Ebook Classical Continuum Mechanics Applied and Computational Mechanics 2Nd Edition Surana Online PDF All Chapter

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 69

Classical Continuum Mechanics

(Applied and Computational


Mechanics) 2nd Edition Surana
Visit to download the full and correct content document:
https://ebookmeta.com/product/classical-continuum-mechanics-applied-and-computat
ional-mechanics-2nd-edition-surana/
More products digital (pdf, epub, mobi) instant
download maybe you interests ...

Hamilton s Principle in Continuum Mechanics Bedford


Anthony

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

Continuum Mechanics for Engineers 4th Edition G. Thomas


Mase

https://ebookmeta.com/product/continuum-mechanics-for-
engineers-4th-edition-g-thomas-mase/

Classical Mechanics in Geophysical Fluid Dynamics 2nd


Edition Osamu Morita

https://ebookmeta.com/product/classical-mechanics-in-geophysical-
fluid-dynamics-2nd-edition-osamu-morita/

Classical Mechanics: with MATLAB Applications 2nd


Edition Javier E Hasbun Phd

https://ebookmeta.com/product/classical-mechanics-with-matlab-
applications-2nd-edition-javier-e-hasbun-phd/
Computational Mechanics of Fluid-Structure Interaction
Jaiman

https://ebookmeta.com/product/computational-mechanics-of-fluid-
structure-interaction-jaiman/

Quantum Mechanics: Inception or Conception? A Classical


Preamble to Quantum Mechanics 1st Edition Giuseppe
Pileio

https://ebookmeta.com/product/quantum-mechanics-inception-or-
conception-a-classical-preamble-to-quantum-mechanics-1st-edition-
giuseppe-pileio/

Computational Fluid Mechanics and Heat Transfer Dale


Arden Anderson

https://ebookmeta.com/product/computational-fluid-mechanics-and-
heat-transfer-dale-arden-anderson/

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/

Classical Beam Theories of Structural Mechanics Andreas


Öchsner

https://ebookmeta.com/product/classical-beam-theories-of-
structural-mechanics-andreas-ochsner/
Classical Continuum
Mechanics
APPLIED AND COMPUTATIONAL MECHANICS
A Series of Textbooks and Reference Books
Founding Editor
J.N. Reddy

Continuum Mechanics for Engineers, Forth Edition


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

Dynamics in Engineering Practice, Eleventh Edition


Dara W. Childs, Andrew P. Conkey

Advanced Mechanics of Continua


Karan S. Surana

Physical Components of Tensors


Wolf Altman, Antonio Marmo De Oliveira

Continuum Mechanics for Engineers, Third Edition


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

Classical Continuum Mechanics, Second Edition


Karan S. Surana

For more information about this series, please visit: https://www.crcpress.com/


Applied-and-Computational-Mechanics/book-series/CRCAPPCOMMEC
Classical Continuum
Mechanics

Karan S. Surana
Department of Mechanical Engineering
The University of Kansas
Lawrence, Kansas
Second edition published 2022
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

© 2022 Karan S. Surana

First edition published by CRC Press 2014

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 it 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: 978-0-367-61296-2 (hbk)


ISBN: 978-0-367-61521-5 (pbk)
ISBN: 978-1-003-10533-6 (ebk)

DOI: 10.1201/9781003105336

Typeset in CMR10
by KnowledgeWorks Global Ltd.
To

My beloved family

Abha, Deepak, Rishi, Yogini, and Riya


Contents

PREFACE xvii

ABOUT THE AUTHOR xix

LIST OF ABBREVIATIONS xxi

1 INTRODUCTION 1

2 CONCEPTS AND MATHEMATICAL PRELIMINARIES 9


2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Notations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2.1 Einstein notations . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2.2 Index notations and Kronecker delta . . . . . . . . . . . . . 11
2.2.3 Vector and matrix notations . . . . . . . . . . . . . . . . . . 12
2.2.4 Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.3 Permutation tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.4  -δδ identity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.5 Operations using vector, matrix and Einstein notations . . . . . . . 16
2.5.1 Multiplication . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.5.2 Substitution . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.5.3 Factoring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.5.4 Contraction (or trace) . . . . . . . . . . . . . . . . . . . . . . 17
2.6 Reference frame and reference frame transformation . . . . . . . . . 19
2.7 Coordinate frames and transformations . . . . . . . . . . . . . . . . 21
2.7.1 Cartesian frame and orthogonal coordinate
transformations . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.7.2 Curvilinear coordinates (or frame) . . . . . . . . . . . . . . . 25
2.8 Curvilinear frames, covariant and contravariant bases . . . . . . . . 26
2.8.1 Covariant basis . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.8.2 Contravariant basis . . . . . . . . . . . . . . . . . . . . . . . 29
2.8.3 Alternate way to visualize covariant and contravariant bases
g̃ gi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
g i , g̃ 30
2.9 Scalars, vectors and tensors . . . . . . . . . . . . . . . . . . . . . . . 31
2.10 Coordinate transformations, definitions and operations . . . . . . . 32
2.10.1 Coordinate transformation T . . . . . . . . . . . . . . . . . . 33
2.10.2 Induced transformations . . . . . . . . . . . . . . . . . . . . 34

vii
viii CONTENTS

2.10.3 Isomorphism between coordinate transformations and induced


transformations . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.10.4 Transformations by covariance and contravariance . . . . . . 36
2.10.5 The tensor concept: Covariant and contravariant tensors . . 40
2.11 Tensors in three-dimensional x-frame, tensor operations, orthogonal
coordinate transformations and invariance . . . . . . . . . . . . . . 47
2.11.1 Tensors in Cartesian x-frame . . . . . . . . . . . . . . . . . . 47
2.11.2 Tensor operations . . . . . . . . . . . . . . . . . . . . . . . . 48
2.11.3 Transformations of tensors defined in orthogonal frames due
to orthogonal coordinate transformation . . . . . . . . . . . 51
2.11.4 Invariants of tensors . . . . . . . . . . . . . . . . . . . . . . . 54
2.11.5 Hamilton-Cayley theorem . . . . . . . . . . . . . . . . . . . 59
2.11.6 Differential calculus of tensors . . . . . . . . . . . . . . . . . 60
2.12 Some useful relations . . . . . . . . . . . . . . . . . . . . . . . . . . 65
2.13 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

3 KINEMATICS OF MOTION, DEFORMATION AND THEIR


MEASURES 71
3.1 Description of motion . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.2 Material particle displacements . . . . . . . . . . . . . . . . . . . . . 72
3.3 Lagrangian, Eulerian descriptions and descriptions in fluid
mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
3.3.1 Lagrangian or referential description of motion . . . . . . . . 73
3.3.2 Eulerian or spatial description of motion . . . . . . . . . . . 74
3.3.3 Descriptions in fluid mechanics . . . . . . . . . . . . . . . . . 75
3.3.4 Notations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
3.4 Material derivative . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
3.4.1 Material derivative in Lagrangian or referential
description . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
3.4.2 Material derivative in Eulerian or spatial description . . . . 77
3.5 Acceleration of a material particle . . . . . . . . . . . . . . . . . . . 79
3.5.1 Lagrangian or referential description . . . . . . . . . . . . . 79
3.5.2 Eulerian or spatial description . . . . . . . . . . . . . . . . . 79
3.6 Deformation Gradient Tensor . . . . . . . . . . . . . . . . . . . . . . 80
3.7 Continuous deformation of matter, restrictions on the description of
motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
3.8 Change of description, co- and contra-variant measures . . . . . . . 84
3.9 Notations for covariant and contravariant measures . . . . . . . . . 85
3.10 Deformation, measures of length and change in length . . . . . . . . 85
3.10.1 Covariant measures of length and change in length . . . . . 85
3.10.2 Contravariant measures of length and change in length . . . 87
3.11 Covariant and contravariant measures of finite strain in Lagrangian
and Eulerian descriptions . . . . . . . . . . . . . . . . . . . . . . . . 88
3.11.1 Covariant measures of finite strains . . . . . . . . . . . . . . 90
3.11.2 Contravariant measures of finite strains . . . . . . . . . . . . 94
3.12 Changes in strain measures due to rigid rotation of frames . . . . . 98
3.12.1 Change in covariant Lagrangian descriptions of strain when
xi are changed to x0i due to rigid rotation . . . . . . . . . . . 100
CONTENTS ix

3.12.2 Changes in contravariant Eulerian measures of strains when


xi are changed to x0i due to rigid rotation . . . . . . . . . . . 102
3.12.3 Change in covariant Lagrangian measures of strain when x̄i
are changed to x̄0i by rigid rotation [Q̄] . . . . . . . . . . . . 103
3.12.4 Changes in Eulerian measures of strain when x̄i are changed
to x̄0i by rigid rotation [Q] . . . . . . . . . . . . . . . . . . . 105
3.13 Invariants of strain tensors . . . . . . . . . . . . . . . . . . . . . . . 107
3.14 Expanded form of strain tensors . . . . . . . . . . . . . . . . . . . . 108
3.14.1 Green’s strain measure: [ε[0] ] . . . . . . . . . . . . . . . . . . 108
3.14.2 Almansi strain measure : [ε̄[0] ] . . . . . . . . . . . . . . . . . 109
3.15 Physical meaning of strains . . . . . . . . . . . . . . . . . . . . . . . 109
3.15.1 Extensions and stretches parallel to ox1 , ox2 , ox3 axes in
the x-frame: covariant measure of strain in Lagrangian
description using ε [0] , Green’s strain in Lagrangian
description . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
3.15.2 Extensions and stretches parallel to x̄-frame axes
using [ε̄[0] ] . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
3.15.3 Angles between the fibers or material lines . . . . . . . . . . 116
3.16 Small deformation, small strain deformation physics . . . . . . . . . 119
3.16.1 Green’s strain: Lagrangian description . . . . . . . . . . . . 119
3.16.2 Almansi strain tensor: Eulerian description . . . . . . . . . . 121
3.17 Additive and multiplicative decompositions of deformation gradient
tensor [J] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
3.17.1 Additive decomposition of [J] . . . . . . . . . . . . . . . . . 123
3.17.2 Multiplicative decomposition of [J]: Polar decomposition
into stretch and rotation tensor . . . . . . . . . . . . . . . . 124
3.17.3 Strain measures in terms of [Sr ], [Sl ] and [R] . . . . . . . . . 131
3.18 Invariants of [C[0] ], [B [0] ], [Sr ] and [Sl ] in terms of principal stretches
of [Sr ] and [Sl ] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
3.18.1 Principal stretches of [Sr ] and [Sl ] . . . . . . . . . . . . . . . 132
3.18.2 Principal invariants of [C[0] ] in terms of λri . . . . . . . . . . 133
3.18.3 Principal Invariants of [Sr ] . . . . . . . . . . . . . . . . . . . 134
3.18.4 Principal invariants of [B [0] ] in terms of λli . . . . . . . . . . 135
3.18.5 Principal Invariants of [Sl ] . . . . . . . . . . . . . . . . . . . 136
3.19 Deformation of areas and volumes . . . . . . . . . . . . . . . . . . . 136
3.19.1 Areas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
3.19.2 Volumes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
3.19.3 Integral form of {dĀ} over ∂ V̄ . . . . . . . . . . . . . . . . . 139
3.20 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139

4 DEFINITIONS AND MEASURES OF STRESSES 147


4.1 Concept of stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
4.2 Cut Principle of Cauchy . . . . . . . . . . . . . . . . . . . . . . . . . 147
4.3 Definition of stress on area dĀn . . . . . . . . . . . . . . . . . . . . 148
4.4 Cauchy stress tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
4.4.1 Force balance . . . . . . . . . . . . . . . . . . . . . . . . . . 151
4.4.2 Moment of Forces . . . . . . . . . . . . . . . . . . . . . . . . 153
4.4.3 Cauchy principle . . . . . . . . . . . . . . . . . . . . . . . . . 154
x CONTENTS

4.5 Stress measures: finite deformation, finite strain . . . . . . . . . . . 155


4.5.1 Contravariant Cauchy stress tensor σ (0) and σ̄ σ (0) in
Lagrangian and Eulerian descriptions . . . . . . . . . . . . . 156
4.5.2 Covariant Cauchy stress tensor in σ̄ σ (0) and σ (0) in Eulerian
and Lagrangian descriptions . . . . . . . . . . . . . . . . . . 159
4.5.3 Mixed stress tensors: Jaumann stress tensor . . . . . . . . . 160
4.6 Contravariant Second Piola-Kirchhoff stress tensor σ [0] or σ̄ σ [0] . . . 161
4.7 Contravariant First Piola-Kirchhoff or Lagrange stress tensor σ ∗ . . 162
4.8 Covariant Second Piola-Kirchhoff stress tensor σ [0] or σ̄ σ [0] . . . . . 163
4.9 General Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
4.10 Summary of stress measures . . . . . . . . . . . . . . . . . . . . . . 165
4.10.1 Cauchy stress tensors . . . . . . . . . . . . . . . . . . . . . . 165
4.10.2 Jaumann stress tensors . . . . . . . . . . . . . . . . . . . . . 165
4.10.3 Second Piola-Kirchhoff stress tensors . . . . . . . . . . . . . 165
4.10.4 First Piola-Kirchhoff stress tensor . . . . . . . . . . . . . . . 166
4.11 Conjugate strain measures . . . . . . . . . . . . . . . . . . . . . . . 166
4.12 Relations between different stress measures and some other useful
relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
4.13 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168

5 RATES, CONVECTED TIME DERIVATIVES AND


OBJECTIVITY 171
5.1 Rate of deformation . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
5.1.1 Lagrangian description . . . . . . . . . . . . . . . . . . . . . 171
5.1.2 Eulerian description . . . . . . . . . . . . . . . . . . . . . . . 172
5.2 Additive decomposition of velocity gradient tensor . . . . . . . . . . 173
5.3 Interpretation of the components of [D̄] . . . . . . . . . . . . . . . . 174
5.3.1 Diagonal components of [D̄] . . . . . . . . . . . . . . . . . . 174
5.3.2 Off diagonal components of [D̄]: physical interpretation . . . 175
5.4 Rate of change or material derivative of strain tensors [C[0] ]
and [ε[0] ] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
5.5 Physical meaning of spin tensor [W̄ ] . . . . . . . . . . . . . . . . . . 179
5.6 Vorticity vector and vorticity . . . . . . . . . . . . . . . . . . . . . . 180
5.7 Rate of change of [J], i.e., material derivative of [J] . . . . . . . . . 181
5.8 Rate of change of [J], ¯ i.e., material derivative of [J]¯ . . . . . . . . . 182
5.9 Rate of change of det[J], i.e material derivative of |J| . . . . . . . . 182
5.10 Rate of change of det[J],¯ i.e., material derivative of det[J] ¯ . . . . . . 183
5.11 Rate of change of volume, i.e., material derivative of volume . . . . 184
5.12 Rate of change of area: material derivative of area . . . . . . . . . . 185
5.13 Convected time derivatives of stress and strain tensors . . . . . . . . 186
5.13.1 Stress and strain measures for convected time derivatives . . 187
5.13.2 Convected time derivatives of the Cauchy stress tensor:
compressible matter . . . . . . . . . . . . . . . . . . . . . . . 188
5.13.3 Convected time derivatives of the Cauchy stress tensor:
incompressible matter . . . . . . . . . . . . . . . . . . . . . . 192
5.13.4 Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
5.13.5 Convected time derivatives of the strain tensors . . . . . . . 194
CONTENTS xi

5.14 Conjugate pairs of convected time derivatives of stress and strain


tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
5.15 Objective tensors and objective rates . . . . . . . . . . . . . . . . . 197
5.15.1 Galilean transformation: tensors . . . . . . . . . . . . . . . . 198
5.15.2 Euclidean or Non-Galilean transformations: tensors . . . . . 202
5.15.3 Objective rates . . . . . . . . . . . . . . . . . . . . . . . . . 205
5.15.4 Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
5.16 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209

6 CONSERVATION AND BALANCE LAWS IN EULERIAN


DESCRIPTION 213
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
6.2 Localization theorem . . . . . . . . . . . . . . . . . . . . . . . . . . 214
6.3 Mass density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
6.4 Conservation of mass: CM . . . . . . . . . . . . . . . . . . . . . . . 217
6.5 Transport theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
6.5.1 Approach I . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
6.5.2 Approach II . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
6.5.3 Continued development of transport theorem . . . . . . . . . 220
6.6 Conservation of mass: CM . . . . . . . . . . . . . . . . . . . . . . . 221
6.6.1 Integral form of CM . . . . . . . . . . . . . . . . . . . . . . . 221
6.6.2 Differential form of CM . . . . . . . . . . . . . . . . . . . . . 221
6.7 Kinematics of continuous media: BLM . . . . . . . . . . . . . . . . 222
6.7.1 Preliminary considerations . . . . . . . . . . . . . . . . . . . 223
6.7.2 Derivation of equations of BLM . . . . . . . . . . . . . . . . 223
6.7.3 Approach I: Integral and differential forms of BLM . . . . . 226
6.7.4 Approach II . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
6.8 Kinematics of continuous media: BAM . . . . . . . . . . . . . . . . 230
6.8.1 Integral form of BAM . . . . . . . . . . . . . . . . . . . . . . 230
6.8.2 Differential form of BAM . . . . . . . . . . . . . . . . . . . . 231
6.9 First law of thermodynamics: FLT . . . . . . . . . . . . . . . . . . . 232
6.9.1 Integral form of FLT . . . . . . . . . . . . . . . . . . . . . . 235
6.9.2 Differential form of FLT . . . . . . . . . . . . . . . . . . . . 235
6.10 Second law of thermodynamics: SLT . . . . . . . . . . . . . . . . . . 237
6.10.1 Integral form of SLT . . . . . . . . . . . . . . . . . . . . . . 238
6.10.2 Differential form of SLT . . . . . . . . . . . . . . . . . . . . 239
6.11 Summary of mathematical model from CBL . . . . . . . . . . . . . 240
6.11.1 Differential form of the CBL . . . . . . . . . . . . . . . . . . 240
6.11.2 Integral form of the CBL . . . . . . . . . . . . . . . . . . . . 242
6.12 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243

7 CONSERVATION AND BALANCE LAWS IN LAGRANGIAN


DESCRIPTION 247
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
7.2 Mathematical model for deforming continua in
Lagrangian description . . . . . . . . . . . . . . . . . . . . . . . . . 247
7.3 Conservation of mass: (CM) . . . . . . . . . . . . . . . . . . . . . . 248
7.3.1 Integral form of CM . . . . . . . . . . . . . . . . . . . . . . . 248
xii CONTENTS

7.3.2 Differential form of CM . . . . . . . . . . . . . . . . . . . . . 248


7.4 Balance of linear momenta: (BLM) . . . . . . . . . . . . . . . . . . 249
7.4.1 Differential form of BLM . . . . . . . . . . . . . . . . . . . . 251
7.5 Balance of angular momenta: (BAM) . . . . . . . . . . . . . . . . . 253
7.5.1 Differential form of BAM . . . . . . . . . . . . . . . . . . . . 253
7.6 First law of thermodynamics: (FLT) . . . . . . . . . . . . . . . . . . 254
7.6.1 Differential form of the FLT . . . . . . . . . . . . . . . . . . 254
7.6.2 Rate of mechanical work conjugate pairs in the energy
equation (differential form) . . . . . . . . . . . . . . . . . . . 257
7.6.3 Energy equation in equivalent rate of work conjugate
measures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259
7.7 Second law of thermodynamics: (SLT) . . . . . . . . . . . . . . . . . 260
7.7.1 Differential form of SLT . . . . . . . . . . . . . . . . . . . . 260
7.8 Second law of thermodynamics using Gibbs potential (differential
form) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
7.8.1 Using σ [0] and ε [0] as conjugate pair . . . . . . . . . . . . . 263
7.8.2 Using σ [0] and C [0] as conjugate pair . . . . . . . . . . . . . 264
7.9 Summary of differential form of CBL . . . . . . . . . . . . . . . . . 264
7.10 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266

8 CONSTITUTIVE THEORIES 271


8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
8.2 Axioms of constitutive theory . . . . . . . . . . . . . . . . . . . . . 273
8.3 Approaches of deriving constitutive theories . . . . . . . . . . . . . 278
8.3.1 Thermodynamic approach . . . . . . . . . . . . . . . . . . . 278
8.3.2 Other approaches (not strictly thermodynamic) . . . . . . . 278
8.4 Considerations in the constitutive theories . . . . . . . . . . . . . . 278
8.4.1 Common deformation physics . . . . . . . . . . . . . . . . . 279
8.5 Thermodynamic approach of deriving constitutive theories . . . . . 281
8.5.1 Entropy inequality in Eulerian description . . . . . . . . . . 281
8.5.2 Additive decomposition of Cauchy stress tensor (0)σ̄ σ . . . . 282
8.5.3 Constitutive tensors, their argument tensors and SLT . . . . 282
8.5.4 Constitutive theory for equilibrium Cauchy stress tensor (0)eσ̄ σ
(volumetric deformation physics): Eulerian description,
Helmholtz free energy density . . . . . . . . . . . . . . . . . 284
8.5.5 Constitutive theories for equilibrium stress (0)eσ (volumetric
deformation): Lagrangian description . . . . . . . . . . . . . 286
8.5.6 Constitutive theories for equilibrium contravariant second
Piola-Kirchhoff stress tensor eσ [0] . . . . . . . . . . . . . . . 287
8.6 Representation Theorem . . . . . . . . . . . . . . . . . . . . . . . . 290
8.6.1 Def: Representation theorem . . . . . . . . . . . . . . . . . . 291
8.6.2 Constitutive theory for a symmetric tensor of rank two . . . 292
8.6.3 Constitutive variable is a tensor of rank one . . . . . . . . . 299
8.7 Other approaches of deriving constitutive theories . . . . . . . . . . 302
8.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
CONTENTS xiii

9 CONSTITUTIVE THEORIES FOR THERMOELASTIC


SOLIDS 305
9.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 305
9.2 Thermodynamic approach . . . . . . . . . . . . . . . . . . . . . . . 306
9.2.1 Finite deformation, finite strain, compressible,
non-isothermal . . . . . . . . . . . . . . . . . . . . . . . . . . 306
9.2.2 Finite deformation, finite strain, compressible,
isothermal: reversible deformation . . . . . . . . . . . . . . . 310
9.2.3 Finite deformation, finite strain, incompressible,
non-isothermal . . . . . . . . . . . . . . . . . . . . . . . . . . 311
9.2.4 Finite deformation, finite strain, incompressible,
isothermal: reversible deformation . . . . . . . . . . . . . . . 312
9.2.5 Small deformation, small strain . . . . . . . . . . . . . . . . 312
9.3 Other approaches of deriving constitutive theories (not necessarily
thermodynamic) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 314
9.3.1 Constitutive theory for σ [0] using Φ: reversible . . . . . . . . 315
9.3.2 Constitutive theory for σ [0] using strain energy density π . . 318
9.3.3 Constitutive theory for σ [0] using π = π(ε ε[0] ) and Taylor
series expansion: non-isotropic, non-homogeneous solid
continua . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
9.3.4 Constitutive theory for σ using π(ε ε): small deformation,
small strain . . . . . . . . . . . . . . . . . . . . . . . . . . . 328
9.4 Constitutive theory for heat vector . . . . . . . . . . . . . . . . . . . 329
9.4.1 Constitutive theory for q using entropy inequality . . . . . . 330
9.4.2 Constitutive theories for q using representation theorem . . 331
9.5 General remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 332
9.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 333

10 CONSTITUTIVE THEORIES FOR THERMOVISCOELASTIC


SOLIDS WITHOUT MEMORY 337
10.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 337
10.2 Finite deformation, finite strain . . . . . . . . . . . . . . . . . . . . 338
10.2.1 Equilibrium stress tensor eσ [0] . . . . . . . . . . . . . . . . . 338
10.2.2 Deviatoric stress dσ [0] . . . . . . . . . . . . . . . . . . . . . . 339
10.3 Small strain, small deformation . . . . . . . . . . . . . . . . . . . . . 342
10.3.1 Equilibrium stress tensor eσ . . . . . . . . . . . . . . . . . . 342
10.3.2 Deviatoric stress dσ . . . . . . . . . . . . . . . . . . . . . . . 343
10.4 Kelvin-Voigt Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 345
10.5 1D wave propagation in viscoelastic solid continua . . . . . . . . . . 346
10.5.1 Alternate phenomenological model for dissipation . . . . . . 347
10.5.2 Model Problem: Numerical Studies . . . . . . . . . . . . . . 348
10.6 Constitutive theories for heat vector q . . . . . . . . . . . . . . . . . 351
10.7 General remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351
10.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 358
xiv CONTENTS

11 CONSTITUTIVE THEORIES FOR THERMOVISCOELASTIC


SOLIDS WITH MEMORY 361
11.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 361
11.2 Finite deformation, finite strain . . . . . . . . . . . . . . . . . . . . 362
11.2.1 Constitutive theory for dσ [0] . . . . . . . . . . . . . . . . . . 362
11.3 Small deformation, small strain . . . . . . . . . . . . . . . . . . . . . 367
11.3.1 Constitutive theory for dσ . . . . . . . . . . . . . . . . . . . 367
11.4 Memory modulus or relaxation modulus . . . . . . . . . . . . . . . . 371
11.5 Zener constitutive model . . . . . . . . . . . . . . . . . . . . . . . . 372
11.6 Model problem: numerical studies . . . . . . . . . . . . . . . . . . . 372
11.7 Constitutive theories for heat vector q . . . . . . . . . . . . . . . . . 375
11.8 General remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 375
11.9 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 380

12 CONSTITUTIVE THEORIES FOR THERMOVISCOUS


FLUIDS 383
12.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 383
12.2 Preliminary considerations . . . . . . . . . . . . . . . . . . . . . . . 384
12.3 Constitutive theory for equilibrium Cauchy stress . . . . . . . . . . 385
12.4 Constitutive theory for deviatoric Cauchy stress . . . . . . . . . . . 385
12.4.1 A constitutive theory of order one (n = 1) for (0)dσ̄ σ:
Newtonian and generalized Newtonian fluids . . . . . . . . . 387
12.4.2 Linear constitutive theory of order n for (0)dσ̄σ . . . . . . . . 388
12.4.3 Linear constitutive theory of order one (n = 1): Newtonian
and generalized Newtonian fluids . . . . . . . . . . . . . . . 389
12.4.4 Generalized Newtonian fluids: variable transport properties 390
12.5 Constitutive theory for heat vector, Eulerian description . . . . . . 394
q using entropy inequality . . . . . .
12.5.1 Constitutive theory for q̄ 394
q using representation theorem . . .
12.5.2 Constitutive theory for q̄ 395
12.6 General remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 396
12.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 397

13 CONSTITUTIVE THEORIES FOR THERMOVISCOELASTIC


FLUIDS 401
13.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 401
13.2 Preliminary considerations . . . . . . . . . . . . . . . . . . . . . . . 402
13.3 Considerations in the constitutive theories . . . . . . . . . . . . . . 403
13.4 Constitutive theory for equilibrium stress (0)eσ̄
σ . . . . . . . . . . . . 404
13.5 Constitutive theory for deviatoric stress (0)dσ̄
σ . . . . . . . . . . . . . 404
13.5.1 Simplified constitutive theories for (m)dσ̄
σ : m = 1, n = 1 . . . 407
13.5.2 Maxwell constitutive model . . . . . . . . . . . . . . . . . . 409
13.5.3 Giesekus Constitutive Model . . . . . . . . . . . . . . . . . . 410
13.5.4 Discussion on the Giesekus constitutive model derived here
and the constitutive model used currently . . . . . . . . . . 412
13.5.5 Simplified constitutive theory for deviatoric Cauchy stress
σ (0) : m = 1, n = 2 . . . . . . . . . . . . . . . . . . . 413
tensor dσ̄
13.5.6 Oldroyd-B constitutive model (m = 1, n = 2) . . . . . . . . 414
CONTENTS xv

13.5.7 A single constitutive theory for dilute and dense polymeric


fluids (m = 1, n = 2) . . . . . . . . . . . . . . . . . . . . . . 416
13.6 Constitutive theory for heat vector . . . . . . . . . . . . . . . . . . . 417
13.7 Numerical studies using Giesekus constitutive model . . . . . . . . . 417
13.7.1 Model Problem 1: fully developed flow between parallel
plates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 418
13.7.2 Model Problem 2: fully developed flow between parallel plates
using 2D formulation . . . . . . . . . . . . . . . . . . . . . . 424
13.8 General remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 428
13.9 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 429

14 CONSTITUTIVE THEORIES FOR THERMO HYPO-ELASTIC


SOLIDS 433
14.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 433
14.2 Preliminary considerations . . . . . . . . . . . . . . . . . . . . . . . 433
14.3 Constitutive theory for equilibrium
Cauchy stress (0)eσ̄
σ . . . . . . . . . . . . . . . . . . . . . . . . . . . . 434
14.4 Constitutive theory for deviatoric Cauchy stress . . . . . . . . . . . 435
14.4.1 Linear constitutive theory of order n for (1)dσ̄ σ . . . . . . . . 437
14.4.2 Linear constitutive theory of order one (n = 1) . . . . . . . . 438
14.5 Constitutive theory for heat vector q̄q . . . . . . . . . . . . . . . . . 438
14.6 General remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 438
14.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 439

15 THERMODYNAMIC RELATIONS AND COMPLETE


MATHEMATICAL MODELS 441
15.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 441
15.2 Thermodynamic pressure: equation of state . . . . . . . . . . . . . . 441
15.2.1 Perfect or ideal gas law . . . . . . . . . . . . . . . . . . . . . 442
15.2.2 Real gas models . . . . . . . . . . . . . . . . . . . . . . . . . 442
15.2.3 Compressible solids . . . . . . . . . . . . . . . . . . . . . . . 443
15.3 Internal energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 447
15.3.1 Compressible matter . . . . . . . . . . . . . . . . . . . . . . 447
15.3.2 Incompressible matter . . . . . . . . . . . . . . . . . . . . . 448
15.4 Differential form of complete mathematical models in Lagrangian
description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 448
15.4.1 CBL: Finite deformation, finite strain . . . . . . . . . . . . . 448
15.4.2 Constitutive theory for eσ [0] : finite deformation, finite
strain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 449
15.4.3 Constitutive theory for dσ [0] : finite deformation, finite
strain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 449
15.4.4 CBL: Small deformation, small strain . . . . . . . . . . . . . 450
15.4.5 Constitutive theory for eσ : small deformation, small strain . 451
15.4.6 Constitutive theory for dσ : small deformation, small strain . 451
15.4.7 Constitutive theory for heat vector q . . . . . . . . . . . . . 452
15.5 Differential form of complete mathematical model in Eulerian
description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 452
15.5.1 CBL: Conservation and balance laws . . . . . . . . . . . . . 453
xvi CONTENTS

15.5.2 Constitutive theory for (0)eσ̄


σ . . . . . . . . . . . . . . . . . . 453
15.5.3 Constitutive theory for (0)dσ̄
σ . . . . . . . . . . . . . . . . . . 453
15.5.4 Constitutive theory for heat vector q̄q . . . . . . . . . . . . . 455
15.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 455

16 ENERGY METHODS, PRINCIPLE OF VIRTUAL WORK,


CALCULUS OF VARIATIONS 457
16.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 457
16.2 Boundary value problems (BVPs) . . . . . . . . . . . . . . . . . . . 457
16.2.1 Mathematical classification of differential operators . . . . . 458
16.2.2 Calculus of variations . . . . . . . . . . . . . . . . . . . . . . 458
16.2.3 Principle of Virtual work: BVPs . . . . . . . . . . . . . . . . 460
16.2.4 Final remarks (BVPs) . . . . . . . . . . . . . . . . . . . . . 461
16.3 IVPs, energy methods and principle of virtual work . . . . . . . . . 461
16.3.1 Energy functional from the differential form of IVP . . . . . 462
16.3.2 Differential form of IVP from energy functional . . . . . . . 463
16.3.3 Principle of virtual work in IVPs . . . . . . . . . . . . . . . 465
16.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 466

Appendix A: Combined generators and invariants 467

Appendix B: Transformations and operations in Cartesian,


cylindrical and spherical coordinate systems 475
B.1 Cartesian frame x1 , x2 , x3 and cylindrical frame r, θ, z . . . . . . . . 475
B.1.1 Relationship between coordinates of a point in x1 , x2 , x3 and
r, θ, z-frames . . . . . . . . . . . . . . . . . . . . . . . . . . . 476
B.1.2 Converting derivatives of a scalar with respect to x1 , x2 , x3
into its derivatives with respect to r, θ, z . . . . . . . . . . . 476
B.1.3 Relationship between bases in x1 , x2 , x3 - and r, θ, z-frames . 478
B.2 Cartesian frame x1 , x2 , x3 and spherical frame r, θ, φ . . . . . . . . . 480
B.2.1 Relationship between coordinates of a point in x1 , x2 , x3 and
r, θ, φ-frames . . . . . . . . . . . . . . . . . . . . . . . . . . . 480
B.2.2 Converting derivatives of a scalar with respect to x1 , x2 , x3
into derivatives with respect to r, θ, φ . . . . . . . . . . . . . 481
B.2.3 Relationship between bases, i.e., unit vectors in x1 , x2 , x3 -
and r, θ, φ-frames . . . . . . . . . . . . . . . . . . . . . . . . 482
B.3 Differential operations in r, θ, z- and r, θ, φ-frames . . . . . . . . . . 483
B.3.1 r, θ, z-frame . . . . . . . . . . . . . . . . . . . . . . . . . . . 483
B.3.2 r, θ, φ-frame . . . . . . . . . . . . . . . . . . . . . . . . . . . 484
B.4 Some examples: r, θ, z-frame . . . . . . . . . . . . . . . . . . . . . . 486
B.4.1 Symmetric part of the velocity gradient tensor D . . . . . . 486
B.4.2 Skew-symmetric part of the velocity gradient tensor W . . . 487
B.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 487

BIBLIOGRAPHY 489

INDEX 499
PREFACE
Phenomenal progress and advances in continuum theories over the last fifty
years necessitate that we precisely define the specific theories of interest and study
within the umbrella of continuum mechanics. Broadly speaking, all continuum the-
ories at present fall into distinct categories: classical continuum mechanics (CCM)
and non-classical continuum mechanics (NCCM). This book is devoted to the study
of CCM based on three fundamental assumptions: (1) The behavior of the matter
at fine-scale is neglected; instead, we employ the concept of infinitesimal volume
referred to as a material point. If we assume the matter to be isotropic and ho-
mogeneous, then a continuum theory derived for a material point is valid for the
entire volume of matter. (2) A material point has only three translational degrees
of freedom (displacements). The displacement and forces are work conjugate in a
deforming volume of matter and always coexist. (3) We assume the deforming mat-
ter to be in thermodynamic equilibrium. This allows us to use laws and principles
of thermodynamics in the development of the CCM theory. CCM theory presented
in this book is the development of the mathematical models of deforming continua
with the limitations of these assumptions.
Non-classical continuum mechanics on the other hand considers continuum the-
ories that are outside the scope of CCM. Couple stress theories, theories based on
internal rotations and Cosserat rotations and non-local theories constitute the sub-
jects of study in NCCM. Non-classical continuum theories are aimed at correcting
some deficiencies of CCM theories and are designed to provide more enhanced con-
tinuum theories that may be better suited for advanced materials and applications.
This classical continuum mechanics book is the second of the Advanced me-
chanics of continua. This book, like its predecessor, is designed to be an advanced
graduate-level study of the basic concepts, principles and the resulting theories
within the umbrella of classical continuum mechanics. The edition retains some
of the basic material from the first edition but in more compact form. Significant
amount of new and more advanced material has been introduced in this edition to
improve and bring completeness to the concepts, principles and theories contained
in the book. Author’s teaching of the subject at the University of Kansas and his
extensive research and publications in NCCM theories has had profound impact in
preparing this highly focused manuscript.
More advanced concepts and theories are introduced as a natural progression
to the simple and commonly used concepts and theories. Curvilinear coordinates
and transformation follow orthogonal frames and orthogonal transformation before
tensors are introduced using curvilinear frames. Throughout the book, covariant
and contravariant bases and measures in Lagrangian and Eulerian descriptions are
clearly distinguished in the kinematics of deformation and stress measures. The or-
thogonal transformations are now introduced as Galilean as well as Euclidean or
non-Galilean and their significance and applications in the convected time deriva-
tives of covariant and contravariant tensors are presented and illustrated. Indepen-
dent derivations of convected time derivatives of co- and contra-variant tensors of
rank two are presented using non-Galilean orthogonal transformation. Objective
tensors of various ranks are defined and the objectivity of various kinematic and
stress measures are established using Galilean and non-Galilean transformations.
Objectivity of rates is derived using Galilean and non-Galilean transformations.

xvii
xviii PREFACE

Integral as well as differential or local form of the conservation and balance laws
(CBL) are derived in Eulerian description. Assumptions are necessary to derive the
differential form of the CBL are clearly stated and discussed. It is shown that the
integral form of the CBL in Lagrangian description is only possible for conservation
of mass; thus, the differential forms of the CBL in Lagrangian description are only
derived within the restriction of the localization theorem. The total deformation
of a deforming volume of matter is decomposed into volumetric and distortional
deformations. It is shown that volumetric deformation physics is independent of
the matter and its constitution and can be described by the constitutive theory for
equilibrium stress tensor (Chapter 8). This has enabled significant reduction and
streamlining the constitutive theory derivations in Chapters 9–14. Role of energy
methods, the principle of virtual work and the fundamental lemma of the calculus
of variations for continuous matter are considered in Chapter 16. It is shown that
CBL of CCM, mathematical classification of differential operators in the CBL and
the elements of the calculus of variations is a complete mathematical framework
for all BVPs and IVPs and that the energy methods and the principle of virtual
work are very limited and extremely restricted subset of this framework. It is shown
that Hamilton’s principle can only be realized within the assumptions of neglecting
integrals on the open boundary of the space-time domain of the IVPs, hence it
is approximate. Append A provides a list of generators and invariants for various
combinations of argument tensors. Appendix B is helpful in transforming informa-
tion and mathematical models from Cartesian to cylindrical or spherical coordinate
systems. Reorganization of the material in the first edition and the new material in-
troduced in this edition provides more comprehensive and more complete treatment
of CCM concepts, principles and theories.
The author is grateful to many of his past and present graduate students: Dr.
Daniel Nunez, Dr. Tristan Moody, Yusshy Mendoza, Dr. Aaron Joy, Dr. Michael
Powell, Sai Mathi, Celso Carranza, Stephen Long, Jacob Kendall, Elie Abboud,
Michael Kitchen, Thomas Ezell whose Ph.D. and M.S. thesis research in various
areas of CCM and NCCM have helped me immensely in bringing the subject matter
of this book to its present level of maturity. I am thankful to my graduate students
Sai Mathi, Celso Carranza, Michael Kitchen, Thomas Ezell, and Elie Abboud for
their efforts in proofreading the manuscript of the book. My very special thanks to
my current Ph.D. student Celso H. Carranza for his interest in the subject, for his
own research on NCCM, for his commitment and interest in this book project, hard
work, for many discussions and suggestions and above all, for typing and retyping
many times the book manuscript, single-handedly to bring it to completion. My
special thanks are also to my current Ph.D student Sai Mathi whose intellectual
curiosity and enormous appetite for new knowledge have always encouraged me to
seek new frontiers in my research work and writings. I am very thankful for his
contribution in various areas of CCM and NCCM that have resulted in significant
growth in the research activities of my research group at the University of Kansas.
This book contains many involved equations, derivations and mathematical de-
tails and it is hardly possible to avoid some typographical and other errors. The
Author would be grateful to those readers who are willing to draw attention to the
errors using the email: kssurana@ku.edu

Karan S. Surana, Lawrence, KS


ABOUT THE AUTHOR

Karan S. Surana, born in India, went to undergraduate school at Birla Insti-


tute of Technology and Science (BITS), Pilani, India, and received a B.E. degree
in Mechanical Engineering in 1965. He then attended the University of Wisconsin,
Madison, where he obtained M.S. and Ph.D. degrees in Mechanical Engineering
in 1967 and 1970, respectively. He worked in industry, in research and develop-
ment in various areas of computational mechanics and software development, for
fifteen years: SDRC, Cincinnati (1970–1973), EMRC, Detroit (1973–1978); and
McDonnell-Douglas, St. Louis (1978–1984). In 1984, he joined the Department of
Mechanical Engineering faculty at the University of Kansas, where he is currently
the Deane E. Ackers University Distinguished Professor of Mechanical Engineering.
His areas of interest and expertise are computational mathematics, computa-
tional mechanics, and continuum mechanics. He is the author of over 350 research
reports, conference papers, and journal articles. He has served as advisor and
chairman of 50 M.S. students and 25 Ph.D. students in various areas of Computa-
tional Mathematics and Continuum Mechanics. He has delivered many plenary and
keynote lectures in various national and international conferences and congresses on
computational mathematics, computational mechanics, and continuum mechanics.
He has served on international advisory committees of many conferences and has co-
organized mini-symposia on k-version of the finite element method, computational
methods, and constitutive theories at U.S. National Congresses of Computational
Mechanics organized by the U.S. Association of Computational Mechanics (US-
ACM). He has organized mini-symposium on classical and non-classical continuum
mechanics at SES (Society of Engineering Science). He is a member of the Interna-
tional Association of Computational Mechanics (IACM) USACM, SES, and a fellow
and life member of ASME.
Dr. Surana’s most notable contributions include: large deformation finite ele-
ment formulations of shells, the k-version of the finite element method, operator
classification and variationally consistent integral forms in methods of approxima-
tions for BVPs and IVPs, and ordered rate constitutive theories for solid and fluent
continua. His most recent and present research work is in non-classical contin-
uum theories for solid and fluent continua and associated constitutive theories. He
is the author of recently published textbooks: Advanced Mechanics of Continua,
CRC/Taylor & France, The Finite Element Method for Boundary Value Prob-
lems: Mathematics and Computations, CRC/Taylor & Francis, The Finite Element
Method for Initial Value Problems: Mathematics and Computations, CRC/Taylor
& Francis, and Numerical Methods and Methods of Approximation in Science and
Engineering, CRC/Taylor & Francis.

xix
LIST OF ABBREVIATIONS

BAM : Balance of Angular Momenta


BLM : Balance of Linear Momenta
CCM : Classical Continuum Mechanics
CBL : Conservation and Balance Laws
CM : Conservation of Mass
FLT : First Law of Thermodynamics
GM : Galerkin Method
GM/WF : Galerkin Method with Weak form
LSM : Least Squares Method
LSP : Least Squares Process
NCCM : Non-classical Continuum Mechanics
PGM : Petrov Galerkin Method
SLT : Second Law of Thermodynamics
STGM : Space-time Galerkin Method
STGM/WF : Space-time Galerkin Method with Weak Form
STLSM : Space-time Least Squares Method
STLSP : Space-time Least Squares Process
STVC : Space-time Variationally Consistent
STVIC : Space-time Variationally Inconsistent
STWRM : Space-time Weighted Residuals Method
TVF : Thermoviscous Fluid
TVEF : Thermoviscoelastic Fluid
TES : Thermoelastic Solid
TVES : Thermoviscoelastic Solid
TVESM : Thermoviscoelastic Solid with Memory
VC : Variationally Consistent
VIC : Variationally Inconsistent
WRM : Weighted Residuals Method

xxi
1
INTRODUCTION

Continuum mechanics is the study of the physics of deformation of continuous


matter constituting physical processes. All deforming physical processes are evolu-
tions in which the states of the processes continuously change as time elapses. All
processes consist of matter; hence the evolutions of the processes are, in fact, the
evolutions of the states of the matter constituting the processes. The mathematical
descriptions or mathematical models describing the evolving state of the deforming
matter as time elapses are called initial value problems. Continuum mechanics is
the study of the development of mathematical models and the evolutions described
by them for the continuously evolving state of the deforming continuous matter. In
continuum mechanics, we consider methodologies, concepts, theories and principles
that are applicable to a large class of matter with minimum possible assumptions
and approximations that may be application or problem-specific. This is obviously
the most general approach to learn about the fundamentals of the entire subject. In
the following, we give basic definitions that will help us in understanding the scope
of study in this book.

1 Def: Mechanics

Mechanics is a much broader study in which we consider motion of nondeforming


matter as well as the matter with motion and deformation. When a process or ob-
ject is disturbed its state changes. The measures of the changes in the state of the
process or the matter or the object may be in terms of the motion, deformation,
velocity, acceleration, density, temperature, etc. Mechanics is the study of such
measures of the evolving state of the processes. This field is obviously too vast and
the study of all such aspects is beyond the scope of this book.

2 Def: Classical mechanics

Classical mechanics is the study of deforming and nondeforming objects or processes


in which the material particles or points only have three translational degrees of
freedom. Hence, in the studies of the deformation of continuous matter in classi-
cal mechanics, the influence of the rotation and rotation rates and their gradients
between neighboring material particles or points is neglected in the mathematical
description of the physics of deformation.

3 Def: Continuum mechanics

Stable matter at the smallest scale consists of molecules that further consist of
atomic and subatomic particles. When the matter is disturbed the changes in
the matter occur at all scales including molecular, atomic and subatomic scales.

DOI: 10.1201/9781003105336-1 1
2 INTRODUCTION

However, there are many aspects of everyday experience regarding the behavior of
matter that can be described and represented accurately with theories that pay no
attention to the molecular structure of the matter or the structure of the matter
at atomic and subatomic scales. The theory that describes the gross phenomenon
of matter, neglecting the structure of the matter at a smaller scale, is known as
continuum theory. The constitution of the matter considered at the molecular level
is obviously not continuous as it consists of discrete molecules separated by the
mean free path. The continuum mechanics consisting of continuum theories consid-
ers matter to be indefinitely divisible. Thus, in this theory, one accepts the concept
of infinitesimal volume of matter referred to as a particle of the continuum. In every
neighborhood of a particle, there are infinitely many particles present.
Whether the continuum theory is valid or not depends upon a given situation
at hand. However, simple guidelines may be considered based on the molecular
dimensions and the mean free path between the molecules. As a general rule, if the
dimensions of the matter are between two to three orders of magnitude compared to
the molecular dimensions or mean free path or larger, we can consider the continuum
theory to be valid in describing the behavior of the matter.
Thus, continuum mechanics as a mathematical idealization of the matter is ap-
plicable to situations in which the fine-scale structure of the matter and its behavior
is ignored. When the behavior of the fine-scale structure of the matter is important,
we must use the principles of particle physics, statistical mechanics or theories that
specifically address smaller scale physics.
All theories are founded on some assumptions that are fundamental in the de-
velopment of the theories. Continuum mechanics theories are no exception to this.
Recent developments in the continuum mechanics theories warrant that we make
some distinction between different continuum theories that consider varied deforma-
tion physics in the studies of continuous media. Broadly speaking, we can classify all
continuum mechanics theories into two categories: classical continuum mechanics
(CCM) and non-classical continuum mechanics (NCCM).

4 Def: Classical Continuum mechanics

The classical continuum mechanics theories are founded on the basic assumption
that a material point of continua has only three translation degrees of freedom, dis-
placements and the corresponding work conjugate quantities are forces. Thus, these
theories do not consider the influence of rotations and rotation rates and their gradi-
ents between the material points in describing the physics of deformation. Rotations
and moments in classical continuum mechanics are consequences of displacements
and forces acting on material points separated by a finite distance, hence these are
not independent of displacements and forces. In CCM theories, displacements and
forces at material points must coexist in a deforming continua.

In CCM, a material point has mass, but it is assumed to have no rotational


inertial properties. Therefore, a material point only experiences linear momenta
and acceleration forces due to velocities and accelerations. However, if the defor-
mation physics results in rotations, angular velocities and angular accelerations at
a material point, in CCM, there is no mechanism for incorporating these in the
mathematical description of the deforming continua. Likewise, the assumption of
INTRODUCTION 3

additional unknown rotational degrees of freedom at a material point is not possible


in CCM. The basic foundation of CCM is based on coexistence of displacements
and forces only at each material point. Any other physics that is not supported by
these conjugate quantities is beyond the scope of CCM.

5 Def: Non-classical Continuum mechanics

Broadly speaking, all continuum theories that fall outside the scope of classical
continuum mechanics can be referred to as non-classical continuum theories. These
constitute the subject of non-classical continuum mechanics. Couple stress theo-
ries, theories based on rotations due to deformation gradient tensor (internal rota-
tions), and theories based on unknown Cosserat rotations as additional degrees of
freedom at material points and non-local theories [116, 117] are some examples of
non-classical continuum theories. The couple stress theories or the theories based
on internal rotations begin with classical continuum theories but additionally incor-
porate influence of internal rotations in the development of the theories arising due
to deformation gradient tensor (neglected in CCM). In these theories, a material
point only has translational degrees of freedom as in CCM due to the fact that
the internal rotations are completely defined by the components of the deformation
gradient tensor, hence cannot be degrees of freedom at a material point. In the non-
classical theories based on Cosserat rotations, a material point has three unknown
rotational degrees of freedom at a material point about the axes of a triad located
at the material point in addition to three translational degrees of freedom. Many
variations of these theories exist that consider internal and/or Cosserat rotations.
In the non-local theories, stress at a material point is assumed to depend upon
the deformation physics at the material point as well as the stresses at the mate-
rial points in its neighborhood. The non-classical continuum theories are believed
to provide a more enhanced mathematical framework that may be meritorious for
modern engineered materials compared to classical continuum theories.
Most textbooks and writings currently available under the title ‘continuum me-
chanics’ only address classical continuum mechanics. This book is also devoted to
the study of classical continuum mechanics, but the distinction between CCM and
NCCM is highlighted throughout this book. In addition, limitations of CCM are
also discussed in various chapters of the book wherever appropriate.

6 Def: Homogeneous matter or Homogeneity

If the physical properties of the material in volume V do not depend upon the
position x of the material points, then the material is called homogeneous. In this
case, for material property P , the following holds
e
x) = P0 ∀ x ∈ V, P0 is constant
P (x
e
7 Def: Heterogeneous or Inhomogeneous or non-homogeneous matter

When P at each material point x ∈ V is not the same, the material is called
e
inhomogeneous or non-homogeneous.
4 INTRODUCTION

8 Def: Isotropic matter

In isotropic matter, the material properties are the same in all directions at a
material point. Thus, isotropic matter can be homogeneous or inhomogeneous.

9 Def: Anisotropic or non-isotropic matter

Anisotropic matter has different material properties in different directions at


each material point. Thus, anisotropic matter can be homogeneous or inhomoge-
neous.

10 Def: Isentropic thermodynamic process

A thermodynamic process in which the entropy density η remains constant is


called isentropic, thermodynamic process.

11 Def: Isothermal thermodynamic process

A thermodynamic process in which temperature θ remains constant is called


isothermal thermodynamic process.

12 Def: Non-isothermal thermodynamic process

A thermodynamic process in which temperature θ does not remain constant is


called non-isothermal, thermodynamic process.

13 Def: Rigid deformation

The deformation of a volume of matter V is said to be rigid if the distance


between any pair of material points remains unchanged during deformation.

14 Def: Potential deformation

When the displacement field can be derived from a potential (the mean rotation
vanishes), the deformation is called potential deformation. For potential deforma-
tion, the necessary and sufficient condition is
J)
S r = grad(J
S r is right stretch and J is deformation gradient.

15 Def: Isochoric deformation, incompressible matter

The deformation under which volume V is preserved after deformation is called


isochoric deformation. The necessary and sufficient conditions for isochoric defor-
mation is given by
J| = 1 ;
|J IIIC[0] = 1
INTRODUCTION 5

The deformation in incompressible matter is isochoric, thus, no volumetric defor-


mation.

16 Def: Compressible matter

Compressible matter has non-zero volumetric deformation, hence results in change


of density for a deforming volume. In the Lagrangian description, conservation of
mass describes density as a function of volumetric deformation.
x) = |J
ρ0 (x J |ρ(x
x, t)

17 Assumptions in CCM theories

In the development of the classical continuum mechanics theory, we employ three


fundamental assumptions: (1) The behavior of the matter at fine scale is neglected.
Instead, we employ the concept of infinitesimal volume referred to as a material
particle or a material point. If we assume the deforming matter to be homogeneous
and isotropic, then the continuum mechanics theory developed for a material point
is valid for the entire deforming volume of the matter. This assumption permits
differential form of the conservation and balance laws. This restriction of homo-
geneity and isotropy can be viewed in a different way by using localization theorem
(Chapter 6), however, the issue of when the continuity of integrand can be ensured
everywhere over the volume of deforming matter remains unresolved if we do not
assume the matter to be homogeneous and isotropic. (2) A material point has only
translational degrees of freedom (displacements). The displacements and forces are
work conjugate and always coexist in deforming continua. (3) In classical continuum
mechanics, we assume that the deforming matter is always in thermodynamic equi-
librium. This assumption allows us to employ conservation and balance laws and
thermodynamic principles in the development of the classical continuum mechanics
theory. The development of the classical continuum mechanics theory presented in
this book is founded on these three fundamental assumptions. Thus, within the
limitations of these three assumptions, the classical continuum mechanics theory is
the development of the mathematical descriptions of the deforming continua.
First, concise, clear, compact and unambiguous notations are necessary for giv-
ing a mathematical form to the physics of the deforming matter. Using these no-
tations we construct basic definitions of quantities related to the deforming matter
and consider various mathematical operations with these quantities. This mate-
rial provides the most fundamental and basic elements to define the kinematics of
motion of the deforming matter. Basic notations, concepts, orthogonal coordinate
systems, orthogonal coordinate transformation, curvilinear coordinate system: co-
and contra-variant bases, induced transformations, the definition of co- and contra-
variant tensors of various ranks, tensors in Cartesian x-frame, orthogonal transfor-
mation of tensors in x-frame, invariants of tensors of various ranks and some other
useful relations are presented in Chapter 2. Galilean and Euclidean (non-Galilean)
transformations are also introduced in Chapter 2. Kinematics of motion, deforma-
tion and their measures for finite deformation are considered in Chapter 3. The
change in length of a material line, change in lengths of an orthogonal group of ma-
terial lines and the change in angle between them are the most fundamental aspects
6 INTRODUCTION

of the physics of deformation. Various definitions and measures are introduced that
eventually lead to measures of finite strain. The definitions of strains are presented
and their measures are derived using co- and contra-variant bases resulting in co-
and contra-variant measures of strains that are presented in Lagrangian as well as
Eulerian descriptions. These definitions and measures are somewhat a departure
from the conventional approach used presently, in which strain measures are classi-
fied as either Lagrangian or Eulerian without regard to the basis. Establishing the
tensorial nature of strain measures, influence of the change of frame on the strain
measures, physical meaning of the components of the strain tensors, polar decompo-
sition of strain tensors into stretch and rotation tensors, deformed and undeformed
areas and volumes that are related to the kinematics of motion and deformation are
all considered in Chapter 3. Definitions and measures of stress for finite deforma-
tion and finite strain, as well as for small deformation, small strain are considered
in Chapter 4. Rate of deformation, area, volume, strain rate tensors, spin tensors
and convected time derivatives of various stress and strain measures in covariant
and contravariant bases are derived in Chapter 5. Definition of objective tensors is
introduced in Chapter 5. Objective tensors and objective rates are presented using
Galilean and non-Galilean transformations. Convected time derivatives of the con-
travariant and covariant tensors of rank two and Jaumann rates are derived using
non-Galilean transformation in Chapter 5. The mathematical model for a deform-
ing volume of matter consisting of conservation and balance laws (CBL) is derived
in Chapter 6 in the Eulerian description. Development of mathematical model in
Lagrangian description consisting of conservation and balance laws is derived in
Chapter 7.
In this book, we only consider the thermodynamic equilibrium processes. Hence,
the conservation laws and thermodynamic principles must be satisfied by the evolu-
tion of a deforming volume of matter. Since conservation of mass, balance of linear
momenta, balance of angular moment and the first law of thermodynamics are in-
dependent of the specific constitution of the matter, hence are applicable to solids,
liquids as well as gases, i.e., they remain valid for any continuous matter.
Chapter 8 contains axioms of constitutive theory, various approaches of deriving
constitutive theories, use of representation theorem in deriving constitutive theories,
including sample examples of the derivation of constitutive theories for constitutive
tensors of rank two and one. Constitutive theories for finite deformation, finite strain
as well as for small strain, small deformation physics for thermoelastic, thermovis-
coelastic (no memory) and thermoviscoelastic solids with memory are presented in
Chapters 9-11. Constitutive theories for thermoviscous and thermoviscoelastic flu-
ent continua for compressible and incompressible physics are presented in Chapters
12 and 13. Maxwell, Oldroyd-B and Giesekus constitutive models for polymeric liq-
uids are derived in Chapter 13. Constitutive theories for thermohypo-elastic solids
are considered in Chapter 14. The equations of state, specific form of deformation
dependent material coefficients, complete mathematical models in Lagrangian and
Eulerian descriptions including constitutive theories are given in Chapter 15. Chap-
ter 16 presents a short overview and discussion of energy methods, the principle of
virtual work and calculus variations with the objective of establishing relationships
between them. A comprehensive and general framework consisting of CBL, calculus
of variations and differential operator classification provides critical assessment of
the usefulness of energy methods and the principle of virtual work.
INTRODUCTION 7

In the development of the mathematical models as well as in obtaining their


solutions in structural mechanics such as for rods, beams, plates, shells, membranes,
the energy methods and the principle of virtual work have played a significant role
and continue to be used dominantly at present also. In Chapter 16, investigation of
the consistency and validity of the mathematical models and the solution methods
for boundary value problems (BVPs) and initial value problems (IVPs) based on
energy methods and the principle of virtual work for homogeneous and isotropic
and non-homogeneous and non-isotropic solid continua is presented. Specific model
problems are also considered for illustrative purposes.
The entire book uses curvilinear coordinate systems: covariant and contravariant
and fixed Cartesian frames. Use of cylindrical or spherical coordinate systems is in-
tentionally avoided throughout the book: first to minimize unnecessary duplication
and clutter, and second, the details of various measures, conservation and balance
laws, etc. can be easily converted from Cartesian frame to cylindrical coordinate
system or spherical frame using the material provided in Appendix B. Appendix
A provides details of combined generators and invariants for symmetric and skew-
symmetric tensors of rank two as well as tensors of rank one. This material is useful
in deriving constitutive theories using representation theorem.
The material presented in this book provides a unified treatment of the devel-
opment of mathematical descriptions of deforming solids, fluids, polymeric solids
and fluids, both compressible and incompressible with clear distinction between La-
grangian and Eulerian descriptions as well as co- and contra-variant bases. Based
on this short introduction of continuum mechanics, we note that this subject is
mathematically demanding with many involved derivations but is very precise and
rigorous within the assumptions considered. A thorough understanding and working
knowledge of this subject is of paramount importance and significance in prepar-
ing graduate students for fundamental and basic research work in engineering and
sciences.
2
CONCEPTS AND MATHEMATICAL
PRELIMINARIES

2.1 Introduction
The material contained in this chapter is elementary but its thorough grasp and
good working knowledge are essential in understanding and learning the material
contained in the remaining chapters. First, we consider notations. The main purpose
of notations is to present information related to the deforming matter in mathemati-
cal form. The notations must be compact, concise and consistent so that abstraction
and yet accuracy of the actual information may be maintained in the mathematical
description of the information. The commonly used notations in continuum mechan-
ics are Einstein notations, index notations, matrix and vector notations. Generally,
it is a matter of preference as to which notation one uses. However, there are some
advantages in matrix and vector notations, especially when the solutions of mathe-
matical models resulting from continuum mechanics principles describing evolutions
are obtained numerically using computational mechanics and computational math-
ematics. In this chapter, we consider and use all three notations. In this book, we
do not strictly adhere to one specific form of notations but rather consider a mix of
these for maintaining clarity of presentation and ease of understanding, keeping in
mind that this material must be easily reproducible in the classroom environment
as well.

2.2 Notations
Details of Einstein notations, index notations and vector and matrix notations
are presented in the following sections.

2.2.1 Einstein notations


The basic elements of the Einstein notation are: summation convention, dummy
index (or indices) and free index (or indices).

2.2.1.1 Summation convention


Based on Einstein’s summation convention, whenever an index is repeated once
(and no more) in the same term or one or more terms in an equation, it implies
summation over the specified range of the index. Thus, in such cases, there is no
need to write the expanded form of the sum or use the summation sign. If the index
is not specified, a value of 3 is assumed.
n
X n
X
α = N1 φ1 + N2 φ2 + · · · + Nn φn = Ni φi = Nj φj
i=1 j=1 (2.1)
= Nk φk = Nl φl ; k, l = 1, 2, . . . , n

DOI: 10.1201/9781003105336-2 9
10 CONCEPTS AND MATHEMATICAL PRELIMINARIES

3 X
X 3 3 X
X 3
β= Nij φi φj = Nkl φk φl = Nij φi φj = Nkl φk φl (2.2)
i=1 j=1 k=1 l=1
3 X
X 3 X
3
γ= Nijk φi φj φk = Nijk φi φj φk = Nlmn φl φm φn (2.3)
i=1 j=1 k=1
3 X
X 3 X
3
η= pi qj rk φi φj φk = pi qj rk φi φj φk = pl qm rn φl φm φn (2.4)
i=1 j=1 k=1
3
X 3
X
κ= Ni φi + Pj φj = Ni φi +Pj φj = Ni φi +Pi φi = Nk φk +Pl φl (2.5)
i=1 j=1
N1 φ 1 + N2 φ 2 + N3 φ 3 = Ni φ i = Nj φ j (2.6)
Nii = N11 + N22 + N33 = Njj (2.7)
Equations (2.2)–(2.7) are examples of the use of Einstein’s summation convention
for compact representation of the terms in the sum in each case. Equations (2.1)
and (2.3) contain a single summation term, Equations (2.2) and (2.4) are single
summation terms and (2.5) is an equation containing two summation terms. In
some instances, there may be a need to have an index repeated more than once
in the same term. In such cases, Einstein’s summation convention cannot be used
and hence we must retain the summation sign for it to be meaningful. Three dots
(no more, no less) in (2.1) represent (n − 3) remaining terms. This is a standard
notation in continuum mechanics. Based on Einstein notation, the occurrence of
three (or more) identical indices in each term of an expression necessitates the use
of summation signs or symbols.
3
X
p1 q1 r1 + p2 q2 r2 + p3 q3 r3 = pi qi ri 6= pi qi ri (2.8)
i=1

2.2.1.2 Dummy index and dummy variables

From (2.1), we note that the sum is independent of i and j. These are called
dummy indices. In (2.2), i, j and k, l are dummy indices. Likewise, i, j, k and l,
m and n are dummy indices in (2.3) and (2.4). Equation (2.1) is an example of
a single dummy index, (2.2) contains two dummy indices as it is a double sum,
whereas (2.3) and (2.4) are examples of the use of three dummy indices due to the
fact that they are a triple sum. Thus, dummy indices are like dummy variables in
the integrals.
Zq Zq Zq Zq
G= g(x)dx = g(y)dy = g(z)dz = g(t)dt (2.9)
p p p p

In (2.9), the integrals are independent of the variables x, y, z and t. These are
called dummy variables. Just like the choice of specific dummy variables does not
influence the integral G, the choice of a specific variable for a dummy index also does
not influence the meaning of term or terms in an equation containing the dummy
index or indices.
2.2. NOTATIONS 11

2.2.1.3 Free indices

Free indices appear in equations. A free index appears only once in each term
of the equation or expression. However, by using more than one free index, we
can represent a set of equations as each free index will describe a single equation.
Consider
pi + qi = ri (2.10)
pi + qij bj = ri (2.11)
φ1 = N11 α1 + N12 α2 + N13 α3 = N1i αi = N1j αj
φ2 = N21 α1 + N22 α2 + N23 α3 = N2i αi = N2j αj (2.12)
φ3 = N31 α1 + N32 α2 + N33 α3 = N3i αi = N3j αj
or
φi = Nij αj
φi = Nik αk (2.13)
φl = Nlm αm
3
X 3
X
σij = Rik Rjk = Rik Rjk = Ril Rjl = Ril Rjl (2.14)
k=1 l=1

Index i in (2.10) and (2.11) is a free index as it appears once in each term of
the equation. Index j in (2.11) is a dummy index. Equation (2.12) is a system of
three equations in which j is a dummy index. These can be represented in compact
form using (2.13) in which i, l are free indices but j, k and m are dummy indices.
Equation (2.14) represents a system of nine equations. For each value of i and j,
the sum over the dummy index k or l yields an equation.

2.2.2 Index notations and Kronecker delta


In the index notation, we consider a suitable coordinate frame and represent
quantities of interest in this frame using the basis of the frame. We generally define a
fixed orthogonal Cartesian coordinate system by ox1 , ox2 , ox3 orthogonal directions,
ox1 x2 x3 for short, referred to as x-frame, o being the location or origin of the
coordinate system. The use of x1 , x2 , x3 as opposed to x,y,z is intentional so that
Einstein’s notation can be employed for compact representation of information. We
assume that e i ; i = 1, 2, 3 are unit vectors along ox1 , ox2 and ox3 axes, i.e., unit
vectors e i form an orthonormal basis in the x-frame. If φ is a vector in x-frame
with components φ1 , φ2 , φ3 along ox1 , ox2 and ox3 , then φ has the following
representation in the x-frame:

φ = e 1 φ1 + e 2 φ2 + e 3 φ3 = e i φi = ~ei φi (2.15)
We note that e i form an orthonormal basis. The following hold
e1 · e1 = e2 · e2 = e3 · e3 = 1
(2.16)
e1 · e2 = e2 · e1 = e3 · e1 = 0

and (2.16) can be written in more compact form using a new notation
12 CONCEPTS AND MATHEMATICAL PRELIMINARIES


1 if i=j
e i · e j = δij = (2.17)
0 if i 6= j
where δij is called the Kronecker delta.
If ψ is another vector in x-frame with projections ψ1 , ψ2 , ψ3 along ox1 , ox2 and
ox3 axes, then
ψ = e j ψj (2.18)

Dot product of vectors φ and ψ is defined by


ei φi ) · (e
φ · ψ = (e e j ψj ) (2.19)
and by expanding the right side of (2.19) using (2.17), we can obtain

φ · ψ = e i φi · e j ψj = e i · e j φi ψj = δij φi ψj = φi ψi (2.20)
If we recognize the following properties of the Kronecker delta, δij , then we can
make use of δij in many different compact representations. From (2.16) and (2.17),
we note that
δ11 = δ22 = δ33 = 1
(2.21)
δ12 = δ21 = δ13 = δ31 = δ23 = δ32 = 0
We also note that
δii = δ11 + δ22 + δ33 = 3 (2.22)
δij φj = φi (2.23)
δim δmj = δij (2.24)
δim δmk δkj = δij (2.25)
δim σmj = σij (2.26)

2.2.3 Vector and matrix notations

When the scalars are arranged in the form of a column we may refer to this
column as a vector. From Equation (2.12), φ1 , φ2 and φ3 can be arranged as
 
φ1 
{φ} = φ2 or φ ~ or φ (2.27)
 
φ3
where curly brackets, boldface character or a character with an over arrow are stan-
dard continuum mechanics notation for vectors. At this stage a vector is an ordered
arrangement of quantities and has no other meaning. Just like φi , αi appearing in
(2.12) can also be arranged in the form of a column or a vector.
 
α1 
{α} = α2 or α ~ or α (2.28)
 
α3
2.2. NOTATIONS 13

Nij appearing in (2.12) or (2.13) can be arranged in rows and columns of a two-
dimensional ordered arrangement, a matrix [N ]
 
N11 N12 N13
[N ] = N21 N22 N23  (2.29)
N31 N32 N33

where Nij refers to an element of [N ] located at row i and column j. Using the
notation for vectors in (2.27) and (2.28) and for matrix in (2.29), we can also write
(2.13) as

{φ} = [N ]{α} ~ = [N ]~
or φ α (2.30)
Equations (2.13) and (2.30) are exactly equivalent but written using two different
notations. In this book, we employ both notations.

Transposition of a vector, a matrix and vector, matrix operations, etc. are ele-
mentary and hence are not discussed here.

2.2.4 Remarks
Using Einstein’s notations, index notations and matrix and vector notations, we
can represent various quantities in many different desired forms. Some examples are
given in the following.

(a) The components of the Kronecker delta δij in (2.16) can be arranged in matrix
form (i being row and j being column), then we have identity matrix [I]
   
δ11 δ12 δ13 100
δ21 δ22 δ23  = 0 1 0 = [I] (2.31)
δ31 δ32 δ33 001

(b) Using vector and matrix notations (2.24)–(2.26) can be written as

[I][I] = [I] (2.32)


[I][I][I] = [I] (2.33)
[I][σ] = [σ] (2.34)

(c) We can also write (2.14) as follows using vector and matrix notations
[σ] = [R][R]T (2.35)

in which [R]T is the transpose of [R].


s is a line segment with length ds and components dx1 , dx2 and dx3 in
(d) If ds
x-frame, then
 
dx1 
~ = dx2
{ds} = ds or dss = e i dxi (2.36)
 
dx3

Let (ds)2 be the square of the length of ds


ds, then
14 CONCEPTS AND MATHEMATICAL PRELIMINARIES

(ds)2 = ds s = e i dxi · e j dxj = e i · e j dxi dxj = dxi dxi = {ds}T {ds}


s · ds
(2.37)
Alternatively, since dxi = δij dxj
(ds)2 = dxi (δij dxj ) = δij dxi dxj = dxi dxi (2.38)

(e) If h = h(x1 , x2 , x3 ), then in continuum mechanics it is standard notation to


use
∂h
= h,i (2.39)
∂xi
thus, using (2.39), total differential dh of h(x1 , x2 , x3 ) can be written as
∂h
dh = dxi = h,i dxi (2.40)
∂xi

(f) In continuum mechanics, it is standard practice to use 1, 2, 3 to represent x1 ,


x2 , x3 . This leads to compact representation when 1, 2, 3 are used instead of
x1 , x2 , x3 as subscripts or superscripts. For example, instead of σx1 x1 , σx1 x2 ,
σx2 x3 . . . , we could simply use σ11 , σ12 , σ23 . . .

2.3 Permutation tensor


The permutation tensor or the permutation symbol (also called alternating ten-
sor or Levi-Civita tensor) is denoted by ijk in which if we consider counterclockwise
order of ijk to be even, then it is defined as
   
 1  an even
ijk = −1 if i, j, k are an odd permutation of 1, 2, 3 (2.41)
   
0 not a
If we arrange 1, 2, 3 in counter-clockwise fashion, then the counter clockwise order
of 1, 2, 3 permutations: 1 2 3, 2 3 1, 3 1 2 are even permutations of 1, 2, 3, hence
123 , 231 , 312 each has a value of 1. On the other hand, the permutations: 1 3 2, 3
2 1, 2 1 3 of 1, 2, 3 are clockwise permutations; thus, odd permutations, hence 132 ,
321 , 213 each has a value of −1. When two or more indices are the same in ijk ,
then it is not a permutation, hence its value is 0. This gives rise to the following:
123 = 231 = 312 = 1 (cyclic order : counter clockwise)
132 = 213 = 321 = −1 (cyclic order : clockwise) (2.42)
111 = 211 = 133 = · · · = 0

or in general we can write


ijk = jki = kij = −jik = −kji = −ikj
The permutation tensor is helpful in compact representation in the cross product
2.4.  -δ
δ IDENTITY 15

2 3

of vectors, determinant of a matrix, etc. If we consider orthonormal basis e i , then


e1 × e2 = e3 ; e2 × e1 = −e e3
e2 × e3 = e1 ; e1
e 3 × e 2 = −e (2.43)
e3 × e1 = e2 ; e2
e 1 × e 3 = −e
which can be represented in compact form using the permutation symbol

e i × e j = ijk e k (2.44)
This property in (2.44) can be applied to cross products of vectors.
φ × ψ = e i φi × e j ψj
= ei × ej φi ψj (2.45)
= ijke k φi ψj = e k ijk φi ψj

Consider a 3 ×3 matrix [J]. Let |J| or det[J] be the determinant of [J]. Then, det[J]
can be obtained by using Laplace expansion. If we use the first column of [J], then
we have
|J| = J11 (J22 J33 −J32 J23 )−J21 (J12 J33 −J32 J13 )+J31 (J12 J23 −J22 J13 )
= J11 (1jk Jj2 Jk3 ) − J21 (−2jk Jj2 Jk3 ) + J31 (3jk Jj2 Jk3 ) (2.46)
= Ji1 (ijk Jj2 Jk3 ) = ijk Ji1 Jj2 Jk3

By considering Laplace expansion using the first row of [J] we can show that
det[J] = ijk J1i J2j J3k (2.47)
This is a compact and convenient way to define det[J].

δ identity
2.4  -δ
The product of permutation tensors may be expressed in terms of the Kronecker
δ identity.
delta. This is known as  -δ
ijm klm = δik δjl − δil δjk (2.48)
Due to the sign change property of  for odd permutation, the following holds
ijm klm = mij klm = jmi mkl
(2.49)
= ijm mkl = −mji lmk etc.

Using (2.48) we can show that


16 CONCEPTS AND MATHEMATICAL PRELIMINARIES

ilm klm = 2δik (2.50)


and using k = i in (2.50)
ilm ilm = 6 (2.51)

 -δδ is an important identity that is helpful in the subsequent chapters.

2.5 Basic operations using vector, matrix and


Einstein notations
There are four basic operations that are often encountered in the developments
of continuum mechanics theory: multiplication, substitution, factoring and con-
traction when manipulating mathematical expressions. We consider these in the
following.

2.5.1 Multiplication
Suppose we define scalars φ and ψ by
φ = pi qi = {p}T {q} = p · q (2.52)
T
and ψ = ri si = {r} {s} = r · s (2.53)
and wish to consider the product (multiplication) of φ and ψ, then we must make
sure that the dummy index i in (2.52) and (2.53) is not the same. That is, the
dummy index in φ or ψ must be changed before taking their product.

φ ψ = pi qi rj sj = pj qj ri si = ri si pj qj = ri si qj pj (2.54)
Using vector and matrix notation we have
φ ψ = ({p}T {q})({r}T {s}) = (p
p · q )(r
r · s)

We note that in Einstein notations the position of quantities in the product in (2.54)
does not influence the product.

2.5.2 Substitution
Suppose we define φi and qi by
φi = pik qk or {φ} = [p]{q} (2.55)
and qi = rik sk or {q} = [r]{s} (2.56)
In (2.55) and (2.56), i is the free index and k is the dummy index. We wish to
substitute qi from (2.56) into (2.55). Before doing so, we must ensure that the free
index (for q) in (2.56) is the same as the dummy index (for q) in (2.55). Furthermore,
changing i into k in (2.56) will cause conflict with the dummy index k in (2.56).
Thus, dummy index k in (2.56) must be changed to some other index (say l) first,
followed by changing i to k. Let us change (2.56) to
qk = rkl sl (2.57)
2.5. OPERATIONS USING VECTOR, MATRIX AND EINSTEIN NOTATIONS 17

Now we can substitute for qk from (2.57) into (2.55)


φi = pik rkl sl = rkl pik sl = sl pik rkl (2.58)
and in matrix and vector notation we have
{φ} = [p][r]{s}

Once again, we note that matrix and vector notations result in exactly the same
expressions as the use of Einstein notations does, but in Einstein notations, the
position of quantities in the product in (2.58) does not influence the product.

2.5.3 Factoring
Consider the following vector equation:
[σ]{φ} − λ{φ} = 0 (2.59)

Suppose we wish to factor out {φ} from both terms in (2.59). This requires that
the form of the second term in (2.59) must correspond to that of the first term. We
can rewrite (2.59) as follows
[σ]{φ} − λ[I]{φ} = 0 (2.60)

and now we can factor out {φ} in (2.60)


([σ] − λ[I]){φ} = 0 (2.61)
The same operation of factoring can also be done using Einstein notations. First,
we rewrite (2.59).

σij φj − λφi = 0 (2.62)


Since φi can also be written as
φi = δij φj (2.63)
substitution of φi from (2.63) into (2.62) gives
σij φj − λδij φj = 0
(2.64)
or (σij − λδij )φj = 0

We note that (2.64) could have been written directly using (2.61). Once again, we
note the simplicity of operation in matrix and vector notation.

2.5.4 Contraction (or trace)


The operation of contraction involves setting the two indices the same and there-
fore summing over the index. If we consider σij , then the contraction of σij would
be σii .

σii = σ11 + σ22 + σ33 = tr[σ] (2.65)


The sum of the diagonal elements of [σ], tr[σ] is pronounced trace of [σ]. That is,
contraction results in trace. As another example consider
18 CONCEPTS AND MATHEMATICAL PRELIMINARIES

σij = 2µDij + κδij Dkk (2.66)


then
σii = tr σ = tr[σ]
= 2µDii + κδii Dkk
= 2µDjj + 3κDkk
(2.67)
= 2µDkk + 3κDkk
= (2µ + 3κ)Dkk
= (2µ + 3κ)Djj

Alternatively, in the matrix and vector notation, (2.66) can be written as


[σ] = 2µ[D] + κ[I] tr[D] (2.68)
then
tr[σ] = 2µ tr[D] + κ tr[I] tr[D]
= 2µ tr[D] + 3κ tr[D] (2.69)
= (2µ + 3κ) tr[D]

which is the same as (2.67).


2.5.4.1 Basic properties of trace
In the following, we list some basic properties of the trace of matrices and the
trace of the products of matrices: Consider n × n square matrices A and B .
(a) Trace is a linear mapping. That is
A + B ) = tr(A
tr(A A) + tr(B
B)
A) = c tr(A
tr(cA A) ; ∀c ∈ R
AT )
A) = tr(A
tr(A

(b) Trace of a product

AT B ) = tr(A
tr(A AB T ) = tr(B
B T A ) = tr(B
BA T )
AB ) = tr(B
tr(A BA )
tr(AAB ) 6= tr(A
A) tr(B
B)

AB ) = tr A B T
tr (A T

(c) Cyclic property


AB C D ) = tr(B
tr(A BC D A ) = tr(C
C D AB ) = tr(D
D AB C )

That is trace is invariant of cyclic permutations.


AB C ) 6= tr(A
tr(A AC B ).
AC B ) is not a cyclic permutation of (A
Note that (A AB C ).
2.6. REFERENCE FRAME AND REFERENCE FRAME TRANSFORMATION 19

(d) Trace of product of symmetric matrices


tr(A AB C )T ) = tr(C
AB C ) = tr((A C B A ) = tr(A
AC B )

Thus, when the matrices in a product are symmetric, the trace of the product
is invariant of permutation.

Proofs of these can be simply constructed by expanding or writing the expression


using Einstein notations and then contraction.

2.6 Reference frame and reference frame


transformation
Observing or measuring events related to the evolution of a deforming volume
of matter in a fixed frame called reference frame is perhaps most convenient and
common approach. The fixed frame is generally chosen to be orthogonal and may be
Cartesian, cylindrical or spherical. The fixed frame may be viewed as fixed observer.
However, a general case would be observations made by an observer standing on a
platform that is translating and rotating in time. This notion is not far fetched. An
observer standing on the surface of the earth observing an object falling towards
the surface of the earth is an example. Due to translation and rotation of the earth,
the falling object’s path is not a straight line path. In order to determine the correct
evolution of the object’s motion, the observer’s translation and rotation must be
taken into account. Observing the equivalence between the fixed frame and the
fixed observer, here also we could view the translating and rotating observer as
being equivalent to a frame that is translating and rotating.
Referring to Figure 2.1 x-frame is translated to O0 by c 0 (t) and then rotated
by an orthogonal rotation tensor Q to obtain y-frame. The coordinates x and y of
point P are related by
{x} = {c0 (t)} + [Q(t)]T {y} (2.70)
0
or {y} = {c(t)} + [Q(t)]{x} ; {c(t)} = −[Q(t)]{c (t)} (2.71)
[Q(t)] is proper orthogonal rotation tensor, i.e., det[Q(t)] = +1.
Newtonian mechanics is founded on the assumption that distance, time and mass
are absolute, i.e., the motion is in inertial frame, which neglects the translation and
rotation of the observer in time (i.e., motion of the earth on which the observer
is standing). Thus, in Newtonian mechanics, i.e., inertial frame, an object falling
towards the earth’s surface will follow a straight line path if only gravitational
forces are acting on it. Classical continuum mechanics is founded on Newton’s laws
in inertial frame for which (2.71) reduces to
{y} = {c} + [Q]{x} (2.72)

in which [Q] is a proper orthogonal rotation tensor (and is not a function of time).
Equation (2.72) describes a rotated frame y that differs from x-frame by a rigid
orthogonal rotation and a rigid translation {c} (also not a function of time).
20 CONCEPTS AND MATHEMATICAL PRELIMINARIES

y2
−−0→
O P = [Q(t)]T {y} P

x2
y3
O0

x
y1
0
c (t)

x1
O

x3

Figure 2.1: Translation and rotation of reference frame

Remarks

(1) The kinematic quantities like displacements, velocities, and strain measured in
two different frames (i.e., observed by two different observers) will, in general,
differ since the perspectives differ.
(2) In Chapter 5, we consider the change of frames described by (2.71) as well as
(2.72) to determine which kinematic tensors are independent of the frame of
measurement (i.e., the observer), frame invariant or objective.
(3) Since the constitution of the deforming matter must be independent of the
observer (or frame of measurement), only the tensors that exhibit this property
must be used in describing the constitution of the deforming matter (Chapters
8–14). Thus, the constitutive tensors as well as their argument tensors must
be objective.
(4) For now, we consider transformation T described by (2.72) (valid only in
inertial frame).
(5) The coordinate transformation (2.71) or (2.70) in which the translation of the
origin of the coordinate system (or observer position) and the rigid rotation
of the coordinate system (rotation of the observer) are functions of time is
called Euclidean transformation [63] or non-Galilean transformation. When
the translation and the rigid rotation of the frame (or observer) are not func-
tions of time as in (2.72), the transformation is called Galilean transformation.
Newtonian mechanics is based on Galilean transformation.
2.7. COORDINATE FRAMES AND TRANSFORMATIONS 21

2.7 Coordinate frames and transformations


Orthogonal frames such as Cartesian, cylindrical and spherical are commonly
used. In these frames, the axes are orthogonal and the basis (unit vectors along the
orthogonal axes) is used to describe quantities of interest in the frame. Cartesian
frame or x-frame is simplest of the three. The x-frame can be transformed into
y-frame through a coordinate transformation T . In doing so, we also transform the
quantities defined in x-frame to y-frame, referred to as induced transformation. In
continuum mechanics, the nature of transformation T and the nature of induced
transformation on various quantities defined in x-frame are important classes of
studies. We also consider non-orthogonal curvilinear frames in which the coordi-
nate axes are curved and are not orthogonal. An important case is a Cartesian
x-frame that transforms into a non-orthogonal curvilinear frame due to coordinate
transformation T . Just like orthogonal frames, a curvilinear frame has a basis. We
consider details in the following sections.

2.7.1 Cartesian frame and orthogonal coordinate


transformations
For now, we consider x-frame to be Cartesian frame and transformation T (2.72)
to be such that it causes translation and rigid rotation of x-frame into y-frame. For
simplicity, consider a two-dimensional case.
Consider two sets of Cartesian frames in a plane: the x-frame o-x1 x2 and the
y-frame o0 -y1 y2 . If the y-frame is obtained from the x-frame by a shift of the origin
and without a rotation of the axes, then the transformation from the x-frame to
the y-frame is a translation. If the point P has coordinates (x1 , x2 ) and (y1 , y2 ) in
x-frame and y-frame and if (c1 , c2 ) are the coordinates of the origin of y-frame with
respect to x-frame, then
x1 = y1 + c1
(2.73)
x2 = y2 + c2
or in compact form
x i = y i + ci ; i = 1, 2 (2.74)
If the origin remains fixed and the y-frame is obtained by rotating the x-frame
through an angle θ in the counter clockwise direction about a line perpendicular to
the plane of the x-frame but at its origin (see Figure 2.2), then the transformation
relating the frames can be derived as follows. Consider a fixed point P with coor-
dinates (x1 , x2 ) and (y1 , y2 ) in x-frame and y-frame. From Figure 2.2, it is rather
straightforward to see that
x1 = y1 cosθ − y2 sinθ
(2.75)
x2 = y1 sinθ + y2 cosθ

Let us define
   
Q11 Q12 cos θ sin θ
[Q] = = (2.76)
Q21 Q22 − sin θ cos θ
22 CONCEPTS AND MATHEMATICAL PRELIMINARIES

x2 P = P (x1 , x2 ) = P (y1 , y2 )

y2 y1 y2
y2 cosθ θ y1

x2

y1 sinθ

θ
o
x1 y2 sinθ x1

y1 cosθ
Figure 2.2: Rigid rotation of x-frame to y-frame by an angle θ

Then (2.75) can be written as


   
x1 T y1
= [Q] ; xi = xi (yj ) ; i, j = 1, 2 (2.77)
x2 y2

We note that [Q] defines pure rotation of x-frame to y-frame and is orthogonal, i.e.,
[Q]T = [Q]−1 hence [Q]T [Q] = [Q][Q]T = [I] (2.78)
Using (2.78), we obtain the following from either (2.75) or (2.77).
   
y1 x
= [Q] 1 ; yi = yi (xj ) ; i, j = 1, 2 (2.79)
y2 x2

That is, the coordinate transformation T defined by [Q] in (2.79) is such that x-
frame is transformed into y-frame. We note that both x-frame and y-frame are
Cartesian frames and the transformation T defined by [Q] in (2.79) is orthogonal.
Remarks
(1) In the rotation of the x-frame to y-frame, the point P remains fixed. Thus,
if point P is a material point in a volume of matter, then every such point
in the volume of matter has coordinates in x- and y-frames that are related
through (2.79). It is important to note that in this process, the x-frame is
rotated to y-frame, but the volume of matter containing point(s) P remains
undisturbed.
(2) In contrast to (1), if we rotate the volume of matter through a rigid rotation,
then its coordinate changes from xi to yi . Both are measured in the x-frame.
The relationship between the new coordinates yi and the original coordinates
Another random document with
no related content on Scribd:
a capricious, nervous disposition. She has too refined a taste for
that, and her soul is too full of harmony.
Ibsen’s Nora is hysterical, and only half a woman; and that is what
he, with his poetic intuition, intended her to be. Eleonora Duse’s
Nora is a complete woman. Crushed by want and living in narrow
surroundings, there is a certain obtuseness about her which renders
her willing to subject herself to new misfortunes. There is also
something of the child in her, as there is in every true woman; but
even in her child-like moments she is a sad child. Then the
misfortune happens! But, strange to say, she makes no desperate
attempt to resist it; she gives no hysterical cry of fear, as a meaner
soul would do in the struggle for life. There is something pitiable in a
struggle such as that, where power and will are so disproportionately
unlike. Duse’s Nora hastily suppresses the first suggestion of fear;
but she does not admire her muff meanwhile, like Fru Hennings. She
merely repeats to herself over and over again in answer to her
thoughts: “No, no!” I never heard any one say “no” like her; it
contains a whole world of human feeling. But all through the night
she hears fate say “Yes, yes!” and the next day, which is Christmas
Day, she is overcome with a fatalistic feeling. She dresses herself for
the festival, but not with cheap rags like Nora; she wears an
expensive dark green dress, which hangs down in rich graceful folds.
It is her only best dress, and sets off her figure to perfection; it
makes her look tall and slender, but also very weary. And as the play
goes on, she becomes even more weary and more resigned, and
when death comes, there is no help for it. Then, after the rehearsal
of the tarantella, when Helmer calls to her from the dining-room and
she knows that fate can no longer be averted, she leaps through the
air into his arms with a cry of joy,—to look at her one would think that
she was one of those thin, wild, joyless Bacchantes whose bas-
reliefs have come down to us from the later period of Grecian art.
The third act:—Nora and Helmer return from the mask ball. She is
absent-minded and quite indifferent to everything that goes on
around her. That which she knows is going to happen, is to her
already a thing of the past, since she has endured it all in
anticipation; her actions in the matter are only mechanical.
When Helmer goes to empty the letter-box, she does not try to stop
him with a hundred excuses, she scarcely makes a weak movement
to hold him back; she knows that it must come, nothing that she can
do will prevent it. While Helmer reads the letter, she stands pale and
motionless, and when he rushes at her, she throws on her mantle
and leaves the room without another word.
He drags her back and overwhelms her with reproaches, in which
the pitiful meanness of his soul is laid bare. Now Duse’s acting
begins in earnest, now the dramatic moment has come—the only
moment in the drama—for the sake of which she took the part.
She stands by the fireplace, with her face towards the audience, and
does not move a muscle until he has finished speaking. She says
nothing, she never interrupts him. Only her eyes speak. He runs
backwards and forwards, up and down the room, while she follows
him with her large, suffering eyes, which have an unnatural look in
them, follows him backwards and forwards in unutterable surprise,—
a surprise which seems to have fallen from heaven, and which
changes little by little into an unutterable, inconceivable
disappointment, and that again into an indescribably bitter, sickening
contempt. And into her eyes comes at last the question: “Who are
you? What have you got to do with me? What do you want here?
What are you talking about?”
The other letter drops into the letter-box, and Helmer loads her with
tender, patronizing words. But she does not hear him. She is no
longer looking at him. What does the chattering creature want now?
She does not know him at all. She has never loved him. There was
once a man whose sympathy she possessed, and who was her
protector. That man is no more, and she has never loved any one!
She turns away with a gesture of displeasure, and goes to change
her clothes, anxious to get away as quickly as possible. He stops
her. What then? The woman is awake in her. She is a woman in the
moment of a woman’s greatest ignominy,—when she discovers that
she does not love. What does he want with her? Why does he raise
objections? He——? Tant de bruit pour une omelette! She throws
him a few indifferent words, shrugs her shoulders, turns her back
upon him, and goes quickly out at the door. Presently we hear the
front door close with a bang. There is no mention at all about the
“miracle.”
That is how Duse united Nora’s double personality. Make it up! There
is no making it up between the man and wife, except the kiss and the
shrug of the shoulders. She ignores Ibsen’s principal argument.
Reason, indeed? Reason has never settled anything in stern reality,
least of all as regards the relationship between husband and wife.
One day Nora wakes up and finds that Helmer has become
loathsome to her, and she runs away from him with the instinctive
horror of a living person for a decomposed corpse. Of course nothing
“miraculous” can happen, for that would mean that the living person
should go mad and return to the corpse.
Eleonora Duse treats all her parts in the same independent manner
that she treats the text of Nora. When we are able to follow her, and
that is by no means always, we notice how she alters it to suit
herself, how another being comes to the front,—a being who has no
place in the written words, and whom the author never thought of,
whom he, in most cases, could certainly not have drawn from his
own views of life and his own inner consciousness. Duse’s heroine is
more womanly, in the deeper sense of the word, than the society
ladies in Ibsen’s and Sardou’s dramas, and she is not only more
simple than they are, but also far greater. Eleonora Duse is not a
dialectician like Ibsen and Sardou; their hair-splitting logic is no
concern of hers, and it certainly was not written for her. She has an
instinctive, unerring intuition of what the part should be, and she
throws herself into it and acts accordingly. She does not vary much;
she is not a realist who makes a careful note of every little
peculiarity, and arranges them in a pattern of mosaic; she is truthful
to a reckless extent, but not always true to the letter; sometimes like
this, sometimes like that, she differs in the different parts. She is true,
because she is proud and courageous enough to show herself as
she really is. There is no need for her to be otherwise. There is
danger of uniformity in this great simplicity of hers, and she would
not escape it if it were not for her emotional nature, and an intense,
almost painful sincerity, which was perhaps never represented on
the stage before her time, and which was certainly never before
made the groundwork of a woman’s feelings. She comes to meet us
half absorbed in her own thoughts, a complete woman,—complete in
that indissoluble unity which is the basis of a healthy woman’s
nature: woman-child and also woman-mother, a woman with the
stamp which is the result of deep, vital experience, with a woman’s
tragedy ineffaceably engraved on every feature,—this same
woman’s tragedy which she reproduces upon the stage. It is the fact
of her not troubling herself about anything else that imbues her
acting with an air of simplicity, and because she is such a complete
woman herself, there is an air of indescribable stateliness about her
acting. She not only simplified all that she took in hand, but she also
improved it. For all these characters which she created were the
result of the completeness of her womanly nature, and that is why
they never had but the one motive, for all the evil they did, and for
their hate: they revenged themselves for the crimen læsæ
majestatis, which sin was committed against their womanly nature,
and which a true woman never forgives, as when the priceless pearl
of her womanhood has been misused. That is why they made no
pathetic gestures, no noise or tragic screams, but acted quietly and
silently, as we do a thing which is expected of us, with a quiet
indifference, as when intact nature bows itself under and assists fate.
That is how Duse acted Nora, but she acted Clotilde in “Fernande” in
the same mood, also Odette in the play, called by the same name,
both by Sardou, and that was more difficult. Clotilde and Odette are
a couple of vulgar people. Clotilde, a widow of distinction, revenges
herself upon a young man of proud and noble family, who has been
her lover for many years, but has broken his marriage vows, by
encouraging his attachment for a dishonored girl, whom she
persuades him to marry, and afterwards triumphantly tells him his
wife’s history.
Odette’s husband finds her one night with her lover, and he turns her
out of the house in the presence of witnesses. For several years she
leads a dissolute life, dishonoring the name of her husband and
grown-up daughter. This stain on the family makes it almost
impossible for the latter to marry, and the husband offers the fallen
woman a large sum of money to deprive her of his name. She
agrees, on condition that she shall be allowed to see her daughter.
She is prevented from making herself known to the latter, and when
she comes away after the interview, she drowns herself in a fit of
hysterical self-contempt. Such are the contents of the two pieces into
which Duse put her greatest and best talent.

IV
She comes as Clotilde into the gambling saloon, to inquire after the
young girl whom she had nearly driven over. She is simply dressed,
and has the appearance of a distinguished lady, with a happy and
virtuous past. The manner in which she receives the girl in her own
house, talks to her and puts her at her ease, was so kind and hearty
that the audience, very unexpectedly in this scene, broke into a
storm of applause before the curtain had gone down. Her lover
returns from a journey which arouses her suspicion, and she,
anxious not to deceive herself, elicits the confession that he no
longer cares for her, and is in love with some one else. That some
one is Fernande. He goes to look for her, finds her in the same
house, and returns immediately. Clotilde thinks that he has come
back to her. Her speechless delight must be seen, for it cannot be
described; her whole being is suffused with a radiant joy, she
trembles with excitement. When it is all made plain to her, and there
is no longer any room for doubt, she bows her head over his hand
for an instant, as though to kiss it, as she had so often done before,
then she strokes it softly with her own.... She will never look into his
face again, yet she cannot cease to love the clear, caressing hand,
which calls to mind her former happiness.
She lets things take their course, and when it is over she has the
scene with Pomerol, when she defends her conduct. Duse has a
form of dialectic peculiar to herself, which is neither sensible nor
deliberate, but impulsive. When she does wrong she does it—not
because she is bad, but because she cannot help herself. A part of
her nature, which was the source of her life, is wounded and sick
unto death, and a gnawing, burning pain compels her to commit
deeds as dark and painful as her own heart. She goes about it
quietly, doing it all as a matter of course; to her they seem inevitable
as the outer expression of a hidden suffering.
She is at her best in the passionate “Fédora,” when she represents
this state of blank amazement, mingled with despair, taking the place
of what has been love. If she afterwards comes across the French
cynic, she reasons with him too—but like a woman, i.e., she drowns
his arguments in an extraordinary number of interjections, with or
without words. She never crosses the threshold of her life as an
actress, she never once attains to the consciousness of objective
judgment.
When the man whom she loves is married to the dishonored girl,
Clotilde comes to bring him the information which she has reserved
until now. Suddenly she stands in the doorway, and sees that he is
alone, and there comes over her an indescribable expression of
dumb, suppressed love. She seems to be making a frantic appeal to
the past to be as though it had never taken place, and in the emotion
of the moment she has forgotten what brought her there. Not until he
has unceremoniously shown her the door, and opened the old
wound, does she tell him who his wife is.
The same with “Odette.” She is in love, and she receives her lover.
At that moment her husband comes home. (Andó, Duse’s partner, is
almost as good an actor as she is.) He is a shallow, restless, hot-
tempered little man, who seizes her by the shoulders as she is about
to throw herself into the other man’s arms. She collapses altogether,
and stands before him stammering and ashamed. He thrusts her out
of the house, although it is the middle of the night, and she is lightly
clad. In a moment she has drawn herself up to her full height,—a
woman deprived of home and child, on whom the deadliest injury
has been inflicted in the most barbarous manner; in the presence of
such cruelty, her own fault sinks to nothing, and with a voice as
hoarse as that of an animal at bay, she cries, “Coward!” and leaves
him.
Many years have gone by, and we meet Odette once more, this time
as a courtesan in a gambling saloon. She is very much aged,—a
thin, disillusioned woman, for whom her husband is searching
everywhere, with the intention of depriving her of his name. There is
still something about her which bears the impress of the injured
woman. She recalls the past as clearly as though it happened only
yesterday; for she can never forget it, and time has not lessened the
disgrace. She treats him with wearied indifference, and her voice is
harsh like an animal’s, and she chokes as though she were trying to
smother her indignation.
Then follows the last act, when she meets her daughter. She comes
in, dressed like an unhappy old widow, shaking with emotion, and
scarcely able to contain herself. Her eyes are aglow with excitement,
as she rushes forward, ready to cast herself into her daughter’s
arms. But when she sees the fresh, innocent girl, she is overcome
with a feeling of shyness, and shrinks from her with an awkward,
anxious gesture. She speaks hesitatingly, like one who is ill at ease;
she raises her shoulders and stoops, and holds her thin, restless
hands clasped together, lest they should touch her daughter. The girl
displays the various little souvenirs that belonged to her mother, and
plays the piece which was her favorite, and talks about her “dead
mother.” Then this man and woman are stirred with a deep feeling,
which is the simple keynote of humanity, which they never
experienced before in the days when they were together. And they
sit and cry, each buried in their own sorrow, and far apart from one
another. After that she puts her trembling arms round the girl, and
kisses her with an expression in her face which it is impossible to
simulate, and which cannot be imitated,—which no one understands
except the woman who is herself a mother. She gazes at her
daughter as though she could never see enough of her; she strokes
her with feverish hands, arranges the lace on her dress, and you feel
the joy that it is to her to touch the girl, and to know that she is really
there. Then she becomes very quiet, as though she had suffered all
that it was possible for her to suffer. As she passes her husband, she
catches hold of his outstretched hand, and tries to kiss it. Then she
tears herself away, overcome with the feeling that she can endure it
no longer.
Eleonora Duse prefers difficult parts. She was nothing more than an
ordinary actress in “La Locandiera,” and the witty dialogue in
“Cyprienne” and “Francillon” had little in common with her nature.
Even the part of “La Dame aux Camélias” was an effort to her. The
silly, frivolous cocotte, with her consumptive longing to be loved, was
too exaggerated a part for Eleonora Duse. A superabundance of
good spirits is foreign to her nature, which is sad as life itself. Pride
and arrogance she cannot act, nor yet the trustfulness which comes
from inexperience. She gave the impression of not feeling young
enough for “La Dame aux Camélias’” happy and unhappy moods.
Eleonora Duse’s art is most at home where life’s great enigma
begins:—Where do we come from? Why are we here? Where are
we going to? We are tossed to and fro on the waters in a dense fog;
we suffer wrong, and we do wrong, and we know not why. Fate! fate!
We are powerless in the hands of Fate! When Duse can act the
blindness of fatalism, then she is content.
She was able to do so in “Fédora.”
The pretty, fashionable heroine does not change into a fury when the
man whom she loves is brought home murdered. When we meet her
again she is quite quiet,—a calm, cold woman of the world, with only
one object in life, which is to punish the murderer. It is a task like any
other, but it is inevitable, and must be undertaken as a matter of
course. She makes no display of anger, and takes no perverse
pleasure in thoughts of vengeance. The murderer is nothing to her,—
he is a stranger. But she has been rendered desolate in the flower of
her youth; the table of life, which is never spread more than once,
has been upset before her eyes at the very moment of her
anticipated happiness, and this is an injury which she is going to
repay. She is proud, and has no illusions; she is a just judge, who
recompenses evil with evil and good with good. This “Fédora” is
reserved and unreasoning.
The scene changes. She loves the man whom she has been
pursuing, and she discovers that the dead man has been false to
both of them, and she realizes that now for the first time life’s table is
spread for her, while the secret police, to whom she has betrayed
him, are waiting outside, and she clings to him terrified, showers
caresses upon him, kisses him with unspeakable tenderness. There
is something in her of the helplessness of a little child, mingled with a
mother’s protecting care, as she implores him to remain, and entices
him to love, and seeks refuge in his love, as a terrified animal seeks
refuge in its hole.
There are two other features of Eleonora Duse’s art which deserve
notice. These are, the way in which she tells a lie, and the way she
acts death. As I have said already, she is not a realist, and she
frames her characters from her inner consciousness, not from details
gathered from the outward features of life. Her representation of
death is also the outcome of her instinct. A death scene has no
meaning for her unless it reflects the inner life. As a process of
physical dissolution, she takes no interest in it. She has not studied
death from the side of the sick-bed, and she makes short work of it in
“Fédora,” as also in “La Dame aux Camélias.” In the first piece, the
point which she emphasizes is the sudden determination to take the
poison; in the second, it is her joy at having the man whom she loves
near her at the last.
Then her manner of lying. When Duse tells a lie, she does it as if it
were the simplest and most natural thing in the world. Her lies and
deceptions are as engaging, persuasive, and fantastic as a child’s.
Lying is an important factor in the character of a woman who has
much to fight against, and it is a weapon which she delights to use,
and the use of it renders her unusually fascinating and affectionate.
Even those who do not understand the words of the play, know when
Duse is telling a lie, because she becomes so unusually lively and
talkative, and her large eyes have an irresistible sparkle in them.
“Cavalleria Rusticana” was the only good Italian play that Duse
acted. She was more of a realist in this piece than in any other,
because she reproduced what she had seen daily before her eyes,—
her native surroundings, her fellow-countrymen,—instead of that
which she had learned by listening to her own soul. Her Santuzza—
the poor, forsaken girl with the raw, melancholy, guttural accents of
despair—was life-like and convincing, but the barbaric wildness of
the exponent was something which was as startling in this stupid,
pale weakly creature as a roar from the throat of a roe deer.
V
And now to sum up:—Eleonora Duse goes touring all round the
world. She is going to America, and she is certain to go back to
Berlin and St. Petersburg and Vienna, and other places where she
may or may not have been before. She will have to travel and act,
travel and act, as all popular actresses have done before her. She
will grow tired of it, unspeakably tired,—we can see that already,—
but she will be obliged to go on, till she becomes stereotyped, like all
the others.
When we see her again, will she be the same as she is now? Her
technical power is extraordinary, but her art is simple; melancholy
and dignity are its chief ingredients. Will Duse’s womanly nature be
able to bear the strain of never-ending repetition? This fear has been
the cause of my endeavor to accentuate her individuality as it
appeared to me when I saw her. Hers is not one of those powerful
natures which always regain their strength, and are able to fight
through all difficulties. Her entire acting is tuned upon one note,
which is usually nothing more than an accompaniment in the art of
acting; that note is sincerity. In my opinion she is the greatest woman
genius on the stage.
Nowadays we are either too lavish or too sparing in our use of the
word genius; we either brandish it abroad with every trumpet, or else
avoid it altogether. We are willing to allow that there are geniuses
amongst actors and actresses, and that such have existed, and may
perhaps continue to exist, but I have never observed that any
attempt is made to distinguish between the genius of man and
woman on the stage. This may possibly be accounted for by the fact
that the difference was not great. The hero was manly, the heroine
womanly, and the old people, whether men or women, were either
comic or tearful, and the characters of both sexes were usually bad.
The difference lay chiefly in the dress, the general comportment, and
the voice: one could see which was the woman, and she of course
acted a woman’s feelings; tradition ruled, and in accordance with it
the actress imitated the man, declaimed her part like him, and even
went as far as to imitate the well-known tragic step. Types, not
individuals, were represented on the stage, and I have seldom seen
even the greatest actresses of the older school deviate from this
rule.
The society pieces were supposed to represent every-day life;
therefore it was necessary before all else that the actress should be
a lady, and where a lady’s feelings are limited, hers were necessarily
limited too. To every actress, the tragedian not excepted, the
question of chief importance was how she looked.
But Duse does not care in the least how she looks. Her one desire is
to find means of expressing an emotion of the soul which
overwhelms her, and is one of the mysteries of her womanly nature.
Her acting is not realistic; by which I mean that she does not attempt
to impress her audience by making her acting true to life, which can
be easily attained by means of pathological phenomena, such as a
cough, the cramp, a death-struggle, etc., which are really the most
expressive, and also, in a coarse way, the most successful. She will
have none of this, because it is the kind of acting common to both
sexes. What she wants is to give expression to her own soul, her
own womanly nature, the individual emotions of her own physical
and psychical being; and she can only accomplish that by being
entirely herself, i.e., perfectly natural. That is why she makes
gesticulations, and speaks in a tone of voice which is never used
elsewhere upon the stage; and she never tries to disguise her age,
because her body is nothing more to her than an instrument for
expressing her woman’s soul.
What is genius? The word has hitherto been understood to imply a
superabundance of intelligence, imagination, and passion, combined
with a higher order of intellect than that possessed by average
persons. Genius was a masculine attribute, and when people spoke
of woman’s genius, their meaning was almost identical. A finer
spiritual susceptibility scarcely came under the heading of genius; it
was therefore, upon the whole, a very unsatisfactory definition.
There can be no doubt that there is a kind of genius peculiar to
women, and it is when a woman is a genius that she is most unlike
man, and most womanly; it is then that she creates through the
instrumentality of her womanly nature and refined senses. This is the
kind of productive faculty which Eleonora Duse possesses to such a
high degree.
A woman’s productive faculty has always shown a decided
preference for authorship and acting,—the two forms of art which
offer the best opportunity for the manifestation of the inner life, as
being the most direct and spontaneous, and in which there are the
fewest technical difficulties to overcome. A woman’s impulses are of
such short duration that she feels the need for constant change of
emotion. The majority of women are attracted by the stage, and
there is no form of artistic production which they find more difficult to
renounce. Why is this? We will leave vanity and other minor
considerations out of the question, and imagine Duse shedding real
tears upon the stage, enduring real mental and maybe physical
sufferings, experiencing real sorrow and real joy.
And now, putting aside all question of nerves and auto-suggestion,
we would ask what it is that attracts a woman to the stage?
Sensation.
A productive nature cannot endure the monotony of real life. To it,
real life means uniformity. Uniformity in love, uniformity in work,
uniformity in pleasures, uniformity in sorrows. To break through this
uniformity—this half sleep of daily existence—is a craving felt by all
persons possessed of superfluous vitality. This vitality may be more
or less centred on the ego, and for such,—i.e., the persons who are
possessed of the largest share of individual, productive vitality,—
authorship and acting are the two shortest ways of escape from the
uniformity of daily life. Of these two, the last-named form of artistic
expression is best suited to woman, and the woman who has felt
these sensations, especially the tragic ones, can never tear herself
away from the stage. For she experiences them with an intensity of
feeling which belongs only to the rarest moments in real life, and
which cannot then be consciously enjoyed. But the artificial
emotions, which can scarcely be reckoned artificial, since they cause
her excited nerves to quiver,—of these she is strangely conscious in
her enjoyment of them; she enjoys both spiritual and physical horror,
she enjoys the thousand reflex emotions, and she also enjoys the
genuine fatigue and bodily weakness which follow after. For the
majority of women our life is an everlasting, half-waking expectation
of something that never comes, or it may be nothing more than a
hard day’s work; but life for a talented actress becomes a double
existence, filled with warm colors—sorrow and gladness. She can do
what other women never can or would allow themselves to do, she
can express every sensation that she feels, she can enjoy the full
extent of a woman’s feelings, and live them over and over again. But
because this life is half reality and half fiction, and because the strain
of acting is always followed by a feeling of emptiness and
dissatisfaction, great actresses are always disillusioned, and that is
perhaps the reason why Duse’s attractive face wears an expression
of weariness and hopeless longing. But the warm colors—the colors
of sorrow and passion—are always enticing, and that is why great
tragedians can never forsake the stage, although gradually, little by
little, the intensity of their feelings grows less, and the colors become
pale and more false.
IV
The Woman Naturalist
I
It is a well-known peculiarity of Norwegian authors that they all want
something. It is either some of the “new devilries” with which Father
Ibsen amuses himself in his old age, or else it is the Universal
Disarm-ment Act and the peace of Europe, which Björnson, with his
increasing years and increasing folly, assures us will come to pass
as a result of “universal morality;” or else it is the rights of the flesh,
which have been discovered by Hans Jaeger; but whatever they
want, it is always something that has no connection with their art as
authors. All their writings assume the form of a polemical or critical
discussion on social subjects; yet in spite of their boasted
psychology, they care little for the great mystery which humanity
offers to them in the unexplored regions lying between the two poles:
man and woman; and as for physiology, they are as little concerned
about it as Paul Bourget in his Physiologie de l’Amour Moderne,
where there is no more physiology than there is in the novels of
Dumas père.
“When the green tree,” etc. That is the style of the Norwegian
authors; and as for the authoresses of the three Scandinavian
countries,—they are all ladies who have been educated in the high
schools. They cast down their eyes, not out of shyness,—for the
modern woman is too well aware of her own importance to be shy,—
but in order to read. They read about life, as it is and as it should be,
and then they set themselves down to write about life as it is and as
it should be; but they really know nothing of it beyond the little that
they see during their afternoon walks through the best streets in the
town, and at the evening parties given by the best bourgeois society.
This is the case with all Scandinavian authoresses, with one
exception. This one exception can see, and she looks at life with
good large eyes, opened wide like a child’s, and sees with the
impartiality that belongs to a healthy nature; she can grasp what she
sees, and describe it too, with a freshness and expressiveness
which betray a lack of “cultured” reading.
II
A lady of remarkable and brilliant beauty may sometimes be seen in
the theatre at Copenhagen, or walking in the streets by the side of a
tall, stout, fair gentleman, whose features resemble those of
Gustavus Adolphus. Any one can see that the lady is a native of
Bergen. To us strangers, the natives of Bergen have a certain
something whereby we always recognize them, no matter whether
we meet them in Paris or in Copenhagen. Björnson’s wife has it as
decidedly as the humblest clerk whom we see on Sundays at the
table of his employer at Reval or Riga. Their short, straight noses
lack earnestness, their hair is shiny and untidy, their eyes are black
as pitch, and they have the free and easy movements that are
peculiar to a well-proportioned body; it is as though the essence of
the vitality of Europe had collected in the old Hanseatic town of the
North. I do not think that the inhabitants of Bergen are remarkable for
their superior intelligence; if they were it might hinder them from
grasping things as resolutely, and despatching them as promptly as
they are in the habit of doing. But among Norwegians, who are
known to have heavy, meditative natures, the people of Bergen are
the most cheerful and light-hearted,—in as far as it is possible to be
cheerful and light-hearted in this world.
The lady who is walking by the side of the man with the Gustavus-
Adolphus head is a striking phenomenon in Copenhagen. She is
different from every one else, which a lady ought never to be.
Compared with the flat-breasted, lively, and flirtatious women of
Copenhagen, she, with her well-developed figure and large hips, is
like a great sailing-ship among small coquettish pleasure boats. She
is always doing something which no lady would do; she wears bright
colors, which are not the fashion; and I saw her one evening at an
entertainment, where there were not enough chairs, sitting on a table
and dangling her feet,—although she is the mother of two grown-up
sons!

III
When the woman’s rights movement made its appearance in
Norway, authoresses sprang up as numerous as mushrooms after
the rain. Women claimed the right to study, to plead, and to legislate
in the local body and the state; they claimed the suffrage, the right of
property, and the right to earn their own living; but there was one
very simple right to which they laid no claim, and that was the
woman’s right to love. To a great extent this right had been thrust
aside by the modern social order, yet there were plenty of
Scandinavian authors who claimed it; it was only amongst the lady
writers that it was ignored. They did not want to risk anything in the
company of man; they did not want any love on the fourth story with
self-cooked meals; they preferred to criticise man and all connected
with him; and they wrote books about the hard-working woman and
the more or less contemptible man. The two sexes were a
vanquished standpoint. These were completed by the addition of
beings who were neither men nor women, and, in consequence of
the law of adaptability, they continued to improve with time, and
woman became a thinking, working, neutral organism.
Good heavens! When women think!
Among the group of celebrated women-thinkers,—Leffler, Ahlgren,
Agrell, etc.,—who criticised love as though it were a product of the
intelligence, followed by a crowd of maidenly amazons, there
suddenly appeared an author named Amalie Skram, whom one
really could not accuse of being too thoughtful. It is true that in her
first book there was the intellectual woman and the sensual man,
and a seduced servant girl, grouped upon the chessboard of moral
discussion with a measured proportion of light and shade,—that was
the usual method of treating the deepest and most complicated
moments of human life. But this book contained something else,
which no Scandinavian authoress had ever produced before: her
characters came and went, each in his own way; every one spoke
his own language and had his own thoughts; there was no need for
inky fingers to point the way; life lived itself, and the horizon was
wide with plenty of fresh air and blue sky,—there was nothing
cramped about it, like the wretched little extract of life to which the
other ladies confined themselves. There was a wealth of minute
observation about this book, brought to life by careful painting and
critical descriptions, a trustworthy memory and an untroubled
honesty; one recognized true naturalism below the hard surface of a
problem novel, and one felt that if her talent grew upon the sunny
side, the North would gain its first woman naturalist who did not write
about life in a critical, moralizing, and polemical manner, but in whom
life would reveal itself as bad and as stupid, as full of unnecessary
anxiety and unconscious cruelty, as easy-going, as much frittered
away and led by the senses as it actually is.
Two years passed by and “Constance Ring,” the story of a woman
who was misunderstood, was followed by “Sjur Gabriel,” the story of
a starving west coast fisherman. There is not a single false note in
the book, and not one awkward description or superfluous word. It
resembles one of those sharp-cut bronze medallions of the
Renaissance, wherein the intention of the artist is executed with a
perfected technical power in the use of the material. This perfection
was the result of an intimate knowledge of the material, and that was
Fru Skram’s secret. Her soul was sufficiently uncultured, and her
sense of harmony spontaneous enough to enable her to reproduce
the simplest cause in the heart’s fibre. She describes human beings
as they are to be found alone with nature,—with a raw, niggardly,
unreliable, Northern nature; she tells of their never-ending, unfruitful
toil, whether field labor or child-bearing, the stimulating effect of
brandy, the enervating influence of their fear of a harsh God,—the
God of a severe climate,—the shy, unspoken love of the father, and
the overworked woman who grows to resemble an animal more and
more. Such are the contents of this simplest of all books, which is so
intense in its absolute straightforwardness. The story is told in the
severest style, in few words without reflections, but with a real
honesty which looks facts straight in the face with unterrified gaze,
and is filled with a knowledge of life and of people combined with a
breadth of experience which is generally the property of men, and
not many men. We are forced to ask ourselves where a woman can
have obtained such knowledge, and we wonder how this
unconventional mode of thinking can have found its way into the
tight-laced body and soul of a woman.
A second book appeared the same year, called “Two Friends.” It is
the story of a sailing vessel of the same name, which travels
backwards and forwards between Bergen and Jamaica, and Sjur
Gabriel’s grandson is the cabin boy on board. This book offers such
a truthful representation of the life, tone of conversation, and work on
board a Norwegian sailing vessel, that it would do credit to an old
sea captain. The tone is true, the characters are life-like, and the
humor which pervades the whole is thoroughly seamanlike. The
description of how the entire crew, including the captain, land at
Kingston one hot summer night to sacrifice to the Black Venus, and
the description of the storm, and the shipwreck of the “Two Friends”
on the Atlantic Ocean, the gradual destruction of the ship, the state
of mind of the crew, and the captain’s suddenly awakened piety;—it
is all so perfectly life-like, so characteristically true of the sailor class,
and so full of local Norwegian coloring, that we ask ourselves how a
woman ever came to write it,—not only to experience it, but to
describe it at all, describe as she does with such masterly
confidence and such plain expressions, without any affectation,
prudery, or conceit, and without any trace of that dilettantism of style
and subject which has hitherto been regarded as inseparable from
the writings of Scandinavian women.

IV
Whence comes this sudden change from the dilettante book,
“Constance Ring,” with its Björnson-like reflections, to the matured
style of “Sjur Gabriel” and “Two Friends”?
I could not understand it all at first, but the day came when I
understood. Amalie Skram as a woman and an author had come on
to the sunny side.
I have often wondered why it is that so few people come on to the
sunny side. I have studied life until I became the avowed enemy of
all superficial pessimism and superficial naturalism. I have
discovered a secret attraction between happiness and individualism,
—an attraction deeper than Zola is able to apprehend; it is the
complete human beings who, with wide-opened tentacles, are able
to appropriate to their own use everything that their inmost being has
need of; but whether a person is or is not a complete human being,
that fate decides for them before they are born.
Fru Amalie Skram was, in her way, one of these complete women.
She passed unscathed through a girl’s education, was perhaps
scarcely influenced by it, and with sparkling eyes and glowing
cheeks she gazed upon the world and society with the look of a
barbaric Northern woman, who retains the full use of her instinct.
When quite young she married the captain of a ship, by whom she
had two sons. She went with him on a long sea voyage round the
world; she saw the Black Sea, the Sea of Azof, and the shores of the
Pacific and Atlantic oceans. She saw life on board ship, and life on
land,—man’s life. Her mind was like a photographic plate that
preserves the impressions received until they are needed; and when
she reproduced them, they were as fresh and complete as at the
moment when they were first taken. These impressions were not the
smallware of a lady’s drawing-room; they represented the wide
horizon, the rough ocean of life with its many dangers. It was the
kind of life that brings with it freedom from all prejudice, the kind of
life which is no longer found on board a modern steamer going to
and fro between certain places at certain intervals.
But it was not to be expected that the monotony of the life could
satisfy her. She separated herself from her husband, and remained
on shore, where she became interested in various social problems,
and wrote “Constance Ring.”
It was then that she made the acquaintance of Erik Skram.
The man with the head of Gustavus Adolphus is Denmark’s most
Danish critic. His name is little known elsewhere, and he cannot be
said to have a very great reputation; but this may be partly
accounted for by the fact that he has no ambition, and partly
because he has one of those profound natures that are rendered
passive by the depth of their intellect. He is a man of one book, a
novel called “Gertrude Colbjörnson,” and he is never likely to write
another. But he contributes to newspapers and periodicals, where
his spontaneous talent is accompanied by that quiet, delicate, easy-

You might also like