - ycp - (.1p) Os : C s = O (18.61)

where:
.1r Cr +CD.1P
.1p = (1- D) and cp == 1- D

The notation:
C(e) == (e)(it+!) _ (e)(it)

designates an iterative 'correction'. Each variable that does not have the
subscript (n) is computed at 'time' t n +! and plastic iteration (it). The general
procedure involves matrix inversions or -in less general cases- the resolution of
456 18. Damage mechanics

linear systems for each local iteration (it). Although those computations are
carried out at the integration point level, their cost may become prohibitive.
Despite the complexity of the constitutive equations considered here, we shall
show that the iterative updating is explicit. Unlike the modified 'cutting plane
algorithm' (Ortiz and Simo, 1986; Simo et al., 1987), no approximations will
be made. Combining Eqs. (18.61a, c) gives:

[ ~] '
B:(cs-cX)=- k s -kX+2GoNc p +(aN-bXn)(I+bLlr)2- -
(18.62)
where:

B -
aLlr ) aN
1 + ( 2GoLlp + 1 + bLlr Os = (1 + "23 g )1 - --
gN ® N

aN
Os = 1 (321 - -N -)
h(ii - X) N = - ax ®
aN

1 ( aLlr ) (18.63)
9 - h(ii - X) 2Go Llp + 1 + bLlr > o

Using lemma (1.1), it is found that B is invertible if and only if 1+(3/2)g =1= O,
which is always verified by an acceptable iterative approximation. The inverse
of B is given by:
-1 1 - -
B = 1 + (3/2)g (1 + gN ® N) (18.64)

Using these results, (18.62) is equivalent to the following equation:


-1 - -
cs - Cx = 1 + (3/2)g[k s - kX + gN: (k s - kX)N]

- [ 2GOCp + (1 + :Llr)2 Cr ] N + 1 + (~/2)9 (1 + !Llr)2


x[X n + g(N: Xn)Nl~ (18.65)
Noting that:
- - 3 - aN
N : N = 2' and N: as = 0,
the fol1owing result follows easily:
aN -3 1
Os : (c s - cX) = 2J2(ii - X) 1 + (3/2)g

x {k s - kX - ~N: (k s - kX)N

- (1 + :Llr)2 [Xn - ~(N : Xn)N] Cr } (1.8.66)


18.8 Return mapping algorithm 457

From (18.61a) an expression of C s as a function of Cr and Cp is found:

- 3Go L1p
C s = -k s - 2GoN cp C X) 1+32g
+ J2(T- (/) { ... } (18.67)

Using (18.61d), an expression of CD as a function of Cr and Cp is found:

ây [ -
+ J2((;3G_ X)
o
CD = -kD + yCp + (L1p) âs: -k s - 2GoNc p

L1p ] (18.68)
x 1 + (3/2)g {... }

In both (18.67, 68), the expression between brackets is that of (18.66).


Using (18.61b), an expression of cp as a function of Cr is found:

_ 1 -. dR "2a-b(N:Xn)
3 - ]}
cp - 3Go { 1- N . (k s - k X) - [ dr + (1 + bL1r)2 Cr
(18.69)

Finally, the definition of Cp is used to find the expression of Cr:

C --
N (18.70)
r - hatg

The numerator is given by the following expression:

N = [(1- D) - yL1p+ 2Go(L1p)2~ : N] [1 - N: (k s - kxl


âY ] 3Go
+3GoL1p [kD + (L1p) âs 2
: k S - 3Go(L1p) J2((; _ X)

L1p ây [ 2- -]
x[I+(3/2)g]âs: ks-kX-3N:(ks-kX)N (18.71)

The denominator of Cr is given by the following expression:

hatg = 3Go + [(1 - D) - yL1p + 2Go(L1p)2 ~i N] :


dR (3/2)a - b(N : Xn)] 2 3Go L1p
x [ dr + (1 + bL1r)2 - 3Go(L1p) J2((; - X) [1 + (3/2)g]

b ây [ 2- -] (18.72)
x (1 + bL1r)2 âs: X n - 3(N: Xn)N
Note that if the plasticity increment is very small, Le. if L1p -t O, then
hatg -t h, where (h) is the 'continuum' modulus given by Eq. (18.29).
458 18. Damage mechanics

18.8.1 Corrections over the elastic predictor

For the elastic predictor, the increments of r, X, D and p are zero. The
residual functions (18.60) -except (f)- also vauish. The corrections over the
elastic predictor have the following simple expressions.

Cr =
(18.73)

Results (18.73) show that the corrections over the elastic predictor corre-
spond to a discretization of the rate constitutive equations (18.22), i.e., a
replacement of p with cpjdt, r with crjdt, idem for b, i and X.

18.8.2 Summary of the algorithm

The time-integration algorithm can be summarized as follows:

-Elastic predictor. Compute iT tr and u tr from Eqs. (18.56).


-Updatejinitialize solution at t n+1: (u, r, X, p, D) = (u tr , r n , X n , Pn, Dn).
-If ftr ~ O, then exit.
-EIse:
-Compute corrections c(.) over elastic predictor, Eqs. (18.73).
-Loop over plastic iterations:
-Update variables (.) according to: (.) t- (.) + c(.).
-Compute residuals from Eqs. (18.60).
-If tolerauce criteria on residuals are satisfied, then exit the loop.
-EIse: compute corrections c(.) as follows:
.C r from Eq. (18.70),
.cp from Eq. (18.69),
.cs from Eq. (18.67),
.(cs - cx) from Eq. (18.65),
.cX = C s - (c s - cX),
.CD from Eq. (18.68).
-End if.
-End loop.
-End if

18.8.3 Non-damaged case

If damage is inactive during the time increment under consideration, let us


check that we do retrieve the results of Sec. 17.7. If D = O, then the following
simplifications hold: ilD = O, D = Dn, Y = O, kD = O, 8y j Os = O aud
CD = O. If, in addition we have, Dn = O is zero, then p = r and '8 = s. In
18.9 Consistent tangent operator 459

that case, it is easily verified that one needs to compute corrections for two
variables only: (r) and (s - X). Equations (18.70-72) simplify to:

Cr= __~f_-_N __:~(k_s_-~kX~)=- (18.74)


3G + dR + (3/2}a-b(N:Xn)
o dr (1 +bLlr )2

The cor rect ion over (s-X) is given by (18.65). Equations (18.74) and (18.65)
are indeed identical to Eqs. (17.36-37) and (17.42).

18.9 Consistent tangent operator

When the local integrat ion algorithm described in Sec. 18.8 has converged,
the corresponding consistent tangent operator Halg is determined by letting
alI the variables, including the total strain tensor € vary slightly around the
converged solution at t n +l, so that: fJu = Halg : &. For an elastic incre-
ment (Le., if the elastic predictor is the solution), then Halg = (1 - Dn)Eo,
otherwise Halg is computed as in (Doghri, 1995).
Differentiation of the incremental relations (18.59) gives:

d - âN
fJs = 2GoI ev : fJ€ - 2GoNfJp - 2G o(..1p) 8s : (Os - fJX)
- dR
N : (fJs - fJX) - ( dr )fJr = O
- fJr a.1r 8N _
fJX = (aN - bX n) (1 + bLlr)2 + (1 + bLlr) âs : (fJs - fJX)

fJD = yfJp + (Llp) ;~ : fJu


fJr = (1 - D)fJp - (Llp)fJD (18.75)

AU the results of Sec. 18.8 can be used if we note that:

ây ~_ ây ~_ 8y ~( _) ~( _) ~
âu : oU = Os : oS + â(tr u) o tr u, o tr u = 311:0 1 : o€, (18.76)

and if we consider the folIowing analogy relations:

f ti O,
8y
kX ti 0, kD ti -(Llp) 8(tr u) (311:0)1 : fJ€ (18.77)

By analogy with (18.70), (fJr) is found to be:


460 18. Damage mechanics

n alg : o€
or halg

where: n alg [
(1 - D) - yt1p + 2GO (t1p)2 âs : N âY-] (2G - oN)

-3Go(t1p)
2 [ 2
Eo - (2G o) [1
t1p
+ (3/2)g] âuâu
f] : âu'
â2 ây

J2 (u-X)
1 (3"2 d
Iev --)
- N ®N (18.78)

Note that if the plasticity increment is very small, i.e. if t1p --* O, then
8r --* i'8t, where i' is the rate given by Eq. (18.30).
The variation of (P) is found by analogy with (18.69):

_ 1 { -. [dR
8p - 3Go 2G oN. 8€ - dr +
!a-b(N:Xn)]
(1 + bt1r)2 8r
} (18.79)

The variat ion of s is


found by analogy with (18.67). Using (18.76b), the
variation of the effective stress u is found to be 8u = H alg : 8€, where:

The variation of the true stress u is now found by differentiation of the


relation u = (1 - D)u,

8u = (1 - D)8u - u8D (18.81)

The variat ion of the damage D is found by analogy with (18.68). After a
rather lengthy algebra, it is found that 8u = Halg : 8€, where the consistent
(or algorithmic) tangent operator can be written in the following farm:

1 -alg { - a l g ay 2-
Ha 9 = (1 - D)H - u ® (t1p)H : au + '3yN
y [dR !a - b(N : Xn)] n alg }
(18.82)
- 3Go dr + (1 + bt1r)2 halg

When the plasticity increment is very small, i.e. when t1p --* O" then it can
be checked that Halg --* H, where H is the 'continuum' tangent modulus
given by (18.32).
18.10 Numerical simulations 461

18.9.1 Non-damaged case

If damage is inactive during the time increment, then the following simpli-
fications hold: fjD = O, D = Dn, Y = O, 8y/8ii = O. If, in addition, Dn is
zero, then p = r and ii = u. In that case, the consistent tangent operator
simplifies to:

Hmod (2G)2 fjp 82f 3Go fjp


- o [1 + (3/2)g]8u8u J2 (u - X) [1 + (3/2)g]

x (1
b [2
+ bfjp)2 X n - "3(N : Xn)N
]® (2G oN)
halg' (18.83)

where h alg is the denominator in (18.74) and H mod is the 'modified' contin-
uum tangent modulus defined by

H mod == Eo - h:'9 (2GoN) ® (2GoN) (18.84)

Equations (18.83) and (18.84) are indeed identical to those of Sec. 17.8, Eqs.
(17.52) and (17.53).

18.10 Numerical simulations

The constitutive equations presented in Sec. 18.2 have been implemented in


the general-purpose finite element program SPECTRUM using the algorithm
developed in Secs. 18.8 and 18.9. More numerical results will be presented in
Chap.19.
Note that several special cases, corresponding to different models, can be
treated with the same source code, by setting appropriate values to different
material parameters (in SPECTRUM, however, each special case was coded
separately, for performance reasons).
- (a) Damage can be turned off by setting the damage threshold, PD, to a
very large positive value.
- (b) Kinematic hardening can be turned off by setting a = O and b = O.
- (c) Kinematic hardening can be made linear by setting b = O.
- (d) Isotropic hardening can be turned off by choosing power law and k = O
or exponentiallaw and Roo = O.
- (e) Isotropic hardening can be made linear by choosing power law and
m=l.
- (f) Plasticity can be turned off alI together by setting the initial yield stress,
ay, to a very large positive value.

Note also that two cases can be treated without any modification to the
algorithm presented in Secs. 18.8 and 18.9.
462 18. Damage mechanics

- (i) Isotropic hardening models of the form R = R(r), where R(r) is an


arbitrary differentiable scalar function of (r).
- (ii) Ductile damage models of the form b = y(u)p, where y(u) is an
arbitrary differentiable isotropic function of (u).

18.10.1 Ductile failure under uniaxial tension

A uniaxial tension test (O'u the only non-zero stress component) is simulated
by submitting a cube to normal displacements on one of its faces. The mate-
rial data is that of Fig. 18.4. The prescribed displacement increases linearly
with time. Equal-size load increments are considered, each corresponding to a
total strain increment LlEu = 1 %, which is ten times the initial yield strain
(O'Y / Eo = 0.1 %). The robustness of the proposed algorithm is illustrated
in Figs. 18.11a-b where the numerical results are plotted together with the
analytical solutions developed in Sec. 18.7. The numerical accuracy can be
improved by choosing smaller load increments. This is illustrated in Figs.
18.11a-b where the results of a simulation corresponding to LlEll = 0.5 % are
plotted.

18.10.2 Ductile failure under simple shear

A simple shear test (0'12 = 0'21 the only non-zero stress components) is sim-
ulated by submitting a cube to tangential displacements on one of its faces.
The material data is that of Fig. 18.4. The prescribed displacement increases
linearly with time. Equal-size load increments are considered, each corre-
sponding to a total strain increment LlE12 = 1 % (Le., an engineering shear
strain increment Ll-Y12 = 2 %), which is ten times the initial yield strain
(O'Y / Eo = 0.1 %). The robustness of the proposed algorithm is illustrated
in Figs. 18.12a-b where the numerical results are plotted together with the
analytical solutions developed in Sec. 18.7. The numerical accuracy can be
improved by choosing smaller load increments. This is illustrated in Figs.
18.12a-b where the results of a simulat ion corresponding to LlE12 = 0.5 % are
plotted.

18.10.3 A post-processor for crack initiation

Based on a simplified vers ion of the model studied in this chapter, a post-
processor, DAMAGE 90, was developed. This computer code takes as input a
strain history and is able to predict ductile failure in one or more dimensions,
brittle failure, low and high cycle fatigue with the nonlinear accumulation and
multiaxial fatigue.
A complete description of the model, several numerical simulations as well
as the listing of the source code are given in (Lemaitre and Doghri, 1994) and
(Lemaitre, 1992).
18.10 Numerical simulations 463

uniaxi.l ten.ion te.t aimulation


1
analyljlcal -
numerical. inc~ 'b'l 5\ o
0.9 numerical, inc. l' ..
O

0.8

0.7

0.6
O
O)
+0

J'" 0.5 O

0.4

0.3

0.2

0.1

O
O 0.04 0.06 0.08 0.1 0.12 0.14
iaotropic hardening strain

uniaxial ten.ion test simulat ion


4.5r------r------r------r~~~~====~----~----~
analytical -

......
~----IWllerical. inc. -0.5\ o
4 o +0 0.0 o.. rical, inc.-l' +
+ o
II o
LI + o
3.5
" o

.....
LI +
c:

~
j
.
OI
2.5

':il"
g 1.5
'ti
u
...
N

ig o

+0
0.5
O

O~------~----~------~------~------~------------~
O 0.02 0.04 O.O~ 0.08 0.1 0.12 0.14
isotropic hardening strain

Fig. 18.11. Ductile failure under uniaxial tension. Comparison between analytical
and numerical results. Material parameters as in Fig. 18.4. Variat ion of two variables
as a function of the isotropic hardening strain, r: (a) damage, Dj (b) von Mises
equivalent stress, h(u), divided by Uy.
464 18. Damage mechanics

eimple ahear test aimulatlo~


1
_lyti~a! -
~Wllilrieal. i~e .• O. 541 o
0.9
~...... dcal. i~c.:~ +

0.8

0.7

0.6

III

J 0.5

0.4

0.3

0.2

0.1

O
O 0.02 0.04 0.06 0.08 0.1 0.12 0.14
laotropie hardeni~g atraln

al~le sheer te.t almul.tion


4.5
• o analytieal -
auAJlie lne •• O• 5' •
• " nullletiS • lnc.-1' +

.•..••
. 3.5

....•
CI
3
...~
~ 2.5
••
...•
li: 2

~ 1.5
iN
....
~,

1
a8 o
0.5
+ •
o
o 0.02 0.04 0.06 0.08 0.1 0.12 0.14
iaotroplc hardening strain

Fig. 18.12. Ductile failure under simple shear. Comparison between analytical and
numerical results. Material parameters as in Fig. 18.4. Variation of two variables
as a function of the isotropic hardening strain, r: (a) damage, Di (b) von Mises
equivalent stress, 12(0"), divided by Uy.
18.11 Further reading 465

Uniaxial stress cases are simulated by giving as input data Ell(t); the
code computes E22(t) and E33(t) such that 0"22(t) = 0"33(t) = O. The results
of a low cyde fatigue simulat ion corresponding to a lI-strain amplitude of
7 % are shown in Figs. 18.13 and 18.14. The material data are those of an
aluminum alloy. The figures show the decay of the stress-carrying capability
of the material as the damage increases. The number of cycles to failure was
found to be N R = 40 cycles.

18.11 Further reading


For ductile damage, besides the Lemaitre-Chaboche model which was studied
in this chapter, another popular model was proposed by Gurson in 1977. In
that model, damage is porosity-based, in order to better describe the physics.
Indeed, ductile fracture in metals is explained by the nucleation, growth and
coalescence of micro-voids. In Gurson' model, the yield function depends on
hydrostatic stress, while that function is always deviatoric in the Lemaitre-
Chaboche model. A comparative study of different ductile models (Lemaitre-
Chaboche, Gurson, Thomason, etc.) is presented in (Pardoen et al., 1998).
For brittle materials (e.g., concrete under tension) damage models have
been developed by Mazars, Bazant, Krajcinovic and their co-workers.
For composite materials, anisotropic damage models were developed by
Ladeveze and his co-workers.
Damage mechanics has enjoyed a great deal of interest since the 1970's
and there is today a large body of literature on the subject. The list of refer-
ences would be too long to include in this book. References to the works cited
in this section and in Sec. 18.1, as well as to many other works can be found
in the following books: (Lemaitre and Chaboche, 1990), (Lemaitre, 1992),
(Kachanov, 1986), (Krajcinovic, 1996), (Skrzypek and Ganczarski, 1999); see
also (Thomason, 1990) for physical and modeling aspects of the ductile frac-
ture of metals.
Once a macro-crack initiates, one can use Fracture Mechanics (FM) in
order to predict its propagation. After the pioneering work of Griffith in
the 1920's who introduced an energy approach, this branch of mechanics
started really to take off in the 1950's and the 1960's with major concepts
such as the path-independent integral. Seminal contributions were made by
Eshelby, Irwin, Rice, Hutchinson, Rosengren, and many others. For references
to those works and others, as well as several applications, see (Anderson,
1995), (Bolotin, 1996) and (Knockaert et al., 1996).
Generally speaking, if plasticity is confined to small regions around crack
tips, then FM gives good predictions. Otherwise, Le. for rather generalized
plasticity, a local approach to fracture -such as damage mechanics- may
lead to better agreement with experimental results. However, as explained
in Chap. 19, there are some inherent problems with the numerical simulation
of the propagat ion of damage.
466 18. Damage mechanics

8
0 .04 II

0 .02

0 .00
N

-0.02

-0.04

- 0 ,06 +.-,rT'T'1"""'TTTTT'T"rTT"rTT1rTT'1rT'T'1"T'T'T"T'TT"TTTTT"rTTT"rTT'1rTT'T'l
0.00 10,00 20.00 30.00 40.00 50.00

"I'e
600.00

400.00

200.00

0.00
E"

-200.00

-400.00

-600.00

-800.00
6

Fig. 18.13. Simulat ion of one-dimensionallow cyde fatigue. (a) lI-strain history
(amplitude: 7%) ; (b) stress-strain loops. Mat erial parameters: Eo = 72 GPa, v =
0.32 , ay = 440 M Pa , So = 6 M Pa.
18.11 Further reading 467

'l(G) H".
500.00

400.00

300.00

5 .00

4.00

3 .00

2.00

1.00

0 .00 4'Mr"f'M;;""''''I'T''I'''I'T''I'T'n'''l'T''l"T'T'T"T'T'T"T'T'TT'T"I"T'T'T"T'T'T'TT1'''1'T''1''T'T1r'TT'\
0.00 10.00 20.00 30.00 40.00 50.00
N
Fig. 18.14. Simulat ion of one-dimensional low cyde fatigue . (a) von Mises stress
versus number of cycles; (b) accumulated plastic strain (P) and damage (D) versus
number of cycles.
19. Strain localization

For strain-softening material models such as the ductile damage model of


Chap. 18, the boundary value problem may become ill-posed and bifurcation
might occur. As a consequence, finite element (FE) simulations involving such
models are faced with problems of loss of uniqueness and strain localization.
If localization occurs, the size of the region where the strains concentrate is
determined by the mesh size and no convergence can be obtained with mesh
refinement. In order to circumvent those difficulties, nonlocal ar internal-
length models have been proposed in the literature. In this chapter, we study
the above-mentioned issues and give analytical and numerical examples, lim-
iting ourselves to the small-strain case for simplicity.

19.1 Motivation: a one-dimensional example


In order to illustrate bifurcation issues with local, strain-softening models,
we consider a lD example from (Knockaert and Doghri, 1999). An elastic-
damageable material is described by the folIowing constitutive equations:
1
a = E(l- D)E, D(t) = -suPr<tE(T), (19.1)
Ef -

where E is Young's modulus of the virgin material, Ef the failure strain and
D a damage variable which ranges from zero for the undamaged material
to 1 for a Representative Volume Element (RVE) with null stress-carrying
capability. The corresponding stress-strain response is shown in Fig. 19.1.
Variants of this model have been used in order to model the damage behavior
of laminated composites (see (Florez, 1989) and references therein).
A rod of length L made of a material obeying Eqs. (19.1) is subjected to
a uniaxial tension test. Imagine now that the rod is virtually divided in two
parts of lengths LI and L 2 (see Fig. 19.2 with L 2 = Ld, LI = L - L 2 and
17 = 1).
Before the load peak, the solution is unique: in loading, the strain and
damage increase in both parts of the rod and if unloading occurs, it will also
take place in the whole structure. After the peak of the load-displacement
curve, the stress has to decrease and besides elastic unloading of the whole
rod, several other solutions are possible:
470 19. Strain localizat ion

(J

EEr
4

Stored energy
(reversible)
A C Er E
Fig. 19.1. Elasticity with damage, no threshold. Stress vs. strain, stored and dis-
sipated energies.

TIE E u

L
Fig. 19.2. Rod under uniaxial tension.
19.1 Motivation: a one-dimensional example 471

- loading in both parts of the rod: fi, f2, ih, D2 > O


- loading in one part of the rod and elastic unloading in the other: f2 >O
and D2 > O; f·l < O and D1 = O.
Those solutions may appear at any time after the peak and for arbitrary
combinations of LI and L 2 (such that LI + L 2 = L). So, there is an infinite
number of solutions.
The following equations hold:

(19.2)

(19.3)

(19.4)

(19.5)

If we as sume that bifurcation occurs at the load peak and that the first part of
the rod unloads elastically, we have after the peak: D 1 = 1/2 and a = EfI!2.
Using ( 19.2) and ( 19.3), it is possible to compute f2 as a function of a:

(19.6)

As f2 has to increase (loading ofthe second part ofthe rod) while a decreases,
the solution with a" +" is chosen above. Using equation ( 19.5), f is given as
a function of a by:

(19.7)

The results obtained for several values of (L 2 / L) are given in Fig. 19.3
Since f is given by equation ( 19.5), it is possible to obtain f ~ O if L 2 / L
is small enough with respect to LI! L, because we have fI < O and f2 > O.
This phenomenon is called "snap-back". We compute f from a in ( 19.7) and
not the opposite because, for a given value of f, there may be two values of
a. Several problems similar to the one considered in this section have been
studied by other authors, see (Florez, 1989) and references therein.
472 19. Strain localizat ion

! Homogeneous -
L21L = 1/4 ----

0.8
/ L2IL = 1/2 ...... .
L21L = 3/4 _ ...-

/
0.6
/
><
tU
E
/
t
l.L t
t
li:: i
0.4 i
/
0.2
i
/
i

O
/
O 0.2 0.4 0.6 0.8
U/Umax
Fig. 19.3. Bifurcation problem for a strain-softening local model: normalized load-
displacement curves for several values of Ld L

19.2 Uniqueness and ellipticity

In this section as well as in Sec. 19.3, we will generalize the one-dimensional


ideas of Sec. 19.1 to the three-dimensional case.
For general elasto-plastic models with internal variables V, it was shown
in Sec. 12.12 that the stress and strain rates are related as follows:

iT = L(€, V) : €, (19.8)

where L is a tangent operator defined by:


b: E: €
L=E ifl < O or 1=0 and h ~ O,
b: E: €
L = E - ~ (E : a) ® (b : E) if I = O and h > O, (19.9)
, ... ,
H
where E is the elasticity operator and h is assumed to be positive.
In this section, we give a few uniqueness results regarding the rate (or
incremental or linearized) problem. We refer to Hill (1958) and (Benallal et
al, 1993) for proofs and more details.
Consider a solid body which occupies an open set il of JIt3 at time t,
and is subject~ to rates of forces per unit volume i in il, rates of forces
per unit area F on a part rF of the boundary and imposed velocities v on
r
the complementary part u of the boundary. Knowing the state at time t,
19.2 Uniqueness and ellipticity 473

the rate problem consists in finding the velocity field v which satisfies the
following equations:

v=v onru
div [L : €(v)] +i =O in il
L: €(v)·n =F
€(v) = ~(V'v + V'T v) in il (19.10)

Knowing v, ir and V can be computed from the constitutive equations. A


"weak" form of problem (19.10) can be constructed in a classical way, see
Chap. 2. Note that because operator L depends on the stress-strain state
-Eqs. (19.9)- problem (19.10) is nonlinear.
A sufficient condition for existence and uniqueness of a solution to the
rate problem is the following:

ir : € > O, Vx E il (19.11)

This also can be interpreted as a non-softening requirement.


If we consider the rate problem for a solid whose constitutive equation is
iT = H : €, i.e. we only consider the plastic branch of the tangent operator
(this is Hill's linear comparison solid), then the associated rate problem is
well posed if the following conditions are met:
- (i) Operator G : w .......-t -div [H : €(w)] is elliptic Vx E il,
- (ii) Complementary conditions are satisfied.
Condition (i) can be rewritten equivalently as follows:

det(n . H . n) :f. O, Vn:f. O and Vx E il (19.12)

In other words, loss of ellipticity corresponds to the time for which the so-
called acoustic tensor (n· H . n) becomes singular for a direction n in a point
of il:

det(n· H· n) = O, for n:f. O in a point of il (19.13)

Similar conditions were obtained and analyzed by Hadamard as early as


1903. We shall see in Sec. 19.3 that Eq. (19.13) corresponds to "continuous
bifurcation" in Rice's terminology. The corresponding mode is a strain rate
jump l€J across a localization surface (8) which has the general form:

l€J = ~(g®n+n®g), (19.14)

where n is a unit normal to (8) and 9 a localization mode.


474 19. Strain localizat ion

It is possible that bifurcation modes appear ruter (19.11) is violated and


before (19.13) is satisfied; they correspond to diffuse modes, see (Nielsen and
Schreyer, 1993). Before we close this section, we mention that complementary
criteria (ii) are related to surface or boundary conditions. They can be impor-
tant in certain cases where it can be shown that complementary conditions
are violated (leading to so-called surface or Rayleigh waves) while the in-bulk
ellipticity condition (19.12) is stiH satisfied. For informat ion, see (Benallal et
al, 1993).

19.3 Strain localization


In this section, we give a presentation of strain localizat ion which was made
popular by Rice (1976) and IDee and Rudnicki (1980). We consider the same
constitutive equations (19.8-9) as in the previous section.
Suppose that there exists a surface (S) across which the strain rates are
discontinuous. We designate by (1) and (2) the two regions from each side
of (S) and n a unit normal to (S); see Fig. 19.4. Actually in that figure, a
"shear band" designated by (1) and delimited by two such surfaces (S), is
depicted.

Jump of the velocity Continuity of the traction at the


gradient interface

2
n 2
-
- L-.----------.:~L_ _ __.J_

Fig. 19.4. Localization with discontinuity of the velocity gradient


19.3 Strain localizat ion 475

A jump of a quantity (.) across (S) is designated by:

(19.15)

We have the following jump conditions:

la- . nJ = O and l VuJ = g ® n, (19.16)

which express continuity of the stress vector rate and Maxwell's geomet-
ric compatibility condition, respectively. Vectors u and g represent the dis-
placement and a localization mode, respectively. Since it is assumed that:
L ijkl = L ij1k , (19.8) can be rewritten as follows:

a- = L: VU

Equation (19.16a) becomes:

L(2) : lVuJ . n + lLJ : VU(l) . n = O,

or, thanks to (19.16b):

(n· L(2) . n) . g +n 'lLJ : Vu(1) = O (19.17)

19.3.1 Continuous bifurcation

This is the case when L(l) = L(2). Equation (19.17) then becomes:
(n· L· n) . g =O (19.18)

If L = E, it is generally not possible to find a solution g f:. O for Eq. (19.18).


An example is given in Sec. 19.11.1. If L = H -the plastic branch of the
tangent operator- then Eq. (19.18) becomes:

(n . H . n) . g =O (19.19)

A non-trivial solution g f:. O to the equation corresponds to a unit vector n


for which the "acoustic tensor" (n· H· n) becomes singular, Le.

det(n· H· n) =O (19.20)

19.3.2 Discontinuous bifurcation

This corresponds to the case when L(l) f:. L(2), Le. we have elastic unloading
on one side of (S) and plastic yielding on the other. Note that in Eq. (19.17),
upper-scripts (1) and (2) play symmetric rolesj therefore we can choose:
476 19. Strain localization

. b: E: E(2)
L(2) =E wlth h ~ O,

1 b E '(1)
L(I) =H =E - -(E : a) ® (b : E)
h
with : h: e >O (19.21)

Equation (19.17) then gives:

(n . E· n) . g + ~n . [(E : a) ® (b : E)] : i(l) =O (19.22)

We generally have that (n· E . n) is invertible (an example is given in Sec.


19.11.1). Equation (19.22) then gives:

g = -.!.(n. E· n)-1 . (n· E : a)(b: E: E(I)) (19.23)


h
Since i = ("\7it + \jT it)/2, compatibility condition (19.16b) gives:

LEJ = ~(g®n+n®g), (19.24)

from which it is deduced -using the minor symmetries of E- that:

b: E: i(1) = b: E: i(2) - (b: E· n)· g (19.25)

Using (19.23), Eq. (19.25) can be rewritten as follows:

(b : E: i(I)) [1- ~(b : E· n)· (n· E· n)-1 . (n· E: a)] = b : E : t(2)


(19.26)

Using the inequalities contained in (19.21), Eq. (19.26) implies:

1- ~(b: E· n) . (n· E· n)-1 . (n· E: a) ~O (19.27)

We now introduce the following notation:

M == n . E . n, A == M- 1 . (n . E : a), B == b : E . n, h c == A . B
(19.28)

With these notations, inequality (19.27) becomes:

(19.29)

where h* is the hardening modulus defined by h* =h - b : E : a (see Sec.


12.12). From (19.21b), it is deduced that:

n·H·n=M· (l-~A®B) (19.30)


19.4 Analytical results for initially homogeneous plane problems 477

where we used notations (19.28). Using lemma (1.2), we conclude that (n·
H . n) is invertible if and only if h ::f. h e , and in that case:

(n.H.n)-l = (1+-1-A®B) ·M-


h- h e
1 (19.31)

If h = h e , then there exists 9 ::f. O such that:

(n· H· n)· 9 = O (19.32)

19.3.3 Summary

The study of discontinuous bifurcation showed that when h = he , Eq. (19.32)


is found, which is identical to (19.20), the condition for continuous bifurca-
tion. Now, taking into account (19.29b) and assuming that for increasing plas-
tic deformation, the hardening modulus h* decreases (an example is given in
Sec. 18.7, Fig. 18.6), we may conclude that continuous bifurcation occurs be-
fore the discontinuous one. Consequently, loss of ellipticity condition (19.13)
for Hill's linear comparison solid is also a sufficient condition for continuous
or discontinuous bifurcation (in Rice's terminology). This result can be inter-
preted as follows: bifurcation is continuous at first (Le., plastic/plastic), then
right after that, we may have elastic unloading on one side of the localization
surface and plastic yielding on the other: bifurcation becomes discontinuous.
Finally, using (19.9b) and lemma (1.2), it is found that conditions (19.13),
(19.20) or (19.32) can be written equivalently in the following form:

h = (b : E· n) . (n . E . n)-l . (n . E : a) (19.33)

19.4 Analytical results for initially homogeneous plane


problems

19.4.1 General strain-softening models

In this section, we present some closed-form solutions due to Doghri and Bil-
lardon (1995) for strain-localizat ion in initially homogeneous plane problems.
Earlier results appeared in (Billardon and Doghri, 1989). We consider general
strain-softening models of the rate form:

where H is a tangent operator. We designate by n the normal to a possible


localization surface and by 9 a localizat ion mode. For plane stress ar strain
problems, we have n3 = O. We assume that 93 = O. Equation (19.14) implies
that Lij3J = O, j = 1,2,3. We introduce the following notations:
478 19. Strain localization

n = [ Si~O
cosO 1 g=g
,
cosX
[ Si~X
1 , (19.34)

where O and X E] - 7f /2, 7f /2] and 9 E III Continuity of the stress vector rate
across the localizat ion surface, LUijJnj = O, leads to:
(19.35)
Combining the two equations, we obtain:

LUu J cos2O = LU22J sin2O, (19.36)


which implies that the jumps in Uu and U22 always have the same sign.
This is the case even if prior to localization, Uu and U22 have opposite signs.
Equations (19.35-36) also show the following:
- (a) If O = 7f/2, then LU22J = LU12J = O, Le., there are no jumps in U22 or
U12.
- (b) If () = O, then LuuJ = LU12J = O, Le., there are no jumps in Uu or U12.
- (c) If () :F 7f /2, the orientation of the localization surface is related to jumps
in U11 and U22 by the relation:

(19.37)
With the localization condition written in the form Hijkl Lik,Jnj = O, system
(19.35) becomes:
(Huu LiuJ + H U22 Li22J + 2H1112 Lt12J) cos O
+ (Hl2l1 LtuJ + H 1222 Lt22J + 2H1212 LtI2J) sin () = O,
(H12U LtuJ + H 1222 Lt22J + 2H1212 LtI2J) cos O
+ (H22U LtuJ + H2222 Lt22J + 2H2212 LtI2J) sinO = O (19.38)
In the remainder of this section, we consider a biaxial loading along the
(1,2) axes, Fig. 19.5. Since there is no shear, mixed terms such H 12U vanish
and system (19.38) simplifies to:

(Huu LtuJ + H1122 Lt22J) cos 0+ 2H1212 Lt12J sin () = O,


2H1212 Lt12J cos () + (H22U LtuJ + H2222 Lt22J) sin () = O (19.39)
We can distinguish three cases:
- (a) If () = 7f/2, then Lt12J = O, since we usually have H 1212 = E 1212 > O.
So in that case, there are no jumps in t12, U12 or U22. The jumps in tu
and Uu are related by:

LuuJ = (Huu-Hu22HH22U) LtuJ (19.40)


2222
19.4 Analytical results for initially homogeneous plane problems 479

Eu=~lI

Fig. 19.5. Strain localization for an initially homogeneous plane state under biaxial
loading. Normal N == n to localization surface makes an angle 8 with horizontal
axis.

- (b) If () = 0, there are no jumps in t12, a12 or an. The jumps in t22 and
a22 are related by:

(19.41)

- (c) The third case is when () fi. {O,7r/2}; this is what we shall study from
now ono
Before strain localization , we can write:

where a is a strain rate ratio. We assume that when localization occurs,


the reIat ion between strain rates stiH holds on each side (+) and (-) of the
localization surface, i.e.
)+) - ,..;(+). ;(-) - at(-)
~22 - ~~11 , ~22 - 11

This implies the following equality between strain rate jumps:

Lt22J = a LtllJ (19.42)

The compatibility condition (19.14) gives:


480 19. Strain localization

2lt12J = 91n2 + n192 = ltuJ tanO + lt22J tan- 1 O, (19.43)

or, introducing the biaxiality ratio a, Eq. (19.42),

. a +tan 2 0 .
2lf12J = tan O lfllJ
(19.44)

We now consider two cases.


(a) t12 = o. This is possible only if a < O. In this case, the orientation O is
related to the ratio a by the simple relation:

tan 2 0 =-a (19.45)

System (19.39) gives the localization condition as:

H llU + aH1122 = H 2211 + aH2222 = O (19.46)

Equations (19.45-46) show that the orientat ion of the localization surface de-
pends only on the biaxiality ratio a, while the localization condition depends
on the material model and the stress state.
As a first application of formula (19.45), consider a uniaxial tension test
in the 2-direction. Neglecting the elastic strains before the plastic strains, we
have a ~ -2; the formula then gives:

tanO = ±v'2, Le. O ~ ±54.7° (19.47)

This angle between the localizat ion surface and the loading direction is in 900d
agreement with experimental observations. Results (19.47) were also found
by Hill (1950, Chap. XII) using a different approach based on the method of
characteristics.
For a second application of formula (19.45), we consider a loading such
that t22 = -tu. The formula gives tanO = ±1, Le. O = ±45°.
For plane strain situations, relations (19.46) show that tTn = tT22 = O.
Since (T12 = O, it is concluded that in plane strain, localization corresponds
to the peaks of the in-plane stress-strain curves.

(b) The second case derived from Eq. (19.44) is when tl2 i- O. This hap-
pens if and only if a + tan2 O i- O, e.g. a > O. Localization conditions (19.39)
become as follows, using Eqs. (19.42, 44),

Huu + aH1122 + H 1212 (a + tan 2 O) o,


H 1212 (a + tan 2 O) + (H2211 + aH2222 ) tan 2 O O, (19.48)

which leads ta:

Hnn + aH1122 = (H2211 + aH2222 tan 2 O (19.49)

For plane strain situations, relation (19.49) gives:


19.5 Numerical results for a ductile damage model 481

an = a22 tan 2 () (19.50)

System (19.48) shows that we must have au =J. o and a22 =J. O, since
(o + tan 2 ()) =J. o and H 1212 = E 1212 > O, in general. So in this case, strain
localizat ion does not coincide with the peaks of the in-plane stress-strain
curves, and should appear in the strain-softening regime. Relation (19.50)
also shows that when localization occurs, an and a22 must have the same
sign.

19.4.2 A ductile damage model

In this section, we consider the Lemaitre-Chaboche ductile damage model of


Sec. 18.2 with isotropic hardening but without kinematic hardening. Strain-
softening was studied in closed-form in Sec. 18.7. For strain-Iocalization, an-
alytical results were obtained by Doghri and Billardon (1995) using the 10-
calization condition under the form (19.33). Three stress states were studied
without any additional assumption. The main findings are the foHowing (()
designates the orientation of the localization surface):
- • Simple shear, such that U12 =J. O and alI other stress components vanish.
Result: () = ±45°, the orientation is fixed and does not depend on the
material parameters or the stress-strain values.
- • Tension-compression, such that Un = -U22 and aH other stress com-
ponents are nil. Results (without making assumption (19.42)): (a) fixed
orientation: () = ±45°, (b) strain localization and strain softening occur
simultaneously.
- • Biaxial tension, such that the only non-zero stress components are Un =
U22. Result (without assumption (19.42)): () is arbitmry, any orientat ion is
possible.

19.5 Numerical results for a ductile damage model

In this section, we present numerica! results from (Doghri and Billardon,


1995). Earlier results appeared in (Billardon and Doghri, 1989).

19.5.1 Biaxialloadings in plane stress

As in Sec. 18.7, we consider the Lemaitre-Chaboche ductile damage model


of Sec. 18.2 with power-law isotropic hardening (R(r) = kr n ) and without
kinematic hardening. Three materials are defined, each corresponding to a
different value of the hardening exponent (n). The material responses under
uniaxial tension and the parameter values are given in Fig. 19.6. In the follow-
ing, €* denotes the value of the strain at the peak of the uniaxial stress-strain
curve for each material.
482 19. Strain localizat ion

rr HP"

2()()().

1000.

200-
OL-__ ~ __ ~ __ ~ __ ~ __ _ L_ _~~~

.00 .10 t!

.5

.2

.10 t!
Fig. 19.6. Material responses under uniaxial tension for ductile damage and power-
law hardeningj (a) Stress versus strain, (b) damage versus strain. Material param-
eters: Eo = 200 GPa, v = 0.3, Uy = 200 M Pa, k = 10 GPa, n = 0.3, 0.6, 1,
So = 0.5 MPa.
19.5 Numerical results for a ductile damage model 483

The uniqueness and localization criteria (19.11,13) are applied to initially


homogeneous plane stress states under various monotonie biaxial loadings.
The problem is modeled with an 8-node isoparametric plane stress element
with 3 x 3 Gauss integration points (Fig. 19.7) which is subjected to imposed
displacements Ul and U2 = aUI, where a is a constant and Ul increases
monotonically with time. The element has dimensions of 1 x 1, which implies
that the (initially homogeneous) in-plane strains are El = Ul and 102 = U2 •
Besides the usual F.E. approximations, the only assumption that is made
in this section is that the out-of-plane components of the normal n to the
localizat ion surface and the localization mode g vanish.
The material model was simulated via the numerical algorithms of Chap.
18, and the localization condition (19.13) was implemented as described in
Sec. 19.11.1.
As a general remark, we mention that a good agreement between the
results reported in this section and the analytical results of Sec. 19.4 was
found, especially for a < O.

·.
lZl
. .
..
...::t> UI/Z

Fig. 19.7. Initially homogeneous plane stress states under various monotonie bi-
axialloadings: (a) finite element modeling, (b) localized modes.

The locus of points for which the localizat ion criterion is first satisfied is
plotted in strain space in Fig. 19.8a for strain rations EI/E2 = l/a varying
from (-1) to (+ 1). The values of the strains are normalized for each of the
three materials by the corresponding values of 10*. Note that the curves of
Fig. 19.8a have the same shape as the limit curves for metal sheet formingj
e.g. (Cordebois and Ladeveze, 1986).
Figure 19.8b gives the localizat ion orientat ion (). When a = -1, we have
a symmetric tension-compression loading (102 = -El and a2 = -al). We find
() = ±45°, and since gl cx: n2 and g2 cx: n1, we conclude that we have a pure
shear localization mode.
The critical values of damage for the localized bifurcation mode are plot-
ted in Fig. 19.8c where D* denotes D(E*). The uniaxial tension case in the
2-direction corresponds to a strain ratio EI/E2 ~ -1/2, because elastic strains
are small compared to plastic strains. Notice that according to the plot, only
biaxial tests such that EI/E2 > O would enable to obtain diffuse (or non-
484 19. Strain localization

localized) damaged states with higher values of damage than those of uniaxial
tension.
The loci of points corresponding to the following criteria: loss of unique-
ness, strain localization and complete failure (D ~ 1) are plotted in Fig.
19.9a. The results show that for a < O, loss of uniqueness and localizat ion
happen at the same time, while for a > O, the second criterion is satisfied
later. This result is also illustrated in Figs 19.9b-c for a = -1 and +1, re-
spectively. Figure 19.9a also shows that the the loss of uniqueness and strain
localization criteria are satisfied long before the complete failure criterion.
In order to check the so-called objectivity of the localizat ion criterion,
the one-element mesh used in this section was re-meshed into five elements
(Fig. 19.10) and alI the simulations were run anew. It was found that up
to the localizat ion time, the results given by the two meshes are identical.
For loadings such that a < O, this is no surprise since bifurcation does not
occur before localization. For a > O, the finding suggests that only diffuse
bifurcation modes appear before localization. Those modes can be captured
by mesh refinement, while localized modes are incompatible with classical
finite elements.
Figure 19.11 shows that the stress-strain solutions are no longer homo-
geneous inside the finite element after localizat ion. Furthermore, the figure
shows that for two B.C. configurations which are equivalent, the F.E. results
after localizat ion are different. Figures 19.12-13 show that after localization,
the finite element results are completely unreliable. The arrows on those fig-
ures correspond to the first fulfillment of the localization criterion.
In summary, it appears from the preceding that it is not possible for
experiments where a < O (e.g., uniaxial tension of tension-compression) to
identify the softening behavior of a material obeying the Lemaitre-Chaboche
ductile damage model. For those stress states, as soon as we have soften-
ing, we also have localization and the deformation becomes discontinuous.
However, we may reach that identificat ion objective with experiments where
a is positive (e.g., biaxial tension). A second conclusion is that localizat ion
problems appear long before the theoretical failure value of D ~ 1. Finally,
the localizat ion criterion indicates the reliability limit of conventional F.Es.

19.5.2 Notched plate with a macro-defect


The geometry of aflat 2 mm-thick notched specimen is given in Fig. 19.14a. A
through hole 1 mm in diameter in the vicinity of the notch plays the role of a
macro-defect or stress concentrator. The specimen is made of 2024 aluminum
alloy and is subjected to monotonically increasing displacements through
bearings which allow free in-plane rotation of each head of the specimen.
The original square grid marked on the specimen consists in points 0.18 mm
in diameter at a mean distance of 0.50 mm.
For the numerical simulat ion of the experiment, a plane stress assumption
is made. The mesh used and the B.Cs are shown in Fig. 19.14b.
19.5 Numerical results for a ductile damage model 485

\
"'-.

,.
pur'-;plitting lIIod. -90-

• _ _o pu,,' sh.irmodi""

tOO

_1. O 1. ~1/e2

0/0*
n:t 1.0

I 3
j// n=.6
·~i . /
/
."
/

~I 2 // n=.3
~.
h."
~! ,ţ"/

-1. O 1.

Fig. 19.8. Strain localizat ion for various values of hardening exponent (n) and
bi-axiality ratio (n = E2/fl): (a) limit curves in strain space, (b) orientation of
localization surface vs. (lin), (c) critical damage vs. (lin).
486 19. Strain localization

uz HPa

.01 El

Fig. 19.9. Comparison of three criteria: loss of uniqueness, strain localization and
complete failurej (a) limit curves in strain space, (b) stress-strain curve for Q = -1,
(c) stress-strain curve for Q = +1.
19.5 Numerical results for a ductile damage model 487

U2
T
Fig. 19.10. The element of Fig. 19.7a is re-meshed into five elements in order to
check the objectivity of the strain localizat ion criterion.

,....,.--.,......
HP,
.,.,.--......'.
" ...•. ....
J'

...,.
"
6(1).
.:
."
60D
......
"
..•..
. ..' "
.
I •

.:'
400.
..:' 1,()()
'. B

(\(
A~C
200. 200.
.
A

o .05 ~22 O .05 e22

n
~U2 ~ U2
<>=
U,/2 _
G1 B- c:::::::C>
A- U,/2

Fig. 19.11. Sensitivity of the post-localization F.E. results to boundary conditions.


B-
Âe_
==<>
U,
488 19. Strain localizat ion

* I02MPa
..~..
..-.
0=1
._ e.
.-..
. ...
~. ~

.
6

:.
. ..
~

.
.. ...
....
4

.. ,\-'
2
.' !
I
!
. I
* 10.2
OL---~---L--~--~~--~--~~~--
O 2 4 6

* I02MPa

..-.-."".--......-.-. 0=1

....
Il= I
..•.
6

"."
. " ..
..... ....
4
..... ..
' ~

....
..- \

O~--~--~--~--~------~--~--~
O 2 4 6 8

el

Fig. 19.12. Non-homogeneous post-localizat ion F.E. stress-strain responses.


19.5 Numerical results for a ductile damage model 489

8
n .. 1
),
..........'
6

... ...
........
4
....-
...
..' .....
.......
..
2

........
..
......
2 4 6

.. ./
0.12 nal
aasn

.,'
."
0.08 ./,

......,,'
0.04 ••••
.... ..
._ •• e
.' .. 10-2
o~------~------~------~------~
o 1.0 2.0 3.0 4.0

Fig. 19.13. Non-compatible post-localization F.E. strain states.


490 19. Strain localizat ion

The Lemaitre-Chaboche ductile damage model of Sec. 18.2 is used with


exponentiallaw isotropic hardening and nonlinear kinematic hardening. The
behavior under uniaxial tension, and the material parameters are given in
Fig. 19.15.
The numerical computation was stopped when the localization crite-
rion was satisfied for the first time. Figure 19.16 shows a good agreement
between the experimentally measured and the numerically computed load-
displacement curves. We note the good accuracy with which the maximum
load is predicted. The damaged zone in the vicinity of the macro-defect is
depicted in Fig. 19.17 at the load step for which the localization criterion
is first fulfilled. Two possible localizat ion surfaces are predicted. They are
->
depicted in Fig. 19.17 together with their normal vectors N (3= n(3 ((3 = 1,2)
and are computed as follows:

nI = (0.567, -0.823) cx: Y2' el ~ -55°,


n2 = (0.256, -0.966) cx: Yl' e2 ~ +76°, (19.51)

Photographs in (Doghri and Billardon, 1995) and (Billardon and Doghri,


1989) indicate that the experimentally-observed macro-crack initiates at the
hole and towards the notch, in a direction normal to nI.
Just prior to the fulfillment of the localization criterion, the strain rate
computed at the critical point is approximately proportional to (y (3 ~ n(3 +
n(3~Y(3) -(no sum)- Le., the strain rate jump which takes place upon localiza-
tion according to (19.14). The experiment described here is such that strain
localization can be considered as a macro-crack initiation criterion (this can-
not be a general conclusion however). lndeed, for this example, the situation
seems to be such that the rate of the relative displacements of the faces of
the initiating crack are compatible with the strain rate already prevailing.
The amplitude of theses displacement rates are arbitrary, but the opening
and sliding of the final crack faces correspond to the first equilibrium state
of the structure along the bifurcated path (viz. approximately fixed global
displacement in the discussed experiment).
Figures 19.18-19 give some numerical responses at the most damaged
point. Figure 19.18a shows that the peak in the global load-displacement
curve corresponds, locally, to the peak in the stress-strain curve, in the most
damaged point. Figure 19.18b shows that the critical value of damage is
De ~ 0.2 ~ D*, where D* corresponds to the peak of the stress-strain curve in
a uniaxial tension test. This confirms results for Q < O in Sec. 19.4. Actually,
Fig. 19.19c shows that 1:22 = QEu, with Q ~ -5. Furthermore, applying
e
the analytical result (19.45) of Sec. 19.4 gives ~ ±65.9°, which does not
compare badly with actual values (19.51) if we recall that in the analytical
derivat ion there is no shear in the (1,2) plane.
19.5 Numerica! results for a ductile damage model 491

11/2 lin

Il

lin

Fig. 19.14. Notched plate with a macro-defect: (a) Geometry and loading, (b)
mesh and boundary conditions.
492 19. Strain localization

<100 ~tll "1'1' ,,!!r~

o
o
300g
:zoo

100

&

0.8

0.6

o
0.4

0.2

oo 0.1 0.2 0.3 0.4


&

Fig. 19.15. Notched plate: material behavior under uniaxial tension. (a) Stress vs.
strain, (b) damage vs. strain. Material model of Sec. 18.2 with parameter values:
Eo = 72 GPa, v = 0.33, oy = 273.5 M Pa, Roo = 275 GPa, n = 1.861, a =
3302 MPa, b = 37.2,80 = 1.3 MPa.
19.5 Numerical results for a ductile damage model 493

.103·
20 ••• NumericId
,. 1·
.............
, 1·
L,. ••

...Y
,•
..
g
! 10

..
.......... .... ... ....-
:~

°0~--~0~.S~--~ID~---I~.S--~~~0----U~--~)D~~~).~5
U(mm)

Fig. 19.16. Notched plate: experimental and numericalload-displacement curves.

-o
Fig. 19.17. Notched plate: contours of damage D at localization and predicted
localization surface directions
494 19. Strain localizat ion

MPa
000 0:0 0 CXlOOOO o
400 0
o00
o
o
300 o

J2(0")
200

100

O
O 0.05 0.10 0.15 0.20
r

0.25

o
0.20
o
o
0.15 o
D 0
o0
o
0.10
of'
00
o
0.05 o
ocfJ
o
O
O 0.05 0.10 0.15 0.20
r

MPa
2.5
o
00
2.0 #0

aP°
00
1.5 o
Y o
0
o0
1.0 9
o
o
0.5

O
O 0.05 0.10 0.15 0.20
r
Fig. 19.18. Notched plate: local responses at the critical point. (a) von Mises
equivalent stress vs. hardening strain T, (b) damage vs. T, (c) strain energy release
rate vs. T.
19.5 Numerical results for a ductile damage model 495

0.5 000 =0000000 o


000 0000

0.4

°.lJ2(0) 0.3

0.2

0.1

O
o O.OS 0.10 0.15 0.20

0.2

0.15 o
I
#
0.10
,p0
J'
O.S o
00
o
00
O o
O 0.15 O.SO 0.75 1.00

U(nun)

o 0.16
o
o
o
~ 0.12
o
% ~
0.08
\
o
~ 0.04

O.S
-0.16 -0.12 -0.08 -0.04 o
eli

Fig. 19.19. Notched plate: local responses at the critical point. (a) Triaxiality ratio
vs. hardening strain r, (b) r vs. global displacement, (c) 22-strain vs. lI-strain.
496 19. Strain localization

19.6 Nonlocal or internal-Iength models

The problem with local strain-softening models is that there is no internal


length to determine the size of the strain localization zone. The only length
scale is given by the mesh size and it is not possible to converge to a solution
with mesh refinement. Figures ( 19.20) and ( 19.21) illustrate this in the case
of a plane strain compression test for an elastic-plastic material with negative
hardening.

,./
...-/ -
/ r- -
r-r- V"I
,-VI I
Fig. 19.20. Impossibility for local models to converge with mesh refinement when
localization occurs; after Pamin (1994) .

Several solutions have been proposed in the literature in order to include


an internal length in the constitutive relations. A first solution is to use
the theory proposed by the Cosserat brothers in the 1900's, e.g. (De Borst,
1991). In this theory, each material point has six degrees of freedom: three in
translation: Ul, U2, U3 and three in rotation: Wl, W2, W3. SO, there is a kinematic
enrichment in comparison with the classical continuum models. There are not
only strains but also curvatures given by the gradient of the rotation field. The
static quantities associated with these strains and curvatures are stresses and
couple-stresses [Nm/m 2 ] . The constitutive parameters linking the curvatures
and the couple- stresses have the dimension [Pa.m 2 ]. So, internallengths are
introduced in the material behavior. The use of the Cosserat theory may be
justified in the case of granular materials like sand. Some authors also propose
to use it in the case of polycrystalline or crystalline metals, the rotations being
linked to the motion of the grains (Lippmann, 1995) or the curvature of the
lattice, (Mandel, 1973) and (Forest, 1996) . Unfortunately, the regularization
which is provided is totally effective in pure shear only when the rotations
are predominant. Otherwise, only partial regularization is achieved, as shown
in Fig. 19.22 for a compression test.
Some authors have also proposed to use a constrained Cosserat theory
where the rotations correspond to the rotations given by the displacement
field:
19.6 Nonlocal or internal-length models 497

F/(Buy)
1.2,---------------,

0.8

0.6
--·-Homogeneous
0.4 ~x 12 elem .
- 12 x 24 elem .
0.2 -24 x 48 elem.

0.2 0.4 0.6 0.8 1.0 1.2


x 10-2 v,opIH
Fig. 19.21. Load-displacement curves showing the non-convergence with a local
model; after Pamin (1994).

.~ >00:: ~

~
\

X
IX ~e
,x· ~

<1 _ 10 '1

~
12x24 mesh

40 6 x 12mcsh

30 } >< 6 mcsh

20

10

o ylmml
o 120

Fig. 19.22. Partial regularization given by the Cosserat continuum (after (Sluys,
1992)). Deformed meshes (top) and strain profile across the shear band (bottom)
498 19. Strain localizat ion

(19.52)

A good example of the use of such a theory is given in (Fleck and Hutchinson,
1993).
Another way to regularize the simulations is to use an averaging procedure
(Bazant et al., 1984). In that approach, an interaction between neighboring
material points is obtained by replacing some variables by their weighted

=: {
averages over a characteristic volume:

ii aw dV (19.53)
c lv
c

where Ve is the characteristic volume, w a weighting function and a a state


variable (Le. total strain or an internal variable such as the accumulated
plastic strain p in plasticity or the damage variable D). Several problems
remain with such models: the choice of the characteristic volume (e.g., size
and shape), the weighting function and the variable on which averaging is
made.
A third solution is the incorporation of gradients of internal variables in
the constitutive relations (Zbib and Aifantis, 1989). For example, a gradient
plasticity model proposed in the literature is based on a simple modificat ion
of the yield function:

(19.54)

where a eq is the equivalent stress, R(p) the hardening stress, c [Pa.m 2 ) a


material parameter and V'2 the Laplacian operator.
As shown in (De Borst et al, 1993), gradient models can be derived directly
from the non-local integral approach ( 19.53).
The regularization provided by approaches ( 19.53) and ( 19.54) seems
effective for all kinds of loading and convergence is obtained with mesh re-
finement if the size of the mesh is smaller than the characteristic length.
Results obtained with the model described by equation ( 19.54) are shown in
Figs. ( 19.23) and ( 19.24).
Nonlocal models ( 19.53) and ( 19.54) are supposed to reflect the in-
teraction between neighboring material points. However, there is no strong
physical justification for those models and the internal lengths which they
introduce are probably not intrinsic material parameters.

19.7 A two-scale homogenization procedure.

A new approach has been proposed recently by Andrieux et al. (1996a-b) and
studied by Knockaert and Doghri (1999).
19.7 A two-scale homogenization procedure. 499

--
r-I r-

r-r-
- >---1"' I I
Fig. 19.23. Regularization provided by a gradient-plasticity model; after Pamin
(1994).

PI(B(1~)
1.2 . , . - - - - - - - - - - - - ,

1.0

0.8

0.6
_. Homogeneous
0.4 -6x 12R32EG
- 12 x 24R32EG
0.2 -24 x 48R32EG

0~--~----~---'---4

v",/H
Fig. 19.24. Load-displacement curves showing the regularization provided by the
gradient dependent model; after Pamin (1994).
500 19. Strain localization

In continuum mechanics, a (macroscopic) material point corresponds to


a RVE which is the smallest volume of material having the same properties
as the material on a more macroscopic scale, see (Nemat-Nasser and Hori,
1993) for a detailed account.
At a finer (micro) scale, the material inside the RVE can be very het-
erogeneous. For example, for many metals, a RVE is viewed as a polycrystal
containing a large number of crystals ("sub-cells") with random orientations.
The macroscopic strain and stress fields should be viewed as averages of the
corresponding micro variables, see Sec. 20.2.
Andrieux et al. consider that each RVE is divided into N sub-cells Y/L
characterized by their position vector z/L from the center of the RVE and
the values of the strain f./L and the internal variables a/L in the sub-cell. A
2-D representation of a RVE is given in Fig. 19.25. It is assumed that the
internal variables are not uniform within the RVE, they can vary between
the sub-cells. However, the strain is uniform in the whole RVE (Voigt model);
otherwise, the homogenization method would lead to the appearance of extra-
stresses associated with the gradient of the strain components. In this paper,
quantities defined at the microscopic (Le., sub-cell) level are designated by
the superscript JL.
The behavior of each sub-ceH is determined by thermodynamic potentials:
free energy 'IjJ( f./L, a/L) and dissipation pseudo-potential 4J( n/L) which give the
equations of state and the evolution equations, respectively.
The macroscopic variables are the strains, the internal variables and the
gradients of the internal variables: f.(x),a(x), V'a(x) == âa/âx, where x is
the position vector of a macro material point in a fixed Cartesian coordinate
system.
The micro and macro variables are linked by the following relations :

f./L = f.(x), a/L = a(x) + V'a (x) . z/L (19.55)

The first relation in the equations expresses the hypothesis that the strains
are uniform in the RVE and the second one imposes a linear spatial variation
of the internal variables within the RVE. This second relation must be verified
at aH time, which gives the foHowing compatibility condition:

n/L = n(x) + V'n(x) . z/L (19.56)

A one-dimensional case, with three sub-ceHs and a scalar internal variable is


illustrated in Fig. 19.26.
The macroscopic thermodynamic potentials are obtained from an average
of the potentials in the sub-ceHs. For the free energy, a simple arithmetic
average is used:
N
- 1
tJi( f., [a , ... , a ])
N
= N1",
~ 'IjJ( f., a/L) (19.57)
/L=1
19.7 A two-scale homogenization procedure. 501

Zll
1
(

Zll
2

Fig. 19.25. 2D Representation of a RVE

time t +6.t

timet

Fig. 19.26. Linear variat ion of the intern al variable in the RVE in 1D
502 19. Strain Iocalization

A Taylor expansion and relations ( 19.55) give the macroscopic free energy
as follows (details of the derivation are given in Sec. 19.11.2):

â2'ljJ
lP(e, a, V'a) 'ljJ(e,a)+ ( V'a. âa 2 • V'a ) :J,

J -2N1 LZ
N

Jl=l
Jl ®zJl (19.58)

(The symbol "." has the same meaning as in Sec. 12.11.2). So there appears
a second order symmetric tensor J which depends on the micro-structure of
the material through the microscopic positions zJl of the sub-cells within the
RVE and their number N. The components of J have dimensions of square
lengths. Note that in other formulations, internal lengths appear as scalars
and not as a tensor as it is the case here. Note also that internallengths (via
J) appear naturally here as the result of a homogenization procedure and
are not introduced ad-hoc as in other models (e.g., Vc in Eq. (19.53) or c in
Eq. (19.54)).
As regards the macroscopic dissipation potential, it must be defined in
such a way that the rate of the intern al variables in the sub-cells derived from
the microscopic or macroscopic potentials are equivalent. An arithmetic mean
does not allow this requirement if, for example, the microscopic dissipation
potential is homogeneous of degree 1. In this case, Eq. ( 19.58) shows that the
macroscopic dis si pat ion potential would not include any higher order term
and would not be different from the microscopic potential. So, a quadratic
mean is defined:

(19.59)

and, with a Taylor expansion and relation ( 19.56):

4>(0, V'a) = 4J(a) + [V'a. ~:~ • V'a + ~ (;: • V'a) ® ( ; : • V'a)] : J


(19.60)
Details of the derivation are given in Sec. 19.11.3. The macroscopic variables
can be derived from the macro. potentials according to:

(19.61)

(19.62)

where (T is the macroscopic stress and Aa, AVa the macroscopic thermody-
namic forces associated with a and V'a, respectively.
19.8 Numerical algorithms 503

An evident constraint is that V'o must be the spatial gradient of o. En-


forcing this constraint and the previous evolution equations may lead to a
lack of solutions. A possibility is to preserve the generalized standard ma-
terial formalism at the global (or the structural) scale only (Nguyen, 1987).
A similar procedure -although simpler since there are no spatial gradients
involved- was also used in (Simo et al, 1989).
The thermodynamic forces A associated with the internal variables o are
defined through two global constitutive equations:

(19.63)

(19.64)

The right-hand sides of the equations are the directional derivatives of the
global potentials with respect to o and a, respectively.
Under certain conditions, the problem can be rewritten as follows: given
Aa == -8tPj8o and AVa == -8tP/8V'o, find a which minimize the func-
tional:

(19.65)

According to Andrieux et al(1996a-b) , the existence of a solution can be


proved under an assumption of coercivity on P. For a positive homogeneous
potential with degree one (material behavior with threshold), the problem
remains open. Finally, an extended Clausius-Duhem inequality is ensured:

(19.66)

19.8 Numerical algorithms

Numeri cal algorithms have been developed and implemented by Knockaert


and Doghri (1999) in order to study the non local formulation of Sec. 19.7 in
the uniaxial case. Material models with a single scalar macroscopic internal
variable o: are considered. A finite-element-based procedure was developed
with discretization of both the displacement and the internal variable fields:
M M'
U(x, t) ~ L Ui(t)~i(X); o:(x, t) ~ L O:i(t)~:(x)
i=l i=l
(19.67)
504 19. Strain localization

where M and M' are the number of nodes used for the discretizat ion of the
displacement and damage fields respectivelYj ~i(X) and ~~(x) are the corre-
sponding shape functionsj .1t == t n+! - tn is the time increment.
With the discretization presented above, minimization of the functional
( 19.65) leads to a system of nonlinear equations. Suppose that the variable
o: cannot decrease with time (e.g. non-healing damage), then enforcing the
constraint a ;::: o with, for example, a penalty method, implies that the
following functional is considered instead of ( 19.65):

(19.68)

where the penalty parameter C is equal to zero when a is positive and has a
positive value when a is negative. Asking for the stationarity of Te gives:

8Fc(o+iJ1J,Vo+iJV1J) I -
8iJ iJ=O -

In [~~77 + 80:0 \177 - Aa77 - AVa \177 + Ca77] dfl = O (19.69)

Using Galerkin's method, the test functions 77 and \177 are replaced by the
shape functions ~i (77 -+ ~i, \177 -+ \1~i)' So, the following system of equations
is obtained:

(19.70)

Using the discretizat ion of the displacement and internal variable fields, the
unknowns are the nodal values of o: at time tn+!. A fully implicit time inte-
gration scheme is used (backward Euler, (J = 1). The system is solved with
a Newton-Raphson method (it is assumed here that Aa and AVa do not
depend on a).
Before we were able to obtain the numerical results which will be presented
in sections 19.9 and 19.10, we had to solve several numeric al problems, some
of which are discussed hereafter.
One problem is that the functional ( 19.68) is very "nasty" and the meth-
ods we tried initially (penalty, Lagrangian, augmented Lagrangian) in order
to enforce the constraint a ;::: O were not robust enough. The method used in
the following is simple and gives good results but necessitates sometimes a
large number ofiterations (~1O). The idea behind this method is illustrated
in Fig. 19.27.
Starting from known values of the strain field f(X, tn) and the nodal val-
ues of o: at time tn (O:i(t n ), i = 1, number of nodes), we need to corn pute
O:i(tn+d corresponding to a strain field f(X, tn) + .1f(X). For that purpose,
O:i(tn+d is initialized at (O:i(t n ) + .1) where .1 is a positive constant. In parts
of the rod where O:i tends to increase, there is no problem. However, if O:i
tends to decrease (o::-l(t n +!) + .1o:~ < O:i(t n ), which is possible dur ing the
19.8 Numerical algorithms 505

damage

--.._t--__ t---_-_-_------~ ai(t n )

Position along the rod


Fig. 19.27. Iterative update of nodal damage variables Oi at tn+l between iter-
ations (k - 1) and k: Oi (t n+l) = O~-l(tn+t} + <: Llof, where <: is chosen so that
Of(tn+l) ~ Oi(t n ).

iterations), instability occurs and no convergence can be obtained. So the


increments ..::lai given by Newton-Raphson are truncated in order to always
keep a slightly increasing damage:

(19.71)

where ( is chosen so that the condition a~(tnH) 2: ai(t n ) is satisfied for alI
iterations. If the damage has to remain constant (in case of elastic unloading
or if the damage threshold is not reached), the corresponding residuals cannot
be brought to zero and the iterations are stopped when the error is stationary.
Another problem arises from the fact that the tangent (âa/âf) cannot be
computed analyticalIy for the nonlocal formulation of Sec. 19.7. Numerical
computation of this tangent has been investigated without success. In the
nonlocal case, the best results were obtained with the initial elastic stiffness.
Numerical investigation has also shown that the non-homogeneous 1D
solution seems to exhibit a "snap-back" in many cases with the nonlocal for-
mulation. We have implemented a modified Riks algorithm (Crisfield, 1981)
in order to overcome this difficulty and solve the equilibrium problem. The
implementation was checked against the analytical results of the local case
(Sec. 19.1, Fig. 19.3).
506 19. Strain localization

19.9 Elasticity with damage- Model without threshold

A first application presented in (Knockaert and Doghri, 1999) deals with a


damageable elastic model without a threshold.

19.9.1 Local constitutive equations

The behavior is described by the following equations in 1D:


1
(7 = E(l - D)f; D(t) = -SUPT<tf(r)
fr -
(19.72)

where E is Young's modulus of the virgin material, fr the failure strain and
Da damage variable which ranges from zero for the undamaged material to
1 for a RVE with null stress-carrying capability. The corresponding stress-
strain response is shown in Fig. 19.1. The model was used in Sec. 19.1 in
order to illustrate bifurcation phenomena exhibited by strain-softening local
models.
The stored energy (per unit volume) corresponds to the area of the tri-
angle ABC in Fig. 19.1. It is equal to:
1 1
'IjJ(f, D) = "2 (7 f = "2 E (1 - D) f2 (19.73)

The dissipated energy is computed for a strain cycle O -t i(t) -t O as in


(Florez, 1989):

(19.74)

Since i(t) = ffD(t), we obtain:

Dissipated energy = Efi D 3 (t) (19.75)


6
Finally, time derivation of the previous expres sion gives the dissipated power:

(19.76)

19.9.2 Nonlocal macroscopic formulation

Using the potentials in Eqs. ( 19.73) and ( 19.76) as microscopic potentials


for the sub-cells and relations ( 19.58) and ( 19.60), the following macroscopic
potentials are obtained (in 1D):
19.9 Elasticity with damage- Model without threshold 507

lJi( E, D) 'lj;(E, D),

~(D,iJ, \lD, \liJ) = fjJ(D,iJ) + ~ (âfjJ \l D + â~ \l iJ) 2 J


fjJ âD âD
â2fjJ 2 2
+ ( âD2 (\lD) + 2 â fjJ. \lD\liJ) J
âDâD
(19.77)

E,Dare the (scalar) macroscopic strain and damage, respectively; J is also


a scalar equal to the square of a characteristic (or internal) length Le = VJ;
and \l D == â D / âx. Note that since fjJ depends on D, a generalizat ion of result
( 19.60) was needed.

19.9.3 Numerical simulations

The problem treated here is a 1D rod under tension; the geometry is illus-
trated in Fig. 19.2. Values of the material parameters we used are: E = 200
GPa, Ef = 0.2.
The load-displacement curves for a rod of length L = 100 with a weakened
part of length Ld = 30 (E multiplied by a factor TI = 0.9 in order to trigger
localization) and characteristic lengths Le of 5,10 and 30 are shown in Fig.
19.28.

. .......~~:.'--.:.;~:::::.~:~:,....•-.....
0.9 .....

F
) ".
0.8

(A Etf /4) 0.7

0.6 lc=Ldl6 -
lc = Ldl3 '''.
Lc = Ld ..
0.5

0.4

0.3

0.2

0.1

o
o 0.1 0.2 0.3 0.4 05 06 07 08

u
tf L
Fig. 19.28. Load-displacement curves with the non-local model

Note that these curves exhibit a snap-back. This behavior, which may
appear as physically unrealistic (immediate failure of the structure for a load
or displacement-controlled system), is obtained when the localization zone is
508 19. Strain localization

small with respect to the size of the structure. As the size of the localization
zone is controlled by Le, snap-back may be obtained or avoided depending
on the value of this parameter (see also next section).
In the local case, the damage profile is given by a step function, while the
non local formulat ion smoothes out the discontinuity (see Figs. 19.29 and
19.30).

Le = Ld/6 -

0.8

0.6
E::===::::::;::

0.4

0.2

20 40 60 80 100
position along the rod

Fig. 19.29. Damage profiles (Le = Ld/6) .

If localization occurs, the damage profile is controlled by Le; this is il-


lustrated in Fig. 19.30. In that figure, the maximum value of damage (in
x = O) is the same in ali cases. This corresponds to different values of the
normalized load: 0.71 in the local case and 0.82, 0.87, 0.78 in the non-local
case for Le = Ld/6, Ld/3 and Ld, respectively.
When Le is smaller than the length Ld of the weakened part, localizat ion
is possible. If Le is close to Ld, localizat ion does not occur and damage and
strain remain nearly homogeneous along the rod despite the weakening (see
Fig. 19.30).
Convergence with mesh refinement has also been investigated and proved.
This is illustrated for another material model in the following section.
The computations were stopped before reaching D = 1 in the most dam-
aged part of the rod because of numerical instabilities (D fina1 = 0.72, 0.79, 0.76
and afinaI/(E€r/4) = 0.82,0.75,0.72 for Le = Ld/6,Ld/3 and Ld, respec-
tively). Similar difficulties were also encountered with other constitutive mod-
els which include the gradient of the damage variable; see for example (Geers,
1997). The problems may be due to some inaccuracy in the computation of
19.10 Elasticity with damage- Model with threshold 509

Le=Ld/6 -
Le = Ld/3 ------
Le = Ld ----
Local model -..........
0.8

0.2

o L -_ _ _ _ _ _ ~ ______ ~ _ _ _ _ _ _ _ L_ _ _ _ _ _ _ L_ _ _ _ _ _ ~

O 20 40 60 80 100
posilion along Ihe rod

Fig. 19.30. Comparison of the damage profiles for a given value of the maximum
damage and several values of the characteristic length

the damage field, and their appearance depends on the material model (see
next section).

19.10 Elasticity with damage- Model with threshold

In a second application presented in (Knockaert and Doghri, 1999), a dam-


ageable elastic model with a threshold is simulated.

19.10.1 Local constitutive equations

The approach of section 19.7 has also been tested for another model of
elasticity with damage but with a threshold. With this model, damage begins
to grow progressively after some amount of deformation f y . Variants of this
model have been used in order to model the damage behavior of concrete (see
(Florez, 1989) and references therein). The behavior is depicted in Fig. 19.31
and is described by the following relations linking stress, strain and damage:

(1 = Efexp( -D) and D(t) = 21n C:)


= sUPo:'ST9 f (r) (19.78)

Exercise: using a procedure similar to that of Sec. 19.9.1, show that the
potentials are given as follows:
510 19. Strain localization

1
Free energy: 'IjJ(f, D) = 2Ef2 exp( -D)
Ef; .
Dissipation : rp(D) = -D (19.79)
2

0.8

0.6

0.4

0.6 0.8 1.2 1.4 1.6 1.8

f.JE y
Fig. 19.31. Elastic behavior with damage, model with threshold

19.10.2 Nonlocal macroscopic formulation

Exercise: using Eqs. ( 19.79), ( 19.58) and ( 19.60), show that the macro-
scopic potentials are given by:

tJ!(f,D, \lD)

p(D, \l D) = (19.80)

Le is again a scalar characteristic (or internal) length equal to /J.

19.10.3 Numerical simulations

The example of the rod of length L = 100 is treated again with a weakened
part of length Ld = 30 (Young's modulus E decreased by an amount of 10
19.10 Elasticity with damage- Model with threshold 511

percent) and several values of the internal Iength Le. Material parameters
are: E = 200 GPa, f.y = O.Ol.
Load-displacement curves are shown in Fig. 19.32 for Le = 5 and element
sizes 1,5 and 10 which correspond to a number of elements equal to 100, 20
and 10, respectively. The figure shows that convergence is obtained if the
element size is smaller than Le.

0.9
F
AEEy 0.6

0.7

0.6

0.5

0.4 number of elements =10 -


number of elements = 20 . __ ._-
number of elements =100 -
0.3

0.2

0.1

O
O 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.6 0.9

U
EyL
Fig. 19.32. Convergence with mesh refinement (Le = Ld/6).

Comparison between nonlocal and local models is made in Fig. 19.33.


The solution with the local model exhibits a snap-back. This is also the case
with the nonlocal model for a small value of Le. For higher values of this
parameter, the solution tends towards a homogeneous solution and damage
localization is prevented. An interesting fact is that the threshold is modified
by the use of the nonlocal model and the maximum load reaches nearly the
load corresponding to a rod without weakening (95 percent of the maximum
load given by AEf. y where A is the cross section area of the rod). Note that
the convergence obtained with this model is much better than the conver-
gence obtained with the previous model. The calculations were stopped for
a maximum damage close to 1.
Damage profiles for the same value of the maximum damage (D = 0.996)
are given in Fig. 19.34. Those profiles correspond to the following values of
the normalized load: F/(AEf. y ) = 0.545 in the local case and 0.66, 0.65, 0.6
for Le = Ld/6, Ld/3 and Ld, respectively. The damage distribution becomes
smoother and smoother with an increasing value of Le. For a small Le, damage
512 19. Strain localization

local model
0.9 Le = Ld/6 .+ •..
Le = Ld/3 o
Le = Ld •
0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

o
o 02 0.4 0.6 0 .8 1.2 1.4 1.6 1 .8

-1L
EyL

Fig. 19.33. Load-displacement curves for the non-local and the local model (Ld =
30)

0.9

0.8

0.7

0.6 local model -


a:> Lc=Ldl6 •••• •.
Ol
OI
0.5 '. Lc=Ld13 .
E Lc=Ld
OI
"ti
0.4
\ ...............
0.3
......
0.2 .... ....
...... ......
0 .1
....... . . .
O
O 20 40 60 80 100
position along the rod

Fig. 19.34. Damage profiles for the same value of the maximum damage
19.11 Appendices 513

increases in the weakened part and even a little bit further, the right end of
the rod remaining undamaged. For Le = Ld, damage develops in the whole
rod and is nearly uniform in the whole structure. For intermediate values, the
whole rod begins to damage and localizat ion occurs later. The same remarks
apply for the strain distribution (see Fig. 19.35).

2.2
local model -
Le=Ld/6 .....
2 Le=Ld/3 ...... .
Lc=Ld ........_.
1.8

1.6

E/E y 1.4

1.2

0.8

0.6

0.4 L--_ _ _-'----_ _ _-'--_ _ _-'--_ _ _--'-_ _- - - - '

O 20 40 60 80 100
posilion along Iha rod

Fig. 19.35. Strain profiles for the same value of the maximum damage

19.11 Appendices
19.11.1 Strain localizat ion criterion in 2D

The strain localization criterion det(n . H . n) = O takes a simple form for


plane strain or stress, shells or plates. lndeed, in those cases we have:

Al1
n = (nI, n2, O), n~ + n~ = 1, n· H . n = [ A~1 (19.81)

where:

n~ H + nln2(H1112 + H 121l ) + n~HI212,


l111

n~H1212 + nln2(H1222 + H 2212 ) + n~H2222,


n~H1112 + nln2(H1122 + H l212 ) + n~HI222,
n~HI211 + nln2(H1212 + H 221 d + n~H2212,
n~ Hl3l3 + n~H2323 (19.82)
514 19. Strain localization

We generally have: C > o. For instance, for the damage model of Sec. 18.2,
C = (1 - D)/.lo and is positive if D < 1. Therefore, it suffices to study the
sign of (det A = Al1A22 - A 21 A 12 ). The expressions of Aa,6 give:

(19.83)

with:

ao Hl1l1H1212 - H1112H1211 ,

al + H 2212 ) - Hl1I2H221I - HlI22H121I


H lIlI (H1222

a2 Hl1l1H2222 + H1112H1222 + H121IH2212 - H 1212 (H lI22 + H 22lI )

- H1I22H221I

a3 H 2222 (H 1112 + H 121I ) - H1I22H2212 - H1222H2211

(19.84)

We introduce the following notation:


1f 1f
nI = cosO, n2 = sinO, O E)- - +-)
2' 2
(19.85)

If 0= 1f /2, then (det A = O) if a4 = O. If Of. 1f /2, then (det A = O) if:

(19.86)

The equation has no solution in elasticitYj an example is given hereafter:


Exercise: for the damage model of Sec. 18.2, prove the following results

n·E·n (1 - D) [(Ao + /.lo)n @ n + /.l01] ,


det(n· E· n) (1 - D)/.l6(Ao + 2/.l0) > 0,

(n·E·n)-l 2(1 + v) ( 1- 1 n@n ) , (19.87)


(1 - D)Eo 2(1 - v)

where 1 is the second-order identity tensor, (A o, /.lo) the initial Lame coeffi-
cients and (Eo , v) the initial Young's modulus and Poisson's ratio.

When there is plastic loading, Ortiz et al. (1987) make the following as-
sumption: before strain localization, f(x) > O. Therefore, it suffices to find
the minima of f(x)j localization corresponds to the instance when one (or
more) minima Xm is (are) such that f(x m ) = O. When a4 f. 0, the minima of
f(x) are the roots of the following equation:

(19.88)

which can be solved in closed form using Cardan's formulae (see Appendix B).
The numeric al examples of Sec. 19.5.1 (biaxial plane stress with f22 = aflI)
19.11 Appendices 515

showed that the above-mentioned assumption (i.e., finding the minima X m of


f(x) for which f(x m ) = O) is valid for states such that o: < O. Figure 19.36
shows indeed that f(x) is positive at the beginning and then moves down until
one of its minima encounters the x-axis: strain localization occurs. For cases
where o: > O however, Fig. 19.37 shows that f(x) > O before localization,
but its behavior at localization is more brutal: it goes upside down. Figures
19.36-37 are taken from (Doghri, 1989).
In practice, we can proceed as follows in order to detect the onset of
localization. Let Xm be the real roots of Eq. (19.88). If a4 > O and f(x m ) > O
for all m, then there is no localization. Otherwise, we can find solutions for
f(x) ~ O. A solution refinement can be achieved by restarting the time step
with a smaller time (or load) increment.
For three dimensional states, a procedure is proposed in (Ortiz et al, 1987)
which consists in minimizing det(n· H . n) subject to linii = 1. Again, as in
the 2D case, the procedure may not be general enough.

19.11.2 Macroscopic free-energy potential

Starting from the microscopic free-energy potential of each sub cell 'l/J(e, 0:1'),
the macroscopic potential corresponding to the RVE is given by an arithmetic
mean:
N
- 1 N
lP( e, [o: , ... , o: ]) = N1",
L...i {'l/J( e, 0:1')} (19.89)
1'=1

where N is the number of sub-cells in the RVE. Using relations (19.55) be-
tween micro and macro variables and assuming that 11V'0:' zl'll « 110:1'11, a
second-order Taylor expansion of Eq. ( 19.89) gives:

lP(e, 0:, V'o:)

(19.90)

Developing the sum gives :

Assuming a symmetric distribution of the sub-cells inside the RVE so that


L::=1 zI' = 0, we finally obtain:
516 19. Strain localization

f(x)

.(\
5. _ ::: 1f'.v,lv
:

·· .· , ·
· ·· ··· .·
...
··· ·.. .·•·
,

··.
~

1.. x

Fig. 19.36. Biaxial plane stress loading of Sec. 19.5.1 in the case a = -8 and
n = 1. Evolution of the localization function f(x) of Sec. 19.11.1.
19.11 Appendices 517

f(x)

··
10.

:
!
5.

-20. o. 20.
x
......

-5 .
.
..
-10.
)tE 10 11

Fig. 19.37. Biaxial plane stress loading of Sec. 19.5.1 in the case Q = 2 and n = 0.6.
Evolution of the localizat ion function f(x) of Sec. 19.11.1.

(19.92)

where J is the tensor describing the "micro-structure" of the material

1 N
J = - "'"' zlJ. ® zlJ. (19.93)
-2N~
1J.=1

19.11.3 Macroscopic dissipation potential

Starting again from the microscopic potential q;(ixlJ.) , the macroscopic poten-
tial is obtained by a quadratic averaging :

(19.94)

Using the compatibility condition (19.56), and assuming II\7ix' zlJ.ll « IlixlJ.lI,
a second-order Taylor expansion of Eq. ( 19.94) gives:
518 19. Strain localizat ion

p(o, V'o) {~ t. [4>(0) + 8:~) • (V'o· z/J)


(19.95)

After developing the terms of the sum, assuming that '2::=1 z/J = O and
neglecting the terms of order 1IV'0 . z/Jlln for n > 2, we obtain :

Jo(";',
~....
r7";')
V <A
~ '+',1.(.)
•- a
{1 + 4>2 2 (84)(0)
(o) ~. va
r7.)!O. (84)(0)
v
r7.) J
80. v a :

_2 (r7. 8 4>(0) r7.). J} î


2
(19.96)
+ 4>(0) va. 80 va. 2 •

Using the Taylor expansion of the function J(1 + x) truncated after the first
terms, we finally obtain:

p(o, V'o) :::::: 4>(0) + [V'o. 8;1~) • V'o


_1 (84)(0).
+ 4>(0) 80
r7va.) ® (84)(0).
80
r7')]
va
:J (19.97)
20. Micro-mechanics of materials

In all previous chapters (except Sec. 19.7), we presented phenomenologi-


cal and macmscopic material models. Those models have the advantage of
being naturally integrated into continuum mechanics, structural mechanics
(theories of beams, plates and shells) and analytical or numerical methods
(e.g., finite elements) which are used by engineers. The approach however
does not incorporate any information from the micro-structural level and
thus does not permit to understand and quantify the influence of the micm-
structure on macroscopic properties. A solution consists in developing a so-
called micro/macm approach. This chapter gives an introduction to the gen-
eral methodology as well as a detailed development for semi-crystalline poly-
mers.

20.1 Micro/macro approach

In a phenomenological and macroscopic constitutive modeling approach, el-


ementary experiments (tension, compression, bending, torsion, constant or
variable loading rates, constant or variable temperature, etc.) are carried
out on specimens of macroscopic dimensions. Later, constitutive mathemat-
ical models (elasticity, viscoelasticity, plasticity, viscoplasticity, etc.) are de-
veloped in order to describe the experimentally-found results; they contain
parameters which can be identified from macro. experiments. Since the ap-
proach does not include any micro-structural information, it cannot answer
the following question:
"Which micmscopic properties (e.g., for semi-crystalline polymers: crys-
tallinity, molecular weight, thickness of lamellae, etc.) should be modified,
and how, in order to improve, or in general modify a given macmscopic prop-
erty (e.g., initial yield stress, ductility, tensile strength, toughness, etc.)?"
This question has scientific and industrial relevance. An answer is pro-
vided by a so-called micro/macro appmach which can be described as follows:

- (1) A macro. material point is viewed as the center of a representative


volume element (RVE).
520 20. Micro-mechanics of materials

- (2) At a smaller -micro.- scale, a RVE contains a finite number of con-


stituents (e.g., crystals, inclusions in matrix). A constitutive model for
each of the constituents is constructed.
- (3) The macro. constitutive behavior of a RVE is found via a micro./macro.
transition.
- (4) Classical (almost!) continuum mechanics computations at macro. scale
are performed with macro. constitutive equations from step (3).
- (5) Macro/micro. transition: at each time and at each macro. material
point, a "numerical zoom" can be done in order to see what happens at
the micro. level (e.g., texture development).

The approach is illustrated in Fig. 20.1 in the case of semi-crystalline


polymers (Sec. 20.3).
Knowing the micro. behavior, the overall (macro.) behavior of a RVE -
step (3)- is usually obtained via a homogenization procedure (Voigt, Reuss,
self-consistent, Mori-Tanaka, etc.); this is explained in the next section.

lamellae
amorphous
layers
~i555iil
" inclusion"
Fig. 20.1. Micro/macro model for the small deformation of semi-crystalline poly-
mers. Body at the macro. scale, representative volume element (RVE), inclusion.

Before we close this section, let us remark that multiple scales always
co-exist in reality; working at one scale or another depends depends on the
20.2 Homogenization schemes 521

problem we want to solve or/and on one's viewpoint. In order to illustrate


this, we consider an example taken from (Bisplinghoff et al., 1965): a metallic
cylinder. On a macro. scale (that of continuum mechanics), this is a homoge-
neous and isotropic solid; "to the metalIurgist, a collection of randomly ori-
ented grains and crystals; to the chemist, a collection of atoms and molecules;
and to the solid-state-physicist, a swarm of nuclei and electrons" .

20.2 Homogenization schemes


Our aim in this section is to present a brief introduction to the theories of
homogenization as applied to material modeling. A wealth of information and
details can be found in numerous articles and textbooks, e.g. (Nemat-Nasser
and Hori, 1993).

20.2.1 Average strains and stresses

Suppose that we carry out a classical continuum mechanics computation. At


each macro. point, we know the macro. strain E or the macro. stress u. If
we know E, we wish to compute u, and vice-versa. We have seen in previous
chapters how to compute the desired quantity for macro. models, we now
need to carry similar calculations for micro-mechanicalIy-based models. First
of alI, we establish two basic results, and we shall see in the next subsections
how to relate E and u.
Remember that a macro. point is viewed as the center of a RVE w. Let
us start with the first case: E given. This means that at micro. level, the
boundary aw of the RVE is subjected to imposed displacements:

Ui(X) = fijXj(X)j X E âw (20.1)

Now, we define the average strain in the RVE as follows:

(20.2)

where V is the volume of the RVE and tij(X) the (infinitesimal) strain field
insi de the RVE:
1 aUi aUj
tij(X) == - ( - + -), (20.3)
2 aXj aXi

where Ui(X) is the displacement field insi de the RVE. Integration by parts
gives (n being the outward unit normal to aw):

(20.4)
522 20. Micro-mechanics of materials

where we used Eq. (20.1). Another integration by parts gives:

(20.5)

where 8kj is Kronecker's delta. Combining Eqs. (20.4-5), we find:

(20.6)

From Eqs. (20.2-3), it is readily deduced that:

(20.7)

In other words, the average strain equals the macro. strain. Note that for
equality (20.7) to make sense, the strain tensors have to be properly rotated
so that the equality is expres sed in the same basis, e.g. a global one (see Ap-
pendix C). The same remark holds every time we compute average quantities
in the remainder of this chapter.
We now consider the case where the stress (T at a macro. point is given.
Since the point is actually the center of a RVE w, the boundary 8w of the
RVE is subjected to an imposed traction:

(20.8)

where n is the outward unit normal to 8w. We use the identities:

and integration by parts to obtain the following result:

l (Tij dV

(20.9)

where we used B.C. (20.8) and assumed the stress field inside the RVE to be
self-equilibrated. Finally, using the previously established result (20.5), the
stress average over the RVE is found to be:

(20.10)

i.e., the ave rage stress equals the macro. stress.


It appears from (20.7, 10) that the problem of relating macro. strains and
stresses l and (T can be transformed onto the problem of relating average
20.2 Homogenization schemes 523

strains and stresses < € > and < u >. In linear elasticity, the problem
becomes: find the macro. stiffness e such that:

< u >=e:< € > (20.11)


This means that the heterogeneous micro-structure of a RVE can be replaced
at the macro. scale with a homogeneous material with macro. stiffness e.
In order to find e, it is conceptually possible to compute the micro. stress
field ifE is given (or the micro. strain field ifu is given), then compute < u >
(or < € » and finally find e solution of linear system (20.11). In practice,
however, this is computationally prohibitive, because we would have to solve
a boundary-value-problem (BVP) for each RVE. For example, in a finite-
element (F.E.) setting one solves a BVP at the macro. level (requiring a
macro. F.E. mesh, etc.), and then each Gauss integration point is viewed as
the center of a RVE where another BVP must be solved (a F.E. mesh for
each RVE is also needed). Conceptually, one could also work directly at the
micro-Ievel, but this is even more expensive than the two-Ievel approach.
What is done in practice is to use a homogenization model which allows
to compute e in a much cheaper way. The final result depends on the as-
sumptions underlying each homogenization scheme.

20.2.2 Voigt model

Consider that a macro. strain E is imposed. Voigt model assumes a uniform


strain field inside the RVE:

€(a:) = E, Va: E w (20.12)


Consequently, Eq. (20.7) is trivially satisfied. Considering linear elasticity, we
obtain the average stress as follows:
< u >=< C : € >=< C >: ...............
E , (20.13)
<€>
where c is the micro. stiffness field. Equation (20.13) implies that the macro.
stiffness e is given by:
e =< c >, (20.14)
Le. the average stiffness. In 1D, Voigt scheme corresponds to the well-known
model of bars in parallel. In 3D however, Voigt model has been used for
general multi-axial situations as well as material and geometric nonlinearities.
The model has been used in the famous theory which was developed by G.I.
Taylor in the 1930's for polycrystalline metals, where each RVE is seen as
a collection or "aggregate" of crystals. Plastic deformation in metals occurs
via dislocation mot ion over well defined slip planes and directions. Taylor was
able to predict macro. stress-strain responses from crystals'- based behavior.
See Sec. 20.4 for references.
524 20. Micro-mechanics of materials

20.2.3 Reuss model

Consider now that a macro. stress (i is imposed. Reuss model assumes a


uniform stress field inside the RVE:

u(x) = (i, 't/x E w (20.15)


As a consequence, Eq. (20.10) is trivially satisfied. In linear elasticity, the
average strain is obtained as follows:

< € >=< d: u >=< d >: (i


~
, (20.16)
<u>
where d is the micro. compliance field. Equation (20.16) implies that the
macro. compliance d and stiffness care given by:
d =< d >, i.e. c- 1 =< c- 1 > (20.17)
In lD, Reuss scheme corresponds to the well-known model of bars in series.
In 3D however, the model has been used for general multi-axial situations as
well as material and geometric nonlinearities . For polymers for example, the
model can give decent predictions.

20.2.4 Self-Consistent model

We now consider the case where a RVE is an aggregate containing a


large number of "grains" or "inclusions" . The basic assumption in the self-
consistent scheme (S.C.) is the following:
The state of each inclusion is equivalent to that of the inclusion alone in
a homogeneous matrix having the same stiffness as that -unknown- of the
whole aggregate (Fig. 20.2) .

Fig. 20.2. Ellipsoidal inclusion in a homogeneous matrix.

The basic results are due to Eshelby (1957) who has solved the problem
of a single ellipsoidal inclusion in an infinite body submitted to a remote
uniform strain €. His assumptions were:
20.2 Homogenization schemes 525

- (i) Linear elastic inclus ion of uniform stiffness el,


- (ii) Linear elastic matrix of uniform stiffness e M .

The main results he found were:


- (1) Strain EI in the inclusion is uniform.
- (2) Inclusion strain EI is related to the imposed remote strain e as follows:
E1 = AI( e,e
1 M ,a.r. ) :E,
- (20.18)

where AI is an operator which depends on the stiffnesses el and e M ofthe


inclusion and the matrix, respectively, and the aspect ratio (a.r.) of the
ellipsoidal inclusion.

Now, using Eshelby's results within the context of the S.C. scheme (the
size of each inclusion must be small compared to that of the RVE), we obtain
for each inclusion (1):

EI =AI(e I , C ,a.r.):e; (71 =eI:E I (20.19)


'--'
unknown!

The average stress is given by:

< (71 >=< el :A I >: e = C:< EI >, (20.20)

where the overall (macro.) stiffness of RVE is given as follows:

(20.21)

For "aggregates" (single-phased RVE's, e.g. polycrystals) the S.C. scheme


usually gives macro. predictions which are close to experimental data. How-
ever, Eq. (20.21) shows a serious difficulty associated with the S.C. scheme:
even in linear elasticity, C obeys a nonlinear problem.
If a uniform macro. strain € is applied to the boundary of RVE, then
< (71 >= C: €. If a uniform macro. stress u is applied, then < EI >= d: u.
For a S.C. scheme, it can be shown that: d = C -1, which explains the name
" self-consistent" .
Voigt model (uniform strain in the inclusions) is a particular case of the
S.C. scheme:
EI = e, AI = 1, c =< el >,

where 1 is the fourth-order identity tensor. It can be shown that Voigt and
Reuss (uniform stress) models give upper and lower bounds, respectively. In
practice, however, these two limits are too far away one from the other. There
are better (closer) bounds, e.g. Hashin-Shtrikman.
526 20. Micro-mechanics of materials

20.2.5 Mori-Tanaka model

We now consider the Mori-Tanaka (M.T.) homogenization scheme which is


used in two-phase materials where the RVE consists of dilute inclusions in a
matrix (Fig. 20.3). The basic assumption in the model is the following:
The state of each inclusion is equivalent to that of the inclusion alone in the
matrix (Fig. 20.2).

Fig. 20.3. Dilute inclusions in a matrix.

Again, Eshelby's results are used (for this, the size of a each inclusion
must be much smaller than that of the RVE). Within the context of the
M.T. model, we obtain for each inclusion (I):

(20.22)

The stress ave rage over the RVE is given by:

<u > X < u l > +(1 - X) < u M >


= X < el : Al >: € + (1 - X) < e M : ('M >, (20.23)

where X is the volume fraction of inclusions. The strain average over the RVE
is:

< t: > = X < Al >: € + (1 - X) < t: M >, (20.24)


------

where, by definition, we have:

(20.25)

where VI is the volume of an inclusion (I) and operator A· was introduced


for convenience. We recall again that in the previous equation, operators Al
have to be properly rotated so that they are expres sed in the same basis (see
Appendix C). The same remark holds for other sums or integrals of tensors
20.3 Micro/macro constitutive model for semi-crystalline polymers 527

in other equations. Combining Eqs. (20.24-25) the strain average over the
matrix phase is found to be:

< €M >= - 11 (1 - xA *) :€ (20.26)


-x
We also have the folIowing result:

< el : A l >= C* : A * , (20.27)

where, assuming that A * is invertible,

(20.28)

We now assume that the matrix is isotropic. Substitut ing Eqs. (20.26-27) into
(20.23), we obtain:

<U>=C:€, with c=eM+x(C*-eM):A* (20.29)

The homogenized stiffness c has a much simpler expres sion if the inclusions
are spheres or ellipsoids with the same orientation (thus A * = A l ) and have
identical and isotropic stiffnesses (thus C* = el). If these conditions are met,
then:

20.3 Micro/macro constitutive model for


semi-crystalline polymers
Semi-crystalline polymers such as High Density Polyethylene (HDPE) are
now widely used as structural materials. The micro-structure of HDPE con-
sists of closely packed crystalline lamellae separated by layers of amorphous
polymer (Lee et al., 1993). The lamellae of melt solidified PE are most of-
ten radialIy arranged in spherulitic structures, thus making the material
macroscopicalIy isotropic at small strains. The macro. stress-strain behav-
ior is strongly non-linear and depends on such characteristics as the over-
alI crystallinity, the molecular weight, the molecular branch content, etc. A
micro-mechanicalIy-based constitutive model for the small deformations of
HDPE was recently proposed by Nikolov and Doghri (2000). The model is
valid in the temperature range where the amorphous phase state is above its
glass transition temperature and the crystallinity level is high enough so that
the notion of amorphous layer makes sense. It can be used for other semi-
crystalline polymers as long as these conditions are met. The basic ideas of
the model are described hereafter and detailed in the upcoming subsections.
528 20. Micro-mechanics of materials

- • The geometric model ing is depicted in Fig. 20.1. Each macro. material
point is supposed to be the center of a RVE, which is an "aggregate" of
randomly oriented composite "inclusions". Each inclus ion consists of a
stack of parallellamellae with their adjacent amorphous layers.
- • For the crystalline phase, permanent deformat ion occurs by slip over slip
planes along slip directions. A visco-plastic model is proposed (Sec. 20.3.1).
- • For the amorphous phase, a non-linear viscoelastic model is proposed
(Sec. 20.3.2). The main concepts underlying the model are:
- (a) The amorphous phase is assumed to possess a polydomain structure
(after Marrucci(1984) and Wissbrun(1985)).
- (b) The dis tort ion of each micro-domain is modeled with equations from
the continuum theory of liquid crystal nematics (De Gennes and Prost,
1993).
- (c) In addition to shrinking, micro-domains slide past each other. A shear
stress versus shear strain model is proposed.
- (d) Stretching of tie molecules (passing from one lamella to another) and
entangled chains (anchored on lamellae surfaces) induces elastic stresses
in the amorphous phase. A rubber-like elastic model is proposed.
- • The dependence of the macro. yield stress on crystallinity level is ex-
plained through a membrane model for the intermediate phase (linking
the crystalline and amorphous phases), see Sec. 20.3.3.
- • With the above-mentioned models at hand, a constitutive model for a
single inclus ion considered as a laminated composite is obtained in Sec.
20.3.4.
- • Finally, a macro. stress-strain model for the whole "aggregate" (i.e.,
RVE) is built via a homogenization procedure (Sec. 20.3.5). Numerical
simulations show that the model is able to correctly represent the macro.
stress-strain behavior and its dependence on strain rate and crystallinity
ratio (Sec. 20.3.6).

20.3.1 Crystalline phase

semi-crystalline polymer!crystalline phase


The PE crystals are formed by association and folding of long polymer
molecules. They possess an orthorombic symmetry with lattice parameters
a = 7.4Ă, b = 4.93Ă and c = 2.54Ă. We designate by the unit vector c
the molecular chain direction. The main deformation mode in crystals is
crystallographic slip (Lin and Argon, 1994), which takes place eithcr in the
(±c) directions or in transverse directions; see Fig. 20.4. Each slip system
(o:) is determined by the normal n(a) to the slip plane and the slip direction
7n(a) in that plane, see Fig. 20.5. The different slip systems of a crystal are
depicted in Fig. 20.6, see also Table 20.1. Because chains are inextensible in
the (±c) directions, it can be shown that therc are only four independent slip
systems.
20.3 Micro/macro constitutive model for semi-crystalline polymers 529

Chain slip Transverse slip


Fig. 20.4. Deformation mechanisms in crystalline lamellae. Left: chain slip, right:
transverse slip.

,...--
~~(a J
-
I
I I
E
Fig. 20.5. Shear deformat ion of a single crystal

~ view A
/1 1
• /
c ,,----i---/b
a

Fig. 20.6. Slip systems in crystalline phase. Left: orthorombic crystal, right: view
"A" of the slip planes and their normal directions

The deformat ion in polymer crystals can be described by the constitutive


relations used for small-molecule crystals. For simplicity we assume that the
crystals are rigid-viscoplastic. In a single crystal the resolved shear stress T(a)
act ing on a particular slip system (a) is given by the relation:

(20.30)

where (Te is the stress in the crystal and R(a) the Schmid tensor associated
with slip system (a):

R(a) = ~ (n(a) l8i mia) + mia) l8i n(a)) (no sum) (20.31)
530 20. Micro-mechanics of materials

For rate-dependent materials, the resolved shear stress 7(0) can be related
to the corresponding shear rate 1'(0) via a power law expression as in (Asaro
and Needleman, 1985):

(o) I /
= 1'0 sign (7(0») I:(0)
I m
1'(0) (20.32)

where 1'0 is a reference strain rate, g(o) the shear strength of slip system (a)
and m the strain rate sensitivity.
Equation (20.32) suggests that plastic flow is always present on slip system
(a) as long as the shear stress 7(0) is not identically equal to zero, but if
17(0) I < g(o) the viscoplastic shear rate 1'(0) is negligible.
The total strain rate in the crystal is:

Ee = L 1'(0) R(a). (20.33)


o

The slip systems and corresponding resistances for HDPE crystals at room
temperature are given in Table 20.1.

Slip system glO<} [MPa)


(100)[001) 8
(010)[001) 20
{UO} [001) 20
(100)[010) 13.3
(010)[100) 20
{11O}(IIO} 17.6

Table 20.1. Slip systems and corresponding resistances for HDPE (Lee et al., 1993)

The slip resistances of polymer crystals depend on temperature, lamellae


thickness and the normal stress acting on the slip planes (Lin and Argon,
1994; Young, 1988; Crist et al., 1989). As in (Lee et al., 1993) the strain-
induced hardening of slip systems' strength is neglected because the thickness
of lamellae (5 - 25 nm) suggests that dislocations must be quickly evacuated
on lamellae surfaces.
The temperature dependence is easily understood as the crystallographic
slip is a thermally activated process. As for the normal stress dependence, it
has been established that the slip resistances are proportional to the applied
normal stress, although the exact mechanism of this behavior is not clear
(Lin and Argon, 1994).
There have been attempts to explain the yield behavior of PE as de-
termined by the nucleation and propagation of screw dislocations along the
chain axis c in the lamellae (Shadrake and Guiu, 1976; Young, 1988; Crist
et al., 1989). In dislocation theory the average lamellar thickness is directly
20.3 Micro/macro constitutive model for semi-crystalline polymers 531

related to the weakest resistance 9(0.) (thicker lamellae would have higher re-
sistances 9(0.»). The theory gives reasonable results at room temperature but
deviates significantly from experiments at low temperatures and above the
Q-relaxation temperature (about 60°C). Our latest work supports the idea
first suggested in (Crist et al., 1989) that the length of the Burgers vector
along c axis decreases at higher temperatures.
In Sec. 20.3.3, we consider the intermediate phase linking the lamellae
and the amorphous layers and its role in the yield behavior of PE.

20.3.2 Amorphous phase

Two deformat ion mechanisms of amorphous layers have been identified: in-
terlamellar shear and inter-Iamellar separat ion (Fig. 20.7). Experiments show
that inter-Iamellar shear is the dominant deformation mode at small strains
of PE (Lin and Argon, 1994). PE is nearly an incompressible material, with
Poisson's ratio v = 0.41 (Dahoun et al., 1995). Therefore, in order to have in-
terlamellar separation in a given inclusion, there must be a flow of amorphous
material into the inter-Iamellar space, which would lead to micro-cavities. At
small strains, no cavitation has been registered for PE and therefore this de-
formation mode should have negligible contribution to the overall strain. Here
inter-Iamellar shear is taken into account with the choice of a basic structural
unit consisting of a stack of parallel lamellae and their adjacent amorphous
layers (Fig. 20.1). In that figure, n denotes the unit vector normal to the
lamellae surfaces and c the unit vector of the direction of PE molecules in
the crystals. Experimental observations show that the chain direction c forms
an angle with the normal vector n ranging between 17° and 40° (Lee et al.,
1993). In our model (n, c) = 30°. For convenience when dealing with the
overall behavior, we call a stack of lamellae and their adjacent amorphous
layers inclusion. The crystalline phase content Xc in each inclus ion is taken
to be equal to the overall crystallinity of the polymer.
The amorphous layers' micro-structure is not fully understood at present,
mainly because of the experimental difficulties stemming from the fact that it
cannot be isolated and studied separately from the bulk material. A complex
micro-structure with ordered micro-domains has been proposed in (Bartczak
et al., 1996) in order to explain the large viscoplastic deformation of the
amorphous phase of HDPE. Here we use a similar idea (though in a different
context) in the case of small deformat ion behavioL
In the frame of our model the relation between the shear viscosity and
shear rate in each inclusion can be extracted (at room temperature the amor-
phous phase of PE is in a molten state, that is why we use the shear viscosity
as a key parameter). In Sec. 20.3.6 we show that it is of a power law type
even at very small strain rates. This behavior is fundamentally different for
Polyethylene melt where a Newtonian plateau is reached at small strain rates
(Larson, 1988).
532 20. Micro-mechanics of materials

III·.· ...............-.... _ .... ........ :".~ _.~_

Fig. 20.1. Deformation of amorphous phase: (a) non-deformed state, (b) inter-
lamellar shear, (c) inter-Iamellar separationj after (Dahoun, 1992).
20.3 Micro/macro constitutive model for semi-crystalline polymers 533

On the other hand, the polymer chains in the amorphous phase are
subjected to severe constraints from the lamellae and are situated in ex-
tremely thin layers with thickness of order 10 nm. We suggest that under
these conditions the viscoelastic stress in the amorphous layers is due to
the resistance of polymer molecules to dis tort ion during deformation. Micro-
mechanical interpretations of viscoelasticity of homogeneous polymer melts
suppose that stress is caused either by simultaneous stretching and release of
polymer strands or reptation of polymer molecules in a" tube" of surrounding
molecules (Larson, 1988). These processes are not likely to operate in shearing
of thin layers where most molecules are anchored in lamellae, chain ends are
situated mostly on the lamellae surfaces (Lin and Argon, 1994) and entangle-
ments are quite stable defects. We introduce a new parameter which reflects
a different mechanism of viscoelasticity, Le. the distortion elastic constant of
molecules.

lamella

amo~housr-~~~~~~~~~~, Ro
layer

lamella
Fig. 20.8. Amorphous layers' micro-structure

Our basic assumption is that each amorphous layer can be modeled as


a polydomain structure (Fig. 20.8). Following Marrucci (1984) and Wiss-
brun (1985), it is assumed that each micro-domain is in a local free energy
minimum. The chain segments in a given micro-domain do not have a net
orientation prior to loading. According to the continuum theory of the ne-
matics class of liquid crystals, the distortion free energy Jd of the chains per
unit volume in a micro-domain with radius R can be approximated by the
following expression (De Gennes and Prost, 1993)

(20.34)

where K is the average elastic constant of distortion of a single micro-domain.


It is related to distortion elasticity of polymer molecules and depends on their
molecular structure and temperature.
When a shear rate 'Ya is applied to the amorphous layer, the domains
will change their radius from Ro in the unloaded state to R in the steady
state (Ro > R) and the distortion energy in the layer will increase. From
conservation of volume, the number of micro-domains grows as their radius
534 20. Micro-mechanics of materials

diminishes. According to Marrucci (1984), the in crease of the distortion en-


ergy is equal to the steady state shear stress in the layer (Wissbrun, 1985):
K K .
R2 - R2 = TJara (20.35)
o
where TJa is the layer's viscosity.
From the above considerations, the domain radius in steady state shear
R should depend on the applied shear rate i'a. In order to derive a reiat ion
between Ro, R and i'a we introduce the probability for a micro-domain,
subjected to distant shear rate i'a of having radius R as P = R/ Ro and require
that in equilibrium state PI'Ya=o = 1. Interaction between micro-domains is
neglected. Then, according to statistical mechanics, P can be expres sed as:

(20.36)

where W is the energy (or work) necessary to shrink the micro-domain to


radius R, kB the Boltzmann constant and T the absolute temperature. As-
suming that W = ai'::,
Eq. (20.36) can be rewritten as:

a i'n)
R = Ro exp ( - kB; (20.37)

where n is the "shear rate sensitivity" of the micro-domains and a a propor-


tionality constant.
Introducing equation (20.37) in (20.35) we obtain the shear viscosity of
the amorphous phase as:

K [ (2akBTi'::) - 1]
TJa = (RoF i'a exp (20.38)

It is found numerically that the above equation is equivalent to a shear


thinning power-Iaw function (TJa = Ci'::') up to very high strain rates, where
a plateau of TJa (i'a) is reached.
In addition to shrinking, the micro-domains slide past each other. Wiss-
brun (1985) has derived the compliance modulus arising from mutual slip
of two adjacent micro-domains assuming linear relation between shear stress
and strain. Similarly, we find the elastic modulus Ca resulting from micro-
domains' slip:

(20.39)

where L is the persistent length of the polymer molecule (the maximum


chain length which can be considered as part of a straight line). Note that
despite the assumption of a linear reiat ion between shear stress and strain,
Ca depends on i'a through R.
20.3 Micro/macro constitutive model for semi-crystalline polymers 535

In our viscoelastic modeling it is assumed that upon shear, the micro-


domains' radius drops instantaneously from Ro to R. The low energy required
to shrink micro-domains (see Sec. 20.3.6) suggests that this is a reasonable
approximation. On the contrary, when deformation is suddenly frozen in a
relaxation test, R will in crease to its equilibrium value Ro through thermal
diffusion, but this process would take more time than shrinking upon external
forces.
It is well known that two neighboring lamellae are linked up through nu-
merous tie molecules passing from one lamella to another as well as through
entangled chains anchored on lamellae surfaces (Lin and Argon, 1994). Upon
deformation the rubber-like stretching of the polymer molecules will produce
elastic stress in the amorphous phase. At small strains, the entropic elastic-
ity can be modeled by neo-Hookean constitutive equations. Because of the
incompressibility, the only material parameter needed is the rubbery shear
modulus Ea, which at microlevel reads:
(20.40)
with Va being the number of active polymer strands per unit volume.
In order to derive the constitutive equations, we shall make some addi-
tional hypotheses backed by experimental results.
A viscoelastic constitutive model for the amorphous phase is developed
based on the well-known three-element rheological model of Fig. 20.9 (it is
shown in (Kitagava and Zhou, 1995) that viscoelasticity of semi-crystalline
polymers is well described by over-stress theory, a slight modification of this
model). Making the usual assumptions, the following differential equation

TIa Ga

Ea
Fig. 20.9. Three-element rheological model for the amorphous phase behavior

relating the shear stress Ta and strain "fa acting on the amorphous phase is
found:

Ta 8
+ "Ia 8t (Ta)
Ca = Ea"fa + "Ia [."fa 8
+ 8t Ca "fa )]
(Ea (20.41)

where a superposed dot denotes a time derivative. If the coefficients Ea and


Ca were constants, then equation (20.41) reduces to the three-element linear
536 20. Micro-mechanics of materials

viscoelastic model:

(20.42)

However, the elastic modulus G a is not constant but depends on the strain
rate i'a according to equation (20.39); hence the differential equation in its
general form (20.41) should be used. Unless the material is loaded with (i'a =
constant) from stress- and strain-free state, equation (20.41) does not reduce
to the linear case.
The multi-axial generalization of equation (20.41) is stiH under investiga-
tion. One could develop a 3D model along the same lines as the 1D model
and find the following differential equation relating the total stress and strain
tensors (O' a and €a) acting on the amorphous phase:

O'a+ M : :t (C- I :O'a) =E:€a+ M : [ea + :t (C- I :E:€a)]


(20.43)

where E and C(I2 ) are 4th order stiffness tensors ofthe linear and non-linear
elastic moduli, respectively, and M(I2) is a 4th order tensor of the viscous
modulii 12 is the second invariant of the strain rate tensor ea.
Equation (20.43) by it self does not translate the fact that at small strains
the dominant mode of deformation of the amorphous layers is inter-Iamellar
shear. We can take this into account by another possible extension of the
constitutive model from 1D to 3D as follows:
(20.44)

where Ha is the Schmid tensor constructed from the pair of the lamellae
normal vector n and the unit vector m(a) along the projection of the stress
vector in the inclusion on the lamellae surface:
t(a)
m(a) = - - . t(a) = O' . n - [(O' . n) . n]n (20.45)
IIt(a)11 '

20.3.3 Intermediate phase

In this section, we consider the intermediate phase linking the lamellae and
the amorphous layers (Fig. 20.10) and its role in the yield behavior of PE.
The conformational entropy of a polymer system with homogeneous molec-
ular structure and spatial inhomogeneity of density and orientation of poly-
mer links has been treated in (Grosberg and Khokhlov, 1994). We make
use of the basic results of Lifshitz (1969) on the entropy losses because of
the spatial inhomogeneity in density or/and orientat ion which give rise to a
density (orientation) gradient. Obvious sources of such a gradient in semi-
crystalline polymers are the chain folding on lamellae surfaces where chains
20.3 Micro/macro constitutive model for semi-crystalline polymers 537

bend more often in one direction than in others and the density difference
between crystals and amorphous material. Thus in semi-erystalline polymers
each crystalline lamella should be wrapped with a surface layer (membrane)
of intermediate phase.

intermediate phase
layer

lamella
lint
Fig. 20.10. Intermediate phase in PE

The free energy of the intermediate layer F int for a given lamella is a
fu net ion of the density difference between two phases Lln, the intermediate
layer volume V and the thickness of the layer lint:

D. _ aTV Lln
rmt - 12 (20.46)
int

where a is a proportionality coefficient and T the temperature.


Assuming a membrane stress state in the intermediate layer we can eval-
uate its surfaee tension a as:
aTLln
a = F int IV = -1-2- (20.47)
int

For more details see (Grosberg and Khokhlov, 1994).


The intermediate phase volume fraction as measured from (Mandelkern
et al., 1990) changes with crystallinity and we as sume that it is directly re-
lated to the intermediate layer thickness. The results of those authors show
that an increase in overall erystallinity leads to decrease in the intermediate
volume fraction and, therefore, surface tension will increase due to thinner
intermediate layers. Here we suggest that the "failure" of the intermediate
phase can explain the yield behavior at the second yield point where local-
ized (" eoarse") slip and, eventually, lamellar-to-fibrillar transition take place
whereas the first yield point (which is more sensitive to crystallinity) can be
explained with the help of dislocat ion theory. The role of intermediate phase
in this case is to maintain homogeneous slip in the lamellae and prevent them
from loealization of deformation.
In eonclusion, the introduction of the intermediate phase can explain the
double yield phenomenon in HDPE reported by several authors (e.g. (Lucas
et al., 1995). At the first yield point the intermediate phase prevents the
lamellae from localization of deformation. The second yield point eorresponds
to the failure of the intermediate phase membrane.
538 20. Micro-mechanics of materials

20.3.4 Single inclus ion

We call a stack of parallel lamellae and their adjacent amorphous layers


inclusion (see Fig. 20.1). The crystalline phase content Xc in each inclus ion
is taken to be equal to the overall crystallinity of the polymer.
It is noted that an inclus ion can accommodate an arbitrary deformation
whereas a single polymer crystal cannot because polymer crystals are in-
extensible in a direction parallel to polymer chains. Thus our inclusion has
five independent slip systems which enable it to accommodate an arbitrary
loading.
The constitutive equations of the inclus ion can be derived with the help
of the assumption of uniform stress in it. This approximation is suitable
for laminated composites subjected to off-plane shear. Let Xc be the volume
fraction of the crystalline phase and Xa the volume fraction of the amorphous
layers. Then the total shear rate in the inclusion is:

(20.48)

with Ee given by Eqs. (20.33) and Ea by (20.44).

20.3.5 Overall behavior

In small deformations, we model each macro. point of HDPE as an aggregate


of randomly oriented inclusions (Fig. 20.1) whose behavior was defined in the
previous subsections. The spherulitic morphology is neglected because inves-
tigations have shown no influence of different morphologies on the mechanical
properties (Kennedy et al., 1994). Once the constitutive behavior of a single
inclus ion is specified (Sec. 20.3.4) we have to find the overall behavior of an
aggregate. This can be done via a homogenization procedure and the results
obtained will be an approximation of the macro. behavior of the materialj
see Sec. 20.2.
In an early version of our model, we considered a fully viscoplastic model
for each inclusion and used a self-consistent scheme to find the overall behav-
iorj see Appendix in (Nikolov and Doghri, 2000). The results unveil that the
Voigt model is not appropriate for semi-crystalline polymers and the reason
for this is that the interaction between lamellae stacks is much weaker than
that between grains in metals. This conclusion is supported by the exper-
imental evidence of freely rotating lamellae stacks during deformation (Lin
and Argon, 1994) as well as by other results obtained by other authors (Lee
et al., 1993). Thus, the overall behavior of HDPE should be closer to the
Reuss lower bound.

20.3.6 Numerical simulations

The micro/macro constitutive model of the previous subsections has been


numerically implementedj see (Zealouk al., 1999). Firstly, we explain how
20.3 Micro/macro constitutive model for semi-crystalline polymers 539

the different material parameters have been identified. Numerical results will
be presented later.
Uniaxial tensile experiments were performed on a INSTRON 4204 ma-
chine. The material is HDPE for gas pipes; the crystallinity Xc measured by
DEC has a value of Xc = 0.67. The standard tensile specimens have width
of 9.6 mm and thickness 3.95 mm in their thinner part. The initial gauge
length is 50.5 mm, and the temperature is 23°C.
We first consider the strain rate dependence of the overall stress before
yielding (at strain € = 0.08) and its micro-mechanical interpretation. Five
different macro-strain rates have been imposed on the specimens in tensile
tests: ~ = 6 x 10- 6 ; 3 x 10- 5 ; 3 x 10- 4 ; 3 x 10- 3 ; 3 x 10- 2 [8- 1]. According
to our model, the overall stress U splits into two parts:

(20.49)

where uv(l, T) is the rate-dependent, or viscous stress and ue(e, T) the elastic
stress. They are both temperature-dependent but the influence of tempera-
ture has not been studied.
Prior to yield, the rate-dependent contribution to the overall stress comes
from the inter-Iamellar shear in the randomly oriented inclusions. The re-
versible part ue(e) of the overall stress is obtained via relaxat ion experiment
as the stress after 24 h ofrelaxation at imposed strain of 8 % (efe ~ 8 M Pa).
We subtract ue(e) in order to obtain uv(l) in Eq. (20.49). We assume that
the stress is uniform in alI inclusions and consider the case of most unfavor-
able loading of the amorphous layers in a given inclusion, i.e. when the layers
are oriented at 45° to the tensile axis. The rate-dependent part of the shear
stress in the amorphous phase is then T~v) = O'v/2.
In order to identify the parameter values in a first approximation, we
assume that the strain rate is also uniform throughout the material and,
prior to yield, the plastic strain rate is negligible. Then, for inclusions with
amorphous layers inclined at 45° w.r.t. the tensile axis the average value of
the shear rate is 'tI = 3~/ 4. The shear rate in the amorphous phase will be
'ta ~ 'tl/Xa (an important remark is that our model works not only in the
reversible small-strain region but also beyond the yield limit of the material).
We take Xa ~ 1 - Xc = 0.33.
At steady state shear the viscosity is equal to:
(v)
Ta
'TJa= -.- (20.50)
"'ta
We are then able to convert the experimental macro stress-strain data in
terms of viscosity vs. shear rate in the amorphous phase of a single inclusion
(Fig. 20.11). It appears that the observed shear-thinning is of a power-Iaw
type and there are no signs of Newtonian plateau for viscosity even for very
small strain rates.
540 20. Micro-mechanics of materials

1~~----~~----~~----~~--~~
10~ 4 10 10~ 10~ 1~
shear rate [1/s1
Fig. 20.11. Amorphous phase shear viscosity TIa vs. shear rate i'a

Some micro-mechanical parameters in our model cannot be directly mea-


sured despite their physical meaning. They can be fixed in a narrow range con-
sidering the restrictions imposed by experimental evidence and the physics.
The average thickness of a single amorphous layer for HDPE is about 10 nm
(Kennedy et al., 1994). The hypothesis of a micro-domain structure requires
at least 2 micro-domains over this thickness. We take the initial micro domain
radius to be Ro ~ 2 nm.
On the other hand, the persistent length L for PE molecule is of the order
of 1 nm (Grosberg and Khokhlov, 1994). The micro-domain radius R cannot
be smaller than that value at any strain rate because L is the smallest possible
radius of curvature (Wissbrun, 1985). At high strain rates R is controlled by
the rate parameters a and n in equation (20.37). In addition, K has to be
estimated before setting the values of a and n. For flexible molecules K
should be of order 10- 11 -;- 10- 12 N (Wissbrun, 1985) and n must be small
in order to meet the requirement of minimum micro-domain radius.
With these indications we can proceed to a more precise parameter iden-
tification. We make use of shear thinning experimental results and numerical
simulations with imposed simple shear on a single inclusion. The direction of
shear is perpendicular to the lamellae normal n (depicted in Fig. 20.1). This
particular choice of the shear direction enables us to eliminate the plastic de-
formation prior to an imposed shear value of 0.1 and thus to obtain indicative
results for the macroscopic behavior. The slip resistances of crystalline phase
are taken from Table 20.1. After tuning, the following set of parameters is
obtained:
Ro = 1.8 nm; n = 0.1; K = 1.6 X 10- 11 N; a = 2.9 x 10- 21 Js n ;
T = 296 K; kB = 1.38 X 10- 23 JK- 1 , L = 0.5 nm, Ea = 2 MPa
With these values, the shear thinning of amorphous phase is compared
to the experimental data in Fig. 20.11. The evolution of micro-domain's ra-
dius with strain rate is shown in Fig. 20.12. The stress-strain curves for a
20.3 Micro/macro constitutive model for semi-crystalline polymers 541

2r-----~----~----~----~-----

1.5

E I~--______________~
E..1
CI:

0.5

OL---~----~----~----~--~
0.02 0.04 0.06 0.08 0.1
sheRr rRte r1/s1
Fig. 20.12. Micro-domain radius R vs. shear rate ia

single inclusion subjected to simple shear (with Xc = 0.67) at three different


strain rates are plotted in Fig. 20.13 (tensor components are W.r.t. a local
orthonormal basis (e1,e2,e3) such that e3 = n). The initial shear modulus
does not change when different shear rates are applied. This is due to the
elastic stress in the amorphous layers. However, the yield stress diminishes
with strain rate as observed in macroscopic experiments.

b 3

o L -_ _- L_ _ _ _ ~ __ ~ _ _ _ _- L_ _ ~

O 0.03 0.06 0.09 0.12 0.15


1'/(13)

Fig. 20.13. Strain-rate dependence of inclusion's shear stress

We also performed simulations changing the overall crystallinity Xc in the


inclusion (Fig. 20.14). The imposed shear rate is (/(13) = 0.001 s-1. The
weakest slip system resistances for different crystallinities (Xc = 0.67, 0.55,
0.43) are taken to be 8, 6.5, 5 M Pa, respectively. The other slip resistances
have been proportionally related to the weakest ones for each crystallinity
542 20. Micro-mechanics of materials

Xc =0.67--
Xc = 0.55
b Xc = 0.43 - -
2

o L -_ _~_ _ _ _~_ _ _ _~_ _~~_ _~

O 0.03 0.06 0.09 0.12 0.15


1'/(13)

Fig. 20.14. Inclusion's shear stress as a function of crystallinity Xc

level. From Fig. 20.14 it is seen that both initial shear modulus and shear yield
stress decrease with decreasing crystallinity. The results obtained are very
similar to those obtained in macroscopic tensile experiments. The relationship
g(O:) /Xc = constant is imposed a priori but what is more important is that
the link between the initial shear modulus and the yield stress is obtained as
a direct result of our constitutive modeling keeping all the parameters of the
amorphous phase model unchanged.
The model allows us to compute the energy vVa necessary to shrink
a micro-domain to radius R. With the above parameter set we obtain
W a ~ 2 X 10- 21 J at strain rate "'Ia = 0.005 S-l. At a first glance it seems to
be excessively low. However, the energy necessary to move a dislocation in a
PE crystal in order to set up plastic fiow is about 1.6 x 10- 19 J (Lin and Ar-
gon, 1994). Therefore, the proposed mechanism of micro-domains' shrinking
should be active well before the propagat ion of dislocations in crystals.
Next, we consider the elastic contribution to the overall stress {Ie. The
rubbery shear modulus obtained from numeric al simulations is Ea = 2 M Pa
which gives the number of active strands per unit volume as: lIa = 0.5 X
10 27 m- 3 . A more representative estimation is to find the number of crystal's
unit cells which can occupy the volume necessary for one strand to be formed.
Simple computation gives that one strand occupies a volume where 20 unit
cells can be situated, which is a reasonable value (a small number of cells
implies that the strands' conformat ion is near crystalline state and cannot
be stretched much further, which is in contradiction with the experimental
evidence).
Uniaxial tension tests at the macro scale were simulated at constant
macro. stress rate. Note that although the macro. stress is uniaxial, the stress
and strain fields inside each RVE are multiaxial. Each RVE is an aggregate
containing 250 randomly oriented inclusions. Reuss homogenization scheme
20.4 Further reading 543

was used. Macro. stress versus macro. strain curves are plotted in Figs. 20.15
and 20.16. The former figure shows the inftuence of the lading rate, while the
latter shows the effect of crystallinity ratio. Figure 20.16 is a typical example
of the benefits one gets by using a micro/macro model: a micro-structural
information such as the crystallinity ratio is not contained in a phenomeno-
logical macro. model and therefore cannot be predicted by it.

40
xXxx
xxx
35 xX
XX
<il XX
o.. 30
••• •••••
xX

•• ••
X
~ 25 x
X
••
+++++++
••
rJ.l
rJ.l x
...,...
Q) +++
20 x. ++
rJ.l x· +++
• ++
...ci 15 xe++

,
U x·+
.+ 0.1 MPa / s
c<l 10 x+
+
:::E
5
1 MPa / s •
10 MPa / s x
O
O 0.04 0.08 0.12 0.16
Macro. strain
Fig. 20.15. Macroscopic tension tests at constant stress rates. Infiuence of loading
rate.

20.4 Further reading


The model presented in Sec. 20.3 is restricted to small strains. For large
deformations, micro/macro models which simulate the texture development
and the stress-strain response of solid polymers do exist in the literature.
Examples are (Wu and Van der Giessen, 1992, 1994) for amorphous polymers,
and (Dahoun et al., 1991) and (Lee et al., 1993) for semi-crystalline polymers.
For the latter class of materials however, important issues such as lamellar-to-
fibrillar transition and fibrils' properties are not addressed. Correct modeling
of these phenomena will bring the simulat ion of texture development much
closer to physical reality and is also a crucial step towards understanding and
predicting crack initiation and growth in a damaged material.
The body of literature on micro/macro modeling of solid polymers is
small compared to that for other materials such as metals. Perhaps, one
reason is the broad range of polymeric materials and the complexity of their
micro-structure. Indeed, "if the architecture of metal crystals is thought of as
classical, then that of polymers is baroque" (Ashby and Jones, 1986, Chap.
22).
544 20. Micro-mechanics of materials

35
30
7
o... 25
~

'"'"<ll 20
..-....
'" 15
....ci
u
ro 10 Xc = 0.43 +
~ Xc = 0.55 •
5 Xc = 0.67 x

O
O 0.04 0.08 0.12
Macro. strain
Fig. 20.16. Macroscopic tension test at constant stress rate = 1 M Pal s. Influence
of crystallinity ratio.

The prediction of macro. properties from the micro-structure has been


enjoying a great deal of interest since the pioneering work of G.1. Taylor in
the 1930's and many great names in Mechanics have made significant con-
tributions to this fascinating subject. For references, the reader can consult
the following books: (Mura, 1987), (Havner, 1992), (Nemat-Nasser and Hori,
1993), (Suquet, 1997), (Teodosiu, 1997) and (Kocks et al., 1998). See also
(Lielens, 1999) and (Van Houtte, 1998).
A. Cylindrical coordinates

Many problems are such that it is advantageous to use cylindrical (r, O, z)


instead of Cartesian (x, y, z) coordinates. Cylindrical basis vectors (er, eo, e z )
are expressed in the Cartesian basis (e x , e y, e z ) as follows:

e r = cos(O)e x + sin(O)e y , eo = - sin(O)e x + cos(O)e y (A.l)

See Fig. 8.1 for an illustration. Both cylindrical and Cartesian bases are
orthonormal. The position vector of a point M(r,O,z) W.r.t. the frame
(O,ex,ey,e z ) is:
~

x -=OM= re r + ze z (A.2)

Differentiation of (A.l) gives:

de r = (dO)eo,de e = -(dO)e r (A.3)

Let v be a vector field defined in the cylindrical basis as follows:

v(r, O, z) = F(r, O, z)e r + G(r, O, z)eo + H(r, O, z)e z (A.4)

The gradient (V' v) of v is defined by:

dv = (V'v) . (dx) (A.5)

Differentiating (A.4) and using (A.3), we obtain:

dv = ( âF
âr dr
âF âF)
+ âO dO + âz dz er + F(dO)eo
âG âG âG)
+( âr dr + âO dO + âz dz eo - G(dO)e r
âH âH âH)
+ ( a:;:dr + ao dO + az dz ez (A.6)

Differentiation of (A.2) gives:

dx = (dr)e r + (rdO)ee + (dz)e z (A.7)


546 A. Cylindrical coordinates

Using (A.6) and (A.7), Eq. (A.5) can be written in the following matrix form:

[
8r dr
8F + !(8F
8C dr +
r 88 - G)rd() 8z dz
8F
!(8C + F)rd() + 8C dz
+ 1
8r r 88 8z (A.8)
8H dr + ! 8H (rd(}) 8H dz +
8r r 88 8z , .
dX

where the 3 x 3 matrix (V'v) is given in the cylindrical basis as follows:

V'v = [
8F
8r
~~
!(8F
r 88
!(8C
r 88
-G)
+F)
8z
8F
8C
8z
1 (A.9)
8H 18H 8H
ar r7fi[ {fi

The divergence of v is given by:


. 8F l 8G 8H
dlV v == tr(V'v) = 8r +;:-( 8(} + F) + 7);' (A.lO)

using (A.9). Now let g(r,(},z) be a scalar field. The gradient (V'g) of 9 is
defined by:

dg = (V'g) . (dx), (A.l1)

where the 3 x l array (V' g) is given in the cylindrical basis as follows:

V'g = [ ~~8z
l, (A.12)

using (A.7). The gradient (V'V'g) of (V'g) is found from (A.9) and (A.12) as
follows:
!t..s.
8r 2
! ( h _ !!!s.)
r 8r8 r 88
V'V'g = [ .!l..(!!!s.)
8r 5 88
!(!~
r r 88
+ !!s.)
8r
(A.13)
b.. !h
8r8z r 888z

The Laplacian (Llg) of 9 is defined by:

_ 8 2g l 182g 8g 8 2g
Llg = tr (V'V'g) = 8r 2 + ;:-(;:- 8(}2 + 8r) + 8z 2 (A.14)

This can be rewritten as follows:


l 8 8g l 8 2g 8 2g
Llg = ;:- 8r (r 8r) + r2 8(}2 + 8z2 (A.15)

Consequently, the following result follows:


A. Cylindrical coordinates 547

Consider a second-order symmetric tensor a (e.g., stress (1" or strain 1:) and a
vector u. In Cartesian coordinates, the following result is easily established:

(A.17)

This can be written in the following intrinsic form which is valid in cylindrical
coordinates for instance

div(a . u) = (diva) . u + a: (\lu), (A.18)

where div designates the divergence operator. We now apply the result to the
cylindrical basis by taking u to be equal to e r , e() and ez, successively,

(diva) . e r div(a· e r ) - a: (\le r )

ar + ~ 7iiJ + arr ) + fu
oarr l (oaer oazr l
- ~aee,
(diva) . ee div(a· e()) - a: (\lee)

ar
oa r() l oa()()
+ ~(7iiJ + are)
oaze l
+ fu + ~ar(),
(diva)· e z div(a· e z ) - a: (\le z )
oarz l (oae oa zz
ar + ~ 7iiJ +a rz
z )
+ fu' (A.19)

For the projection along e r ) we used (A.9) with v = er , Le. F = l and


G = H = O, and Eq. (A.lO) with

Le. F = arn G = aer and H = azr. The projections along e() and e z are
obtained in a similar fashion. As an application of results (A.19), equilibrium
equations (div (1" + f = O) are obtained in cylindrical coordinates by setting
a = (1", i.e.

oarr ~ oar()oazr arr - ae() f - O·


ar + r of) + oz + r + r- ,
oare
- loa()()
- + ---
ar r of)
oaze
+- 2
- + -are
oz r
+ 1,() = O
oarz loa()z oa zz arz f O
--+---+--+-+
OZ r of) OZ r
z= (A.20)

The infinitesimal strain tensor is defined by: 1: = (\lu + ,;r u)/2, where u
is the displacement. Using (A.9), the components in the cylindrical basis are
given as follows:
548 A. Cylindrical coordinates

âUr
Err ,
âr
1 (âUO
Eee = ;: â() + Ur ) ,

EOr ~ (~âUr _ Ue + âUo) = ErO,


2râC r âr
âuz
Ezz ,
âz
Erz ~ (âur âU z )_
2 âz + âr - Ezr

Eez ~ (âue + ~ âU z ) = Eze (A.21)


2 âz r â(}
We can go from Cartesian to cylindrical coordinates via the following 3 x 3
tmnsformation matrix:

cos(}
[P] = [ - sin() cos() O
sin(} O 1 (A.22)
O O 1

The rows correspond to the components of e r , eo and e z in the Cartesian


hasis, respectively. Matrix [P] has the following properties:
[p][p]T = [pf[p] = [8], det[P] = 1,
where [8] is the 3 x 3 identity matrix. A vector v transforms from one or-
thonormal hasis (eI, e2, e3) to another (ei, e2' e3) according to:
{v*} = [P]{v}, {v} = [Pf{v*} (A.23)
An application to cylindrical versus Cartesian coordinates gives:
vr = (cos ())v", + (sin ())vy, Vo = - (sin ())v", + (cos ())vy (A.24)
A second-order tensor a transforms from one orthonormal hasis (ei) to an-
other (ei) according to:
[a*]" = [P][a][Pf, [a] = [p]T[a*][P] (A.25)
If a is symmetric (e.g., stress U or strain f.), then applying results (A.25) to
cylindrical versus Cartesian hases, we ohtain:
arr a",,,, cos 2 () + ayy sin 2 () + 2a",y sin(fJ) cos(())
aee = a",,,, sin 2 () + ayy cos 2 () - 2a",y sin(()) cos(())
arO (a yy - a",,,,) sin () cos () + a",y (cos 2 () - sin2 9) = aOr
arz = a",z cos () + ayz sin () = a zr
aez = -a",z sin() + ayz cos() = aze (A.26)
A. Cylindrical coordinates 549

In order to obtain the expressions of the Cartesian components in terms of


the cylindrical ones, we need not to develop (A.25)b, it suffices to substitute
8 with (-8) in Eqs. (A.26). Note also that a rr +aoo + a zz = a xx +ayy + a zz ,
'-"'" '-"'"
because (tr a) is an invariant.
B. Cardan's formulae

Consider the following cubic equation where A, B and Care given parameters:

x3 + Ax 2 + Bx + C = O (B.1)

A closed-form solution exists and is named after Cardan (1501-1576). A typ-


ical example of (B.l) is the search for the eigenvalues (TJ1, TJ2, TJ3) of a second-
order symmetric tensor 7] (e.g., stress U or strain €). lndeed, as mentioned in
Sec. 1.4, those eigenvalues satisfy the following equation:

where I J (7]) , J = 1, 2, 3, are the principal invariants of 7],

Actually, it is better to first compute the eigenvalues r;J of the deviatoric part
of 7] defined by:
Il (7])
dev 7] = 7] - -3- 1,

because we have h(dev 7]) = O and thus A = O in (B.l). Once the r;/s are
found, the eigenvalues of 7] are computed as follows:

Another example where the solution of Eq. (B.1) is needed is the strain
localization condition in 2D (Sec. 19.11.1).
There are various ways of presenting Cardan's formulae. We shall give
hereafter an implementat ion which proved to be computationally robust but
has one problem which will be explained at the end of the appendix.
The following change of variable:
A
X=x+- (B.2)
3
transforms (B.l) into the following equation:
552 B. Cardan's formulae

X 3 +pX +q = O, (B.3)

where p and q are defined as follows:

(B.4)

We only consider the case when p < O (see comments at the end). We intro-
duce the following notation:

(B.5)

If it is found numerically that C</> > 1, then it is reset to 1. Likewise, if


C</> < -1, then it is reset to (-1). Next, we compute the following angle:

1
cP3 == :3 arccos CrjJ (B.6)

Finally, the solutions of the original equation (B.I) are the following:

A
-3 + 2P3 COScP3,
-~ - P3[COScP3 + (sin cP3)v'3J,
- 3A - .
P3[cos cP3 - (SIn cP3) v'3] (B.7)

The solution procedure given here is robust but there is one problem:
In theory, if (4 p 3 + 27 q 2 > O) then there is one real solution and two complex
ones. This case is not detected with our algorithm, which always returns three
real solutions. A workaround is to always check whether those solutions are
physically acceptable or not.
c. Matrices for the representation of second-
and fourth-order tensors

C.l Storage
Let a,b be second-order symmetric tensors (e.g., stress and strain tensors),
and e, D, E fourth-order tensors (e.g., Hooke's operator in linear elastic-
ity, tangent operators in elasto-plasticity, etc.) For computations, a or bare
stored in 6 x 1 arrays as follows:

{a} = [ an (C.l)

This is not the traditional way of storing stress and strain tensors. That
method distinguishes between stress and strain: shear components of strain
are multiplied by a factor of 2 while the shear components of stress are kept
as they are. 1 However, we shall show hereafter that definit ion (C.l) has some
nice properties which the traditional method does not possess.
The inner product of a and b is given by the scalar:

(C.2)

The traditional storage does not lead to such expressions if both tensors are
stresses or both of them are strains.
If e relates a and b in the following way (e.g., linear elasticity or incre-
mental elasto-plasticity, etc.)

(C.3)

then it must have the following symmetries (because a and bare symmetric),

(C.4)

and it is stored in a 6 x 6 matrix according to:

{t} [1011 1022 1033 21012 21023 210 31] T

{O"} [0"11 0"22 0"33 0"12 0"23 0"31]T


554 C. Matrices for the representation of second- and fourth-order tensors

al1 Cl111 C1l22 C1l33 Cll12 V2 C1l23 V2


a22 C2211 C2222 C2233 C2212 V2 C2223 V2
a33 C 3311 C 3322 C 3333 C3312 V2 C3323 V2
a12V2 C12l1 V2 C1222 V2 C1233 V2 2C1212 2C1223
a23V2 C 2311 V2 C2322 V2 C2333 V2 2C2312 2C2323
a31V2 C3111 V2 C3122 V2 C3I33 V2 2C3112 2C3123
C l131 V2 bl1
C2231 V2 b22
C 3331 V2 b33
(C.5)
2C1231 b12 V2
2C2331 b23 V2
2C3131 b31 V2

If C has the additional symmetries (e.g., Hooke's operator in linear elasticity):

(C.6)

then Eq. (C.5) shows that the 6 x 6 matrix [C] is symmetric.


The fourth-order tensor E = C : D, defined in component form by:

(C.7)

is stored as a 6 x 6 matrix [E] which is the product of the 6 x 6 matrices [C]


and [D], i.e.

[E] = [C][D] (C.8)

The re ader can check for example that:

[E]ll = E l111 , [Eh4 = E 1l12 V2, [E]55 = 2E2323 , etc.

The fourth-order identity tensor 1, defined in component form by:


1
I ijk1 = 2(c5 ik c5j1 + c5il c5jk ) (C.9)

is stored as the 6 x 6 matrix [1]:

1 O O O O O
O 1 O O O O
O O 1 O O O
(C.lO)
O O O O 1 O
O O O O 1 O
O O O O O 1

i.e., the 6 x 6 identity matrix (if one stores the components Iijkl as they are,
one finds the last 3 terms in the diagonal equal to 1/2 instead of 1).
C.I Storage 555

The isotropic elasticity tensor C defined by (A and ţi being Lame's coef-


ficients) :

C = 2fJ,l + Al 01, i.e. Cijkl = J1.(rS ik rSjl + rSi/rSjk ) + ArSijrSkl (C.lI)

is stored as the 6 x 6 matrix [C] computed by:

[C] = 2J1.[I] + A{I}{I}T, i.e. (C.12)

l O O O O O l
O l O O O O l
O O l O O O l
[C] 2J1. +A [ l l l O O O]
O O O l O O O
O O O O l O O
O O O O O l O
A + 2J1. A A O O O
A A+2J1. A O O O
A A A + 2J1. O O O
(C.13)
O O O 2J1. O O
O O O O 2J1. O
O O O O O 2J1.
This is consistent with storing the stress u and the strain € as the 6 x l
arrays defined in Eq. (C.I); one can also check that the more general 6 x 6
matrix defined in Eq. (C.5) reduces to that of Eq. (C.13) in the isotropic case
defined by Eq. (C.lI). (Note: with a traditional storage, the last three terms
in the diagonal are simply J1.).
If a fourth-order tensor C is the the tensor product of two symmetric
second-order tensors a and b,

(C.14)
then it is stored as a 6 x 6 matrix [C] (where {a} and {b} are 6 x l arrays
defined in (C.I)):

[C] = {a}{b}T, i.e. (C.15)

al1 bl1 allb n au b33 allb12 V2 allb23 V2 all b31 V2


a22 bll a22 bn a22 b33 a22 b12V2 a22 b23V2 a22 b31V2
a:n b11 a33 b22 a33 b33 a33 b12V2 a33 bz3V2 a33 b31 V2
[C]=
a1Z bll V2 a12 b22V2 a1Z b33 V2 2a12b12 2a12b23 2a12b31
aZ3 b11 V2 a23 b22V2 a23 b33V2 2a23b12 2a23b23 2a23b31
a31 bl1 V2 a31 b22 V2 a31 b33 V2 2a31 b12 2a31 b23 2a31 b31
(C.16)
556 C. Matrices for the representation of second- and fourth-order tensors

This is consistent with the matrix storage of fourth-order tensors as defined


in Eq. (C.5), and unlike the traditional storage, there is no problem and no
distinction to be made when both a and bare stress tensors or both are
strain tensors.

C.2 Change of coordinates


Consider two orthonormal bases (ei) and (ei), i = 1,2,3. The change of
coordinates from the first basis to the second is determined by a 3 x 3 matrix
[P] defined by:
(C.17)

Since:
Oij = e; . ej = (Pikek) . (Pjlel) = PikPjlOkl,
it appears that the matrix [P] verifies:

(C.18)

Let v be a vector represented by 3 x 1 arrays {v} in the first basis and {v*}
in the second. We have:

Therefore, the following transformation rules hold:

{v*} = [P]{v}, {v} = [p]T{v*} (C.19)

A second-order tensor acan be represented by 3 x 3 matrices [a] in the first


basis and [a*] in the second. We have:

a:j = (a· ej) . e; = PjkPil(a· ek) . el = PjkPilalk

Consequently, the following transformation rules hold:

[a*] = [p][a][p]T, [a] = [p]T[a*][P] (C.20)

We now assume that a is symmetric. In order to write the transformat ion


rules in component form, one can use Eq. (C.20), or use a procedure which
is explained hereafter (and which will be useful for the transformation of
fourth-order tensors).
In component form, Eq. (C.20)a gives:

a;j PikaklPjl
PilPjlall + Pi2Pj2a22 + Pi3Pj3a33
+ (Pi1 Pj2 + Pi2 Pj1 )a12 + (Pi2 Pj3 + Pi3 Pj2 )a23 + (Pi3 Pj1 + Pi1 Pj3 )a31
C.2 Change of coordinates 557

(Note that we do have aij = aii)' Equations (C.20) can then be written in
the following matrix forms:

{a*} = [q){a}, {a} = [qf {a*} (C.21)

where {a*} and {a} are 6 x 1 arrays defined as in (C.1),

{a*} = a~2 a 33 ai 2 V2 ahV2 a 31V2 (


T
{a} [ all a22 a33 a12V2 a23V2 a31 V2 ]

and [q] is a 6 x 6 matrix given by:

PllPll P12 P12 P13 P13 Pll PI2 V2


P21 P21 P22 P22 P23 P23 P21 P22 V2
P31 P31 P32 P32 P33 P33 P31 P32 V2
Pll P21 V2
[q]
P12 P22 V2 P13 P23 V2 PU P22 + P12 P21
P21 P31 V2 P22 P32 V2 P23 P33 V2 P21 P32 + P22 P31
Pl1 P31 V2 P12 P32 V2 P13 P33 V2 Pl1 P32 + P12 P31
P12 P13 V2 P11 P13 V2
P22 P23 V2 P21 P23 V2
P32 P33 V2 P31 P33 V2
(C.22)
P12 P23 + P13 P22 Pu P23 + P13 P21
P22 P33 + P23 P32 P21 P33 + P23 P31
P12 P33 + PI3 P32 PU P33 + P13 P31

Note that unlike the traditional storage, the expres sion of [q] is unchanged,
whether we transform stress or strain tensors. The reader can count the
number of operations (especially the multiplications) to see which one of
transformations (C.20) or (C.21) is computationally cheaper.
From Eqs. (C.21) it is easy to check that:

[q][q]T = [qf[q] = [1] (C.23)

In linear elasticity, the relation (T = C : f: is derived from a potential w


(strain energy density):

âw . 1
(T= - , wlth w = -f: : C : f: (C.24)
âf: 2
We wish to find a matrix transformation rule for the fourth-order tensor C.
We follow the method of Lekhnitskii (1981). In the basis (ei), we can write
Eq. (C.24)b as:

(C.25)
558 C. Matrices for the representation of second- and fourth-order tensors

where {E} is a 6 x 1 array defined as in Eq. (C.l), and [C] a 6 x 6 matrix


defined as in Eq. (C.5).
In the basis (ei), we can write Eq. (C.24)b as:

(C.26)

where {f*} is a 6 xl array defined as in Eq. (C.l), and [C*] a 6 x 6 matrix


defined as in (C.5).
Equating the expressions of w in Eqs. (C.25) and (C.26) and using the
transformation rule (C.2l)a, we obtain:

(C.27)

Since this relation musţ hold for any {f}, we deduce that:

[C] = [qf[C*][q] (C.28)

Using the transformation rule (C.2l)b and equating the expressions of w in


Eqs. (C.25) and (C.26), we obtain:

[C*] = [q][C][q]T (C.29)

Note that transformation rules (C.28) or (C.29) are also valid if C is not
Hooke's operator but has the same symmetries as that tensor. Finally, no
transformation is needed if C is isotropic since [C*) = [C).
D. Zero-stress constraints

In the various stress update algorithms that were presented in Chaps. 12,
13, 15, 16, 17 and 18 it was assumed that alt strain or deformation gradient
components are known. There are cases however where the assumption is not
valid. If we have a (local) plane stress state (as for plates or shells) the out-of-
plane component of strain is unknown. AIso, for a beam, only the axial strain
component is known. This appendix shows how to deal with those cases for
some important classes of material models.

D.I Small-strain J 2 elasto-plasticity

We keep the notations of Sec. 12.10.2. We show how the return mapping
algorithm is modified in plane stress when the out-of-plane strain component
E33 is unknown. A trial stress is defined as follows:

(D.l)

where all components of €tr are equal to those of € except 4'3 which is com-
puted such that (Ţ~3 = o. Using a well-known technique (see Chap. 7), the
following value is found:

(D.2)

The stress-elastic strain relation reads:

(D.3)

The reader can check that this can be rewritten as in (Doghri, 1995):

(D.4)

where:
560 D. Zero-stress constraints

8ij = It~3 = c5i3 c5j3 - ~c5ij,


1
O -3
1
O -3
2
1 3
"6 = dev = (D.5)
O O
O O
O O
using the 6 x 1 array notat ion (Appendix C). The volumetric and deviatoric
parts of the stress-strain relations are:
tr (7 tr (7tr + 3K(t33 - t~~);
S = str - 2GLlep + 2G(t33 - t~~)"6 (D.6)
Combining the first equation with the incremental plastic flow relation
3 s
Lle P = --Llp,
20"eq
(D.7)

we find:

(D.8)

Taking the inner product of each side of the equation and using the yield
condition:
O"eq = O"y + R(p), (D.9)
the following scalar equation is found:
(O"eq + 3GLlp)2 = (0";;)2 + 4G 2(t33 - t~~)2 + 6G(t33 - t~~)8~~,
which can be rewritten as follows:
k 1 (p, t33) = 3GLlp + O"y + R(p) - {(0";;)2 + 2G(t33 - t~~)[38~~
+2G(t33 - t~~)]} 1/2 =O (D.lO)
The extra-unknown t33 is such that 0"33 = O, Le.
1
833 + "3tr (7 =O (D.11)

This can be rewritten as follows, using (D.6)a and (D.8),

k 2 (p,t33) = (K + 4~) (t33 - t~~)[O"y + R(p)]


+G [tr + 3K(t33 - t~~)] Llp =O
,
(7tr
.
3Ktr e
'
(D.12)
D.2 General small-strain models 561

In summary, the problem is reduced to finding two scalar unknowns: p and t:33
which satisfy two scalar nonlinear equations (D.lO) and (D.12). This system
can be solved iteratively using Newton's method:

( âk 1 )(it)[p(iHl) _
p
(it)] + (âk
â
1 )(it)[f (it+l) _ (it)] -_ -k1 (it) (it))
P , f 33
âP f33
33 f 33

(ââk2 )(it)[P(it+l) _ p(it)] + (ââk 2 )(it)[f~~+l) _ f~~)] = -k2(p(it),f~~)) (D.13)


P f33

The reader can check that the four partial derivatives are given as follows:
âk 1 dR
âp 3G + dp'
âk 1 G[3sM; + 4G(f33 - 4'3)]
ât:33 k 1 - 3GL1p - [O'y + R(p)]
âk2 (K4+G
âp
= 3) (f33 - t:33 )
tT dR
dp + 3GKtr t:

âk2
âf33
(K + 4~) [O'y +R(p)] +3GKL1p (D.14)

The algorithm presented here is the same as in (Aravas, 1987). Another,


projection-based algorithm is proposed in (Simo and Taylor, 1986).

D.2 General small-strain models

As we have seen in the previous section, enforcing a plane stress condition for
h elasto-plasticity, which is perhaps the simplest nonlinear material model,
is rather involved. For more sophisticated models, the algorithm can become
very tedious. That is why it may be advantageous to use the following pro-
cedure which is both very simple and general.
The plane stress condition is considered as a nonlinear equation where
the unknown is f33:

(D.15)

This equation is solved iteratively with Newton's method:

(it) + (â0'33 (it)]_ O

------
)(it) [ (it+l) _
0'33 â f 33 f 33 - (D.16)
t:33
(it )
c 3333

So, at each iteration (it), the constitutive routine is called with aU strain
components given, including f~~). The routine computes the stress (T and
the material (consistent or algorithmic) tangent c at tn+l' If the plane stress
562 D. Zero-stress constraints

condition (D.15) is not satisfied, then we iterate again by computing E~~+1)


from (D.16).
For a beam with axis 1, only Ell is known. Components E22 and 1:33 are
found from the conditions:

(D.17)

Similarly to the plane stress case -(D.16)- these two nonlinear equations are
solved iteratively using Newton's method.

D.3 General finite-strain models

In this section, we extend the algorithm of the previous section to the finite-
strain regime. In plane stress, component F33 of the deformation gradient is
unknown and should be computed such that the 33-component of the Cauchy
stress u is nil. As we shall see hereafter, it is easier to write the condition as
follows:

(D.18)

where T = (det F)u is the Kirchhoff stress. This equation is solved iteratively
with Newton's method:

(it)
7 33
+ ( 8733 )(it) [F,(it+1)
8F33 33
_ F,(it)j
33
=O (D.19)

The problem is to compute (8733/8F33) knowing the spatial tangent operator


c. The reader can assume hyperelasticity for now, but the results we shall
find are also valid for elasto-plasticity.
The following reiat ion between Kirchhoff and first Piola-Kirchhoff stresses
hold:

(D.20)

Taking the partial derivative w.r.t. FpQ and making extensive use of results
found in Sec. 15.1.3, we obtain the following successive equalities:
8PiK
8FpQ F jK + PiK8jp8KQ
AKiQpFjK + PiQ8jp
= (S KQ 8ip + FiMFpNCKMQN)FjK + P iQ 8 jp
(PjQ8ip + P iQ8jp ) + FjKFiMFrRFpNCKMRN(F-l)Qr
(PjQ8ip + P i Q8jp ) + Cjirp(F-l)Qr, (D.21)

where c is the spatial tangent operator. Now, using the relation between T
and P again,
D.3 General finite-strain models 563

(D.22)

we can recast (D.21) in the following simple format:

(D.23)

where asum over r is assumed. The Jacobian needed in (D.19) is found as


an application: (i, j) = (3,3) and (p, Q) = (3,3),

(D.24)

In summary, for plane stress, at each iteration (it), the constitutive routine is
called with aU deformation gradient components given, including Fi;t). The
routine computes the Kirchhoff stress T and the spatial tangent operator c
at tn+l. If the plane stress condition (D.18) is not satisfied, then we iterate
again by computing Fi;t+l) from (D.19) and (D.24)
For a beam with axis 1, only Fll is known. Components F 22 and F33 are
found from the conditions:

(D.25)

These two nonlinear equations are solved iteratively using Newton's method.
The four partial derivatives which are needed are given directly by (D.23).
Exercise: Consider a finite-strain elasto-plastic model based on a multi-
plicative deformat ion gradient decomposition and hyperelasticity and show
that result (D.23) still holds. Hint: use Eq. (16.35) and review the above
calculations carefully.
References

ABAQUS, a general-purpose finite element program, Ribbitt, Karlsson & Sorensen,


Inc., Pawtucket, RI, U.S.A.
Anderson (1995) Anderson, T.L. (1995), Fracture mechanics- Fundamentals and
applications, 2nd edition, CRC Press, Boca Raton, Florida.
Andrieux, S., M. Joussemet & E. Lorentz (1996a), "A Class of constitutive relations
with internal variable derivatives" , Proc. French national MECAMAT colloquium
, Aussois, France, pp. 116-123.
Andrieux, S., M. Joussemet & E. Lorentz (1996b), "A Class of constitutive relations
with internal variable derivatives. Derivation from homogenization and initial
value problem", Journal de Physique IV, 6, 463-472 (supplement to Journal de
Physique III, 10).
Aravas, N. (1987), "On the numerical integrat ion of a class of pressure-dependent
plasticity models", Internat. Journal for Numerical Methods in Engineering, 24,
1395-1416.
Armstrong, P.J. & C.O. Frederick (1966), "A mathematical representation of the
multiaxial Baushinger effect", Report no. RD/B/N731, General Electricity Gen-
erating Board, Berkeley Nuclear Laboratoires.
Ashby, M.F. & D.R.R. Jones (1980), Engineering Materials 1: An introduction to
lheir properties and applications, Pergamon Press, Oxford.
Ashby, M.F. & D.R.R. Jones (1986), Engineering Materials 2: An introduction to
microslructures, processing and design, Pergamon Press, Oxford.
Auricchio, F. & R.L. Taylor (1995), "2 Material models for cyclic plasticity- Non-
linear kinematic hardening and generalized plasticity", [nt. J. Plasticity, 11(1),
65-98.
Bartczak, Z., A. Galeski, A.S. Argon, & R. Cohen (1996), Polymer, 37, 2113.
Bathe, K.J. (1982), Finite element procedures in engineering analysis, Prentice-Rall,
Englewood Cliffs, N.J., U.S.A.
Bazant, Z.P., T. Belytschko & T.P. Chang (1984), "Continuum theory for strain
softening", Journal of engineering mechanics 110-12, 1666-1693.
Benallal, A., R. Billardon & G. Geymonat (1993), "Bifurcation and localization in
rate-independent materials", in Bifurcation and stability of dissipative systems,
Q.S. Nguyen (editor), CISM Lecture notes no. 327, Udine, ltaly, published by
Springer-Verlag, New York.
Billardon, R. & I. Doghri (1989), "Prevision de I'amon;age d'une macro-fissure par
localisation de l'endommagement", Comples Rendus de l'Academie des Sciences
de Paris, 308(11), 347-352.
Bisplinghoff, R.L., J.W. Marr & T.R.R. Pian (1965), Stalics of deformable solids,
Addison-Wesley Publishing Co., Reading, Mass., republished by Dover, New
York, 1990.
Boley, B.A. & J.R. Weiner (1960), Theory of thermal slress, J. Wiley & Sons, New
York, republished by Dover, New York, 1997.
566 References

Bolotin, V.V. (1996), Stability problems in fracture mechanics, John Wiley & Sons,
New York.
Bonet, J. & RD. Wood (1997), Nonlinear continuum mechanics for finite element
analysis, Cambridge University Press, New York.
Calladine, C.R (1969), Engineering plasticity, Pergamon Press, Oxford.
Chou, P.C. & N.J. Pagano (1967), Elasticity. Tensor, dyadic and engineering ap-
proaches, D. Van Nostrand Co., Princeton, New Jersey, republished by Dover,
New York, 1992.
Coleman, B.D. & D. Gurtin (1967), "Thermodynamics with internal state vari-
ables", Journal of Chemical Physics, 47, 597-613.
Cordebois, J.-P. & P. Ladeveze (1986), "Sur la prevision des courbes limites
d'emboutissage", J. de Meca. Theorique et Appliquee, 5(3), 341-370.
Courbon, J. (1971), Resistance des materiaux, Tome 1, Dunod, Paris.
Crisfield, M. A. (1981), "A fast incrementaljiterative solution procedure that han-
dles snap-through", Computers & Structures 13, 55~62.
Crisfield, M.A. (1997), Nonlinear finite element analysis of solids and structures.
Voi. 1: Essentials, John Wiley & Sons, U.K.
Crisfield, M.A. (1998), Nonlinear finite element analysis of solids and structures.
Voi. 2: Advanced topics, John Wiley & Sons, U.K.
Crist, B., C. Fisher, & P. Howard (1989), Macromolecules, 22, 1709.
Crochet, M. (1993), Cours d'elasticite, Lecture notes, report, Unite de Mecanique
Appliquee, Universite catholique de Louvain, Belgium.
Crochet, M. (1994), Plaques et coques elastiques, Lecture notes, report, Unite de
Mecanique Appliquee, Universite catholique de Louvain, Belgium.
Dahoun, A. (1992), Comportement plastique et textures de deformation des
polymeres semi-cristallins en traction uniaxiale et en cisaillement simple, Ph.D.
dissertation, Institut National Polytechnique de Lorraine, France.
Dahoun, A., G.R Canova, A. Molinari, M.J. Philippe & Ch. G'Sell (1991), "The
modeling of large strain textures and stress-strain relations of polyethylene",
Texture and Mierostruetures, Vols. 14-18, 347-354.
Dahoun, A. et al. (1995), Polym. Eng. Se., 35, 317.
De Borst, R. (1991), "Simulation of strain localization: a reappraisal of the Cosserat
continuum" , Eng. Comput. , 8, 317-332.
De Borst, R., L.J. Sluys, H.B. Muhlhaus & J. Pamin (1993), "Fundamental issues in
finite element analysis of localization of deformation", Eng. Comput. 10,99-121.
De Gennes, P.G. & J. Prost (1993), The physics of liquid crystals, 2nd edition,
Oxford University Press, Oxford.
Dellus, P. (1961), Resistance des materiaux, Technique et Vulgarisation, Paris.
Den Hartog, J.P. (1949), Strength of materials, McGraw-Hill Book Co., New York,
republished by Dover, New York, 1961.
Den Hartog, J.P. (1952), Advaneed strength of materials, McGraw-Hill Book Co.,
New York, republished by Dover, New York, 1987.
Desai, C.S. & H.J. Siriwardane (1984), Constitutive laws for engineering materials,
with emphasis on geologic materials, Prentice Hali, Englewood Cliffs, New Jersey,
U.S.A.
De Vathaire, M. & A. Faessel (1981), "Contraintes internes dans les bobines de
tâles laminees fi froid", Revue de Metallurgie, CIT, Mai, 405-419.
Doghri, 1. (1989), Etude de la loealisation de l'endommagement, Ph.D. dissertation,
Universite Paris 6.
Doghri, 1. (1993), "Fully implicit integration and consistent tangent modulus in
elastoplasticity", Internat. Journal for Numerical Methods in Engineering, 36,
3915-3932.
References 567

Doghri,1. (1995), " Numerical implementation and analysis of a class of metal plas-
ticity models coupled with ductile damage", Internat. Journal for Numerical
Methods in Engineering, 38, 3403-343l.
Doghri, 1. & R. Billardon (1995), "Investigation of localization due to damage in
elasto-plastic materials", Mechanics of Materials, 19/2-3, 129-149.
Doghri, 1., S. Jansson, F.A. Leckie & J. Lemaitre (1994), "Optimization of coating
layers in the design of ceramic fiber reinforced metal matrix composites" , Journal
of Composite Materials, 28/2, 167-187.
Doghri, 1. & F. A. Leckie (1994), "Elasto-plastic analysis of interface layers for
fiber-reinforced metal-matrix composites", Composites Science and Technology,
51, 63-74.
Doghri, 1., A. Muller & R.L. Taylor (1998), "A general three dimensional contact
procedure for implicit finite element codes", Engineering Computations, 15/2,
233-259.
Dreyfuss, G. (1962), Resistance des materiaux des recipients sous pression, Editions
Technip, Paris.
Dumontet, H., G. Duvaut, F. Lene, P. Muller & N. Turbe (1994), Exercices de
mecanique des milieux continus, Masson, Paris.
Duvaut, G. (1990), Mecanique des milieux continus, Masson, Paris.
Dym, C.L., I.H. Shames (1973), Solid mechanics: a variational approach, McGraw-
Hill Book Co., New York.
Eshelby, J.D. (1957), "The determination of the elastic field of an ellipsoidal inclu-
sion, and related problems", Proc. Roy. Soc. London, Ser. A, 241, pp 376-396.
Feodossiev, V. (1976), Resistance des maUriaux, Mir, Moscow.
Ferencz, R.M. & T.J.R. Hughes (1998), "Iterative finite element solutions in non-
linear solid mechanics", In Handbook of numerical analysis, P.G. Ciarlet and J.L.
Lions (editors), VoI. VI, pp 3-178, Elsevier Science, Amsterdam.
Filonenko-Borodich, M. (1958), Theory of elasticity, Foreign Languages Publishing
House, Moscow, republished by Dover, New York, 1965.
Fleck, N.A. & J.W. Hutchinson (1993), "A phenomenological theory for strain
gradient effects in plasticity", J. Mech. Phys. Solids 41-12, 1825-1857.
Florez J. (1989), ElasticiU couplee ci l'endommagement: formulation, analyse
theorique et approximation numerique , Ph.D. dissertation, Vniversite Paris 6.
Flugge, W. (1973), Stresses in shells, Springer-Verlag, Berlin.
Forest, S. (1996), Modeles mecaniques de la deformation heUrogime des
monocristaux, Ph.D. dissertation, Ecole nationale Superieure des Mines de Paris.
Franl.<ois, D., A. Pineau & A. Zaoui (1993), Comportement mecanique des
materiaux. Viscoplasticite, endommagement, mecanique de la rupture, mecanique
du contact, Hermes, Paris.
French, S.E. (1995), Fundamentals of structural analysis, West Publishing Co., St.
Paul, MN.
Friaa, A. (1982), Theorie des structures, Lecture notes, Departement de Genie Civil,
Ecole Nationale d'Ingenieurs de Tunis, Tunisia.
Geers, M. (1997), Experimental analysis and computational model/ing of damage
and fracture, Ph.D. dissertation, Eindhoven Vniversity of Technology.
Germain, P. & P. Muller (1980), Introduction ci la mecanique des milieux continus,
Masson, Paris.
Germain, P., Q.S. Nguyen & P. Suquet (1983), "Continuum thermodynamics",
Trans. ASME, J. Appl. Mech., 50, 1010-1020.
Goldstein, H. (1980), Classical mechanics, 2nd edition, Addison-Wesley, Reading,
MA, V.S.A.
Green, A.E. & W. Zerna (1968), Theoretical elasticity, 2nd edition, Oxford Vniver-
sity Press, Republished by Dover, New York, 1992.
568 References

Grosberg, A.Y. & A.R Khokhlov (1994), Statistical physics of macromolecules,


American Institute of Physics (AIP) Press, New York.
Guo, Z.H. (1981), "Representations of orthogonal tensors", Solid Mechanics
Archives, 6, 451-466.
Gurtin, M.E. (1981), An introduction to continuum mechanics, Academic Press,
San Diego, California.
Halphen, B. & Q.S. Nguyen (1975), Sur les materiaux standards gem\ralises, Journal
de Mecanique, 14/1, 39-63.
Hartmann, S. & P. Haupt (1993), "Stress computation and consistent tangent op-
erator using non-linear kinematic hardening models", Internat. Journal for Nu-
merical Methods in Engineering, 36, 3801-3814.
Havner, K.S. (1992), Finite Plastic DeJormation of Crystalline Solids, Cambridge
University Press, London.
Hertzberg, RW. (1989), Deformation and Jracture mechanics of engineering mate-
rials, 3rd edition, Wiley, New York.
Hibbeler, RC. (1997), Structural analysis, 3rd edition, Prentice Hali, New Jersey.
Hill, R (1950), The mathematical theory of plasticity, Clarendon Press, Oxford
University Press, England, reprinted, 1989.
Hill, R (1958), "A general theory of uniqueness and stability in elastic-plastic
solids", J. Mech. Phys. Solids 6.
Hill, R (1968), "On constitutive inequalities for simple materials-I", J. Mech. Phys.
Solids, 16, 229-242.
Hill, R (1983), "On intrinsic eigenstates in plasticity with generalized variables",
Math. Proc. Camb. Phil. Soc., 93,177-189.
Hodge, P.G., Jr. (1959), Plastic analysis of structures, McGraw Hill Book Co., New
York.
Hoger, A. & D.E. Carlson (1984), "Determination of the stretch and rotation in the
polar decomposition of the deformation gradient", Quarterly of Applied Mathe-
matics, 42, 113-117.
Hughes, T.J.R. (1984), "Numerical implementation of constitutive models: rate-
independent deviatoric plasticity" , in Theoretical foundation for large-scale com-
putations for non-linear material behavior, S. Nemat-Nasser et al. (eds.), Marti-
nus Nijhoff Publishers, Dordrecht, The Netherlands.
Hughes, T.J.R (1987), The finite element method, Prentice-Hall, New Jersey,
U.S.A.
Hughes, T.J.R. & K.S. Pister (1978), "Consistent linearization in mechanics of
solids and structures", Computers and Structures, 8, 391-397.
Hughes, T.J.R & J. Winget (1980), "Finite rotation effects in numeric al integration
of rate constitutive equations arising in large-deformation analysis" , Internat. J.
Numer. Methods Engrg., 15(9), 1413-1418.
Hutchinson, J.W. (1974), "Plastic buckling", in Advances in Applied Mechanics,
voI. 14, Acad. Press, New York, 67-144.
Johnson, K.L. (1987), Contact mechanics, Cambridge University Press, Cambridge,
U.K.
Johnson, W. & P.B. Mellor (1962), Plasticity for mechanical engineers, D. Van
Nostrand Co., London.
Kachanov, L.M. (1971), Foundations of the theory of plasticity, North-HolJand Pub-
lishing Co., Amsterdam.
Kachanov, L.M. (1986), Introduction to continuum damage mechanics, M. Nijhoff
Publ., Dordrecht, The Netherlands.
Kennedy, M., A. Peacock & L. Mandelkern (1994), Macromolecules, 21, 5297.
References f;69

Kestin, J. & J.M. Rice (1970), "Paradoxes in the application of thermodynamics


to strained solids", in A critical review of of thermodynamics, E.B. Stuart et al.
(Editors), pp. 275-298, Mono Book Co., Baltimore, MD, U.S.A.
Kitagava, M. & D. Zhou (1995), Polym. Eng. Se., 35/22, 1725.
Kleiber, M. (Editor) (1998), Handbook of computational solid mechanics- Survey
and comparison of contemporary methods, Springer-Verlag, Berlin.
Knockaert, R., 1. Doghri, Y. Marchal, T. Pardoen & F. Delannay (1996), "Ex-
perimental and numerical investigation of fracture in double-edge notched steel
plates", Internat. Journal of Fracture, 81/4, 383-399.
Knockaert, R. & 1. Doghri (1999), "Nonlocal constitutive models with gradients
of internal variables derived from a micro/macro homogenization procedure",
Computer Methods in Applied Mechanics and Engineering, 174, 121-136.
Kocks, U.F., C.N. Tome & H.-R. Wenk (1998), Texture and anisotropy: preferred
orientations in polycrystals and their etJect on material properties, Cambridge
University Press, Cambridge, U.K.
Krajcinovic, D. (1996), Damage mechanics, Elsevier, Amsterdam.
Ladeveze, P. (1986), Mecanique des structures, Lecture notes, report no. 64, written
by O. Allix and J.-P. Pelle, Laboratoire de Mecanique et Technologie, Ecole
Normale Superieure de Cachan, France.
Lanczos, C. (1970), The variational principles of mechanics, 4th edition, University
of Toronto Press, Toronto, Canada, republished by Dover, New York, 1986.
Larson, R. (1988), Constitutive Equations for Polymer Melts and Solutions, But-
terworth.
Lee, E.H. (1969), "Elastic-plastic deformations at finite strains", J. Appl. Mech.,
36, 1-6.
Lee, B.J., A.S. Argon, D.M. Parks, S. Ahzi, & Z. Bartczak (1993), Polymer, 34,
3555.
Lekhnitskii, S.G. (1981), Theory of elasticity of an anisotropic body, Mir Publishers,
Moscow (translated from the revised 1977 Russian edition).
Lemaitre, J. (1992), A course on damage mechanics, Springer-Verlag, Germany.
Lemaitre, J. & J.-L. Chaboche (1990), Mechanics of solid materials, Cambridge
University Press, England, translation of the French edition (Dunod and Bordas,
Paris, 1985).
Lemaitre, J. & 1. Doghri (1994), "DAMAGE90: A post processor for crack ini-
tiation", Computer Methods in Applied Mechanics and Engineering, 115/3-4,
197-232.
Lielens, G. (1999), Micro-macro modeling of structured materials, Ph.D. disserta-
tion, Faculte des Sciences Appliquees, Universite catholique de Louvain, Belgium.
Lifshitz, I.M. (1969), "Some problems of the statistical theory of biopolymers", Sov.
Phys. JETP, 28, 1280.
Lin, L. & A.S. Argon (1994), J. Mater. Se., 29, 294.
Lippmann, H. (1972), Extremum and variational principles in mechanics, 2nd print-
ing, publication no. 54, CISM, Udine, Italy.
Lippmann, H. (1995), "Cosserat plasticity and plastic spin", Appl. Mech. Rev. 48,
11 part 1, 753-762.
Lubliner, J. (1990), Plasticity theory, Macmillan Publishing Company, New York.
Lucas, J., M. Failla, F. Smith & L. Mandelkern (1995), Pol. Eng. Se., 35/13, 1117.
Luenberger, D.G. (1989), Linear and nonlinear programming, Addison-Wesley Pub-
lishing Co., Reading, MA, U.S.A.
Love, A.E.H. (1927), A treatise on the mathematical theory of elasticity, Fourth
edition, Cambridge University Press, republished by Dover, New York, 1944.
Mandel, J. (1966), Cours de Mecanique des milieux continus. II- Mecanique des
solides, Gauthier-Villars, Paris, France.
570 References

Mandel, J. (1973), Equations constitutives et directeurs dans les milieux plastiques


et viscoplastiques, Int. J. Solids Structures 9, 725-740.
Mandelkern, L., R.G. Alamo & M.A. Kennedy (1990), Macromolecules, 23, 4271.
Marrucci, G. (1984), "Remarks on the viscosity of polymeric liquid crystals", Frac.
IX Intl. Congress on Rheology, Acapulco, Mexico, pp. 441-448.
Marsden, J.E. & T.J.R Hughes (1983), Mathematical foundations of elasticity,
Prentice-Hall, New Jersey, republished by Dover, New York, 1994.
Mase, G.E. (1970), Theory and problems of continuum mechanics, Schaum's outline
series, McGraw-Hill, New York.
Mason, J. (1980), Variational, incremental and energy methods in solid mechanics
and shell theory, Elsevier, Amsterdam, The Netherlands.
Massonnet, Ch., Olszak & A. Phillips (1979), Flasticity in structural engineering:
fundamentals and applications, publication no. 241, CISM, Udine, Italy.
Massonnet, Ch. & Save, M. (1967), Calcul plastique des contructions. 1- Structures
dependant d'un parametre, 2nd edition, Centre Belgo-Luxembourgeois de l'acier,
Brussels, Belgium
Maugin, G.A. (1992), The thermomechanics of plasticity and fracture, Cambridge
University Press, Cambridge, England.
Mura, T. (1987), Micromechanics of defects in solids, 2nd edition, Martinus Nijhoff
Publishers, Dordrecht, The Netherlands.
Nadai, A.L. (1963), Theory of fiow and fracture of solids. Volume II, McGraw Hill
Book Co., New York.
Nagtegaal, J.C. & J.E. de Jong (1981), "Some aspects of non-isotropic work-
hardening in finite deformat ion plasticity", Froc. workshop on plasticity of met-
als at finite strain: theory, computation and experiment, E.H. Lee & R Mallet
(eds.), Division of Applied Mechanics, Stanford University, California.
Nemat-Nasser, S. & M. Hori (1993), Micramechanics: overall praperties of hetera-
geneous materials, Elsevier Science publishers, Amsterdam.
Nguyen, Q.S. (1987), "Bifurcation and post-bifurcation analysis in plasticity and
brittle fracture", J. Mech. Fhys. Solids 35-3, 303-324.
Nguyen, Q.S. & H.D. Bui (1974), "Sur les materiaux elastoplastiques il, ecrouissage
positif ou negatif', Journal de Mecanique, 13/2, 321-342.
Nielsen, M.K. and H.L. Schreyer (1993), Bifurcations in elastic-plastic materials,
Int. J. Solids Structures 30, 521-544.
Nikolov, S. & 1. Doghri (2000), "A micro/macro constitutive model for the smalI-
deformat ion behavior of polyethylene", Folymer, 41, 1883-1891.
Novozhilov, V.V. (1953), Foundations of the nonlinear theory of elasticity, Graylock
Press, Rochester, N.Y., translated from the first (1948) Russian edition.
Oden, J.T. & J.N. Reddy (1976), Variational methods in theoretical mechanics,
Springer-Verlag, Berlin.
Ogden, RW. (1984), Non-linear elastic deformations, Ellis Horwood Limited, West
Sussex, England, republished by Dover, New York, 1997.
Ortiz, M., Y. Leroy & A. Needleman (1987), "A finite element method for localized
failure analysis", Computer Methods in Applied Mechanics and Engineering, 61,
189-214.
Ortiz, M. & E.P. Popov (1985), "Accuracy and stability of integration algorithms
for elastoplastic constitutive relations", Internat. Journal for Numerical Methods
in Engineering, 21, 1561-1576.
Ortiz, M. & J.C. Simo (1986), "An analysis of a new class of integration algorithms
for elastoplastic constitutive equations", In. J. Numer. Methods Eng., 23, 353-
366.
Owen, D.RJ. & E. Hinton (1980), Finite elements in plasticity: theory and practice,
Pineridge Press Ltd., Swansea, U.K.
References 571

Paduart, A., J. Kestens & G. Warzee (1984), Calcul des structures hyperstatiques,
Masson-Editions de l'universite de Bruxelles, Belgium.
Pamin, J.K. (1994), Gradient-dependent plasticity in numerical simulation of local-
ization phenomena, Ph.D. dissertation, Delft University of Technology. .
Pardoen, T., 1. Doghri & F. Delannay (1998), "Experimental and numeric al com-
parison of void growth models and void coalescence criteria for the prediction of
ductile fracture in copper bars", Acta Mater., 46/2, 541-552.
Parton, V. & P. Perline (1983), Equations integrales de la theorie de l'elasticite,
Mir, Moscow.
Parton, V. & P. Perline (1984a), Methodes de la theorie mathematique de l'elasticite,
Volume 1, Mir, Moscow.
Parton, V. & P. Perline (1984b), Methodes de la theorie mathematique de l'elasticite,
Volume 2, Mir, Moscow.
Perzyna, P. (1963), Q. Appl. Math., 20, 32l.
Pissarenko, G., A. Yakovlev & V. Matveev (1979), Aide-memoire de resistance des
materiaux, Mir, Moscow.
Prager, W. & P.G. Hodge, Jr. (1951), Theory of perfectly plastic solids, John Wiley
& Sons, New York.
Przemieniecki, J.S. (1968), Theory of matrix structural analysis, McGaw-Hill Book
Co., New York, republished by Dover, New York, 1985.
Popov, E.P. (1952), Mechanics of materials, Prentice-Hall' New York.
Rice, J.R. (1976) "The localization of plastic deformat ion" , in Theoretical and ap-
plied Mechanics , North-Holland, Amsterdam.
Rice, J.R. & J.W. Rudnicki (1980), "A note on some features of the theory of
localization of deformat ion" , 1nt. J. Solids Structures, 16, 597-605.
Roux, J. (1995), Resistance des materiaux par la pratique- 2: Methodes energetiques,
poutres continues, systemes reticules, calcul des ossatures, methodes matricielles,
Eyrolles, Paris.
Salenc;on, J. (1988a), Mecanique des milieux continus. l-Concepts generaux, EI-
lipses, Ecole Polytechnique, Paris.
Salenc;on, J. (1988b), Mecanique des milieux continus. II-Elasticite. Milieux
curvilignes, Ellipses, Ecole Polytechnique, Paris.
Save, M. & Ch. Massonnet (1972), Calcul plastique des contructions. 11- Structures
dependant de plusieurs parametres, 2nd edition, Centre Belgo-Luxembourgeois
de l'acier, Brussels, Belgium.
Sechler, E.E. (1952), Elasticity in engineering, John Wiley and Sons, New York,
republished by Dover, New York, 1968.
Sedov, L. (1975), Mecanique des milieux continus, Volume 1, Mir, Moscow.
Segel, L.A. (1977), Mathematics applied ta continuum mechanics, Macmillan Pub-
lishing Co., New York, republished by Dover, New York, 1987.
Shadrake, L. G. & F. Guiu (1976), Phyl. Magasine, 34, 565.
Simo, J.C. (1992), "Algorithms for static and dynamic multiplicative plasticity
that preserve the classical return mapping schemes of the infinitesimal theory" ,
Computer Methods in Applied Mechanics and Engineering, 99, 61-112.
Simo, J.C. (1998), "Numerical analysis and simulat ion of plasticity", In Handbook
of numerical analysis, P.G. Ciarlet and J.L. Lions (editors), VoI. VI, pp 183-499,
Eisevier Science, Amsterdam.
Simo, J.C. & D.D. Fox (1989), "On a stress resultant geometrically exact shell
model. Part 1: Formulation and optimal parametrization", Computer Methods in
Applied Mechanics and Engineering, 72, 267-304.
Simo, J.C. & T.J.R. Hughes (1998), Computational inelasticity, Springer-Verlag,
New York.
572 References

Simo, J.C., J.W. Ju, R.L. Taylor & K.S. Pister (1987), "On strain-based continuum
damage models: formulation and computational aspects", in Constitutive laws
for engineering materials: theory and applications. Volume 1, C.S. Desai et al.
(editors), Elsevier, New York, pp. 233-245.
Simo, J.C., J.G. Kennedy & R.L. Taylor (1989), "Complementary mixed finite
element formulations for elastoplasticity", Comput. Meth. Appl. Mech. Eng. 74,
177-206.
Simo, J.C. & C. Miehe (1992), "Associative coupled thermoplasticity at finite
strains: formulation, numerical analysis and implementation", Computer Meth-
ods in Applied Mechanics and Engineering, 98:1, 41-104.
Simo, J.C. & R.L. Taylor (1985), Consistent tangent operators for rate-independent
elastoplasticity, Computer Methods in Applied Mechanics and Engineering, 48,
101-118.
Simo, J.C. & R.L. Taylor (1986), A return mapping algorithm for plane stress
elastoplasticity, Internat. Journal for Numerical Methods in Engineering, 22,
649-670.
Simo, J.C. & R.L. Taylor (1991), "Quasi-incompressible finite elasticity in principal
stretches. Continuum basis and numerical algorithms", Comput. Meth. Appl.
Mech. Eng., 85, 273-310
Skrzypek, J. & A. Ganczarski (1999), Modeling of material damage and failure of
structures- Theory and applications, Springer-Verlag, Berlin.
Sluys, L.J. (1992), Wave propagation, localization and dispersion in softening solids,
Ph.D. dissertation, Delft University of Technology, Delft.
SPECTRUM, a general-purpose finite element program, Centric Engineering Sys-
tems, Inc., Sunnyvale, CA, U.S.A.
Stronge, W.J. & T.X. Yu (1993), Dynamic models for structural plasticity, Springer-
Verlag, London.
Suquet, P. (Editor) (1997), Continuum micromechanics, CISM Lecture notes,
Udine, Italy, published by Springer-Verlag, Berlin.
Temam, R. (1985), Mathematical problems in plasticity, Gauthier-Villars, Paris,
translated from the 1983 French edition, Bordas, Paris.
Teodosiu, C. (Editor) (1997), Large plastic deformation of crystalline aggregates,
CISM Lecture notes, Udine, Italy, published by Springer-Verlag, Berlin.
Thomason, P.F. (1990), Ductile fracture of metals, Pergamon press, Oxford.
Thomsen, E.G., C.T. Yang & S. Kobayashi (1965), Mechanics of plastic deformation
in metal processing, The MacMillan Co., New York.
Timoshenko, S.P. (1956a), Strength of materials- Part 1: Elementary theory and
problems, 3rd edition, D. Van Nostrand Co., Princeton, New Jersey.
Timoshenko, S.P. (1956b), Strength of materials- Part II: Advanced theory and
problems, 3rd edition, D. Van Nostrand Co., Princeton, New Jersey.
Timoshenko, S.P. & J.M. Gere (1961), Theory of elastic stability, 2nd edition, Mc-
Graw Hill Book Co., New York.
Timoshenko, S.P. & J.N. Goodier (1987), Theory of elasticity, McGraw-Hill Book
Co., New York.
Timoshenko, S.P. & S. Woinowsky-Krieger (1982), Theory of plates and shells,
McGraw-Hill Internat. Book Co., Tokyo.
Ting, T.T. (1985), "Determination of C 1/ 2 , C- 1/ 2 and more general isotropic tensor
functions of C", Journal of Elasticity, 15, 319-323.
Tolstov, G.P. (1962), Fourier series, translated from the Russian by R.A. Silverman,
Prentice HalI, Englewood Cliffs, New Jersey, republished by Dover, New York,
1976.
TruesdelI, C. & R.A. Toupin (1960), "The classical field theories", in Handbook der
Physik, III/l, Springer-Verlag, Berlin.
References 573

Turna, J.J. (1988), Handbook of structural and mechanical matrices, McGraw-Hill


Book Co., New York.
Valanis, K.C. (1971), Arch. Mech., 23, 517.
Van Houtte, P. (1998), Micromechanics of polycrystalline materials, Lecture Notes,
report, MTM Department, Katholieke Universiteit Leuven, Belgium.
Whittaker, E.T. (1937), A treatise on the analytical dynamics of particles and rigid
bodies, 4th edition, Cambridge University Press, Cambrdige
Wilkins, M.L. (1964), "Calculation of elasto-plastic flow" , in Methods of computa-
tional physics 3, B. Alder et al. (eds.), Academic Press, New York.
Wissbrun, K.F. (1985), Faraday Discuss. Chem. Soc., 79, 161.
Wu, P.D. & E. Van Der Giessen (1992), "A modified 3-D constitutive model for
glassy polymers and its application to large simple shear of polycarbonate", In
Modelling of Plastic Deformation and Its Engineering Applications, S.1. Andersen
et al. (editors), Proc. 13th Riso Internat. Symposium on Materials Science, 7-11
September, 1992, Roskilde, Denmark, pp. 519-524.
Wu, P.D. & E. Van Der Giessen (1994), "Analysis of shear band propagation in
amorphous glassy polymers", Internat. Journal of Solids and Structures, 31,
1493-1517.
Young, R.J. (1988), Mater. Forum, 11, 210.
Zbib, H.M. & E.C. Aifantis (1989), "A gradient-dependent flow theory of plasticity:
application to metal and soil instabilities", Appl. Mech. Rev. 42-11, 295-304.
Zealouk, L., S. Nikolov & 1. Doghri (1999), "Numerical simulat ion of the micro-
mechanics of semi-crystalline polymers in small strains", Proc. European Con-
ference an Computational Mechanics (ECCM '99), August 31-Scptember 3,1999,
Munich, Germany.
Zienkiewicz, O.C. & R.L. Taylor (1989), The finite element method. Val. 1- Basic
Formulation and linear problems, 4th edition, McGraw Hill Book Co., England.
Zienkiewicz, O.C. & R.L. Taylor (1991), The finite element method. Voi. 2- Solid
and fluid mechanics, dynamics and nonlinearity, 4th edition, McGraw Hill Book
Co., England.
Index

acceleration, 338 - plate, 133


accumulated plastic strain, 302 boundary-value problem, see local or
acoustic tensor, 473 weak formulations , 313
additive decomposition, 321, 413 brittle material, 16, 203, 241
Airy stress function, 167, 196 buckling, 249
algorithmic tangent operator, see bulk modulus, 20, 377
consistent tangent operator
Almansi-Euler strain, 346 Cardan's formulae, 551
- algorithm, 416 Castigliano's theorem, 44
Armstrong-Frederick model, see - beam, 68, 85
kinematic hardening Cauchy stress, 351, 371
axisymmetric problem Cauchy-Green strain
- 2D, 198 - left, 340, 375, 399
- plate, 146 - right, 340, 370, 400
- thermo-elasticity, 239 Cayley-Hamilton theorem, 376, 418
centroid, 50
back stress, see kinematic hardening Chaboche-Marquis model, see
stress kinematic hardening
backward Euler integration, 316 change of coordinates
Baushinger effect, 302 - cylindrical, 545
beam, 49 - matriees for, 556
beam-column, 268 - polar, 193
Beltrami-Mitchell compatibility characteristie volume, 498
equations, 24 circular arch, 214
bending moment circular section
- beam, 58 - torsion, 113, 114
- plate, 130, 143, 148 circular tube
bending problem - torsion, 121
- beam, 51 Clausius-Duhem inequality, 321, 367,
- plate, 131 399
bending stiffness, 130, 297 - ductile damage, 445
bi-harmonie, 25, 167 - kinematic hardening, 428
bifurcation - nonlocal model, 503
- continuous, 475 Codazzi-Gauss compatibility condi-
- discontinuous, 475 tions, 280
- elastic damage, 471 coil winding, 206
Biot's strain, see nominal strain compatibility equations, 12
body force complementary energy, 38
- 2D, 187 composite material, 26, 205, 242
boundary condition, 9, 388 compression of a disk, 230
- beam, 60 conie surface, 283
Index 575

conjugate function, 98, 105 - mechanical, 324, 399


conjugate variables, see dual variables distortion energy, 22
conservat ion of linear momentum, 350 divergence operator
conservat ion of mass, 349 - cylindrical coordinates, 546, 547
consistent tangent operator, 318 Drucker's stability criterion, 312
- ductile damage, 459 dual variables, 33, 69, 322, 356
- kinematic hardening, 433 ductile material, 203, 226, 241
- viscoplasticity, 333
continuity equations, 11 eccentric force, 267
- beam, 60 effective stress, 440, 443
contour lines eigenvalues, see principal values
- torsion, 100 eigenvectors, see principal directions
Cosserat mechanics, 496 Einstein's summation convention, 1
creep, 329 elastic predictor, 315, 404
critical buckling load, 252 elastic unloading, 309
critical buckling stress, 254 elasticity domain, 304, 307, 325, 399
critical damage, 441 - ductile damage, 443
cross section, 49 - kinematic hardening, 426
current configuration, 337 elasticity tensor
curvature - material, 371, 375
- beam, 54 - spatial, 372, 376, 387, 395, 407
- plate, 129, 145 elasto-plasticity, 305
- shell, 289 elliptic section
- yield surface, 320 - torsion, 110
curved beam, 209 ellipticity
curvilinear coordinates, 273 - rate problem, 473
cyclic plasticity, 303, 423, 435 energy method
cylindrical coordinates, see change of - stability, 252, 256, 269
coordinates entropy, 366
- torsion, 109 equation of state, 321, 400
- ductile damage, 443
damage mechanics, 439 - kinematic hardening, 426
damage model equilibrium equations, 8
- ductile, 440, 442 - 2D, 163, 164, 195
- elastic, 469, 506, 509 - beam, 58
damage potential, 444 - cylindrical coordinates, 547
damage threshold, 441 - plate, 126, 144
damage variable, 439 - shell, 292
dead load, 388 Eshelby's results, 524
deformat ion gradient, 339, 416 Euler's method
deviatoric, 6 - buckling, 252
direct method Eulerian description, 338, 351
- stability, 252 exponential algorithm, 404, 419
directional derivative, 385, 503
Dirichlet problem, 99 fatigue, 462
discontinuity fictitious cut
- beam, 63 - beam, 58
- plate, 151 - plate, 147
displacement finite element method, 47, 313
- shell, 285, 287 first moment, 50
dissipation, 322 Flamant's solution, 224
- elastic damage, 506, 510 flow rule, 307, 322, 403
- macroscopic, 502, 517 - viscoplasticity, 331
576 Index

Fourier series influence function


- 2D, 169, 182, 200 - plate, 140
- plate, 136, 157 influence line, 69
Fourier's heat con duct ion law, 235 initial yield stress, 302
fracture Mechanics, 465 internal energy, 366
free energy, 321, 367, 399, 410 internal force, 3
- duc ti le damage, 442 - plate, 130
- elastic damage, 506, 510 internal length, 496, 502
- kinematic hardening, 426 internal load, see stress resultant
- macroscopic, 500, 515 internal variable, 321, 399
fundamental form - ductile damage, 442
- first, 274 - kinematic hardening, 426
- second, 278 isochoric deformat ion, 6, 350, 374, 382,
403
Gâteaux derivative, 385 isotropic material, 18, 374, 379, 399
Galerkin's method, 45
generalized standard material, 323 J 2 flow theory, 305, 411
- nonlocal model, 503 Jaumann rate, 363, 372
geometric stiffness, 389 - algorithm, 420
Gibbs vector, 420
gradient operator kinematic hardening, 425, 442
- curvilinear coordinates, 283 - stress, 423
- cylindrical coordinates, 545, 546 kinematically admissible, 31
gradient plasticity, 498 kinetic energy, 353, 366
Green-Lagrange strain, 344, 345 Kirchhoff stress, 356, 370, 399, 413
- algorithm, 416 Kirchhoff-Love theory
Green-Naghdi-McInnis rate, 364, 417 - plate, 127, 135
- shell, 287
hardening modulus, 311, 327 Kirsh's solution, 223
- ductile damage, 446 Kronecker's symbol, 2
- kinematic hardening, 429 Kuhn-Tucker conditions, 308
hardening stress, 302, 399, 423
harmonic, 96 Levy's method
heat equation, 234, 324 - plate, 136
heat flux, 235, 366 Lagrange multiplier, 373, 390, 392
Helmholtz's free energy, see free energy Lagrangian description, 338, 354, 355,
Hill's criterion, 306 388
Hill's linear comparison solid, 473 Lagrangian multiplier, 308
Hill's maximum dissipation principle, Lame's problem, 201, 205, 207, 222
307, 324, 402 Lame's coefficients, 18
hollow elliptic section Laplacian operator
- torsion, 118, 119 - cylindrical coordinates, 546
homogenization, 521 laws of thermodynamics, 366
- nonlocal model, 498 - first, 366
Hooke's operator, 17, 305 - second, 366
hydrostatic pressure, 19, 373 Lee's multiplicative decomposition, 397
hydrostatic stress, 24 Leibnitz's formula, 220
hyperelasticity, 369, 401 Lejeune-Dirichlet theorem, 252
hypoelasticity, 369, 372, 413 Lemaitre-Chaboche model, 441, 442
length variation, 343
identity tensor, 2 Lie derivative, 364, 371, 400
incompressibility, 20, 307, 340, 359, linear elasticity, 16, 33, 305
373,392 linearization, 385
indicator function, 323 - of constitutive equations, 387
Index 577

- of deformation, 386 - ductile damage, 442


- of elasto-statics, 387 - kinematic hardening, 425
- of pressure, 393 non-conservative load, 258
- of weak formulation, 389 nonlocal model, 496
local formulation, 9, 29, 351, 356 normal force
local state method, 320 - beam, 57
localization mode, 473 normal stress
localization surface, 473 - beam, 64
localization theorem, 349 normality rule, see flow rule , 323
logarithmic strain, 344, 348 - generalized, 325, 427, 443
logarithmic stretch, 396, 401 Norton's power law, 331, 332

macro-crack, 439, 490 objective stress rate, 363, 413


macroscopic approach, 519 objectivity, 359, 370
material description, see Lagrangian - incremental, 422
description octahedral shear stress, 14
material time derivative, 338 Ogden's model, 380
Maxwell's geometric compatibility, 475
Maxwell-Betti theorem, 42, 47 penalty, 378, 380, 392
- beam, 68, 80 Piola identity, 387, 393
- plate, 140 Piola-Kirchhoff stress
mechanical problem, 237 - first, 354
membrane analogy, 102 - second, 357, 370, 375, 379
membrane problem, 245 plane strain, 163, 195
- plate, 130 - generalized, 28, 201, 240, 243
membrane theory, 298 plane stress, 129, 164, 195, 296
micro/macro approach, 519 - h elasto-plasticity, 559
mid-point - algorithms, 559
- rotation, 420 - general finite-strain models, 562
- rule, 415, 420 - general small-strain models, 561
- spin tensor, 420 - generalized, 131, 170
- strain, 416 plastic buckling, 255
mid-surface plastic corrector, 316
- shell, 273 plastic loading, 309
middle fiber, 49 plastic multiplier, 326
mixed formulation, 390 - ductile damage, 445
mixed stress, 391 - kinematic hardening, 428
Mohr's stress circles, 7 - viscoplasticity, 332
moment of inertia, 50 plastic potential, 325
Mooney-Rivlin model, 378 - ductile damage, 443
Mori-Tanaka model, 526 - kinematic hardening, 427
motion,338 plastic power, 308
multi-connected, 12 plastic strain, 302, 408
- torsion, 104 plastic work
- ductile damage, 449
Nanson's formula, 354 Poisson's ratio, 19
Navier equations, 23 polar coordinates, see change of
Navier-Bernoulli assumption, 59 coordinates
necking, 359 polar decomposition, 340
neo-Hookean model, 377 - algorithm, 417
nominal strain, 344, 346 potential energy, 35, 389, 390
nominal stress, 354, 370 - plate, 139, 148
non-associative plasticity, 325 - torsion, 108
578 Index

Poynting effect, 383 - polydomain, 533


Prandtl's stress function, 98 - slip system, 528
Prandtl-Reuss equation, 305 semi-infinite plate, 224, 228
principal axes of inertia, 51 shape variation, 22
principal curvature lines, 273 shear force
principal directions, 340, 379, 401, 417 - beam, 57
principal invariants, 6, 374, 418 - plate, 143, 148
principal stretches, 340, 374, 379, 418 shear modulus, 19, 377
principal values, 6 shear reduced area, 67, 70
proper-orthogonal, 2, 419 shear stress, 7
pull-back, 403 - beam, 64
push-forward, 372, 375 Simo's algorithm, 404, 406, 408, 410
simple shear, 313, 381
radial return algorithm, 317, 411 simply connected, 12
radius of gyration, 254 slenderness ratio, 254
rate of deformation, 347, 370, 399, 413 small-perturbation hypothesis, 10
rate-dependent plasticity, see spatial description, see Eulerian
viscoplasticity description
reciprocity theorem, see Maxwell-Betti specific heat, 234
theorem specific heat supply, 234
rectangular section spectral decomposition, 341, 406, 418
- torsion, 116 spin tensor, 347
reference configuration, 337 - algorithm, 420
reference temperature, 233 square section
Reissner-Mindlin theory - torsion, 118
- plate, 135 stability, 250
representative volume element (RVE), state varia bie, 320
500, 519 statically admissible, 37
return mapping algorithm, 315, 406 statically determinate problem, 66, 71
- ductile damage, 450 statically indeterminate problem, 66,
- kinematic hardening, 430 77
- viscoplasticity, 333 steady state, 236
Reuss model, 524 stored energy, see strain energy
rigid body motion, 35, 200, 360 strain, 5
Ritz's method, 45 - average, 521
- plate, 160 - cylindrical coordinates, 547
- torsion, 117 - plate, 129
Rodrigues formula, 420 - shell, 286, 288
rotation matrix or tensor, 340, 413 strain energy, 21, 40, 369
- algorithm, 417, 419 - beam, 67
rotation of a disk, 215, 218 - plate, 139, 148
strain energy rele ase rate, 441, 443
Saint-Venant's principle, 25, 53, 134 strain hardening, 310, 311
- 2D, 176, 180, 211 strain increment, 415
- thermo-elasticity, 240, 246 strain localizat ion
- torsion, 97 - algorithm, 513
self-consistent model, 524 - analytical results, 477
self-equilibrium, 37 - ductile damage, 481, 483, 490
semi-crystalline polymer, 527 - Rice's presentation, 474
- amorphous phase, 531 strain softening, 310, 311
- inclusion, 528 - ductile damage, 450, 483
- intermediate phase, 536 - elastic damage, 469
- micro-domain, 533 strength criteria, 12
Index 579

stress, 3 uniaxial stress, 312, 358, 384


- average, 522 - macro. scale, 542
stress concentration, 115, 218, 224 - strain localization, 480
stress power, 353 uniqueness, 34, 35, 38
stress resultant, 57 - rate problem, 473
- beam, 62
- plate, 123 variational formulation, 33, 37, 390
- shell, 290 velocity, 338
stress triaxiality, see triaxiality ratio velocity gradient, 347
stretch tensor virtual power theorem, 353
- left, 340 virtual work theorem, 32, 388
- right, 340 - beam, 90
strong formulation, see local formula- viscoelasticity, 535
tion viscoplasticity
structural heating, 324 - function, 329, 332
superposition principle, 25, 42 - Perzyna's formulation, 331
- 2D, 182, 183, 230 viscosity, 330, 534
- plate, 151, 153, 155 viscous stress, 330
surface of revolution, 280 Voigt model, 523
volume variation, 5, 22, 48
tangent operator, 310 von Mises equivalent stress, 13, 306
- ductile damage, 445
- kinematic hardening, 429 warping function, 95
tensor, 1 weak formulation, 32, 353, 355, 388
- matrices for change of coordinates, well posed problem, 10
556 - rate problem, 473
- matrices for storage, 553
thermal conductivity, 235 yield condition or criterion, 302, 399
thermal expansion, 233 - kinematic hardening, 425
thermal problem, 236 yield function, 304
thermal strain, 233 - ductile damage, 443
thermal stress, 237 - kinematic hardening, 426
thermo-mechanical problem, 235 - viscoplasticity, 331
thermodynamic force, 321 yield surface, 304
- ductile damage, 442 Young's modulus, 17
- kinematic hardening, 426
- nonlocal model, 502
thin walled section
- torsion, 121
torsional rigidity, 99, 106
transformat ion rule, 1, see change of
coordinates
- cylindrical coordinates, 548
transient, 245
transport formula, 348
Tresca's criterion, 15
triangular section
- torsion, 113
triaxiality ratio, 441, 447
true stress, see Cauchy stress
Truesdell rate, 364
two-scale approach, see micro/macro
approach

You might also like