Download as pdf or txt
Download as pdf or txt
You are on page 1of 51

Continuum Mechanics for Engineers

4th Edition G. Thomas Mase


Visit to download the full and correct content document:
https://ebookmeta.com/product/continuum-mechanics-for-engineers-4th-edition-g-tho
mas-mase/
More products digital (pdf, epub, mobi) instant
download maybe you interests ...

Tensor Algebra and Tensor Analysis for Engineers With


Applications to Continuum Mechanics Fifth Edition
Mikhail Itskov

https://ebookmeta.com/product/tensor-algebra-and-tensor-analysis-
for-engineers-with-applications-to-continuum-mechanics-fifth-
edition-mikhail-itskov-2/

Tensor Algebra and Tensor Analysis for Engineers With


Applications to Continuum Mechanics Fifth Edition
Mikhail Itskov

https://ebookmeta.com/product/tensor-algebra-and-tensor-analysis-
for-engineers-with-applications-to-continuum-mechanics-fifth-
edition-mikhail-itskov/

Classical Continuum Mechanics (Applied and


Computational Mechanics) 2nd Edition Surana

https://ebookmeta.com/product/classical-continuum-mechanics-
applied-and-computational-mechanics-2nd-edition-surana/

Tensor Analysis for Engineers and Physicists - With


Application to Continuum Mechanics, Turbulence, and
Einstein’s Special and General Theory of Relativity
Schobeiri
https://ebookmeta.com/product/tensor-analysis-for-engineers-and-
physicists-with-application-to-continuum-mechanics-turbulence-
and-einsteins-special-and-general-theory-of-relativity-schobeiri/
Principles of Continuum Mechanics A Basic Course for
Physicists Zden■k Martinec

https://ebookmeta.com/product/principles-of-continuum-mechanics-
a-basic-course-for-physicists-zdenek-martinec/

Hamilton s Principle in Continuum Mechanics Bedford


Anthony

https://ebookmeta.com/product/hamilton-s-principle-in-continuum-
mechanics-bedford-anthony/

Small Animal Diagnostic Ultrasound 4th Edition Thomas


G. Mattoon

https://ebookmeta.com/product/small-animal-diagnostic-
ultrasound-4th-edition-thomas-g-mattoon/

Quantum Mechanics for Electrical Engineers 1st Edition


Isaak D Mayergoyz

https://ebookmeta.com/product/quantum-mechanics-for-electrical-
engineers-1st-edition-isaak-d-mayergoyz/

Quantum Mechanics: For Scientists and Engineers 1st


Edition Harish Parthasarathy

https://ebookmeta.com/product/quantum-mechanics-for-scientists-
and-engineers-1st-edition-harish-parthasarathy/
Continuum Mechanics for
Engineers
APPLIED AND COMPUTATIONAL MECHANICS
A Series of Textbooks and Reference Books
Founding Editor
J.N. Reddy

Con nuum Mechanics for Engineers, Fourth Edi on

G. Thomas Mase, Ronald E. Smelser & Jenn Stroud Rossmann

Dynamics in Engineering Prac ce, Eleventh Edi on

Dara W. Childs, Andrew P. Conkey

Advanced Mechanics of Con nua

Karan S. Surana

Physical Components of Tensors

Wolf Altman, Antonio Marmo De Oliveira

For more information about this series, please visit: h!ps://www.crcpress.com/Applied-and-


Computa"onal-Mechanics/book-series/CRCAPPCOMMEC
Continuum Mechanics for
Engineers
Fourth Edition

G. Thomas Mase
Ronald E. Smelser
Jenn Stroud Rossmann
Fourth edition published 2020
by CRC Press
6000 Broken Sound Parkway NW, Suite 300, Boca Raton, FL 33487-2742

and by CRC Press


2 Park Square, Milton Park, Abingdon, Oxon, OX14 4RN

c 2020 Taylor & Francis Group, LLC

[First edition published by CRC 1991]


[Third edition published by CRC 2009]

CRC Press is an imprint of Taylor & Francis Group, LLC

Reasonable efforts have been made to publish reliable data and information, but the author and publisher
cannot assume responsibility for the validity of all materials or the consequences of their use. The authors
and publishers have attempted to trace the copyright holders of all material reproduced in this publication and
apologize to copyright holders if permission to publish in this form has not been obtained. If any copyright
material has not been acknowledged please write and let us know so we may rectify in any future reprint.

Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmitted,
or utilized in any form by any electronic, mechanical, or other means, now known or hereafter invented,
including photocopying, microfilming, and recording, or in any information storage or retrieval system, without
written permission from the publishers.

For permission to photocopy or use material electronically from this work, access www.copyright.com or contact
the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA 01923, 978-750-8400. For works
that are not available on CCC please contact mpkbookspermissions@tandf.co.uk

Trademark notice: Product or corporate names may be trademarks or registered trademarks, and are used only
for identification and explanation without intent to infringe.

ISBN: 9781482238686 (hbk)


ISBN: 9780429174391 (ebk)

Visit the Taylor & Francis Web site at


http://www.taylorandfrancis.com

and the CRC Press Web site at


http://www.crcpress.com
Contents

Preface to the Fourth Edition ix

Authors xi

Nomenclature xiii

1 Continuum Theory 1
1.1 Chapter Learning Outcomes . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Continuum Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.3 Starting Over . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.4 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

2 Essential Mathematics 5
2.1 Chapter Learning Outcomes . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2 Scalars, Vectors and Cartesian Tensors . . . . . . . . . . . . . . . . . . . . . 6
2.3 Tensor Algebra in Symbolic Notation - Summation Convention . . . . . . 8
2.3.1 Kronecker Delta . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.3.2 Permutation Symbol . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.3.3 ε - δ Identity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.3.4 Tensor/Vector Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.4 Indicial Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.5 Matrices and Determinants . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.6 Transformations of Cartesian Tensors . . . . . . . . . . . . . . . . . . . . . 26
2.7 Principal Values and Principal Directions of Symmetric Second-Order
Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.8 Tensor Fields, Tensor Calculus . . . . . . . . . . . . . . . . . . . . . . . . . 38
2.9 Integral Theorems of Gauss and Stokes . . . . . . . . . . . . . . . . . . . . 43
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

3 Stress Principles 57
3.1 Chapter Learning Outcomes . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
3.2 Body and Surface Forces, Mass Density . . . . . . . . . . . . . . . . . . . . 58
3.3 Cauchy Stress Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
3.4 The Stress Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
3.5 Force and Moment Equilibrium; Stress Tensor Symmetry . . . . . . . . . . 66
3.6 Stress Transformation Laws . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
3.7 Principal Stresses; Principal Stress Directions . . . . . . . . . . . . . . . . . 71
3.8 Maximum and Minimum Stress Values . . . . . . . . . . . . . . . . . . . . 77
3.9 Mohr’s Circles For Stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
3.10 Plane Stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
3.11 Deviator and Spherical Stress States . . . . . . . . . . . . . . . . . . . . . . 87
3.12 Octahedral Shear Stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

v
vi Contents

4 Kinematics of Deformation and Motion 107


4.1 Chapter Learning Outcomes . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
4.2 Particles, Configurations, Deformations and Motion . . . . . . . . . . . . . 107
4.3 Material and Spatial Coordinates . . . . . . . . . . . . . . . . . . . . . . . . 109
4.4 Langrangian and Eulerian Descriptions . . . . . . . . . . . . . . . . . . . . 114
4.5 The Displacement Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
4.6 The Material Derivative . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
4.7 Deformation Gradients, Finite Strain Tensors . . . . . . . . . . . . . . . . . 122
4.8 Infinitesimal Deformation Theory . . . . . . . . . . . . . . . . . . . . . . . 128
4.9 Compatibility Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
4.10 Stretch Ratios . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
4.11 Rotation Tensor, Stretch Tensors . . . . . . . . . . . . . . . . . . . . . . . . . 142
4.12 Velocity Gradient, Rate of Deformation, Vorticity . . . . . . . . . . . . . . . 146
4.13 Material Derivative of Line Elements, Areas, Volumes . . . . . . . . . . . . 151
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156

5 Fundamental Laws and Equations 177


5.1 Chapter Learning Outcomes . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
5.2 Material Derivatives of Line, Surface, and Volume Integrals . . . . . . . . 178
5.3 Conservation of Mass, Continuity Equation . . . . . . . . . . . . . . . . . . 179
5.4 Linear Momentum Principle, Equations of Motion . . . . . . . . . . . . . . 182
5.5 Piola-Kirchhoff Stress Tensors, Lagrangian Equations of Motion . . . . . . 183
5.6 Moment of Momentum (Angular Momentum) Principle . . . . . . . . . . 187
5.7 Law of Conservation of Energy, The Energy Equation . . . . . . . . . . . . 188
5.8 Entropy and the Clausius-Duhem Equation . . . . . . . . . . . . . . . . . . 190
5.9 The General Balance Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
5.10 Restrictions on Elastic Materials by the Second Law
of Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
5.11 Invariance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
5.12 Restrictions on Constitutive Equations from Invariance . . . . . . . . . . . 208
5.13 Constitutive Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214

6 Linear Elasticity 223


6.1 Chapter Learning Outcomes . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
6.2 Elasticity, Hooke’s Law, Strain Energy . . . . . . . . . . . . . . . . . . . . . 223
6.3 Hooke’s Law for Isotropic Media, Elastic Constants . . . . . . . . . . . . . 227
6.4 Elastic Symmetry; Hooke’s Law for Anisotropic Media . . . . . . . . . . . 231
6.5 Isotropic Elastostatics and Elastodynamics, Superposition Principle . . . 236
6.6 Saint-Venant Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
6.6.1 Extension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
6.6.2 Torsion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241
6.6.3 Pure Bending . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
6.6.4 Flexure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
6.7 Plane Elasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
6.8 Airy Stress Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
6.9 Linear Thermoelasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
6.10 Three-Dimensional Elasticity . . . . . . . . . . . . . . . . . . . . . . . . . . 266
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273
Contents vii

7 Classical Fluids 285


7.1 Chapter Learning Outcomes . . . . . . . . . . . . . . . . . . . . . . . . . . . 285
7.2 Viscous Stress Tensor, Stokesian, and Newtonian Fluids . . . . . . . . . . . 285
7.3 Basic Equations of Viscous Flow, Navier-Stokes Equations . . . . . . . . . 287
7.4 Specialized Fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289
7.5 Steady Flow, Irrotational Flow, Potential Flow . . . . . . . . . . . . . . . . 290
7.6 The Bernoulli Equation, Kelvin’s Theorem . . . . . . . . . . . . . . . . . . 295
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297

8 Nonlinear Elasticity 301


8.1 Chapter Learning Outcomes . . . . . . . . . . . . . . . . . . . . . . . . . . . 301
8.2 Nonlinear Elastic Behavior . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301
8.3 Molecular Approach to Rubber Elasticity . . . . . . . . . . . . . . . . . . . 304
8.4 A Strain Energy Theory for Nonlinear Elasticity . . . . . . . . . . . . . . . 309
8.5 Specific Forms of the Strain Energy . . . . . . . . . . . . . . . . . . . . . . . 313
8.6 Exact Solution for an Incompressible, Neo-Hookean Material . . . . . . . 314
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321

9 Linear Viscoelasticity 325


9.1 Chapter Learning Outcomes . . . . . . . . . . . . . . . . . . . . . . . . . . . 325
9.2 Viscoelastic Constitutive Equations in Linear Differential Operator Form . 326
9.3 One-Dimensional Theory, Mechanical Models . . . . . . . . . . . . . . . . 328
9.4 Creep and Relaxation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 332
9.5 Superposition Principle, Hereditary Integrals . . . . . . . . . . . . . . . . . 335
9.6 Harmonic Loadings, Complex Modulus, and Complex Compliance . . . . 337
9.7 Three-Dimensional Problems, The Correspondence Principle . . . . . . . 341
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 347
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348

10 Plasticity 361
10.1 Chapter Learning Outcomes . . . . . . . . . . . . . . . . . . . . . . . . . . . 361
10.2 One-Dimensional Deformation . . . . . . . . . . . . . . . . . . . . . . . . . 362
10.3 Modeling Plasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 366
10.4 Yield Criteria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 369
10.4.1 Tresca-Coulomb Yield Criterion . . . . . . . . . . . . . . . . . . . . 372
10.4.2 von Mises Yield Criterion . . . . . . . . . . . . . . . . . . . . . . . . 374
10.4.3 Kinematic Hardening Yield Criterion . . . . . . . . . . . . . . . . . 376
10.5 Plastic Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 376
10.5.1 Tresca-Coulomb Yield Criterion . . . . . . . . . . . . . . . . . . . . 379
10.5.2 von Mises Yield Criterion . . . . . . . . . . . . . . . . . . . . . . . . 379
10.5.3 Kinematic Hardening Yield Criterion . . . . . . . . . . . . . . . . . 380
10.6 Plastic Modulus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 380
10.6.1 Isotropic Hardening . . . . . . . . . . . . . . . . . . . . . . . . . . . 381
10.6.2 Kinematic Hardening . . . . . . . . . . . . . . . . . . . . . . . . . . 383
10.7 Elasto-Plastic Constitutive Equations . . . . . . . . . . . . . . . . . . . . . . 386
10.7.1 Prandtl-Reuss (J2 ) Elasto-Plastic Equations . . . . . . . . . . . . . . 388
10.7.2 Levy-Mises Flow Equations . . . . . . . . . . . . . . . . . . . . . . . 388
10.7.3 Perfectly Plastic Constitutive Behavior . . . . . . . . . . . . . . . . . 390
10.8 Deformation Theory of Plasticity . . . . . . . . . . . . . . . . . . . . . . . . 390
10.9 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 391
viii Contents

10.9.1 Torsion of a Shaft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 391


10.9.2 Bending of a Beam by a Moment . . . . . . . . . . . . . . . . . . . . 393
10.9.3 Thin-Walled Tube Tension and Torsion . . . . . . . . . . . . . . . . 396
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 398
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 399

Appendix A: General Tensors 403


A.1 Representation of Vectors in General Bases . . . . . . . . . . . . . . . . . . 403
A.2 The Dot Product and the Reciprocal Basis . . . . . . . . . . . . . . . . . . . 405
A.3 Components of a Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 407
A.4 Determination of the Base Vectors . . . . . . . . . . . . . . . . . . . . . . . 408
A.5 Derivatives of Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 410
A.5.1 Time Derivative of a Vector . . . . . . . . . . . . . . . . . . . . . . . 410
A.5.2 Covariant Derivative of a Vector . . . . . . . . . . . . . . . . . . . . 411
A.6 Christoffel Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 414
A.6.1 Types of Christoffel Symbols . . . . . . . . . . . . . . . . . . . . . . 414
A.6.2 Calculation of the Christoffel Symbols . . . . . . . . . . . . . . . . . 415
A.7 Covariant Derivatives of Tensors . . . . . . . . . . . . . . . . . . . . . . . . 416
A.8 General Tensor Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 416
A.9 General Tensors and Physical Components . . . . . . . . . . . . . . . . . . 418
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 421

Appendix B: Viscoelastic Creep and Relaxation 423

Index 427
Preface to the Fourth Edition

First, our thanks to all the users of the previous editions. We are grateful to the faculty
members who have taught from the text, the students who have learned from it, and
those who have found the book a valuable resource for independent study. We hope
that you will find that our updates, emendations, and additions to the text make it a
more valuable introduction to continuum mechanics. The changes made in this edition,
though substantial, did not change the basic concept of the book. We seek to provide
engineering students with a complete, concise introduction to continuum mechanics that
is not intimidating. We have prioritized the accessibility of the text, and emphasized the
relevance of continuum mechanics concepts and mathematics in engineering applications.
Additional discussions and examples particularly highlight applications in the fields of
materials science and bioengineering.
This text has grown from our course notes and problems used to teach the topic to
senior undergraduate or first year graduate students. New in the fourth edition is a list
of the learning outcomes associated with each chapter. To the previous editions’ coverage
of linear elasticity, nonlinear elasticity, and viscoelasticity, we have added discussions
of plasticity and applications of all types of material behavior. We have enhanced and
expanded our explanations of theory where appropriate, and we have added several new
worked examples and end-of-chapter problems to the fourth edition.
One of the things that students find challenging in continuum mechanics and sub-
sequent topics is notation. In the third edition, we had made some changes in nota-
tion making the book more consistent with modern continuum mechanics literature; the
fourth edition maintains the notation of the third, and has been revised to ensure the
book is also self-consistent. Our goal of accessibility informs the text’s organization as
well as content. The extension, torsion, pure bending and flexure subsections provide a
robust foundation for posing and solving basic elasticity problems. We have also added
examples of the application of continuum mechanics to biological materials in light of
their importance.
The student who masters the material here should have the mechanics foundation nec-
essary to be a skilled, thoughtful user of modern computational tools that incorporate
nonlinear kinematics and a range of constitutive relationships. We include some descrip-
tions and examples of writing software to solve complex problems. As has been true
since the first edition, this text is designed as an introduction, providing necessary back-
ground in continuum theory that will prepare early graduate or advanced undergraduate
students for further study. We include References that will build on these fundamen-
tals: advanced texts, mathematically detailed discussions, and applications of continuum
mechanics.
We express our deep gratitude to Rebecca Brannon, of the University of Utah, whose
close reading of the third edition helped guide our efforts in developing the fourth. We
also appreciate the contributions of Ever Barbero of West Virginia University and Luciano
Demasi of San Diego State University. We thank Jonathan Plant and CRC Press, and
Toby Rossmann. Previous editions of this text were strengthened by the contributions of
Ryan Miller, Nicholai Volkoff-Shoemaker, Peter Brennen, Roger Sharpe, John Wildharbor,

ix
x Preface to the Fourth Edition

Kevin Ng, and Jason Luther; Morteza M. Mehrabadi of Tulane University and Charles
Davis of Kettering University; and Sheri Burton. This project originated in the teaching
of George E. Mase, and we remain grateful for his inspiration and insight. Our greatest
thanks go to our families for their patience and support.

G. Thomas Mase
San Luis Obispo, California, USA

Ronald E. Smelser
Charlotte, North Carolina, USA

Jenn Stroud Rossmann


Easton, Pennsylvania, USA
Authors

G. Thomas Mase, Ph.D. is Professor of Mechanical Engineering at California Polytechnic


State University, San Luis Obispo, California. Dr. Mase received his B.S. degree from
Michigan State University in 1980 from the Department of Metallurgy, Mechanics, and
Materials Science. He obtained his M.S. and Ph.D. degrees in 1982 and 1985, respec-
tively, from the Department of Mechanical Engineering at the University of California,
Berkeley. After graduate school, he worked in industry including positions at General
Motors Research Laboratories and Callaway Golf. He has taught or held research posi-
tions at the University of Wyoming, Kettering University, Michigan State University and
California Polytechnic State University. Dr. Mase is a member of numerous professional
societies including the American Society of Mechanical Engineers, American Society for
Engineering Education, International Sports Engineering Association, Society of Experi-
mental Mechanics, Pi Tau Sigma, and Sigma Xi. He received an ASEE/NASA Summer
Faculty Fellowship to work at NASA Lewis (now Glenn) Research Center. While at the
University of California, he twice received a distinguished teaching assistant award in the
Department of Mechanical Engineering. His research interests include mechanics, design
and applications of explicit finite element simulation, with specific foci in golf equipment
design and performance and vehicle crashworthiness.

Ronald E. Smelser, Ph.D., P.E. is Professor and Senior Associate Dean for Academic
Affairs in The William States Lee College of Engineering at the University of North Car-
olina at Charlotte. Dr. Smelser received his B.S.M.E. from the University of Cincinnati
in 1971, a S.M.M.E. from MIT in 1972, and his Ph.D. in mechanical engineering from
Carnegie Mellon University in 1978. His industrial experience includes work at the United
States Steel Research Laboratory, the Alcoa Technical Center, and Concurrent Technolo-
gies Corporation. Dr. Smelser served as a full-time or adjunct faculty member at the
University of Pittsburgh, Carnegie Mellon University, and the University of Idaho, and
was a visiting research scientist at Colorado State University. Dr. Smelser is a member of
the American Academy of Mechanics, the American Society for Engineering Education,
Pi Tau Sigma, Sigma Xi, and Tau Beta Pi. He is also a member and Fellow of the Ameri-
can Society of Mechanical Engineers. Dr. Smelser’s research interests are in the areas of
process modeling including rolling, casting, drawing and extrusion of single and multi-
phase materials, the micromechanics of material behavior and the inclusion of material
structure into process models, and the failure of materials.

Jenn Stroud Rossmann, Ph.D. is Professor of Mechanical Engineering at Lafayette Col-


lege. She earned her B.S. in mechanical engineering and Ph.D. in applied physics from
the University of California, Berkeley. Prior to joining Lafayette, she was a faculty mem-
ber at Harvey Mudd College. Her scholarly interests include the fluid dynamics of blood
in vessels affected by atherosclerosis and aneurysm, the cultural history of engineering,
and the aerodynamics of sports projectiles. She is a member of the American Society for

xi
xii Authors

Engineering Education, the American Society of Mechanical Engineers, the Biomedical


Engineering Society, Pi Tau Sigma, Sigma Xi, Tau Beta Pi, and Phi Beta Kappa. She is
also the author of the novel The Place You’re Supposed to Laugh, and is the co-author, with
Clive L. Dym and Lori Bassman, of the textbook Introduction to Engineering Mechanics: A
Continuum Approach.
Nomenclature

x or xi Spatial or current coordinates

X or XA Material or referential coordinates

u or ui Displacement components or displacement


vector

v or vi Velocity or general vector

a or ai Acceleration or general vector

x∗1 , x∗2 , x∗3 Principal axes

^i
e Unit vectors along coordinate axes
^IA Unit vectors along coordinate axes in refer-
ence configuration

δij Kronecker delta

A or aij Transformation or general matrix

I Identity matrix

εijk Permutation symbol

∂t Partial derivative with respect to time


˙
(·) Derivative with respect to time

∇ or ∂x Spatial gradient operator

∇φ = grad φ = φ,j Scalar gradient

∇v = ∂j vi = vi,j Vector gradient

∇ · v or vi,i Divergence of a vector v

∇ × v or εijk vi,j Curl of a vector v

d/dt = ∂/∂t + vk ∂/∂xk Material derivative operator

b or bi Body force (force per unit mass)

xiii
xiv Nomenclature

p or pi Body force (force per unit volume)

f or fi Surface force (force per unit area)

V, V0 Current and referential total volumes

∆V, dV Small and infinitesimal element of volumes

S, S0 Current and referential total surfaces

∆S, dS Small and infinitesimal elements of surface

ρ Density

^ or ni
n Unit normal in current configuration
^ or NA
N Unit normal in reference configuration
n)
(^
t(^n) or ti Traction vector

σN , σS Normal and shear components of traction


vector

T or tij Cauchy stress or general tensor

T ∗ or t∗ij Cauchy stress referred to principal axes


^ ^
0(N)
p0(N) or pi Piola-Kirchhoff stress vector referred to ref-
erential area

P or PAi First Piola-Kirchhoff stress

s or sAB Second Piola-kirchhoff stress

σ(1) , σ(2) , σ(3) Principal stress values

IT , IIT , IIIT First, second and third stress invariants

σM = 13 tii Mean normal stress

Sij Deviatoric stress components

ηij Deviatoric strain components

J1 = 0, J2 , J3 Deviatoric stress invariants

σoct Octahedral shear stress

F or FiA Deformation gradient tensor

C or CAB Right Cauchy-Green deformation tensor


Nomenclature xv

E or EAB Lagrangian finite strain tensor

c or cij Cauchy deformation tensor

e or eij Eulerian finite strain tensor

 or ij Infinitesimal strain tensor

(1) , (2) , (3) Principal strain values

I , II , III First, second and third infinitesimal strain


invariants

B or Bij Left Cauchy-Green deformation tensor

IB , IIB , IIIB , or
Invariants of left deformation tensor
I1 , I2 , I3
W Strain energy per unit volume or strain
energy density

eN
^
^ direction
Normal strain in the N

γij Engineering shear strain

e = ∆V/V = ii = I Cubical dilatation

ω or ωij Infinitesimal rotation tensor

ω or ωj Rotation vector

ΛN
^ = dx/dX
^
Stretch ratio or stretch in the direction of N

λn^ = dX/dx Stretch ratio in the direction of n


^

R or Rij Rotation tensor

U or UAB Right stretch tensor

V or Vij Left stretch tensor

L or Lij Spatial velocity gradient

D or dij Rate of deformation tensor

W or wij Vorticity, or spin tensor

J = det F Jacobian

P(t) or Pi Linear momentum vector

K(t) Kinetic energy


xvi Nomenclature

P(t) Mechanical power or rate of work done by


forces

S(t) Stress work

Q Heat input rate

r Heat supply per unit mass

q or qi Heat flux vector

θ Temperature or angle

g = grad θ or gi = θ,i Temperature gradient

u Specific internal energy

η Specific entropy or viscoelastic viscosity

ψ Gibbs’ free energy

ζ Free enthalpy

χ Enthalpy

γ Specific entropy production

E Modulus of elasticity or Young’s modulus

G or µ Shear modulus

K Bulk modulus

ν Poisson’s ratio

λ, µ Lamé constants

Cijkl or C General elastic constants

τij Viscous stress tensor

βij Deviatoric rate of deformation

λ∗ , µ ∗ Viscosity coefficients

κ∗ Bulk viscosity coefficient


1
Continuum Theory

1.1 Chapter Learning Outcomes


After mastering the material in this chapter, you will be able to:
• Appreciate the continuum assumption and concepts
• Distinguish between fundamental and constitutive equations
• Contextualize the continuum framework and notation
The atomic/molecular composition of matter is well established. On a small enough
scale, a body of aluminum is really a collection of discrete aluminum atoms stacked
on one another in a particular repetitive lattice. And on an even smaller scale, the atoms
consist of a core of protons and neutrons around which electrons orbit. Thus matter is not
continuous. At the same time, the physical space in which we live is truly a continuum,
for mathematics teaches us that between any two points in space we can always find
another point regardless of how close together we choose the original pair. Clearly then,
although we may speak of a material body as “occupying” a region of physical space, it
is evident that the body does not totally “fill” the space it occupies. It is this “occupying”
of space that will be the basis of our study of continuum mechanics.

1.2 Continuum Mechanics


If we accept the continuum concept of matter, we agree to ignore (or more precisely,
to average) the discrete composition of material bodies, and to assume that the sub-
stance of such bodies is distributed uniformly throughout, and completely fills the space
it occupies. In keeping with this continuum model, we assert that matter may be divided
indefinitely into smaller and smaller portions, each of which retains all of the physical
properties of the parent body. Accordingly, we are able to ascribe field quantities such
as density and velocity to each and every point of the region of space which the body
occupies.
The continuum model for material bodies is important to engineers for two very good
reasons. On the scale by which we consider bodies of steel, aluminum, concrete, etc., the
characteristic dimensions are extremely large compared to molecular distances, so that
the continuum model provides a very useful and reliable representation. Additionally,
our knowledge of the mechanical behavior of materials is based almost entirely upon
experimental data gathered by tests on relatively large specimens.
The analysis of the kinematic and mechanical behavior of materials modeled on the
continuum assumption is what we know as Continuum Mechanics. There are two main

1
2 Continuum Mechanics for Engineers

themes into which the topics of continuum mechanics are divided. In the first, emphasis
is on the derivation of fundamental equations which are valid for all continuous media.
These equations are based upon universal laws of physics such as the conservation of
mass, the principles of energy and momentum, etc. In the second, the focus of attention
is on the development of the constitutive equations characterizing the behavior of specific
idealized materials; the perfectly elastic solid and the viscous fluid being the best known
examples. These equations provide the focal points around which studies in elasticity,
plasticity, viscoelasticity and fluid mechanics proceed.
Mathematically, the fundamental equations of continuum mechanics mentioned above
may be developed in two separate but essentially equivalent formulations. One, the
integral, or global form, derives from a consideration of the basic principles being applied
to a finite volume of the material. The other, a differential, or field approach, leads to
equations resulting from the basic principles being applied to a very small (infinitesimal)
element of volume. In practice, it is often useful and convenient to deduce the field
equations from their global counterparts.
As a result of the continuum assumption, field quantities such as density and velocity
which reflect the mechanical or kinematic properties of continuum bodies are expressed
mathematically as continuous functions, or at least piecewise continuous functions, of
the space and time variables. Moreover, such functions are sufficiently smooth that their
derivatives will be continuous as well.
Inasmuch as this is an introductory textbook, we shall make two further assumptions
on the materials we discuss in addition to the principal one of continuity. First, we require
the materials to be homogeneous, that is to have identical properties at all locations. And
secondly, that the materials be isotropic with respect to certain mechanical properties,
meaning that those properties are the same in all directions at a given point. Later,
we will relax this isotropy restriction to discuss briefly anisotropic materials which have
important meaning in the study of composite materials.

1.3 Starting Over


The topic of continuum mechanics typically comes at the end of an undergraduate or at
the beginning of a graduate program. Continuum mechanics has a reputation of being
a theoretical course more focused on mathematical abstraction than on applications. The
first part of continuum mechanics’ reputation is correct: it is based on fundamental math-
ematics and mechanics. However, the second part is not founded. There are many, many
applications for continuum mechanics, but it is hard to cover the basics and develop the
applications in a single quarter or semester. Continuum mechanics takes all the math-
ematical, physical and engineering principles and casts them in a single structure from
which the student is prepared to pursue advanced engineering topics. After a course
in continuum mechanics many applications become accessible to the student: elastic-
ity, nonlinear elasticity, plasticity, crashworthiness, biomechanics, polymers and more.
Many sophisticated simulation programs such as LS-DYNAr become a playground for
advanced design and analysis once continuum mechanics has been mastered.
Some students find continuum mechanics to be a difficult subject. However, outside
of a new notation, the topics studied should be very familiar to the student. Vectors
have to be written in component form, and we need to be able to use “dot” and “cross”
products. These are skills from the sophomore level statics course. Also needed will be
Continuum Theory 3

a description for conservation of linear and angular momentum. Taking the time rate of
change of these quantities is really no different from what was done in an undergraduate
course in dynamics. Just as in dynamics courses, a description of the energy equation will
be examined. Rather than study only rigid bodies as done in undergraduate statics and
dynamics, deformation is allowed. This requires defining stress and strain that were first
introduced in a mechanics of solids class. Stress and strain are tensors which are an order
more complex than vectors. When looking at strain and the resulting stress one needs to
have a material model. At the undergraduate level students have studied linear elastic,
fluid and gas behavior. But the topics generally are taught in separate courses, and often
the common, underlying theory is not noticed. Also, the methods used to determine
the relationship between stress and strain, and the full spectrum of possible constitutive
relationships, are not considered.
So continuum mechanics is not a new topic. Rather it is a chance to start over and put all
that was studied previously under a single umbrella. A course in continuum mechanics
is a chance to synthesize what was learned during an undergraduate education into a
coherent structure, revealing the connections among seemingly distinct fields of study.
One of the challenges of doing this is developing a common notation, but no new physics
is presented. Continuum mechanics is just the process of confirming the foundation for
all that was done in undergraduate studies.

1.4 Notation

Building a theoretical foundation for the study of continuum mechanics creates notational
difficulties. This is especially true for the student just learning continuum mechanics. In
the pages that follow there are many different symbols used for all the quantities of inter-
est. There are more symbols than a student experiences as an undergraduate because
a general, nonlinear theory is being contructed. For instance, consider stress. There
are different measures of stress that are indistinguishable in the linear theory: Cauchy,
first Piola-Kirchhoff, and second Piola-Kirchhoff. In addition, von Mises and octahedral
stresses are defined to help analyze yield and failure theory. Finally, it is often advanta-
geous to subdivide stress into deviatoric and spherical parts because of the different role
the two have in deformed bodies.
With all these quantities it is hard to come up with unique symbols for each of them.
Often, one symbol is very similar to another symbol, and the context is necessary to fully
understand the meaning. This is one of the things that makes continuum mechanics
difficult for the beginner.
Finally, as students continue beyond this course, they encounter the disheartening real-
ity that not everybody uses the same notation. Often notational marks will have to be
made in margins when reading the literature. The reason there is not a single notation
comes from the fact that people were not flying in airplanes to technical conferences when
this material was being developed. For example, the governing laws set down by Newton
in the 1687 Principia could not be expressed in integral form by Euler in 1776 without the
intervening development of calculus – accomplished independently, and using distinct
notation, by both Newton and Leibniz. And the concept of stress, central to continuum
mechanics, was developed over a century by several thinkers working in various contexts,
from hydraulics to structural beams to the military applications that concerned Cauchy.
4 Continuum Mechanics for Engineers

Even within a single author’s works spanning a decade the notation can change. For
example, Table 1.1 1 shows some historical notation for stress.

TABLE 1.1
Historical notation for stress.
Naghdi, Eringen, Clebsch, Truesdell t11 t22 t33 t12 t23 t31
Cauchy (early work) ABCDEF
Cauchy (later work), St. Venant, Maxwell pxx pyy pzz pxy pyz pzx
F. Neumann, Kirchhoff, Love Xx Yy Zz Xy Yz Zx
Green and Zerna, Russian and German writers τ11 τ22 τ33 τ12 τ23 τ31
Karman, Timoshenko σx σy σz τxy τyz τzx
Some English and American writers σ11 σ22 σ33 σ12 σ23 σ31
σxx σyy σzz σxy σyz σzx

1 Adapted from Nonlinear Theory of Continuous Media, A. Cemal Eringen, McGraw-Hill Inc., (1962)
2
Essential Mathematics

2.1 Chapter Learning Outcomes


After mastering the material in this chapter, you will be able to:

• Articulate the definition of tensors

• Distinguish among scalars, vectors, and higher-order tensors

• Understand tensor algebraic identities and rules of behavior

• Apply the del operator, appreciate its mathematical and physical significance, and
use it in the integral theorems of Gauss and Stokes

• Use indicial and direct notation as well as matrix representation to communicate


tensor relationships and transformations

• Appreciate the utility of computational methods to perform tensor mathematics

Learning a discipline’s language is the first step a student takes toward becoming com-
petent in that discipline. The language of continuum mechanics is the algebra and calcu-
lus of tensors. Here, tensor is the generic name for those mathematical entities which are
used to represent the important physical quantities of continuum mechanics. A tensor is
a linear transformation that assigns any vector v to another vector T · v such that

T · (v + w) = T v + T w , (2.1a)

and
T · (αv) = αT · v (2.1b)
for all v and w. Furthermore, the sum of two tensors and the scalar multiple of a tensor
is defined by
(T + S) · v = T · v + S · v , (2.1c)
and
(αT ) · v = α (T · v) . (2.1d)
Because of these properties, tensors constitute a vector space.
Tensors have a most useful property in the way that they transform from one basis
(reference frame) to another. Having defined the tensor with respect to one reference
frame, it is possible to write the tensor quantity (components) in any reference frame.
For example, we can define stress in principal and non-principal components. Both are
representations of the same stress tensor even though the individual components may
be different. As long as the relationship between the reference frames is known, the
components with respect to one frame may be found from the other.

5
6 Continuum Mechanics for Engineers

Only that category of tensors known as Cartesian tensors is used in this text, and defini-
tions of these will be given in the pages that follow. General tensor notation is presented
in Appendix A for completeness, but it is not necessary for the main text. The tensor
equations used to develop the fundamental theory of continuum mechanics may be writ-
ten in either of two distinct notations; the symbolic notation, or the indicial notation. We shall
make use of both notations, employing whichever is more convenient for the derivation
or analysis at hand but taking care to establish the inter-relationships between the two.
We have made an effort to emphasize indicial notation in most of the text; an introductory
course must teach indicial notation to the student who may have little prior exposure to
the topic.

2.2 Scalars, Vectors and Cartesian Tensors


A considerable variety of physical and geometrical quantities have important roles in
continuum mechanics, and fortunately, each of these may be represented by some form
of tensor. For example, such quantities as density and temperature may be specified com-
pletely by giving their magnitude, i.e., by stating a numerical value. These quantities
are represented mathematically by scalars, which are referred to as zero-order tensors. It
should be emphasized that scalars are not constants, but may actually be functions of
position and/or time. Also, the exact numerical value of a scalar will depend upon the
units in which it is expressed. Thus, the temperature may be given by either 68 ◦ F, or 20 ◦ C
at a certain location. As a general rule, lower-case Greek letters in italic print such as α,
β, λ, etc. will be used as symbols for scalars in both the indicial and symbolic notations.
Several physical quantities of mechanics such as force and velocity require not only an
assignment of magnitude, but also a specification of direction for their complete charac-
terization. As a trivial example, a 20 N force acting vertically at a point is substantially
different than a 20 N force acting horizontally at the point. Quantities possessing such
directional properties are represented by vectors, which are first-order tensors. Geometri-
cally, vectors are generally displayed as arrows, having a definite length (the magnitude),
a specified orientation (the direction), and also a sense of action as indicated by the head
and the tail of the arrow, as seen in Figure 2.1. In this text arrow lengths are not to scale
with vector magnitude. Certain quantities in mechanics which are not truly vectors are
also portrayed by arrows, for example, finite rotations.
Consequently, in addition to the magnitude and direction characterization, the com-
plete definition of a vector requires the further statement: vectors add (and subtract) in
accordance with the triangle rule by which the arrow representing the vector sum of two
vectors extends from the tail of the first component arrow to the head of the second when
the component arrows are arranged ”head-to-tail,” as illustrated in Figure 2.1.
Although vectors are independent of any particular coordinate system, it is often useful
to define a vector in terms of its coordinate components, and in this respect it is necessary
to reference the vector to an appropriate set of axes. In view of our restriction to Cartesian
tensors, we limit ourselves to consideration of Cartesian coordinate systems for designat-
ing the components of a vector. Cartesian coordinates, named for Rene Descartes, may
bear the names (x, y, z) or (x1 , x2 , x3 ), and may also be defined by what they are not: they
are not cylindrical, spherical, or polar coordinates, but coordinate systems whose axes are
perpendicular lines.
Essential Mathematics 7

a
a+b b

-b

a a+(-b)

(a) Addition of vectors (b) Subtraction of vectors

FIGURE 2.1
Addition (a) and subtraction of vectors a and b independent of coordinate system.

A significant number of physical quantities having important status in continuum


mechanics require mathematical entities of higher order than vectors for their represen-
tation in the hierarchy of tensors. As we shall see, among the best known of these are the
stress and the strain tensors. These particular tensors are second-order tensors and are said
to have a rank of two. Third-order and fourth-order tensors are not uncommon in contin-
uum mechanics, but they are not nearly as plentiful as second-order tensors. Accordingly,
the unqualified use of the word tensor in this text will be interpreted to mean second-order
tensor. With only a few exceptions, primarily those representing the stress and strain
tensors, we shall denote second-order tensors by upper-case sans serif Latin letters in
bold-faced print, a typical example being the tensor T . The components of said tensor
will, in general, be denoted by lower-case Latin letters with appropriate indices: tij .
Tensors, like vectors, are independent of any coordinate system, but just as with vectors,
when we wish to specify a tensor by its components, we are obliged to refer to a suitable
set of reference axes. The precise definitions of tensors of various order will be given
subsequently in terms of the transformation properties of their components between two
related sets of Cartesian coordinate axes.
As a quick notation summary, the International Standards Organization (ISO) conven-
tions for typesetting mathematics are summarized below:
1. Scalar variables are written as italic letters. The letters may be either Roman or
Greek style fonts depending on the physical quantity they represent. The following
examples are a partial list of scalar notation:
(a) a – magnitude of acceleration
(b) v – magnitude of velocity
(c) r – radius
(d) θ – temperature or angle depending on context
(e) α – coefficient of thermal expansion
(f) σ – principal value of stress
(g) λ – eigenvalue or stretch
2. Vectors are written as boldface italic. Examples are as follows:
(a) x – position
8 Continuum Mechanics for Engineers

(b) v – velocity
(c) a – acceleration
(d) e
^1 – base vector in x1 direction

3. Second- and higher-order tensors are designated by uppercase fonts. Additionally,


matrices are shown in the calligraphic form to differentiate them from tensors. Ten-
sors can be represented by matrices of their components, but not all matrices are
tensors. In the case of several well known engineering quantities this case conven-
tion will not be accommodated. For example, linear strain has been chosen to be
represented by . Here are some samples of tensor and matrix symbols:

(a) Q – orthogonal matrix


(b) E – finite strain
(c) T – Cauchy stress tensor
(d)  – infinitesimal strain tensor
(e) R – rotation matrix

2.3 Tensor Algebra in Symbolic Notation - Summation Convention


The three-dimensional physical space of everyday life is the space in which many of the
events of continuum mechanics occur. Mathematically, this space is known as a Euclidean
three-space , and its geometry can be referenced to a system of Cartesian coordinate axes.
In some instances, higher order dimension spaces play integral roles in continuum topics.
Because a scalar has only a single component, it will have the same value in every system
of axes, but the components of vectors and tensors will have different component values,
in general, for each set of axes.
In order to represent vectors and tensors in component form we introduce in our physi-
cal space a right-handed system of rectangular Cartesian axes Ox1 x2 x3 , and identify with
these axes the triad of unit base vectors, e
^1 , e
^2 , e
^3 , shown in Fig. 2.2(a). All unit vectors
in this text will be written with a caret placed above the bold-faced symbol. Due to the
mutual perpendicularity of these base vectors they form an orthogonal basis, and further-
more, because they are unit vectors, the basis is said to be orthonormal. In terms of this
basis an arbitrary vector v is given in component form by

X
3
^1 + v2 e
v = v1 e ^2 + v3 e
^3 = ^i .
vi e (2.2)
i=1

This vector and its coordinate components are pictured in Fig. 2.2(b). For the symbolic
description, vectors will usually be given by lower-case Latin letters in bold-faced print,
with the vector magnitude denoted by the same letter. Thus v is the magnitude of v.
A notational device called the summation convention will greatly simplify the writing of
the equations of continuum mechanics. Stated briefly, we agree that whenever a subscript
appears exactly twice in a given term, that subscript will take on the values 1, 2, 3 suc-
cessively, and the resulting terms summed. For example, using this scheme, we may now
write Eq 2.2 in the simple form
^i ,
v = vi e (2.3)
Essential Mathematics 9

x3 x3

v3
^3
e v

x2
O
^2
e v2
O
^1
e v1 x2
x1
x1
(a) Unit vectors in the coordinate directions x1 , (b) Rectangular Cartesian components of the
x2 and x3 . vector v.

FIGURE 2.2
Base vectors and components of a Cartesian vector.

P
and delete the summation symbol . For Cartesian tensors, only subscripts are required
on the components; for general non-orthogonal bases both subscripts and superscripts are
used (see the discussion in Appendix A). The summed subscripts are called dummy indices
since it is immaterial which particular letter is used. Thus vi e
^i is equivalent to vj e
^j , or to
^k , when the summation convention is used. A word of caution, however; no subscript
vk e
may appear more than twice in a singe term. But as we shall soon see, more than one
pair of dummy indices may appear in a given expression that is the summation of terms
(see Example 2.2). Note also that the summation convention may involve subscripts from
both the unit vectors and the scalar coefficients.

Example 2.1

Without regard for meaning as far as mechanics is concerned, expand the


following expressions according to the summation convention:

(a) ui vi wj e
^j (b) tij vj e
^j (c) tii vj e
^j

Solution
(a) Summing first on i, and then on j

^j = (u1 v1 + u2 v2 + u3 v3 )(w1 e
ui vi wj e ^1 + w2 e
^2 + w3 e
^3 )

(b) Summing on i, then on j and collecting terms on the unit vectors.

^i
tij vj e ^1 + t2j vj e
= t1j vj e ^2 + t3j vj e
^3
= (t11 v1 + t12 v2 + t13 v3 ) e^1 + (t21 v1 + t22 v2 + t23 v3 ) e
^2
+ (t31 v1 + t32 v2 + t33 v3 ) e ^3
10 Continuum Mechanics for Engineers

(c) Summing on i, then on j,

^j = (t11 + t22 + t33 ) (v1 e


tii vj e ^1 + v2 e
^2 + v3 e
^3 )

Note the similarity between (a) and (c). Also note the same results would be
obtained by summing first on j then on i.

With the above background in place it is now possible, using symbolic notation, to
present many useful definitions from vector/tensor algebra. There are two symbols
needed prior to writing out all of the vector and tensor algebra. These two symbols are the
Kronecker delta and the permutation symbol. Additionally, there are several useful rela-
tionships between the Kronecker delta and permutation symbol that are used throughout
continuum mechanics. The following three subsections introduce the Kronecker delta,
the permutation symbol, and their relationships. Following that, vector/tensor algebra is
presented.
The Kronecker delta is similar to the identity matrix, so the reader should quickly
embrace this new but familiar entity. On the other hand, the permutation symbol is a
little more abstract than the Kronecker delta since it cannot be represented by a matrix.
In subsequent chapters the Kronecker delta and the permutation symbol play integral
roles in describing how forces are carried by continuum bodies and how the position of
a particle is described.

2.3.1 Kronecker Delta


Since the base vectors e
^i (i = 1,2,3) are unit vectors and orthogonal

1 if numerical value of i = numerical value of j
^i · e
e ^j = .
0 if numerical value of i 6= numerical value of j

Therefore, if we introduce the Kronecker delta defined by



1 if numerical value of i = numerical value of j
δij =
0 if numerical value of i 6= numerical value of j
we see that

^i · e
e ^j = δij (i, j = 1, 2, 3) . (2.4)
Also, note that by the summation convention

δii = δjj = δ11 + δ22 + δ33 = 1 + 1 + 1 = 3 ,


and furthermore, we call attention to the substitution property of the Kronecker delta by
expanding (summing on j) the expression

^j = δi1 e
δij e ^1 + δi2 e
^2 + δi3 e
^3 .
But for a given value of i in this equation, only one of the Kronecker deltas on the right-
hand side is nonzero, and it has the value one. Therefore,
Essential Mathematics 11

^j = e
δij e ^i ,

and the Kronecker delta in δij e^j causes the summed subscript j of e ^j to be replaced by
i reducing the expression to simply ei . This is often called the ”delta substitution” rule
^
for the Kronecker delta. In fact, the Kronecker delta is sometimes called the ”substitution
operator” because for any vector v,

vi δij = vj .

2.3.2 Permutation Symbol


By introducing the permutation symbol εijk defined by



 1 if numerical values of ijk appear as in the sequence 123123123
εijk = −1 if numerical values of ijk appear as in the sequence 321321321


0 if numerical values of ijk appear in any other sequence
(2.5)
we may express the cross products of the base vectors (i=1,2,3) by the use of Eq 2.5 as

^i × e
e ^j = εijk e
^k (i, j, k = 1, 2, 3) . (2.6)

Also, note from its definition that the interchange of any two subscripts in εijk causes a
sign change so that for example,

εijk = −εkji = εkij = −εikj ,

and, furthermore, that for repeated subscripts is zero as in

ε113 = ε212 = ε133 = ε222 = 0 .

2.3.3 ε - δ Identity
The product of permutation symbols εmiq εjkq may be expressed in terms of Kronecker
deltas by the ε - δ identity

εmiq εjkq = δmj δik − δmk δij (2.7a)

as may be proven by direct expansion (see Problem 2.17). This is a most important for-
mula used throughout this text and is well worth memorizing. Also, by the sign-change
property of εijk ,
εmiq εjkq = εmiq εqjk = εqmi εqjk = εqmi εjkq .

Additionally, by setting i = k in Eq 2.7a, it is easy to show that

εjkq εmkq = 2δjm , (2.7b)

and then setting j = m gives


εjkq εjkq = 6 . (2.7c)
12 Continuum Mechanics for Engineers

2.3.4 Tensor/Vector Algebra


Vector addition w = u + v is easily written in indicial form as

^i = (ui + vi ) e
wi e ^i , (2.8)
where the components simply add together.
Vector multiplication can take one of several forms. The specific form depends on the
type of entity multiplying the vector. For now, two forms of vector multiplication can be
defined in symbolic form. Multiplication of a vector by a scalar is written as

^i ,
λv = λvi e (2.9)
and the dot (scalar) product between of two vectors is

u · v = v · u = uv cos θ (2.10)
where θ is the angle between the two vectors when drawn from a common origin.
From the definition of δij and its substitution property the dot product u · v may be
written as

u · v = ui e
^i · vj e ^i · e
^j = ui vj e ^j = ui vj δij = ui vi . (2.11)
Note that scalar components pass through the dot product since it is a vector operator.
The vector cross (vector) product of two vectors is defined by

u × v = −v × u = (uv sin θ) e
^
where θ, is the angle between the two vectors when drawn from a common origin, and
where e^ is a unit vector perpendicular to their plane such that a right-handed rotation
about e
^ through the angle θ carries u into v .
The vector cross product may be written in terms of the permutation symbol (Eq 2.5) as
follows:
u × v = ui e
^i × vj e ei × e
^j = ui vj (^ ^j ) = εijk ui vj e
^k . (2.12)
Again, notice how the scalar components pass through the vector cross product operator.
There are a couple of useful ways three vectors can be multiplied. The triple scalar
product (box product) is

u · (v × w) = (u × v) · w = [u, v, w] ,

or

[u, v, w] ^i · (vj e
= ui e ^j × wk e
^k ) = ui e^i · εjkq vj wk e
^q , (2.13)
= εjkq ui vj wk δiq = εijk ui vj wk

where in the final step we have used both the substitution property of δiq and the sign-
change property of εijk . The triple cross product is similar to the triple scalar product

u × (v × w) ^i × (vj e
= ui e ^j × wk e ^i × (εjkq vj wk e
^k ) = ui e ^q ) (2.14)
^m = εmiq εjkq ui vj wk e
= εiqm εjkq ui vj wk e ^m

which, by Eq 2.7a, may be written as

u × (v × w) ^m
= (δmj δik − δmk δij ) ui vj wk e (2.15)
= (ui vm wi − ui vi wm ) e^m = ui wi vm e
^m − ui vi wm e
^m .
Essential Mathematics 13

Observation of the indices in Eq 2.15 admits

u × (v × w) = (u · w) v − (u · v) w
known as Lagrange’s identity.
In addition to the common vector products above, two vectors can be multiplied together
to yield a tensor. The tensor product of two vectors creates a dyad

^i vj e
uv = ui e ^j = ui vj e
^i e
^j (2.16)

which in expanded form, summing first on i, yields

^i e
u i vj e ^j = u1 vj e
^1 e
^j + u2 vj e
^2 e
^j + u3 vj e
^3 e
^j ,

and then summing on j

^i e
ui vj e ^j = u 1 v1 e^1 e
^1 + u1 v2 e ^1 e
^2 + u1 v3 e ^1 e
^3
^ ^ ^ ^ ^ ^
+ u 2 v 1 e2 e1 + u 2 v 2 e2 e2 + u 2 v 3 e2 e3 (2.17)
+ u 3 v1 e^3 e
^1 + u3 v2 e ^3 e
^2 + u3 v3 e ^3 e
^3 .

This nine-term sum is called the nonion form of the dyad, uv. A sum of dyads such as

u1 v1 + u2 v2 + · · · + uN vN (2.18)

is called a dyadic.
A common, alternative notation frequently used for the dyad product is

a ⊗ b = ai e
^i ⊗ bj e ^i ⊗ e
^j = ai bj e ^j (2.19)

and when this notation is used, the term tensor product is preferred over dyad, despite
their equivalence. The tensor product of vectors a and b is defined by how a ⊗ b maps
all vectors u:
(a ⊗ b) u = a (b · u) . (2.20)
If one takes vectors a, b and u to be vectors from a Euclidian 3-space, the expanded form
for the tensor product may be written as
 
a1
(a ⊗ b) u = a (b · u) = a2  (b1 u1 + b2 u2 + b3 u3 )
 

a3
 
a1 b1 u1 + a1 b2 u2 + a1 b3 u3
= a2 b1 u1 + a2 b2 u2 + a2 b3 u3 
 

a3 b1 u1 + a3 b2 u2 + a3 b3 u3
   (2.21)
a1 b1 a1 b2 a1 b3 u1
= a2 b1 a2 b2 a2 b3  u2 
  

a3 b1 a3 b2 a3 b3 u3
   
a1 h i u1
= a2  b1 b2 b3  u2  .
   

a3 u3
14 Continuum Mechanics for Engineers

The last line of Eq 2.21 shows the matrix representaion of the tensor product which is
sometimes called the outer product. Contrast this matrix outer product form to the inner
product given by
 
h i b1
a · b = a1 a2 a3 b2  = a1 b1 + a2 b2 + a3 b3 .
 

b3

The equivalence of the tensor product and the dyad product for two vectors is demon-
strated by the second line of Eq 2.21. Both notations will be used in the book.
A dyad can be multiplied by a vector giving a vector-dyad product:

1. u · (vw) = ui e
^i · (vj e
^j wk e
^k ) = ui vi wk e
^k (2.22)

2. (uv) · w = (ui e ^j ) · wk e
^i vj e ^k = ui vj wj e
^i (2.23)

3. u × (vw) = (ui e
^i × vj e
^j ) wk e
^k = εijq ui vj wk e
^q e
^k (2.24)

4. (uv) × w = ui e ^j × wk e
^i (vj e ^k ) = εjkq ui vj wk e
^i e
^q (2.25)
(Note that in products 3 and 4 the order of the base vectors e
^i is important.)
Dyads can be multiplied by each other to yield another dyad

(uv) · (ws) = ui e ^j · wk e
^i (vj e ^k ) sq e
^q = ui vj wj sq e
^i e
^q . (2.26)
Vectors can be multiplied by a tensor to give a vector. The reduction in the order of the
tensor is why the “dot” is used in symbolic notation.

1. v · T = vi e
^i · tjk e
^j e
^k = vi tjk δij e
^k = vi tik e
^k (2.27)

2. T · v = tij e ^j · vk e
^i e ^k = tij e
^i δjk vk = tij vj e
^i (2.28)
(Note that these products are sometimes written, less precisely, as vT and T v.)
Finally, two tensors can be multiplied resulting in a tensor

T · S = tij e ^j · spq e
^i e ^p e
^q = tij spq δjp e
^i e
^q = tij sjq e
^i e
^q . (2.29)

Example 2.2

Let the vector v be given by v = (a · n


^) n^ +n ^ × (a × n^ ) where a is an arbitrary
^ is a unit vector. Express v in terms of the base vectors e
vector, and n ^i , expand,
and simplify. (Note that n ^·n ^i · nj e
^ = ni e ^j = ni nj δij = ni ni = 1.)

Solution
^i , the given vector v is expressed by the equation
In terms of the base vectors e

^i · nj e
v = (ai e ^j ) nk e ^i × (aj e
^k + ni e ^j × nk e
^k ) .

We note here that indices i, j, and k appear four times in this line; however, the
summation convention has not been violated. Terms that are separated by a plus
Essential Mathematics 15

or a minus sign are considered different terms each having summation convention
rules applicable within them. Vectors joined by a dot or cross product are not
distinct terms, and the summation convention must be adhered to in that case.
Carrying out the indicated multiplications, we see that

v = (ai nj δij ) nk e ^i × (εjkq aj nk e


^k + ni e ^q )
= ^k + εiqm εjkq ni aj nk e
ai ni nk e ^m
= ^
ai ni nk ek + εmiq εjkq ni aj nk e^m
= ^k + (δmj δik − δmk δij ) ni aj nk e
ai ni nk e ^m
= ^k + ni aj ni e
ai ni nk e ^j − ni ai nk e
^k
= ^j = aj e
ni ni aj e ^j = a .

Since a must equal v, this example demonstrates that the vector v may be
resolved into a component (v · n ^) n
^ in the direction of n ^ , and a component
^ × (v × n
n ^ ) perpendicular to n
^ . This result is the projection theorem for vectors.

Example 2.3

Using Eq 2.7 show that (a) εmkq εjkq = 2δmj and that (b) εjkq εjkq = 6. (Recall
that δkk = 3 and δmk δkj = δmj .)

Solution
(a) Write out Eq 2.7a with index i replaced by k and use the Kronecker substi-
tution property to get

εmkq εjkq = δmj δkk − δmk δkj


= 3δmj − δmj = 2δmj .

(b) Start with the first equation in Part (a) and replace the index m with j,
giving

εjkq εjkq = δjj δkk − δjk δjk


= (3) (3) − δjj = 9 − 3 = 6 .

Example 2.4

Double dot products of dyads are defined by

(a) (uv) · · (ws) = (v · w) (u · s)

(b) (uv) : (ws) = (u · w) (v · s) .


16 Continuum Mechanics for Engineers

Expand these products and compare the component forms.

Solution

(a) (uv) · · (ws) = (vi e


^i · wj e ^k · sq e
^j ) (uk e ^q ) = vi wi uk sk

(b) (uv) : (ws) = (ui e


^i · wj e ^k · sq e
^j ) (vk e ^q ) = ui wi vk sk

Summary of Symbolic Notations

1. Addition of vectors, Eq 2.8:

w=u+v or ^i = (ui + vi ) e
wi e ^i

2. Multiplication:

(a) of a vector by a scalar, Eq 2.9:


^i
λv = λvi e

(b) dot (scalar) product of two vectors, Eq 2.11:

u · v = v · u = uv cos θ = ui vi

(c) cross (vector) product of two vectors, Eq 2.12:

u × v = −v × u = (uv sin θ) e ^k
^ = εijk ui vj e

(d) triple scalar product (box product), Eq 2.13:

[u, v, w] ^i · (vj e
= ui e ^j × wk e
^k ) = ui e^i · εjkq vj wk e
^q
= εjkq ui vj wk δiq = εijk ui vj wk

(e) triple cross product, Eq 2.14:

u × (v × w) ^i × (vj e
= ui e ^j × wk e ^i × (εjkq vj wk e
^k ) = ui e ^q )
^m = εmiq εjkq ui vj wk e
= εiqm εjkq ui vj wk e ^m

(f ) tensor product of two vectors (dyad), Eq 2.16:

^i vj e
uv = ui e ^j = ui vj e
^i e ^i ⊗ vj e
^j = ui e ^i ⊗ e
^j = ui vj e ^j
Essential Mathematics 17

(g) Vector-dyad products, Eqs 2.22–2.25:

1. u · (vw) = ui e
^i · (vj e
^j wk e
^k ) = ui vi wk e
^k

2. (uv) · w = (ui e ^j ) · wk e
^i vj e ^k = ui vj wj e
^i

3. u × (vw) = (ui e
^i × vj e
^j ) wk e
^k = εijq ui vj wk e
^q e
^k

4. (uv) × w = ui e ^j × wk e
^i (vj e ^k ) = εjkq ui vj wk e
^i e
^q

(h) dyad-dyad product, Eq 2.26:


(uv) · (ws) = ui e ^j · wk e
^i (vj e ^k ) sq e
^q = ui vj wj sq e
^i e
^q

(i) vector-tensor products, Eqs 2.27–2.28:

1. v · T = vi e
^i · tjk e
^j e
^k = vi tjk δij e
^k = vi tik e
^k

2. T · v = tij e ^j · vk e
^i e ^k = tij e
^i δjk vk = tij vj e
^i

(j) tensor-tensor product, Eq 2.29:


T · S = tij e ^j · spq e
^i e ^p e
^q = tij sjp e
^i e
^q

(k) double dot product:

1. (uv) · · (ws) = (vi e


^i · wj e ^k · sq e
^j ) (uk e ^q ) = vi wi uk sk

2. (uv) : (ws) = (ui e


^i · wj e ^k · sq e
^j ) (vk e ^q ) = ui wi vk sk
3. A · ·B = Aij Bji
4. A : B = Aij Bij

2.4 Indicial Notation


By assigning special meaning to the subscripts, indicial notation permits us to carry out
the tensor operations of addition, multiplication, differentiation, etc., without the use, or
even the appearance of the base vectors e ^i in the equations. We simply agree that the
tensor rank (order) of a term is indicated by the number of “free,” that is, unrepeated,
subscripts appearing in that term. Accordingly, a term with no free indices represents a
scalar, a term with one free index a vector, a term having two free indices a second order
tensor, and so on. The specific meaning of these symbols are given in Table 2.1.
For tensors defined in a three-dimensional space, the free indices take on the values
1,2,3 successively, and we say that these indices have a range of three. If N is the number
of free indices in a tensor, that tensor has 3N components in three-dimensional space.
18 Continuum Mechanics for Engineers

TABLE 2.1
Indicial form for a variety of tensor quantities
λ = scalar (zeroth-order tensor) λ
vi = vector (first-order tensor) v, or equivalently, its 3 components
ui vj = dyad (second-order tensor) uv, or its 9 components
tij = dyadic (second-order tensor) T , or its 9 components
Qijk = triadic (third-order tensor) Q, or its 27 components
Cijkm = tetradic (forth-order tensor) C, or its 81 components

In indicial notation exactly two types of subscripts appear:

1. “free” indices, which are represented by letters that occur only once in a given term,

2. “summed,” or “dummy” indices which are represented by letters that appear only
twice in a given term.

Furthermore, every term in a valid equation must have the same letter subscripts for the
free indices. No letter subscript may appear more than twice in any given term.
Mathematical operations among tensors are readily carried out using the indicial nota-
tion. Thus addition (and subtraction) among tensors of equal rank follows according to
the typical equations; ui +vi −wi = si for vectors, and tij −vij +sij = qij for second-order
tensors. Multiplication of two tensors to produce an outer tensor product is accomplished
by simply setting down the tensor symbols side by side with no dummy indices appear-
ing in the expression. As a typical example, the outer product of the vector vi and tensor
tjk is the third-order tensor vi tjk . Contraction is the process of setting equal to one another
any two indices of a tensor term. The result of contracting an outer tensor product by one
or more contractions involving indices from separate tensors is still a tensor. We note that
the rank of a given tensor is reduced by two for each contraction. Some outer products,
when contracted, form well-known products as listed in Table 2.2.

TABLE 2.2
Forms for tensor products.
Outer Products: Contraction(s): Contracted Results:
ui vj i=j ui vi (vector dot product)
εijk uq vm j = q, k = m εijk uj vk (vector cross product)
εijk uq vm wn i = q, j = m, k = n εijk ui vj wk (box product)
Aij Brs i = r, j = s Aij Bij (tensor inner product)
Aij Brs j=r Aij Bjs (tensor multiplication)

A tensor is symmetric in any two indices if interchange of those indices leaves the tensor
value unchanged. For example, if sij = sji and cijm = cjim , both of these tensors are said
to be symmetric in the indices i and j. A tensor is anti-symmetric (or skew-symmetric) in
Essential Mathematics 19

any two indices if interchange of those indices causes a sign change in the value of the
tensor. Thus if aij = −aji , it is anti-symmetric in i and j. Also, recall that by definition,
εijk = −εjik = εjki , etc., and hence the permutation symbol is anti-symmetric in all
indices.

Example 2.5

Show that the inner product sij aij of a symmetric tensor sij = sji , and an
anti-symmetric tensor aij = −aji is zero.

Solution
By definition of symmetric tensor sij and skew-symmetric tensor aij , we have

sij aij = −sji aji = −smn amn = −sij aij


where the last two steps are the result of all indices being dummy indices.
Therefore, 2sij aij = 0 , or sij aij = 0.

One of the most important advantages of the indicial notation is the compactness it
provides in expressing equations in three dimensions. A brief listing of typical equations
of continuum mechanics is presented below to illustrate this feature.

1. φ = sij tij − sii tjj (1 equation, 18 terms on RHS)


2. ti = qij nj (3 equations, 3 terms on RHS of each)
3. tij = λδij Ekk + 2µEij (9 equations, 4 terms on RHS of each)

Example 2.6

By direct expansion of the expression vi = εijk wjk determine the components


of the vector vi in terms of the components of the tensor wjk .

Solution
By summing first on j and then on k and then omitting the zero terms, we find
that

vi = εi1k w1k + εi2k w2k + εi3k w3k


= εi12 w12 + εi13 w13 + εi21 w21 + εi23 w23 + εi31 w31 + εi32 w32 .

Therefore,
v1 = ε123 w23 + ε132 w32 = w23 − w32 ,
v2 = ε213 w13 + ε231 w31 = w31 − w13 ,
v3 = ε312 w12 + ε321 w21 = w12 − w21 .
Note that if the tensor wjk were symmetric, the vector vi would be a null (zero)
vector.
20 Continuum Mechanics for Engineers

To end this section, consider a skew–symmetric second order tensor W = wij e ^j .


^i e
All skew–symmetric tensors can be represented in terms of an axial vector by using the
permutation symbol. Let the axial vector for wij be ωi defined by

1
ωi = − εijk wjk . (2.30)
2
It is a straightforward exercise in indicial manipulation to find the inverse of Eq 2.30. We
can write wjk in terms of ωi as follows:

εimn ωi = − 21 εimn εijk wjk


= − 12 (δmj δnk − δmk δnj ) wjk
= − 12 (wmn − wnm ) (2.31)
= − 12 (2wmn )
= −wmn

where Eq 2.7a and the skew–symmetric property of wij were used. Note that W · u =
ω × u for any vector u. In this case, ω is called the axial vector of the tensor W.

2.5 Matrices and Determinants


For computational purposes it is often expedient to use the matrix representation of vectors
and tensors. Accordingly, we review here several definitions and operations of elementary
matrix theory.
A matrix is an ordered rectangular array of elements enclosed by square brackets and
subjected to certain operational rules. The typical element Aij of the matrix is located in
the ith (horizontal) row, and in the jth (vertical) column of the array. A matrix having
elements Aij , which may be numbers, variables, functions, or any of several mathematical
entities, is designated by [Aij ], or symbolically by the kernel letter A. An M by N matrix
(written M × N) has M rows and N columns, and may be displayed as

A11 A12 A1N


 
···
A21 A22 ··· A2N
 
A = [Aij ] =  (2.32)
 
.. .. ..  .
. . .
 
 
AM1 AM2 ··· AMN

If M = N, the matrix is a square matrix. A 1 × N matrix [A1N ] is a row matrix, and an M × 1


matrix [AM1 ] is a column matrix. Row and column matrices represent vectors, whereas a
3 × 3 square matrix represents a second-order tensor. A scalar is represented by a 1 × 1
matrix (a single element). The unqualified use of the word matrix in this text is understood
to mean a 3 × 3 square matrix, that is, the matrix representation of a second-order tensor
or a matrix that is not a tensor.
A zero, or null matrix has all elements equal to zero. A diagonal matrix is a square matrix
whose elements not on the principal diagonal, which extends from A11 to ANN , are all
zeros. Thus for a diagonal matrix, Aij = 0 for i 6= j. The unit or identity matrix I, which,
Essential Mathematics 21

incidentally, is the matrix representation of the Kronecker delta, is a diagonal matrix


whose diagonal elements all have the value one: Iij = δij .
The N × M matrix formed by interchanging the rows and columns of the M × N matrix
A is called the transpose of A, and is written as AT , or [Aij ]T . By definition, the elements
of a matrix A and its transpose are related by the equation ATij = Aji . A square matrix
for which A = AT , or in element form, Aij = Aji , is called a symmetric matrix; one
for which AT = −A , or, Aji = −Aij , is called an anti-symmetric, or skew-symmetric
matrix. The elements of the principal diagonal of a skew-symmetric matrix are all zeros.
Two matrices are equal if they are identical element by element. Matrices having the same
number of rows and columns may be added (or subtracted) element by element. Thus if
A = B + C, the elements of A are given by

Aij = Bij + Cij . (2.33)


Addition of matrices is commutative, A + B = B + A, and associative, A + (B + C) =
(A + B) + C .

Example 2.7

Show that the square matrix A can be expressed as the sum of a symmetric
and a skew-symmetric matrix by the decomposition

A + AT A − AT
A= + .
2 2

Solution
Let the decomposition be written as A = B + C where B = 1
A + AT and

2
C = 12 A − AT . Then writing B and C in element form,


Aij + ATij Aij + Aji ATji + Aji


Bij = = = = Bji = BTij (symmetric) ,
2 2 2
Aij − ATij Aij − Aji ATji − Aji
Cij = = =− = −Cji = −CTij (skew-symmetric) .
2 2 2
Therefore B is symmetric, and C skew-symmetric.

Multiplication of the matrix A by the scalar λ results in the matrix λA, or [λAij ] =
λ [Aij ]. The product of two matrices A and B, denoted by AB, is defined only if the
matrices are conformable, that is, if the prefactor matrix A has the same number of columns
as the postfactor matrix B has rows. Thus the product of an M × Q matrix multiplied by
a Q × N matrix is an M × N matrix. The product matrix C = AB has elements given by

Cij = Aik Bkj (2.34)

in which k is, of course, a summed index. Therefore each element Cij of the product
matrix is an inner product of the ith row of the prefactor matrix with the jth column of
the postfactor matrix. In general, matrix multiplication is not commutative, AB 6= BA,
but the associative and distributive laws of multiplication do hold for matrices. The
22 Continuum Mechanics for Engineers

product of a matrix with itself is the square of the matrix, and is written AA = A2 .
Likewise the cube of the matrix is AAA = A3 , and in general, matrix products obey the
exponent rule

Am An = An Am = Am+n (2.35)
where m and n are positive integers, or zero. Also, we note that
n
(An )T = AT , (2.36)
and if BB = A then
√ 1
B= A = A2 (2.37)
but the square root is not unique.

Example 2.8

Use indicial notation to show that for arbitrary matrices A and B:


(a) (A + B)T = AT + BT
(b) (AB)T = BT AT
(c) IB = BI = B where I is the identity matrix with components Iij = δij .
Solution
(a) Let A + B = C, then in element form Cij = Aij + Bij and therefore CT is
given by
CTij = Cji = Aji + Bji = ATij + BTji ,
or
CT = (A + B)T = AT + BT .
(b) Let AB = C, then in element form

Cij = Aik Bkj = ATki BTjk = BTjk ATki = CTji .

Hence (AB)T = BT AT . Note that exchanging the order ATki BTjk = BTjk ATki
is not necessary when using indicial notation. It is done for clarity. In direct
or matrix notation the order of the terms is critical.
(c) Let IB = C, then in element form

Cij = δik Bkj = Bij = Bik δkj

by the substitution property. Thus IB = BI = B.


Essential Mathematics 23

The determinant of the matrix A is a scalar designated by either det A, or by |Aij |, and
for a 3 × 3 matrix A,

A11 A12 A13


det A = |Aij | = A21 A22 A23 . (2.38)
A31 A32 A33

A minor of det A is another determinant |Mij | formed by deleting the ith row and jth
column of |Aij |. The cofactor of the element Aij (sometimes referred to as the signed minor)
is defined by
(c)
Aij = (−1)i+j |Mij | (2.39)
where superscript (c) denotes cofactor of matrix A.
Evaluation of a determinant may be carried out by a standard method called expansion
by cofactors. In this method, any row (or column) of the determinant is chosen, and each
element in that row (or column) is multiplied by its cofactor. The sum of these products
gives the value of the determinant. For example, expansion of the determinant of Eq 2.38
by the first row becomes

A22 A23 A21 A23 A21 A22


det A = A11 − A12 + A13 (2.40)
A32 A33 A31 A33 A31 A32
which upon complete expansion gives

det A = A11 (A22 A33 − A23 A32 ) − A12 (A21 A33 − A23 A31 ) (2.41)
+A13 (A21 A32 − A22 A31 ) .

Several interesting properties of determinants are worth mentioning at this point. To


begin with, the interchange of any two rows (or columns) of a determinant causes a sign
change in its value. Because of this property and because of the sign-change property of
the permutation symbol, the det A of Eq 2.38 may be expressed in the indicial notation
by the alternative forms (see Prob. 2.15)

det A = εijk Ai1 Aj2 Ak3 = εijk A1i A2j A3k . (2.42)
Furthermore, following an arbitrary number of column interchanges with the accompa-
nying sign change of the determinant for each, it can be shown from the first form of
Eq 2.42 that (see Prob 2.16):

εqmn det A = εijk Aiq Ajm Akn . (2.43)


Finally, we note that if the det A = 0, the matrix is said to be singular. It may be eas-
ily shown that every 3 × 3 skew-symmetric matrix is singular. Also, the determinant
of the diagonal matrix, D, is simply the product of its diagonal elements: det D =
D11 D22 · · · DNN .
The determinant can also be defined using the triple scalar product or box product.
This notation is useful in kinematics and in the definition of volume changes. For three
arbitrary, non-coplanar vectors u, v, and w, the determinant is found from (see Problem
2.39)
[Au, Av, Aw] = det A [u, v, w] ,
24 Continuum Mechanics for Engineers

or
[Au, Av, Aw]
det A = . (2.44)
[u, v, w]

Example 2.9

Show that for matrices A and B, det(AB) = det(BA) = det(A) det(B).

Solution
Let C = AB, then Cij = Aik Bkj and from Eq 2.42

det C = εijk Ci1 Cj2 Ck3


= εijk Aiq Bq1 Ajm Bm2 Akn Bn3
= εijk Aiq Ajm Akn Bq1 Bm2 Bn3 .

But from Eq 2.43

εijk Aiq Ajm Akn = εqmn det A ,


so now

det C = det AB = εqmn Bq1 Bm2 Bn3 det A = det A det B .


By a direct interchange of A and B, det AB = det BA.

Example 2.10

Use Eq 2.40 and Eq 2.41 to show that det A = det AT .

Solution
Since

A11 A21 A31


AT = A12 A22 A32
A13 A23 A33
cofactor expansion by the first column here yields

A22 A32 A21 A31 A21 A31


det A = A11 − A12 + A13
A23 A33 A23 A33 A22 A32

which is equal to Eq 2.41.


Essential Mathematics 25

The inverse of the matrix A is written A−1 , and is defined by

AA−1 = A−1 A = I (2.45)


where I is the identity matrix. Thus if AB = I, then B = A , and A = B
−1 −1
. The
adjugate matrix A∗ is defined as the transpose of the cofactor matrix
h iT
A∗ = A(c) . (2.46)

In terms of the adjugate matrix the inverse matrix is expressed by

A∗
A−1 = (2.47)
det A
which is actually a working formula by which an inverse matrix may be calculated. This
formula shows that the inverse matrix exists only if det A 6= 0, i.e., only if the matrix A is
non-singular. For example, a 3 × 3 skew-symmetric matrix has no inverse.

Example 2.11

Show from the definition of the inverse, Eq 2.45 that


(a) (AB)−1 = B−1 A−1 ,
−1 T
(b) AT = A−1 .

Solution
(a) By pre-multiplying the matrix product AB by B−1 A−1 , we have
(using Eq 2.45),

B−1 A−1 AB = B−1 IB = B−1 B = I


and therefore B−1 A−1 = (AB)−1 .
(b) Taking the transpose of both sides of Eq 2.45 and using the result of
Example 2.8 we have
T T
AA−1 AT = IT = I .
= A−1
T T −1
Hence A−1 must be the inverse of AT , or A−1 = AT . It is
T −1
common to write A −T
for A

.

An orthogonal matrix, call it Q, is a square matrix for which Q−1 = QT . From this
definition we note that a symmetric orthogonal matrix is its own inverse since in this case

R−1 = RT = R . (2.48)
Also, if A and B are orthogonal matrices.

(AB)−1 = B−1 A−1 = BT AT = (AB)T , (2.49)


26 Continuum Mechanics for Engineers

so that the product matrix is likewise orthogonal. Furthermore, if A is orthogonal it may


shown (see Prob. 2.20) that

det A = ±1 . (2.50)
As mentioned near the beginning of this section, a vector may be represented by a row
or column matrix, a second-order tensor by a square 3 × 3 matrix. For computational
purposes, it is frequently advantageous to transcribe vector/tensor equations into their
matrix form. As a very simple example, the vector-tensor product, u = T · v (symbolic
notation) or ui = Tij vj (indicial notation) appears in matrix form as
    
u1 T11 T12 T13 v1
[ui1 ] = [Tij ] [vj1 ] or  u2  =  T21 T22 T23   v2  . (2.51)
    

u3 T31 T32 T33 v3


In much the same way the product
w=v·T or wi = vj Tji
appears as
 
h i h i T11 T21 T31
[w1i ] = [v1j ] [Tji ] or w1 w2 w3 = v1 v2 v3  T12 T22 T32  . (2.52)
 

T13 T23 T33

Note that in Eq 2.52 the subscripts are exchanged since the transpose of the matrix is
being written.

2.6 Transformations of Cartesian Tensors


Although — as already mentioned — vectors and tensors have an identity independent
of any particular reference or coordinate system, the relative values of their respective
components do depend upon the specific axes to which they are referred. The relation-
ships among these various components when given with respect to two separate sets
of coordinate axes are known as the transformation equations. In developing these trans-
formation equations for Cartesian tensors, we consider two sets of rectangular Carte-
sian axes, Ox1 x2 x3 and Ox10 x20 x30 , sharing a common origin, and oriented relative to one
another so that the direction cosines between the primed and unprimed axes are given by
aij = cos(xi0 , xj ) as shown in Fig. 2.3.
The square array of the nine direction cosines displayed in Table 2.3 is useful in relat-
ing the unit base vectors e ^i and e ^i0 to one another, as well as relating the primed and
unprimed coordinates xi and xi of a point. Thus the primed base vectors e
0 ^i0 are given
in terms of the unprimed vectors e ^i by the equations (as is also easily verified from the
geometry of the vectors in the diagram of Fig. 2.3),

^10 = a11 e
e ^1 + a12 e
^2 + a13 e
^3 = a1j e
^j , (2.53a)
^20 = a21 e
e ^1 + a22 e
^2 + a23 e
^3 = a2j e
^j , (2.53b)
e 0
^3 = a31 e^1 + a32 e
^2 + a33 e
^3 = a3j e
^j , (2.53c)
Essential Mathematics 27

x3 x2

x3 cos−1 a13


^3
e ^2
e x1
^3
e ^1
e
cos−1 a12
O ^2
e x2
^1
e
cos−1 a11

x1

FIGURE 2.3
Rectangular coordinate system Ox10 x20 x30 relative to Ox1 x2 x3 . Direction cosines shown for
coordinate x10 relative to unprimed coordinates. Similar direction cosines are defined for
x20 and x30 coordinates.

TABLE 2.3
Transformation table between Ox1 x2 x3 and
Ox10 x20 x30
^1 ,
e x1 ^2 ,
e x2 ^3 ,
e x3
^10 ,
e x10 a11 a12 a13
^20 ,
e x20 a21 a22 a23
^30 ,
e x30 a31 a32 a33

or in compact indicial form


^i0 = aij e
e ^j . (2.54)

By defining the matrix A whose elements are the direction cosines aij , Eq 2.54 can be
written in matrix form as

    
e^10 a11 a12 a13 ^1
e
0
or (2.55)
 0  
ei1
[^ ej1 ]
] = [aij ] [^ e^ =
 2   a21 a22 ^2 
a23   e
 

e^30 a31 a32 a33 ^3


e

where the elements of the column matrices are unit vectors. The matrix A is called the
transformation matrix because, as we shall see, of its role in transforming the components
of a vector (or tensor) referred to one set of axes into the components of the same vector
(or tensor) in a rotated set.
Because of the perpendicularity of the primed axes, e ^j0 = δij . But also, in view of
^i0 · e
Eq 2.54,

^i0 · e
e ^j0 = aiq e
^q · ajm e
^m = aiq ajm δqm = aiq ajq = δij
28 Continuum Mechanics for Engineers

from which we extract the orthogonality condition on the direction cosines (given here in
both indicial and matrix form),

aiq ajq = δij or AAT = I . (2.56)


Note that this is simply the inner product of the i row with the j row of the matrix
th th

A. By an analogous derivation to that leading to Eq 2.54, but using the columns of A, we


obtain

^j0 ,
^i = aji e
e (2.57)
which in matrix form is

 
h i h i a11 a12 a13
^1j [aij ] or a23  . (2.58)
 0 
e1i ] = e
[^ ^1
e ^2
e ^3
e = ^10
e ^20
e ^30
e  a21 a22
 

a31 a32 a33

Note that using the transpose AT , Eq 2.58 may also be written


    
^1
e a11 a21 a31 ^10
e
e1i ] = [aij ]T e
^1j or  e (2.59)
 0   0 
[^ ^2  =  a12 a22 a32   e^2 
  
^3
e a13 a23 a33 ^30
e
in which column matrices are used for the vectors e
^i and e^i0 . By a consideration of the
dot product and Eq 2.57 we obtain a second orthogonality condition

aij aik = δjk or AT A = I (2.60)

which is the inner product of the jth column with the kth column of A.
Consider next an arbitrary vector v having components vi in the unprimed system, and
vi0 in the primed system. Then using Eq 2.57,

v = vj0 e
^j0 = vi e ^j0
^i = vi aji e

from which, by matching coefficients on e^j0 , we have (in both the indicial and matrix
forms),
vj0 = aji vi or v 0 = Av = vAT (2.61)
which is the transformation law expressing the primed components of an arbitrary vector in
terms of its unprimed components. Although the elements of the transformation matrix
are written as aij we must emphasize that they are not the components of a second-
order Cartesian tensor as it might appear. Multiplication of Eq 2.61 by ajk and using the
orthogonality condition Eq 2.60 we obtain the inverse law

vk = ajk vj0 or v = v 0 A = AT v 0 (2.62)

giving the unprimed components in terms of the primed.


By a direct application of Eq 2.62 to the dyad uv we have
0
ui vj = aqi uq0 amj vm = aqi amj uq0 vm
0
. (2.63)

But a dyad is, after all, one form of a second-order tensor, and so by an obvious adaptation
of Eq 2.63 we obtain the transformation law for a second-order tensor, T as
0
tij = aqi amj tqm or T = AT T 0 A (2.64)
Another random document with
no related content on Scribd:
compliance. To SEND DONATIONS or determine the status of
compliance for any particular state visit www.gutenberg.org/donate.

While we cannot and do not solicit contributions from states where


we have not met the solicitation requirements, we know of no
prohibition against accepting unsolicited donations from donors in
such states who approach us with offers to donate.

International donations are gratefully accepted, but we cannot make


any statements concerning tax treatment of donations received from
outside the United States. U.S. laws alone swamp our small staff.

Please check the Project Gutenberg web pages for current donation
methods and addresses. Donations are accepted in a number of
other ways including checks, online payments and credit card
donations. To donate, please visit: www.gutenberg.org/donate.

Section 5. General Information About Project


Gutenberg™ electronic works
Professor Michael S. Hart was the originator of the Project
Gutenberg™ concept of a library of electronic works that could be
freely shared with anyone. For forty years, he produced and
distributed Project Gutenberg™ eBooks with only a loose network of
volunteer support.

Project Gutenberg™ eBooks are often created from several printed


editions, all of which are confirmed as not protected by copyright in
the U.S. unless a copyright notice is included. Thus, we do not
necessarily keep eBooks in compliance with any particular paper
edition.

Most people start at our website which has the main PG search
facility: www.gutenberg.org.

This website includes information about Project Gutenberg™,


including how to make donations to the Project Gutenberg Literary
Archive Foundation, how to help produce our new eBooks, and how
to subscribe to our email newsletter to hear about new eBooks.

You might also like