Вы находитесь на странице: 1из 383

An Introduction to

Tectonophysics
Theoretical Aspects of Structural Geology

Atsushi Yamaji
Division of Earth and Planetary Sciences, Graduate School of Science
Kyoto University, Kyoto, Japan

TERRAPUB, Tokyo
An Introduction to Tectonophysics:
Theoretical Aspects of Structural Geology
Atsushi Yamaji
ISBN No. 4-88704-135-7

Published by TERRAPUB, 2003 Sansei Jiyugaoka Haimu, 5-27-19 Okusawa, Setagaya-ku, Tokyo 158-0083,
Japan. Tel. +81-3-3718-7500, Fax. +81-3-3718-4406. URL http://www.terrapub.co.jp

All rights reserved.

© 2005, 2007 by TERRAPUB, Tokyo

No part of the material protected by this copyright notice may be reproduced or utilized in any form or by any
means, electronic, digital or mechanical, including photo-copying, recording or by any information storage and
retrieval system, without written permission from the copyright owner.

This book is partly supported by Grant-in-Aid for Publication of Scientific Research Results of the Japan Society
for the Promotion of Science (No. 165283).

Printed in Japan
Preface

The purpose of this book is to explain the theoretical studies that deal with geologic structures as
constraints so as to understand the deformation of the lithosphere of terrestrial planets and satellites.
Topics range from tectonic processes at mesoscale to global scales. The term “tectonophysics,”
within the meaning of those studies, is usually regarded as a branch of geophysics. However, this
book is all about the middle ground between geology and geophysics. The audience includes re-
searchers, advanced undergraduate, and graduate students in geology, geophysics, and planetary
science.
Tectonic activity in orogenic belts, island arcs, and rift zones lasted for millions of years, so that
geological observations are necessary for understanding tectonics. The word “tectonics” means the
natural processes that create geologic structures on meso to regional scales or on a global scale. Folds
and faults are not the only objects of tectonics. This realm of science covers not only folds and faults,
but also structures that were created through natural deformation processes in the shallow parts of
terrestrial planets and of icy and rocky satellites. They include the formation of mountain belts,
sedimentary basins, and impact craters. Although microscopic deformations and interplay between
tectonics and climate are important issues today, they are omitted from this book. Geophysical
observations including seismological and geodetic ovservations, which have provided important data
for tectonophysics, are not included.
Geologic structures, stratigraphic sequences, and the surface topography of terrestrial planets
and satellites are record of geologic processes from the past. They are the clue to understanding
the internal evolution of the bodies. They can constrain deep processes, and their orbital evolution.
Mechanical properties depend on rock types, so that the difference in type is visualized as geologic
structures in some cases. If so, we are able to estimate the composition of the heavenly bodies from
their geologic structures.
Theories of faulting, folding, and vertical movements are explained in this book so that the reader
can not only understand the phenomena themselves but also their origins. We shall find that the
answers to many questions on those structures, including the following, depend on circumstances:

• Can we know paleo stress fields from faults?

• Does the subsidence that resulted in the formation of a sedimentary basin simultaneously
indicate crustal thinning?

v
vi

• Does clockwise paleomagnetic rotation in a shear zone indicate a dextral sense of shear?

It is important to know what kind of conditions justify them.


To construct theoretical models for geologic structures, continuum mechanics is a basic and in-
dispensable tool, as rock masses occupy continuous spaces at depth. We call such a continuous
medium a continuum, in that physical variables, such as force, temperature, etc., are continuous.
This assumption makes mathematical treatment very easy. However, continuum mechanics is not
always familiar to geologists and undergraduate students, so that the basic equations of continuum
mechanics are explained with the use of models. In addition, basic formulas of vectors and tensors
for the mechanics is provided in an appendix. This book covers a wide range of topics on geologic
structures with scales from outcrops to planetary bodies, so that I believe that a brief explanation is
necessary for the geological background or implications of each of the topics. Accordingly, expla-
nations on continuum mechanics and geological ones are provided alternatively.
Continuum or solid mechanics has been developed in mechanical engineering to deal with artif-
ical structures which are made of materials with known mechanical and chemical properties. Nev-
ertheless, Y. C. Fang wrote, in the epilogue of his textbook “Foundations of solid mechanics,” that
the reader must have a feeling that the panorama of solid mechanics is somewhat unreal. The ex-
amples selected and discussed in the book are idealized problems. . . The real material behaves in a
complicated way [61]. Tectonic phenomena are still more complex.
What is the benefit of the theoretical treatment of geologic structures and of the use of geological
constraints? Geological data are usually less precise than geophysical data. Can the theories predict
future tectonic movements and help prevent natural disasters? They may be difficult, even if they are
possible, because the crust and mantle are so heterogeneous that their mechanical properties vary by
many orders of magnitude. In addition, the temporal and spatial scales of tectonics are so large that
rocks at great depths are inaccessible and estimation of the parameters that control the properties is
usually difficult. Coupled with several other factors makes it still more difficult. In many instances
we do not know even how to ask questions about outcrops.
Instead, the theories provide insights into tectonic phenomena. Using proper interpretations of
idealized results, we can discuss how and which factors control tectonic phenomena under what
conditions. Theoretical and quantitative models not only provide insights, but also enable us to
verify our understanding of the phenomena. The following statement by Fung [61] is appropriate for
tectonic researchers as well as for engineers: Central questions constantly asked by an engineer are:
how can a problem be idealized . . . And how can an idealization be evaluated as to its adequacy
with respect to the real situation?
Ultimately, the striking feature of this book is the detailed description of analytical models for
geologic structures, stratigraphy, and topography that were the results of tectonics over geological
timescales in the terrestrial planets and satellites. The results of numerical simulation, which is now
an important tool for tectonophysics, are only briefly referred to, mainly in the last chapter.
Geologists have studied the structures as aspects of natural history. When I was a student in
geology, I learned from many geophysical articles that a wealth of geological observations were
vii

used as constraints for tectonophysics. However, there were a few textbooks acting as a go-between
for geologic structures and geophysics. This book is the outcome of my endeavors and aspirations
for bridging the gap from the geological side. This is relevant to the difference between this book
and the recognized textbook by Turcotte and Schubert [245]. There are two points that discriminate
the textbooks: (1) this book takes the scope of the problem areas generally restricted in geology, and
(2) the candid usage of tensorial equations the basics of which are briefly explained in Appendix
C. Tensor calculus in general is not simple. However, only orthogonal coordinates and second-
order tensors are used in this book, making the tensors intelligible to readers who have studied
linear algebra and elementary calculus. Every time an important concept is introduced, topics on
its application to tectonics are presented. The representations give easy access to a comprehensive
understanding of continuum mechanics that, in turn, provides the coherent architecture of this book
with a variety of topics. A few textbooks [186, 190, 233, 184] are similar in ethos, but they deal with
geologic structures ranging from micro- to regional scales. This book covers meso- to global scales,
and takes topics from other planets and satellites as well as from the Earth.
This book consists of three parts. The first and second parts introduce the equations of continuum
mechanics and their applications. Apart from the basics of continuum mechanics, special topics are
discussed in the last part.
In Part I, strain and stress are formulated. Chapter 1 in this part formulates deformations. Con-
trasting to this chapter in which finite deformations are discussed as tectonics for a geological time
scale results in finite deformations, Chapter 2 is limited to infinitesimal deformations, and describes
the time rate of deformation. The limitation allows us to linearize problems and to link to stress in
Part II. Chapters 1 and 2 describe the kinematic aspects of tectonics. Chapter 3 introduces stress
and balance equations to formulate statics in tectonics. Chapter 4 explains the principal stresses and
stress fields.
Part II links stress and strain together for the discussion about the dynamical aspects of tectonics.
The part consists of five chapters. Chapter 5 briefly explains how materials behave under loading
and introduces the representation theorem, that is, a tight mathematical constraint for the stress-strain
relationship of isotropic materials. Chapter 6 explains brittle faulting and the brittle strength of the
crust. Chapter 7 puts forward the equations of linear elasticity and their applications to the state of
stress in the shallow parts of the crust. Based on those equations, Chapter 8 explains the elasticity of
the lithosphere and geological implications. Newtonian fluid is introduced in Chapter 9. Topics in
this chapter includes the viscous relaxation of topographic undulations on the Earth, Mars, and icy
satellites, and the formation of long wavelength undulations as the surface manifestations of mantle
convection. Linear stability theory for Newtonian fluids is applied to the formation of spatially
periodic geologic structures including folds and boudins. Plasticity is introduced in Chapter 10 to
formulate the yielding of flexed sedimentary slabs and the collapse of large-scale impact craters. The
equations of the power-law rheology of rocks and water ice are also presented in this chapter.
Part III is devoted to special topics. Chapter 11 discusses how to estimate paleo stresses from
mesoscale faults. Chapter 12 explains the mechanical strength of the lithosphere.
This book has four appendices. The list of symbols is provided in Appendix A. A formulation of
viii

stress inversion that was developed in the last few years is briefly introduced in Appendix B, which
supplements Chapter 11. Appendix C provides the basic equations of vectors and tensors, which are
referred to in the preceding parts. There are exercises at the end of each chapter. Answers for most
of the questions are presented in Appendix D.

Remarks
Symbols We will use a few kinds of variables, namely, scalars, vectors and tensors. Scalar vari-
ables are denoted by italic symbols such as A. Bold italic and san-serif fonts indicate, respectively,
vectors and tensors, for example A and A.

Coordinates We adopt the right-handed Cartesian coordinates as standard. If we consider mod-


els where gravitational force is significant, we use the coordinates O-xyz with the z axis pointing
vertically downwards. Vector components are denoted in this case as a = (ax , ay , az )T , where the
superscript T indicates matrix transpose. If it is not the case, we use the Cartesian coordinates O-123,
and we write vector components as a = (a1 , a2 , a3 )T .

Sign convention for stress and strain Attention must be paid to the signs of stress and strain, be-
cause Earth science and continuum mechanics have different sign conventions. Tensional stress and
elongation are positive stress and strain, respectively, in continuum mechanics. Structural geologists
define compression and elongation as positive stress and strain. This sign convention is natural for
Earth scientists as the state of stress at depth is commonly compressional. However, tensile stresses
results in elongation. Accordingly the fashion of continuum mechanics is theoretically more natural
than that of Earth science. Both sign conventions have reasons. Consequently, we use both fash-
ions and distinguish them by assigning different symbols as follows. The symbols S = (Sij ) and
r = (σij ) represent tension and compression as positive stresses. They are different only in sign,
so that S = −r. Corresponding to them, E and e represent extension and contraction as positive
infinitesimal strains, respectively.

Acknowledgments
I have lectured on this material to advanced undergraguate and graduate classes at Kyoto University
for several years, and the book is the result of these courses. The author gratefully acknowledges
Prof. D. Stevenson for his excellent lectures at Caltech. I would not have decided to write this book
without attending at his lectures. I would like to thank my colleagues, co-workers, and students who
made comments, corrections, and gave me encouragement. Among them, Prof. F. Masuda acted
as the intermediary between the author and publisher. The author is grateful to Mr. K. Sato for
the preparation of the fault-slip data in Fig. 11.19. I acknowledge the encouragement by Profs. T.
Setoguchi, N. Fujii, T. Nishiyama, K. Kanagawa, T. Eguchi, S. Sasaki, and K. Ishii. During my
ix

student years at Tohoku University, I was lucky enough to have good teachers and classmates. Prof.
M. Seki introduced me to the first course of fluid mechanics. Profs. K. Otsuki, H. Nakagawa, and N.
Kitamura taught me the ABC of structural geology. The textbooks by Shimazu [214] and T. Kakimi
[96] attracted me to the theoretial aspects of structural geology. I am thankful to my parents and
my wife Hatsumi for their unwavering support. The Japan Society for the Promotion of Science is
thanked for its financial support (165283) for publication of this book.
Contents

I Kinematics and Statics 1

1 Finite Strain 3
1.1 Definition of strain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Displacement and deformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Deformation Gradient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.4 Geological observation of deformation . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.5 Tensorial representation of strain ellipsoid . . . . . . . . . . . . . . . . . . . . . . . 12
1.6 Green’s and Almansi’s finite strain tensors . . . . . . . . . . . . . . . . . . . . . . . 16
1.7 Shear zone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.8 Deformation history . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.9 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

2 Infinitesimal Strains and Their Accumulations 31


2.1 Infinitesimal strain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.2 Time rate of change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.3 Strain rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
2.4 Pile up of infinitesimal deformations . . . . . . . . . . . . . . . . . . . . . . . . . . 39
2.5 Rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
2.6 Paleomagnetic rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
2.7 Deformatin of rock mass by the activity of minor faults . . . . . . . . . . . . . . . . 53
2.8 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

3 Stress, Balance Equations, and Isostasy 59


3.1 Definition of stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
3.2 Cauchy’s stress formula . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
3.3 Force balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
3.4 Symmetry of stress tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.5 Conservation of energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
3.6 Self gravitational, spherically symmetric body . . . . . . . . . . . . . . . . . . . . . 70

xi
xii CONTENTS

3.7 Isostasy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
3.8 Balance between ocean and continent . . . . . . . . . . . . . . . . . . . . . . . . . 76
3.9 Sediment load . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
3.10 Quantitative stratigraphy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
3.11 Tectonic force caused by horizontal density variations . . . . . . . . . . . . . . . . . 83
3.12 Thermal isostasy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
3.13 Horizonal stress resulting from topography . . . . . . . . . . . . . . . . . . . . . . 91
3.14 Thermal subsidence of continental rifts . . . . . . . . . . . . . . . . . . . . . . . . . 92
3.15 Kinematic model of syn-rift subsidence . . . . . . . . . . . . . . . . . . . . . . . . 95
3.16 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98

4 Principal Stresses 101


4.1 Principal stresses and principal stress axes . . . . . . . . . . . . . . . . . . . . . . . 101
4.2 Tectonic stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
4.3 Mohr diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
4.4 Boundary conditions of stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
4.5 In-situ stress measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
4.6 Dikes as paleostress indicators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
4.7 First-order regional stress field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
4.8 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121

II Dynamics 123

5 Constitutive Equations 125


5.1 Stress-strain diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
5.2 Representation theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
5.3 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129

6 Faulting and Brittle Strength 131


6.1 Primary fractures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
6.2 Coulomb-Navier criterion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
6.3 Anderson’s theory of faulting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
6.4 Frictional strength of fault . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
6.5 Effect of pore fluid pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
6.6 Reactivation of fault . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
6.7 Depth dependence of brittle strength . . . . . . . . . . . . . . . . . . . . . . . . . . 144
6.8 Joint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
6.9 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
CONTENTS xiii

7 Elasticity 153
7.1 Linear elasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
7.2 Earth Pressure at Rest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
7.3 Stability of elastic rock masses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
7.4 Thermal stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160
7.5 Global thermal changes and surface stress field . . . . . . . . . . . . . . . . . . . . 165
7.6 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169

8 Flexure 171
8.1 Regional isostasy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
8.2 Flexure of thin elastic plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
8.3 Flexure of the oceanic lithosphere . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
8.4 Buckling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
8.5 Rift shoulder uplift . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192
8.6 The degree of compensation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
8.7 Effective elastic thickness estimated from gravity anomalies . . . . . . . . . . . . . 197
8.8 Lunar mare tectonics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
8.9 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203

9 Fluid 205
9.1 Fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
9.2 Stream function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
9.3 Viscous relaxation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
9.4 Self-organizing geological structures . . . . . . . . . . . . . . . . . . . . . . . . . . 215
9.5 Dynamic topography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225

10 Plasticity 237
10.1 Quasilinear Fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
10.2 Principal stress space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
10.3 Yield conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
10.4 Plastic yielding by folding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
10.5 Slip line theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
10.6 Collapse of impact craters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
10.7 Bingham model for complex craters . . . . . . . . . . . . . . . . . . . . . . . . . . 255
10.8 Power-law fluid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259

III Special Topics 269


11 Determination of Stress from Faults 271
11.1 Mesoscale faults . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
xiv CONTENTS

11.2 Wallace-Bott hypothesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273


11.3 Stress inversion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280
11.4 Fry’s formulation of stress inversion . . . . . . . . . . . . . . . . . . . . . . . . . . 282
11.5 Multiple inverse method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285
11.6 Slip tendency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 288
11.7 Faulting controlled by strain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 290
11.8 Reches’ model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 291
11.9 Block rotation and faulting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 300

12 Dynamics of the Lithosphere 303


12.1 Strength profile . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
12.2 Stretching instability of the lithosphere . . . . . . . . . . . . . . . . . . . . . . . . . 307
12.3 Factors controlling te of the continental lithosphere . . . . . . . . . . . . . . . . . . 312
12.4 Periodic deformation of the lithosphere . . . . . . . . . . . . . . . . . . . . . . . . 315

A List of Symbols 323

B Stress Inversion 327

C Basic Equations 329


C.1 Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 329
C.2 Dumy and free indices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331
C.3 Rotation of coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 332
C.4 Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 333
C.5 Eigenvalues and eigenvectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 334
C.6 Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 337
C.7 Polar decomposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342
C.8 Calculus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 344
C.9 Rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 346
C.10 Exercise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 349

D Answers to Selected Problems 351

References 369

Index 379
Chapter 1

Finite Strain
Tectonics is the natural processes that deform the shallow solid part of rocky or icy
bodies of the Solar System. The deformation proceeds with the movements of rock or
ice masses. In this chapter, we will study how to describe the geometric relationship
between deformation and displacement—the kinematic aspect of tectonics.

1.1 Definition of strain


Deformation and strain are similar words. However, their distinction is sometimes important. Strain
is the change of shape or size, or both; deformation includes both strain and rotation. We discuss the
treatment of rotation later, and here we describe how strain is quantified.
Suppose a rock mass with a length L0 was deformed and the length has become L, its strain E
is defined as
L − L0
E= . (1.1)
L0
The ratio L/L0 is called stretch or elongation, and (L/L0 )2 is quadratic elongation. The quantity,
defined by the equation
En ≡ ln(L/L0 ),
is a natural or logarithmic strain. If we know length L every moment in the deformation process,
the logarithmic strain of the final shape is
L
dL L
En = = ln . (1.2)
L0 L L0

The integral indicates that logarithmic strain is the sum of infinitesimal strains during a deformation
process.  
L L − L0
En = ln = ln 1 + = ln(1 + E).
L0 L0
Therefore, En is identical with E if each strain is very small.

3
4 CHAPTER 1. FINITE STRAIN

Figure 1.1: Flattened ammonite Hollandites haradai obtained in folded Triassic strata in the southern
Kitakami massive, northern Japan. The closed circle is 1 cm in diameter.

Geodesy can detect a tiny strain of a rock mass, but strains that are observable as geologic
structures are not so small. Most of them are acknowledged by the naked eye. The ammonite in Fig.
1.1 is an example. The original shape of an ammonite must be round, but this one was flattened by
a few tens of percents.
It is obvious that we should know the original shape of deformed objects to quantify the strain,
although, this is not always possible. We call such objects strain markers. Fossils are often used as
markers1 . Zenolith blocks shown in Fig. 1.2 appear to be elongated. However, we do not know their
initial shapes. Therefore, they are not useful as strain markers2 .
Fine grained sedimentary layers lie almost horizontally, so that they tell us how the strata were
tilted—they are markers of rotation about a horizontal axis. Vertical axis rotations are revealed by
paleomagnetism.
Changes in shape are quantified not only by a variation of lengths, but by that of angles between
marker lines. Angle changes are an important aspect of the deformation resulting in shear strain3 . If
the angle at the corner ∠ABC that was originally at 90◦ decreases by φ and becomes an acute angle,
the angular change is represented by the engineering shear strain (Fig. 1.3)

γ = tan φ.

If an obtuse angle is the result, the engineering shear strain is negative. It is a matter of course that
shear strain depends on the direction of the marker lines in the initial configuration. The dependence
is given by Eq. (1.32).

1 Many kinds of strain markers and procedures for quantification are described in the textbooks [170, 189].
2 Each of the blocks is not a strain marker. However, if it is allowed to assume that they had random orientations before
deformation, we can estimate strain from the present shapes and long-axis orientations (Exercise 1.4).
3 The words ‘shear strain’ is commonly used for the non-diagonal components of infinitesimal strain tensor (§2.1) which

is just the half of engineering shear strain.


1.2. DISPLACEMENT AND DEFORMATION 5

Figure 1.2: Elongated zenolith blocks in granite. Takanuki, Northeast Japan.

Figure 1.3: Definition of engineering shear strain γ ≡ tan φ.

1.2 Displacement and deformation


Geological observations are not so accurate as geodetic ones, so that we should describe both small
and large deformations. Although the accuracy is so, geologic data provide information on long-term
processes.
6 CHAPTER 1. FINITE STRAIN

Rock masses deform and their portions move. If entire parts move in the same orientation by the
same distance, the rock mass moves as a rigid body. If they move in different directions, deformation
occurs. Accordingly we can use the tracks of the portions to describe the deformation. There are
two ways to carry out the task. Suppose that we have to describe population movements. One way
to describe them is to distinguish individuals by their address at time t = 0. We assume in this case
a one-to-one correspondence between people and loci for simplicity. Let the vector x indicate the
address at arbitrary time t = 0 for the person who lived at ξ at time t = 0. The vector acts as the
person’s name. The function x(ξ, t) describes the movement of the person. Taking the vector ξ as
a variable, the function indicates the movements of all the people. The other way to describe the
population movements is to use the inverse of this function, ξ(x, t). If we input an address x and
time t, we get the legal domicile ξ.
Now suppose the vectors ξ and x represent the original and present position of a rock portion,
either of the functions

x = x(ξ, t), (1.3)


ξ = ξ(x, t). (1.4)

Obviously, the displacement is


u = x − ξ.

Equations (1.3) and (1.4) represent the orbit of the portion. The dotted lines in Fig. 1.4 show
the orbit of two nearby portions. If we know the orbit of all portions of a rock mass, we know
the deformation and translation of the mass. For simplicity, we assume that these functions are
continuous at any place and at any time, and we can get derivatives of any order of the functions.
When we describe the variation of a quantity F, there are two ways to describe the variation of
quantity F (ξ, t) and F (x, t). If the quantity is money, the function F (ξ, t) shows the variation of
the fortune of the person who lived at ξ at t = 0. The function shows the economic ups and downs
of the person. On the other hand, the other function F (x, t) describes how great fortune the person
living at x has at time t. That is the fixed point observation of the variation. If T is temperature,
for example, T (ξ, t) indicates the temperature variation for the rock portion that was at ξ. T (x, t) is
the temporal variations of temperature at spatially fixed point at x. We call (x1 , x2 , x3 ), spatial or
Euler coordinates and (ξ1 , ξ2 , ξ3 ), material or Laglangian coordinates. F (ξ, t) is the Laglangian or
material description of the variation of F. F (ξ, t) is the spatial or Euler description of the variation.
Equations (1.3) and (1.4) are those of movement.

1.3 Deformation Gradient


As we have assumed above that Eqs. (1.3) and (1.4) are continuous and smooth, the transformation
between two coordinates (ξ1 , ξ2 , ξ3 ) and (x1 , x2 , x3 ) is in one-to-one correspondence and we obtain
1.3. DEFORMATION GRADIENT 7

Figure 1.4: Deformation of a rock mass can be calculated from the relative position of two nearby
portions in the mass before and after the deformation. Their initial positions are represented by the
vectors ξ and ξ + dξ, and the final positions by x and x + dx. Their tracks during deformation is
indicated by dotted lines.

the following equation from Eq. (1.3):

∂xi ∂xi ∂xi


dxi = dξ1 + dξ2 + dξ3 , (1.5)
∂ξ1 ∂ξ2 ∂ξ3
where i = 1, 2, or 3. The small vectors dx represents the distance between the rock portions that
occupied at nearby points P1 and P2 at t = 0. The vector dξ stands for the distance between the same
portions at time t (Fig. 1.4). The matrix
⎛ ⎞
∂x1 ∂x1 ∂x1
⎜ ⎟
⎜ ∂ξ1 ∂ξ2 ∂ξ3 ⎟
⎜ ⎟
∂x ⎜ ⎟
F≡ =⎜ ∂x2 ∂x2 ∂x2 ⎟
∂ξ ⎜ ∂ξ1 ∂ξ2 ∂ξ3 ⎟


⎜ ⎟
⎝ ∂x3 ∂x3 ∂x3 ⎠
∂ξ1 ∂ξ2 ∂ξ3

is useful to describe deformation. Then we obtain the following formula from Eq. (1.5):

dx = F · dξ, (1.6)

showing that the transformation between dx and dξ is linear, although the transformation between
x and ξ is not necessarily linear. F is called a deformation gradient tensor, and is a measure of
deformation. Such a linear transformation maps a sphere to a ellipsoid and a rectangular solid to a
parallelepiped.

1.3.1 Strain ellipsoid


A state of strain is graphically represented in structural geology by a strain ellipsoid that is the result
of the strain from an unit sphere (Fig. 1.5). The principal radii of the ellipsoid and their directions
are called principal strains and principal axes, respectively. The principal strains are labelled as X,
8 CHAPTER 1. FINITE STRAIN

Figure 1.5: Strain ellipsoid and its principal axes.

Y , and Z, in descending order, and the corresponding principal axes are called X-, Y-, Z-axes. The
axes cross each other at right angles. Pairs of axes define three principal planes of strain. A strain
ellipsoid is symmetric with respect to those three planes, therefore, a state of strain has orthorhombic
symmetry4 Strain markers are often observed on the surface of rocks; the ammonite in Fig. 1.1 was
found in a bedding plane. In two dimensional problems, we use strain ellipse to show the strain. The
ellipsoid that becomes a unit sphere by the same deformation is called a reciprocal strain ellipsoid.
According to linear algebra, |F | = 0 is the necessary and sufficient condition for the tranforma-
tion being linear. This is also the condition for the existence of its inverse matrix. The determinant

J = |F | (1.7)

is called the Jacobian of the deformation. To show that this quantity represents the volume change
during deformation, suppose three vectors dξ (1) = (dξ1 , 0, 0)T , dξ (2) = (0, dξ2 , 0)T , dξ (3) =
(0, 0, dξ3 )T were perpendicular to each other when t = 0. After deformation, the vectors have
become dx1 , dx2 , dx3 . The initial vectors span an initial volume equal to V0 = dξ1 dξ2 dξ3 . As
the initial vectors are infinitesimally small, the final ones are also small. Hence, the volume of the
parallelepiped that is spanned by the final vectors is equal to their triple scalar product

V = dx(1) · dx(2) × dx(3) = dx(1) , dx(2) , dx(3) .

The triple scalar product is identical to the determinant of the matrix that is composed by the com-
ponents of the vectors (see Appendix A), so that

V = dx(1) , dx(2) , dx(3) = F · dξ (1) , F · dξ (2) , F · dξ (3)
 
= F dξ (1) , dξ (2) , dξ (3) = J V0 . (1.8)

Therefore, the Jacobian J = |F | is equal to the volume ratio of the initial and final configurations.
In the case of |F | = 0, a rock mass vanishes. As negative volume is meaningless in tectonics, we
can always assume the inequality
J = |F | > 0. (1.9)
4 This property will be used to consider the strain of a rock mass by many faults therein.
1.3. DEFORMATION GRADIENT 9

Figure 1.6: Displacement field, strain ellipse, and principal strain axes for four deformation gradient
tensors.

If |F | = 1, then the volume of any portion of the mass is retained during deformation. This type of
deformation is called incompressible.
The linearity represented by Eq. (1.6) is useful for structural geology. The Triassic ammonite
shown in Fig. 1.1 originally should have been of round shape, but was flattened by Cretaceous
orogeny that affected the eastern margin of Eurasia. The strata that contained the fossil were folded.
Folding itself is a nonlinear transformation between x and ξ. However, the ammonite suffered
a simple flattening: the round shape became oval, because the ammonite is very small and any
vectors between portions within the fossil acted as the small vector dx. In general, F is a function of
10 CHAPTER 1. FINITE STRAIN

position and the transformation is non-linear: rocks may be distorted. However, F is assumed to be
a continuous and smooth function of position. Therefore, the linear transformation

x=F·ξ (1.10)

is a good approximation of the deformation field for a small region. The whole space is approximated
by tessallation of the regions. If this equation is satisfied, the deformation is called homogeneous for
a specific region.

1.3.2 Special types of deformation


Zero and infinitesimal strains and rigid body rotation

If the deformation gradient tensor is equal to the identical tensor, then dx = F · dξ = I · dξ = dξ.
This means that the initial and final vectors between nearby portions in a rock mass do not change
at all. This further indicates zero deformation, even if the mass was moved. It should be noticed that
zero tensor O does not stand for zero deformation but for disappearance. If FT = F−1 , the length of
the vectors does not change but their directions vary, indicating a rigid body rotation (Fig. 1.6).
Let δF be the difference between F and I:

δF ≡ F − I. (1.11)

In the case of infinitesimal deformation, F ≈ I and δF ≈ O where |δF|  1. This is called a finite
strain if δF is far from the zero tensor.

Plane strain

If the deformation gradient tensor has the following shape by choosing an orientation of orthogonal
coordinates, it is called plane strain:
⎛ ⎞
• • 0
F = ⎝• • 0⎠ ,
0 0 1

where closed circles represent any value. In this case, the movement of every portion of a rock
mass is in translation along the third coordinate axis. We can deal with the deformation by a two-
dimensional problem on the O-12 coordinate plane.
There are many tectonic problems that can be treated as plane strain. For example, orogenic
belts, island arcs, and mid oceanic ridges are linear or elongated tectonic features, so that plane
strain on their cross-section is a good approximation for the movements in the zones.
1.4. GEOLOGICAL OBSERVATION OF DEFORMATION 11

Figure 1.7: Coaxial deformation and principal strain axes. If the fan OAB is deformed to OA B with
OA and OB being parallel to the principal axes, the line segments OA and OB are not rotated, but
OC is rotated to OC . Arrows indicate displacements.

Pure and simple shear

Pure shear and simple shear often appear in the literature on tectonics. They are special cases of
plane plain strain, and are written as
 
p 0
pure shear: F= , (1.12)
0 1/p
 
1 2q
simple shear: F = , (1.13)
0 1

where p and q are arbitrary numbers. The factor of 2 is attached in the upper-right component in
the last matrix only for the convenience of later calculations. Sinple shear is often assumed for
fault zones, where positive qs represent dextral shear (Fig. 1.6(b)). Note that |F | = 1 for both
pure and simple shear, indicating that the volume of any part of a rock mass is kept constant during
deformation. Structural geologists uses the term coaxial deformation in the case of
⎛ ⎞
p 0 0
F = ⎝0 q 0⎠ . (1.14)
0 0 r

Linear markers parallel to the coordinate axes do not rotate by this deformation. Hence, this is also
called irrotational (Fig. 1.7).

1.4 Geological observation of deformation


Structural geologists have a grasp of the deformation of a mappable-scale rock mass by, for example,
air- and space-borne remote-sensing techniques in arid areas, and seismic sounding. For areas where
those remote-sensing techniques are not useful, they measure F at outcrops that are indicated by
strain markers, and plot the data on a map to understand the large-scale deformation.
12 CHAPTER 1. FINITE STRAIN

The movement of a rock mass is characterized by the movement of the center of the mass, rigid
body rotation, and deformation. However, deformation gradient tensor, F, lacks the information on
the first factor, translation. Long-distance translation of a rock mass is determined by, for example,
the change in paleomagnetic declination and paleoclimate indicated by fossils. They indicate latitude
changes. East-west movements are difficult to identify. Offset of old geologic structures across a
fault zone indicates relative horizontal displacement, but does not indicate their absolute movements.
To constrain them, we need some hypothesis, such as the fixity of hotspots.
On the other hand, horizontal-axis rotations are found by the tilting of sedimentary strata. Vertical-
axis rotations are observed by paleomagnetism and twisting of old geologic structures that penetrate
rotating and surrounding blocks. Disorder in the trend of dike swarms are used to infer vertical-axis
rotations in the western United States [199].

Primary structures

Quantification of strain is done by comparing the initial and final shape of a rock mass. So, we need
to know the shape before tectonic deformation, that is called a primary structure.
The offset of key beds shows the displacement of faults. If there are talus or alluvial fan deposits
accompanied by a fault scarp, we are able to know when the faulting occurred by the age of the
deposits.
Sediments lie almost horizontally when they were deposited. Stratification in sediments is often
used to know the tilting. However, there are following exceptions, so that we are careful to deal with
strata as strain markers. Talus breccias at the foot of scarps can lie at a significant angle from the
horizontal. Volcanic rocks accumulates on the slope of volcanoes. In addition, folding sometimes
turns strata upside down.
Lithologic stratification is important not only as a strain marker but also as an agent of tectonic
movements. Mechanical property of rocks is dependent on lithology, therefore, once a stratified
rock mass is loaded, lithologic boundary act as a rheological discontinuity. Bedding planes are often
used as fault planes (§8.4.2). In addition, sediments themselves with a thickness of hundreds of
meters indicate the presence of a basin or subsidence of the basement when they were deposited.
Sedimentary sequence is an indicator of vertical tectonic movements (§3.10).
Sediments are useful indicators of ancient tectonics. This is true not only on the Earth but also
on other planets and satellites. For example, lunar maria are sedimentary basins that are filled with
horizontally lying, very low viscosity, flood basalts and breccias derived from nearby and distant
impact craters, and are covered by a veneer of regolith (Fig. 1.8).

1.5 Tensorial representation of strain ellipsoid


Strain ellipses are represented by tensors. As long as we consider tectonics, we can assume that
|F | > 0 (Eq. (1.9)). Accordingly, using the polar decomposition theorem, F is decomposed into the
1.5. TENSORIAL REPRESENTATION OF STRAIN ELLIPSOID 13

Figure 1.8: Sedimentary layers (L) cropping out at the margin of the Euler crater in the Mare Im-
brium on the Near Side of the Moon. The crater is about 30 km in diameter with a central peak (P).
The marginal slope had collapsed to form terraces (T) the bases of which sunk in the impact melt
with a level surface (IM). Apollo 15 frame P10274. Courtecy of NASA.

product of an orthogonal tensor R and symmetric tensors U and V:

F = R · U = V · R. (1.15)

For given F, the tensors are uniquely determined. Substituting this equation into x = F · ξ, we have
x = R · (U · ξ). Therefore, the deformation indicated by F is equal to the two-stage process: pure
shear represented by U followed by the rigid body rotation R (Fig. 1.9). U and V are called right
and left stretch tensors. As the names suggest, the tensors are the extension stretch defined in p. 3
to three-dimensional strain. Their principal axes (eigenvectors) are not parallel except for coaxial
deformation (R = I).
Tensors U and V are real symmetric tensors, so that they are represented by diagonal matrices
like Eq. (1.14) by taking proper Cartesian coordinates. Therefore, the tensors stand for coaxial
deformations.
Let us consider the relationship between the strain ellipse and these tensors. If |ξ| = 1, the
terminal point of the vector ξ is on the surface of the unit sphere. The vector is transformed to
x = F · ξ = V · (R · ξ) of which the terminal point is on the strain ellipsoid. The transition from
unit sphere to ellipsoid is due to V. Therefore, V represents the strain ellipsoid of which the principal
axes and principal strains are equivalent to the eigenvectors and eigenvalues of V. Note that V also
represents a coaxial deformation.
Let us derive the tensorial representation of strain ellipsoids. Remember that a unit sphere is
transformed into a strain ellipsoid by a deformation represented by F = V · R. The right-hand
side of this equation indicates what the sphere has undergone, firstly by rotation R and secondly by
14 CHAPTER 1. FINITE STRAIN

Figure 1.9: Polar decomposition of the deformation gradient tensor F. The upper and lower deforma-
tion processes have the initial (circle) and final (elipse) shapes in common. The cross in the ellipse
indicates the principal strain axes. White line segments are strain markers.

coaxial deformation V. The rotation does not change the shape or size of the sphere, so that the final
shape is controlled solely by the tensor V. In this case, the deformation is written as the equation
x = F · ξ = V · ξ, where the vector ξ is the radius of the unit sphere so that |ξ| = 1. The strain
ellipsoid of this deformation is represented by the end point of x. As the deformation due to V is
coaxial, the line segments parallel to the principal directions of the ellipsoid do not rotate during the
deformation. Therefore, let λ be the elongation of the segments and ξ be one of the directions, then
the two quantities satisfy the equation V · ξ = λξ. This indicates that we obtain the shape and attitude
of the strain ellipsoid by solving the eigenequation V · ξ = λξ. Since we have assumed |F| > 0,
the eigenvalues are all positive. The principal strains X, Y and Z are calculated as the maximum,
intermediate, and mimimum eigenvalues of V.
Based on these observations, we shall derive the equation of strain ellipsoid. The unit sphere
before deformation
  is described
 by the equation |ξ|2 = 1. Combining the equation ξ = F−1 · x, we
have F−1 · x · F−1 · x = 1. This is the equation of strain ellipsoid (Fig. 1.10). The left-hand
side is rewritten by the components as
     −1 −1    
Fij−1 xj Fik−1 xk = xj Fij Fik xk = xj Fji−1 T Fik−1 xk .
i,j,k i,j,k i,j,k
 
∴ x · F−1 T · F−1 · x = 1. (1.16)

We define the left Cauchy-Green tensor, B, and the right Cauchy-Green tensor, C, by the fol-
lowing equations:

B ≡ F · FT (1.17)
C ≡ F · F.
T
(1.18)
1.5. TENSORIAL REPRESENTATION OF STRAIN ELLIPSOID 15

Figure 1.10: Strain ellipse and tensors that represent the identical finite strain associated with the
displacement field indicated by short arrows. The dilatation of this deformation is 25% (|F| = 1.25).

Since |F| > 0 and |FT | > 0, C and B are symmetric tensors. Combining Eqs. (1.17) and (1.18), we
get the relationship between these tensors:

C = F−1 · B · F, B = F · C · F−1 .

From Eq. (1.18), we have

C = FT · F = (R · U)T · (R · U) = UT · RT · R · U = UT · U,

where Eqs. (C.19) and then (C.15) are used. Likewise, we get B = V · VT . Since U and V are
symmetric,
C = U2 , B = V2 . (1.19)
These equations indicate that the left and right Cauchy-Green tensors are the expansion of quadratic
elongation (§1.1). According to the discussion in Section C.5 and Eq. (1.19), the symmetric tensors
C, U, and U−1 have eigenvectors in common. Three symmetric tensors B, V, and V−1 have common
eigenvectors (Fig. 1.10).
Using Eqs. (C.21) and (C.22), we have
     
B−1 = F · FT −1 = FT −1 · F−1 = F−1 T · F−1 . (1.20)

Thus, the equation of strain ellipsoid (Eq. (1.16)) becomes

x · B−1 · x = 1.
16 CHAPTER 1. FINITE STRAIN

B−1 is a real symmetric tensor, so that its eigenvalues are all positive in sign (see §C.7). Accordingly,
let B1 , B2 , B3 be the eigenvalues of B in a descending order of magnitude, then they are related to
the principal strains by the equations:
  
X = B1 , Y = B2 , Z = B3 .

Since V = B1/2 , the principal strains are identical with the eigenvalues of V. On the other hand, U−1
represents the reciprocal strain ellipsoid (Fig. 1.10) for the given deformation gradient F.

1.6 Green’s and Almansi’s finite strain tensors


Definitions We have seen how F is related to elongation in the previous section. Now, we consider
the relationship between F and strain. Suppose that there were two near-by points P1 and P2 whose
positions are indicated by the vectors ξ and ξ + dξ, respectively. The materials that existed at the
points P1 and P2 moved to the points Q1 and Q2 that are indicated by the vectors x + dx (Fig. 1.4).
Let u be the displacement vectorof the rock portion that moved from P1 to Q1 . The distance between

the points changed from ds0 = dξ · dξ to ds = dx · dx. The square of the former is
 
ds2 = dx · dx = (F · dξ) T · (F · dξ) = dξ · FT · F · dξ = dξ · C · dξ, (1.21)

therefore we have
 
ds2 − ds20 = dξ · C · dξ − dξ · I · dξ = dξ · C − I · dξ.

Accordingly, we introduce the tensor

1 T  1 
G≡ F ·F−I = C−I , (1.22)
2 2
where C is the right Cauchy-Green tensor and is symmetric (CT = C), hence, G is a symmetric
 
tensor (§C.7): G = GT . The length change is then rewritten as

ds2 − ds20 = 2dξ · G · dξ. (1.23)

If the movement of a rock mass is a translation or rigid-body rotation, the distance between any
points in the mass do not change. Only if the mass changes its shape, there are pairs of points with
ds2 − ds20 = 0. Accordingly, G is a measure of strain. Zero strain is indicated by G = O. G is called
Green’s finite strain tensor.
Now, substituting dξ = F−1 · dx into the right-hand side of the equation dξ = F−1 · dx, we have
       
ds2 − ds20 = dx · dx − F−1 · dx · F−1 · dx = dx · I − F−1 T · F−1 · dx
  −1 
= dx · I − F · FT · dx = 2dx · A · dx, (1.24)
1.6. GREEN’S AND ALMANSI’S FINITE STRAIN TENSORS 17

where the tensor A is called Almansi’s finite strain tensor and is defined by the equation

1  −1  1  
A= I − F · FT = I − B−1 , (1.25)
2 2

where B is the left Cauchy-Green tensor. If A = O, strain is nil. This tensor is symmetric, also
 
A = AT , and is related with Green’s finite strain tensor by the equation

G = FT · A · F. (1.26)

The length change was written as the product of the material vector ξ and the strain tensor G (Eq.
(1.22)). On the other hand, the same quantity is written as the product of the spatial vector x and
the strain tensor A (Eq. (1.24)). It follows that G, U, and C indicate the material (Laglangian)
expressions of strain, whereas A, V, and B represent the spatial (Euler) expressions of the strain.

Strain ellipsoid We have seen in Section 1.5 that the parameters of a strain ellipsoid are calcu-
lated as the eigenvalues and eigenvectors of B−1 . Now, let us consider the relationship between the
parameters and the finite strain tensors, G and A. From Eqs. (1.19) and (1.22), we have

2G + I = FT · F = C = U2 (1.27)

and, similarly,
 
2A = I − B−1 = I − V2 −1 . (1.28)

Let A1 , A2 , A3 be the eigenvalues of A in descending order of magnitude, then


     
1 1 1 1 1 1
A1 = 1− , A2 = 1− , A3 = 1− .
2 B3 2 B2 2 B1

Thus, the principal strains X, Y , and Z are related to the eigenvalues as


 1
X = U1 = V1 = B1 =
2G1 + 1 =  , (1.29)
1 − 2A3
  1
Y = U2 = V2 = B2 = 2G2 + 1 =  , (1.30)
1 − 2A2
  1
Z = U3 = V3 = B3 = 2G3 + 1 =  . (1.31)
1 − 2A1

The principal strain axes (X, Y , and Z in Fig. 1.5) are parallel to the eigenvectors of the strain
tensors V, B, and A. Those of the tensors U, C, and G are identical with those of the former three
tensors if the deformation is coaxial.
18 CHAPTER 1. FINITE STRAIN

Shear strain The Cauchy-Green tensors, B and C, not only indicate length changes, but also shear
strains in relation to the initial direction of marker lines. Suppose that the angle between two unit
vectors u(1) ≡ dξ (1) /|dξ (1) | and u(2) ≡ dξ (2) /|dξ (2) | was Θ. After deformation, it becomes θ, and the
vectors dξ (1) and dξ (2) become dx(1) and dx(1) , respectively. Shear strain φ is defined as φ ≡ Θ − θ.
Using Eq. (1.21), we obtain
dx(1) · dx(2) dξ (1) · C · dξ (2)
cos θ = cos(Θ − φ) =   = 
dx(1)  dx(2)  dξ (1) · C · dξ (1) dξ (2) · C · dξ (2)
 (1)  (1)  
dξ  u · C · u(2) dξ (2)  u(1) · C · u(2)
=  =  . (1.32)
dξ (1) · C · dξ (1) dξ (2) · C · dξ (2) u(1) · C · u(1) u(2) · C · u(2)
Given the initial angle Θ = u(1) · u(2) , we can calculate the shear strain for a couple of marker lines in
any direction. For the case of Θ = 90, Eq. (1.32) is equal to sin φ, and the engineering shear strain
(γ = tan φ) is positive and negative if acute and obtuse angles are the results, respectively. Equation
(1.32) is the Laglangian description of the shear strain, whereas that of the Eulerian description is
expressed with B−1 instead of C. Determination of a strain ellipse from angle changes in fossils and
other geological strain markers is explained in great detail in [189].

1.7 Shear zone


1.7.1 Strain tensor for simple shear
In order to study finite deformations in shear zones, consider a simple shear deformation (Fig. 1.11).
In this case, the deformation gradient is
⎛ ⎞
1 2q 0
F = ⎝0 1 0 ⎠ , (1.33)
0 0 1
so that the right and left Cauchy-Green tensors are
⎛ ⎞ ⎛ ⎞
1 + 4q 2 2q 0 1 2q 0
B = V2 = F · FT = ⎝ 2q 1 0⎠ , C = U2 = FT · F = ⎝2q 1 + 4q 2 0⎠ .
0 0 1 0 0 1

In order to calculate the eigenvalues of B, we take the upper-left 2 × 2 submatrix, and wehave its
characteristic equation, λ2 −2(2q 2 +1)+1 = 0. The solutions of this equation are 1+2q 2 ±2q 1 + q 2 ,
therefore, we have the maximum, intermediate, and minimum eigenvalues are, respectively,
 
B1 = 1 + 2q 2 + 2q 1 + q 2 , B2 = 1, B3 = 1 + 2q 2 − 2q 1 + q 2 .

The principal strains are equivalent to the eigenvalues of V = B1/2 . Thus,


   
X = 1 + 2q 2 + 2q 1 + q 2 , Y = 1, Z = 1 + 2q 2 − 2q 1 + q 2 . (1.34)
1.7. SHEAR ZONE 19

Figure 1.11: Finite strain tensors for the case of simple shear (q = 0.8).

The principal strain axes are parallel to the eigenvectors of B, hence, the X, Y , and Z axes lie in the
directions    
 
q + q 2 + 1, 1, 0 T , (0, 0, 1) T , q − q 2 + 1, 1, 0 T

On the other hand, the reciprocal strain ellipsoid has the principal radii, 1/X, 1/Y , 1/Z, and the
corresponding principal axes are
     
−q − 1 + q 2 , 1, 0 T , (0, 0, 1) T , −q + 1 + q 2 , 1, 0 T ,

respectively. Note that the maximum principal radius is not 1/X but 1/Z. The strain and reciprocal
strain ellipsoids are symmetric with respect to the coordinate planes (Fig. 1.11) that are defined so
as to write simple shear deformations in the form of Eq. (1.33).
Now let us calculate U and R for this deformation. We shall determine the components of U and
R from the polar decomposition, F = R · U. To this end, let the orthogonal tensor be
⎛ ⎞
cos ϕ − sin ϕ 0
R = ⎝ sin ϕ cos ϕ 0⎠
0 0 1

where ϕ stands for the angle of rigid-body rotation. Subsitituting this into the equation F = R · U,
20 CHAPTER 1. FINITE STRAIN

we have
⎛ ⎞⎛ ⎞ ⎛ ⎞
cos ϕ sin ϕ 0 1 2q 0 cos ϕ 2q cos ϕ + sin ϕ 0
U = RT · F = ⎝− sin ϕ cos ϕ 0⎠ ⎝0 1 0⎠ = ⎝− sin ϕ cos ϕ − 2q sin ϕ 0⎠ . (1.35)
0 0 1 0 0 1 0 0 1

However, the angle ϕ must satisfy the equation 2q cos ϕ + sin ϕ = − sin ϕ because of the symmetry
U = UT . Thus,
−q = tan ϕ. (1.36)

Therefore, −q is equal to the engineering shear strain (Fig. 1.11), and we obtain
⎛ ⎞
cos(arctan q) sin(arctan q) 0
R = ⎝− sin(arctan q) cos(arctan q) 0⎠ . (1.37)
0 0 1

From Eqs. (1.35) and (1.36), we have


⎛ ⎞ ⎛ ⎞
cos ϕ −2 tan ϕ cos ϕ + sin ϕ 0 cos ϕ − sin ϕ 0
U = ⎝− sin ϕ cos ϕ + 2 tan ϕ sin ϕ 0⎠ = ⎝− sin ϕ cos ϕ + 2 tan ϕ sin ϕ 0⎠ . (1.38)
0 0 1 0 0 1

The eigenvalues are calculated from the matrix in the right-hand side of this equation. They are
sec ϕ ± tan ϕ and 1, and are equal to the eigenvalues of V. These are the principal strains with a
different form from Eq. (1.34). For q > 0, ϕ < 0 . Thus, (sec ϕ − tan ϕ) is the greatest eigenvalue,
and the principal strains are

X = sec ϕ − tan ϕ, Y = 1, Z = sec ϕ + tan ϕ. (1.39)

Consider infinitesimal and infinite simple shear deformations. They are represented by the limits
q → +0 and q → ∞. According to Eq. (1.36), ϕ approaches zero from minusfor the case of
the infinitesimal simple shear. Therefore, R becomes an infinitesimal, clockwise rotation in Fig.
1.11. For the case of infinite simple shear (q → ∞), ϕ → −π/2, indicating that R approaches the
clockwise rotation by 90◦ . What are the components of U at the limints? Using the approximations,
cos ϕ ≈ 1, sin ϕ ≈ ϕ, and tan ϕ ≈ ϕ for an infinitesimal ϕ, Eq. (1.38) becomes
⎛ ⎞
1 −ϕ 0
lim U = ⎝−ϕ 1 0⎠ .
ϕ→−0
0 0 1

This tensor has three eigenvalues, (1 + ϕ)  1  (1 − ϕ). The eigenvectors corresponding to


the greatest and smallest eigenvalues have the components (1, 1, 0)T , and (1, −1, 0)T , respectively.
Therefore, the principal strain axes intersect the displacement vectors at 45◦ for an infinitesimal
simple shear.
1.7. SHEAR ZONE 21

Figure 1.12: Deformation patterns within a shear zone. Note the different sense of en echelon arrays
accompanied by horizontal shortening and extension for the same sense of shear. (a) The patterns
within a right-lateral shear zone. (b) En echelon arrays of folds and reverse faults within a right-
leteral shear zone. (c) En echelon arrays of normal faults within a right- and a left-lateral shear
zones.

As for the infinite simple shear, the angle goes to ϕ → −π/2. Thus, the U22 goes to infinity:
⎛ ⎞
0 1 0
lim U = lim U = ⎝1 ∞ 0⎠ .
q→∞ ϕ→−π/2
0 0 1

The eigenvector paired with the maximum eigenvalue rotates to the direction (1, x, 0)T with x → ∞.

1.7.2 Deformation accompanied by wrench faulting


Using the equations derived above, deformation patterns within a shear zone are investigated here.
Consider an east-west trending dextral shear zone accompanied by wrench faulting5 , and assume
that the shear zone is subject to a simple shear for which the amount of shear is evaluated by q (Fig.
5 A wrench fault is a strike-slip fault with a vertical fault plane or shear zone. We assume a wrench fault here, only because

sections perpendicular to the fault zone are horizontal.


22 CHAPTER 1. FINITE STRAIN

1.12(a)). If q is small, the principal strain axes intersect the fault zone at ∼45◦ . The X and Y axes
are oriented NE–SW and NW–SE, respectively. Due to the strain, sedimentary layers blanketing the
shear zone may be deformed to form a right-step en echelon array of folds (Fig. 1.12(b)). Anderson’s
theory (§6.3) predicts that a right-step array of reverse faults and a left-step normal faults can be
formed there. Namely, folds and reverse faults are the features of horizontal shortening, so that
they are arranged in a right-step en echelon manner. Normal faults are accompanied by horizontal
extension, so that they are arranged in a left-step extension.
En echelon cracks filled with formation water or magma are sometimes formed at depth. Figure
1.13 shows an example; they were filled with quartz that precipitated in the cracks, and were twisted
by incremental shearing. The S-shaped veins are the older and twisted ones.
There are many surface manifestations of geologic structures indicating a horizontal shortening
or extension on the Moon. Lunar mare basins are topographic depressions formed by large impact
cratering about 4 billion years ago. The basins are blanketed by layers mainly of basaltic flows and
breccias ejected by impact cratering. Figure 1.14 shows the surface features in southwestern Mare
Selenitatis. The mare deposits are divided into geologic units I, II and III in this region. Unit I is
older and unit III is the youngest.
The layers lie horizontally, but are folded in places to form topographic ridges, called wrinkle
ridges. Large wrinkle ridges are called arches. The ridges evidence that the deposits were affected
by horizontal shortening. Note that NNW–SSE trending ridges make left-step en echelon arrays.
Long linear rilles are believed to be grabens. There is no rille in unit III, suggesting that the
graben formation had ceased before this unit was deposited in this region. However, ridges were
formed in all the units.

1.8 Deformation history


We understand strain by comparing the present and initial shapes of strain markers. However, even
if we know both shapes, there are many deformation paths from the initial to the final configurations.
Therefore, the configurations are not enough to understand the deformation history. Other lines of
evidence are needed to constrain the path6 .
We have studied the means to quantify strain by small strain markers such as the deformed
ammonite (Fig. 1.1). However, if such strain markers show little deformation, a rock mass that
encompasses the markers can show a map-scale deformation. For example, Neogene formations in
the western side of Northeast Japan were folded and faulted since the late Miocene, resulting in the
horizontal shortening of the crust by ∼10% [201]. However, fossils yielded from the strata exhibit
little deformation.

6 It is a further problem as to how small tectonic movements represented by, e.g., earthquakes are related to geologic

features that are formed by the integral of those movements. See the next chapter for detail.
1.8. DEFORMATION HISTORY 23

Figure 1.13: (a) Schematic pictures showing mineral veins filling en echelon cracks accompanied
by dextral shear. Black arrows indicate the maximum compressive orientation, which is parallel to
each of the cracks. (b) Incremental shear rotates old veins and forms new ones, which cut the old
veins. If the new shear zone is narrower than the older, the ends of old veins may be left unrotated
to form an array of Z-shaped veins. (c) Example of the array in the Sambagawa metamorphic belt,
Wakayama, Southwest Japan. Metamorphic foliation is folded showing the dextral sense of shear
along the en echelon veins.

1.8.1 Stratigraphy
Sedimentary rocks are useful to study tectonics, because they have preserved a variety of information
since they were deposited, and their depositional ages can be easily determined. A sedimentary
sequence can be compared to a magnetic tape. Therefore, a sedimentary basin provides a wide
24 CHAPTER 1. FINITE STRAIN

Figure 1.14: Tectonic features in the southeastern Mare Selenitatis on the Moon. The boundary
between the dark unit I and the somewhat brighter unit III is indicated by wedges. Courtesy of
NASA.

variety of evidence to constrain its tectonic history. The law of superposition and cross-cutting
relationships are the clues. In addition, sedimentary layers are often assumed to be horizontal when
they were deposited. Particular species of fossils are the indicators of paleo-environments. Some
benthic fossils, e.g., molluscs and benthic foraminifera, are used to infer a paleo water depth. If
a sedimentary pile is inferred to have accumulated upon a shallow continental shelf and is a few
kilometers thick, the strata indicates syn-sedimentary subsidence of the basement, because eustatic
sea-level changes have amplitudes smaller than the thickness by an order of magnitude.
Figure 1.15 shows a seismic profile taken from the southwestern margin of the Japan Sea. The
margin was subject to a folding event at the end of the Miocene. To the south of the MITI Tottori-
Oki Well, a north-dipping reverse fault affects the Koura, Josoji, and the lower part of the Furue
Formations, suggesting that the faulting was terminated in the middle of the Furue stage. Folded
1.8. DEFORMATION HISTORY 25

Figure 1.15: Seismic profile through the southwestern margin of the Japan Sea [237]. The Furue,
Josoji and Koura Formations deposited in the Miocene, and were folded at the terminal Miocene
time. Note that Pleistocene sediments are not affected by the folding.

strata are truncated with an erosional surface. This cross-cutting relationship tells us that the folding
is older than the erosion. It is probable that the folding created ridges and troughs, but they were
eroded to form a flat surface on which the Pleistocene blankets. A reverse fault at the middle of this
figure affects not only the Miocene formations, but also the lower Pleistocene. The relative timing
of sedimentation, deformation, and erosion can be determined with this profile. In addition, the ages
of strata that were determined at the MITI Tottori-Oki Well present constraints to the absolute ages
of the events.
Structural geology has a weak point that geologic structures do not tell the ages of their for-
mation. Therefore, some articles implicitly assume that deformation followed immediately after
deposition, and that deformation and depositional ages are virtually the same. Coexisting structures
are sometimes considered to have been formed at the same time. However, evidence other than the
structures is needed to establish the age of deformation. For the case of the off-shore region shown
in Fig. 1.15, the timing of folding was revealed by the fossil and radiometric ages of the strata that
were cored at the well.

1.8.2 Balanced cross-section

The procedure of cross-section balancing [48] has become popular in recent years as a means of
helping to analyze and improve cross-sections through folded or faulted sedimentary layers. Geo-
metrically consistent cross-sections tell not only the present configurations but also the history of the
26 CHAPTER 1. FINITE STRAIN

Figure 1.16: Balanced cross-section (upper panel) through western Pakistan where the convergence
between the Eurasian and Indo-Australian plates is accommodated by folding and faulting [204].
The lower panel shows the configuration of the strata prior to the deformation. Rock masses bounded
by faults are restored to their initial position relative to the right side of the section. Comparison of
the panels indicates a horizontal shortening of this area by 2.6 km.

construction7 and the amount of deformation (Fig. 1.16).


The key steps involved in the procedure is the restoration of the beds depicted in the cross-section
to the relative positions that they had prior to deformation. The original states of the beds are found
with the assumption that sedimentary layers lie horizontally when they were deposited. This is not
valid if their basement did not have an horizontal surface or if sedimentation was coeval with the
deformation. Sedimentological studies reveal the architecture of a sedimentary basin. For example,
it is possible to infer whether the strata in question accumulated upon a slope or vast plain. Syn-
sedimentary tectonics often results in horizontal variations of layer thickness and lithofacies that
were affected by growing paleo basins and swells. Therefore, the validity can be tested. Conse-
quently, we can assume the initial attitude of the beds for some cases. In addition, the conservation
7 See [248, §6.4] for the explanation about the relative timing of the stacking of fault blocks for the case of “hinterland

dipping duplex”. The geometry of the fault blocks and their relative positions constrain the history.
1.8. DEFORMATION HISTORY 27

Figure 1.17: Restoration of faulted soft sediment in the Upper Pleistocene Shimosa Group, central
Japan. The left photograph was taken at an outcrop on which a lens cap about 5 cm across was put
as a scale. Faulting was due to a landslide [11]. The photograph was separated along the faults into
pieces, which were restored to the original positions relative to each other (right panel).

of bed length or bed area on the section is used as a constraint for the restoration. When all the
pieces are restored, key beds should be contiguous across faults on the section. The cross-setions
are constructed by thoughtful analysis of fault shapes and of the conservation and satisfy the initial
condition from the pieces of information on the present attitude and position of strata and faults ob-
served at outcrops or boreholes. The section that satisfies all these requirements are called a balanced
cross-section8 . And, the geologic structures depicted on the section is said to be retrodeformable.
Such sections show the amount of deformation. The section in Fig. 1.16 indicates a 13% short-
ening of a shallow level of the crust. This was estimated by comparing the horizontal displacement
of the upper part of the Maliri Pin and the original distance of the Maliri and Indus Pins.
The retrodeformability is useful to estimate not only shortening but also extensional deforma-
tions, and apply the consept to meso-scale deformations. Meso-scale normal faults that cut Pleis-
tocene sandy soft sediment are shown in Fig. 1.17(a). The photograph was broken into pieces along
the faults and sedimentary layers were restored to their original positions (Fig. 1.17(b)). There are
gaps and overlaps in the latter configuration, possibly due to plastic deformation of the fault blocks
and more probably to the component of displacements of the blocks perpendicular to the section.
Although each fault represents discontinuous movements, coarse graining of the gross deforma-
tion of the sediment allows us to ensure that the deformation is uniform (Fig. 1.18). The strain ellipse
that represent the gross deformation has its major axis subparallel to the lamination. The horizontal
8 See [260] for further reading on cross-section balancing.
28 CHAPTER 1. FINITE STRAIN

Figure 1.18: Strain estimated from a balanced cross-section. (a) The external form of the restored
section (1.17(b)) and a dark gray circle with a radius of 10 cm drawn on the figure. (b) The circle
is broken into arcs by normal faulting. Thin lines indicate the faults and frame of the photograph
(1.17(a)). A light gray ellipse is drawn to fit the arcs, and approximate the strain ellipse representing
the strain of the sediment by fault movements. (c) Parameters of the strain ellipse. The semi-major
and minor are 12 and 8.3 cm long, respectively.

and vertical strains are estimated to be E1 ≈ (12 − 10)/10 = 20% and E2 ≈ (8.3 − 10)/10 = −17%,
respectively.

1.9 Exercises
1.1 The increase of p and q in Eqs. (1.13) and (1.13) indicates progressive pure shear and simple
shear deformations. Determine how engineering shear strain should vary for the two cases.

1.2 Show that the principal radii of reciprocal strain ellipsoid is determined as the eigenvalues of
the tensor U−1 .

1.3 A two-dimensional homogeneous deformation reshapes an ellipse to a different ellipse. As-


sume that the equation ξ · A · ξ = 1 indicates the initial ellipse. Show the equation of the final shape
by the deformation F.

1.4 It is possible to quantify two-dimensional strain from the assemblage of objects such as ooids
and rounded gravels whose section were elliptical before strain (Fig. 1.19), assuming (1) a homo-
geneous deformation for the assemblage, (2) random orientations of their pre-strain major-axis9 and
(3) a variation in grain shapes [188, 120]. That is, the optimal strain ellipse for the assemblage is
determined from their present eccentricity and long-axis orientations (Table 1.1). The eccentricity
9 Actually, sediments have anisotropic grain fabric to some extent: major-axes of grains tend to have one or more dominant

orientations. A mathematical inverse method for determining the optimal strain ellipse from those grains has been made
recently [267]. In addition, the method evaluates the confidence interval of the strain ellipse.
1.9. EXERCISES 29

Figure 1.19: (a) Photomicrograph showing deformed calcareous ooids [94]. (b) Ellipses fitted to
18 grains in the photomicrograph. Long-axis orientations, φf , are measured from the reference
orientation, where the subscript ‘f’ stands for the quantities after strain (final stage).

is defined by the ratio of long and short axes, R, called the aspect ratio. Formulate a mathematical
inversion to determine the aspect ratio Rs and long-axis orientation φs of the optimal strain ellipse
from the aspect ratios and orientations of the deformed grains [120].

Table 1.1: Aspect ratios, Rf , and long-axis orientations, φf , of the 18 grains in Fig. 1.19.

Rf 1.8 2.3 2.1 2.0 1.6 1.6 1.9 1.8 1.7


φf 18◦ 16◦ 24◦ 2◦ 12◦ 28◦ 11◦ 11◦ 18◦
Rf 1.6 1.8 1.5 1.5 1.8 1.4 1.6 1.7 1.7
φf 17◦ 12◦ 18◦ 10◦ 8◦ 12◦ 16◦ 24◦ 15◦
Chapter 2

Infinitesimal Strains and Their


Accumulations
In this chapter we will study infinitesimal deformation and the rate of deformation. Most
geologic structures are the integral of infinitesimal deformations, so that it is important
to understand how infinitesimal and finite deformations are related.

2.1 Infinitesimal strain


The lifetime of a deformation zone or an orogenic belt can be over millions of years, during which
time rock masses undergo large deformations. Tectonism is fairly slow and the incremantal defor-
mation in our usual time-scale is very small. Therefore, the deformations detected by a geodetic
survey are usually very small. Geology detects the result of long-term deformations.
Geological measurement is much less accurate than geodetic one. Thus, it is usually difficult
for geology to detect such small strains as geodesy detects. However, it is possible in some cases1
that geologic structures provide clues to quantify those strains. In addition, it is necessary to know
the theories of infinitesimal strains to understand the following sections in which the dynamics of
tectonic deformatoins is discussed.

2.1.1 Definition
Infinitesimal strain is a very small strain. If a deformation gradient tensor is nearly equal to the iden-
tity tensor (F ≈ I), it is called infinitesimal deformation. Let us first rewrite F with the displacement2 ,

u = x − ξ.
1 Note that strain is defined as the change of length compared to the original one. Given a tiny strain, the associated length

change can be large enough for detection by geological measurement.


2 The famous textbook by Turcotte and Schubert [245] defines the displacement u = ξ − x in order to deal with contraction

and compression as possitive strain and stress, respectively.

31
32 CHAPTER 2. INFINITESIMAL STRAINS AND THEIR ACCUMULATIONS

We assume that the deformation and associated displacement are very small. Namely, ξ ≈ x. Three
components of ξ are independent of each other, consequently ∂ξi /∂ξj = δij . Therefore, we have the
deformation gradient tensor

∂x ∂(ξ + u) ∂u ∂u
F= = =I+ ≈I+ .
∂ξ ∂ξ ∂ξ ∂x

The last term is a tensor indicating the spatial gradient of displacement, so that we use the expression
∇u = ∂u/∂x for it. Taking the limit u → 0, we obtain

F = I + ∇u,

where the small departure from I is indicated by ∇u = ∂u/∂x. The equation F = I indicates no
deformation. Let us write the difference as δF, with which we have the deformation gradient for
infinitesimal deformation
F = I + δF (δF ≈ O), (2.1)

where
δF = ∇u.

In this case,
2 G + I = FT · F = (F + δF) T · (F + δF) ≈ I + δF + δFT ,
 
where the quadratic term δFT · δF was neglected. Geometrical linearity applies if this approxima-
tion is satisfied.
Rearranging the above equation, we have Green’s strain tensor

1  1 
G= δF + δFT = ∇u + (∇u)T , (2.2)
2 2
the components of which are
   
1 ∂uj ∂ui 1 ∂uj ∂ui
Gij = + = + . (2.3)
2 ∂ξi ∂ξj 2 ∂xi ∂xj

Almansi’s strain tensor (Eq. (1.25)) is linearlized to be


 −1    −1
2 A + I = 2 I − F · FT = 2 I − I + δF · I + δFT
  −1
≈ 2 I − I + δF + δFT ≈ I + δF + δFT .

Therefore, Green’s and Almansi’s strain tensors are identical (G = A) for infinitesimal strains. We,
accordingly, definethe infinitesimal strain tensor,
 
1  1 ∂uj ∂ui
E= ∇u + (∇u) T or Eij = + . (2.4)
2 2 ∂xi ∂xj
2.1. INFINITESIMAL STRAIN 33

This is often referred to as a strain tensor. Elongation and contraction are indicated by the pos-
itive and negative components of this tensor (2.1.3). If we use the opposite sign convention, the
infinitesimal strain tensor is written as
 
1  1 ∂ui ∂uj
e = − ∇u + (∇u) T or = − + .
2 2 ∂xj ∂xi
It is obvious that the infinitesimal strain tensors E and e satisfy the following equations,

E = ET , e = eT , E = −e. (2.5)

2.1.2 Relationships among strain tensors for infinitesimal deformations


We have seen that the infinitesimal strain tensor E was derived from G and A by geometrical lin-
earization. Then, how is E related to other strain tensors? By the linealization, the right and left
Cauchy-Green tensors become
    1  1
U = I + δFT · I + δF 2 ≈ I + δF + δFT 2
1 
≈I+ δF + δFT = I + G = I + E = V, (2.6)
2
indicating that both tensors coincide with each other for infinitesimal strains. The orthogonal tensor
R, which is related to the deformation gradient F by the equation F = R · U, becomes a tensor
approximately equal to the unit tensor:
   δF δFT −1    δF δFT 
R = F · U−1 = I + δF · I + + ≈ I + δF · I − −
2 2 2 2
1 
≈I+ δF − δFT = I + X. (2.7)
2
The last term is called the rotation tensor defined by the equation
 
1  1 ∂ui ∂uj
X≡ ∇u − (∇u) T or Ωij = − . (2.8)
2 2 ∂xj ∂xi
E and X are the symmetric and antisymmetric parts of ∇u, respectively. Using Eqs. (2.6) and (2.7),
we obtain
U + R = 2 I + δF = I + F = 2 I + G + X.
Thus
F = U + R − I = V + R − I = I + G + X. (2.9)
Since G approaches E for small strains,

F = I + E + X. (2.10)

For finite strains F = R · U = U · R, as matrix maltiplications are generally non-commutative. By


contrast, matrix summations are commutative, so that Eq. (2.10) indicates that the order of rotation
X and shape-change E are commutative for infinitesimal deformations.
34 CHAPTER 2. INFINITESIMAL STRAINS AND THEIR ACCUMULATIONS

2.1.3 Physical interpretation of infinitesimal strain and spin tensors


To interpret the tensors E and X, we shall see how a small rectangle element deforms. Let us
examine the diagonal components of E, first. If displacement is parallel to the coordinate axis O-1
(Fig. 2.1(a)),
∂ui
u2 = u3 = 0, =0 (i = 1, 2, 3; j = 2, 3).
∂xj
Therefore, all the components of E vanishes except for E11 . In this case, neglecting higher-order
terms, we have the length change
ds ds0

  
 
∂u1 ∂u1
ds − ds0 ≈ u1 + dx1 + dx1 − u1 − dx1 = dx1 = E11 dx1 .
∂x1 ∂x1
Therfore, we obtain (ds − ds0 )/ds0 = E11 . Accordingly, the diagonal terms of E represent the
elongation along the coordinate axes.
Using the angles θ and ψ shown in Figs. 2.1(b) and (c), we have
∂u1 ∂u2
2E12 = 2E21 = + = tan ψ + tan θ.
∂x2 ∂x1
For infinitesimal deformations, the last two terms of this equation become ψ + θ. These two com-
ponents indicate the angular decrease of the corner at which two sides of the rectangle originally
parallel to the axes O-1 and -2 meet. The indices of E12 and E21 correspond to the axes. Conse-
quently, 2E12 and 2E21 represent the engineering shear strain of the corner. The term “shear strain”
often refers to the diagonal components of E, which are half of the engineering shear strains. Com-
paring Figs. 1.11 and 2.1(b), the parameter q in Eq. (1.13) is related to the displacement gradients
as ∂u1 /∂x2 = 2q and ∂u2 /∂x1 = 0. Therefore, q equals shear strain for infinitesimal simple shears.
The infinitesimal strain tensor is symmetric (Eq. (2.5)), so that it has three real eigenvalues, E1 ,
E2 , and E3 , and three mutually perpendicular eigenvectors. Consider a rectangular parallelepiped
whose edges are parallel to the eigenvectors. The length of the edge parallel to the ith eigenvector is
elongated by (1 + Ei ). Therefore the volume change is given by
V
− 1 = (E1 + 1)(E2 + 1)(E3 + 1) − 1 ≈ trace E = EI , (2.11)
V0
where V0 and V stand for the volume before and after deformation. This quantity is called the
dilatation, and is equal to the first basic invariant of the tensor E for infinitesimal strains.
Figure 2.1(d) shows the counterclockwise rotation of a rectangle around the O-3 axis. Two sides
at the corner were parallel to the coordinate axes before deformation. We are considering very small
rotations associated with a infinitesimal deformation, so that the sides that were initially parallel
to the O-1 and -2 axes are rotated by the angles ∂u2 /∂x1 and −∂u1 /∂x2 , respectively. If it is a
rigid-body rotation, the two quantities are equal. The average of these angles is
 
1 ∂u2 ∂u1
Ω12 = − .
2 ∂x1 ∂x2
2.1. INFINITESIMAL STRAIN 35

Figure 2.1: The gradient of u and the deformation of a rectangle (dotted line), the sides of which
were originally parallel to the coordinate axes.

Accordingly, the ijth component of X represents the mean rotation about the kth coordinate axis,
where i = j = k.
Generally an antisymmetric tensor has three independent components. The vector that is made
up of these components is called the axial vector of the antisymmetric tensor. Rotation vector  =
(1 , 2 , 3 )T is defined as the axial vector of the rotation tensor:
⎛ ⎞
0 −3 2
X = ⎝ 3 0 −1 ⎠ . (2.12)
−2 1 0

The above equation is alternatively written as

Ωij = − ijk k . (2.13)

Not only the correspondence of the components, but also the tensor and vector show an important
property—for an arbitrary vector a the dot-product of the tensor and the cross-product of the vector
36 CHAPTER 2. INFINITESIMAL STRAINS AND THEIR ACCUMULATIONS

are identical:
X · a =  × a. (2.14)
This is a vector quantity, and the first component is
Left-hand side = −3 a2 + 2 a3 ,
Right-hand side = 2 a3 − 3 a2 ,
therefore both sides are identical. The identity of Eq. (2.14) is readily demonstrated for other
components. Combining Eqs. (2.8) and (2.12), we obtain
 
1 ∂u3 ∂u2 ∂u1 ∂u3 ∂u2 ∂u1 T
= − , − , − ,
2 ∂x2 ∂x3 ∂x3 ∂x1 ∂x1 ∂x2
which is the expression of the rotation vector by means of displacement. It is obvious from the
right-hand side that
1
 = (∇ × u) .
2

2.2 Time rate of change


When we describe the temporal change of variables such as the density and velocity, the time rate
of change of the variables is important for theoretical considerations. Let us first take a position as
a variable. The time rate of change of position is velocity of a particle. Consider that the particle
occupied the position ξ at time t = 0 and x(ξ, t) at time t. The volocity is the time derivative
∂x(ξ, t)
v = v(ξ, t) = . (2.15)
∂t
This has the Laglangian coordinates (ξ1 , ξ2 , ξ3 ) and time t as the arguments, so that this is the Laglan-
gian description of velocity. The function v(ξ, t) describes the velocity change that would be mea-
sured by an observer traveling with the particle. On the other hand, there is the Euler description of
velocity v = v(x, t) that indicates the velocity change that would be observed at the specific location
x.
There are Euler and Laglangian descriptions for the time rate of change of variables. They have
the independent variables (x, t) and (ξ, t), respectively:

∂ ∂ 

Euler or spatial derivative ≡ , (2.16)
∂t ∂t x=const

D ∂ 

Laglangian or matrial derivative ≡ (2.17)
Dt ∂t ξ=const.

and are generally different from each other. The operator D/Dt stands for the Laglangian derivative,
which is also called the material derivative. Material derivative is often denoted by a dot, for example
Dx
ẋ = = v.
Dt
2.3. STRAIN RATE 37

If the operand is composed of plural symbols, we put the operand in parentheses and put a dot on
the right parenthesis such that
DFT  T ·
= F .
Dt
How are the two time derivatives related? Let F be a scalar function of position and time. The
material derivative of the function is
DF 1    
= lim F x(ξ, t + Δt), t + Δt − F x(ξ, t), t . (2.18)
Dt Δt→0 Δt

Because of x(ξ, t + Δt) = x(ξ, t) + vΔt + · · · , the first-order approximation of the first term in the
square brackets above is

F (x(ξ, t + Δt), t + Δt) ≈ F (x(ξ, t) + vΔt, t + Δt) . (2.19)

Note that the Taylor expansion of a function f (x, y) is


∂f ∂f
f (x0 + Δx, y0 + Δy) = f (x0 , y0 ) + Δx + Δy + · · ·
∂x ∂y
The first-order approximation of the right-hand side of Eq. (2.19) becomes
∂F
F (x(ξ, t) + vΔt, t + Δt) = F (x(ξ, t), t) + ∇F · (vΔt) + Δt.
∂t
Using these approximations we rewrite Eq. (2.18) and get
∂  ∂   
D ∂
F (x, t) = + vi F (x, t) = + v · ∇ F (x, t) (2.20)
Dt ∂t i
∂xi ∂t

or simply
D ∂
= + v · ∇. (2.21)
Dt ∂t
This formula is applicable to vectors and tensors. Let F = (F1 , F2 , F3 )T be a vector field that
depends on time, then it is rewritten by the sum of the products of the scalar functions and base
vectors: F (x, t) = F1 (x, t)e(1) + F2 (x, t)e(2) + F3 (x, t)e(3) . Because of the independency of the base
vectors on time, we have DFi e(i) /Dt = e(i) (DFi /Dt). Therefore, the material derivative of a vector
is obtained by taking the material derivative of each components. We obtain the material derivative
of tensors in the same way. Accordingly, we simply write the material derivatives of the vectors and
tensors such that  
DF ∂
= + v · ∇ F.
Dt ∂t

2.3 Strain rate


The velocity of deformation is a basic quantity for the theory of tectonics. Velocity is the change
during a very short time dt. The deformation in the infinitesimally short period is an infinitesimal
deformation.
38 CHAPTER 2. INFINITESIMAL STRAINS AND THEIR ACCUMULATIONS

2.3.1 Velocity gradient


The time rate of deformation is obtained from the velocity gradient tensor
 
∂v ∂vi
L≡ = ∇v = . (2.22)
∂x ∂xj
The difference in velocity between two positions that are separated by the distance dx is
 
∂v
dv = · dx = L · dx.
∂x
Several important quantities are obtained from the velocity gradient tensor—first, we obtain the
material derivative of the deformation gradient
DFij D ∂xi ∂ Dxi ∂vi  ∂vi ∂xk
= = = = .
Dt Dt ∂ξj ∂ξj Dt ∂ξj ∂xk ∂ξj
k

∴ Ḟ = L · F. (2.23)

The velocity gradient tensor is divided into symmetric and antisymmetric parts
 
1  1 ∂vi ∂vj
D≡ L+L = T
+ , (2.24)
2 2 ∂xj ∂xi
 
1  1 ∂vi ∂vj
W≡ L − LT = − (2.25)
2 2 ∂xj ∂xi
which are called the stretching tensor and spin tensor, respectively. Obviously, we have

L = D + W. (2.26)

Note that the definition of D is similar to that of E (Eq. (2.4)). They have the relation

D = Ė. (2.27)

This is derived as follows:


   
D 1 ∂ui ∂uj 1 ∂ Dui ∂ Duj
Ėij = + = +
Dt 2 ∂xj ∂xi 2 ∂xj Dt ∂xi Dt
   
1 ∂ D(xi − ξi ) ∂ D(xi − ξi ) 1 ∂vi ∂vj
= + = + = Dij .
2 ∂xj Dt ∂xi Dt 2 ∂xj ∂xi

The material derivative of infinitesimal strain tensor Ė has identical components with D, although
they have different definitions. The former is referred to as the strain rate tensor. As the name
suggests, Ė indicates the velocity of strain. This is the infinitesimal strain during a short period dt,
so that the trace of Ė represents the velocity of the volume change (2.11). Namely, the equation

J˙ = (trace L) J = (∇ · v)J (2.28)


2.4. PILE UP OF INFINITESIMAL DEFORMATIONS 39

is obtained (Exercise 2.2) as the time rate of the Jacobian (Eq. (1.7)), Therefore, if ∇ · v = 0,
then we have J˙ = 0. This is a differential equation indicating that J has no temporal change. The
equation has the solution J = const. The constant should be 1, otherwise J˙ = 0 when a volume
change occurs. Consequently, a rock mass has no volume change if ∇ · v = 0. If the velocity field
satisfies the equation
∇ · v = 0, (2.29)

the deformation is called incompressible. In that case, density does not change, either. Incom-
pressibility is often assumed in tectonic models, but is not applicable to phenomena with significant
density changes associated, for example, with metamorphic reactions from low to high pressure
phases.
How fast are actual tectonic movements? A million years is often used as the unit to measure
geologic ages. The lifetime of an orogenic belt is on the order of ten to a hundred million years
during which rock masses deform by hundreds of %. A million years is approximately equal to
3.16 × 1013 s, because a year is approximately 60 seconds × 60 minutes × 24 hours × 365.25 days
= 31,557,600 s. If a rock mass deforms by 100 % in a million years, the strain rate is 1/(3.16 × 1013
s) ≈ 3.17 × 10−14 s−1 .
Tectonics of island arcs is often faster than that of continents [263]. Extensional deformations
in the Early Miocene was as fast as ∼ 10−13 − 10−14 s−1 in western Northeast Japan (§12.2). The
Northeast Japan arc has been subject to compressional tectonics since the Pliocene, and deforma-
tion was accelerated in the Pleistocene. The strain rate representative for Northeast Japan has been
estimated by several methods. A seismological estimate from moment tensor accumulation (§2.7)
indicates the rate to be about 1 × 10−16 s−1 [258]. This is the rate averaged over a few decades. That
over the last ∼100 years was determined by triangulation at (3 − 6) × 10−15 s−1 [160]. Northeast
Japan has undergone compressional deformations, mainly in the Quaternary [202]. Folds and faults
have shortened the crust under the arc. They were combined with chronostratigraphy to infer the
strain rate averaged over the last 1.3 million years [201]. The result was on the order of 10−14 s−1 .
The discrepancy between the estimated rates may shed light to non-steady movements.
Stable continents are not rigid, but are very slowly deforming. As an example, earthquakes
rarely occur in Fenoscandia. The stress field responsible for such seismic activity is consistent with
the combination of plate boundary forces with the plate flexure accompanied by Holocene glacial
rebound [229]. The average strain rate of cratonic lithosphere is on the order of 10−15 s−1 [37]. With
a strain rate below the order of, say, 10−17 , the region is too slow, even for geology, to detect and
may seem rigid.

2.4 Pile up of infinitesimal deformations


Most tectonic structures are the result of finite deformations that are an accumulation of infinites-
imal deformations. Therefore, it is important for tectonics how finite deformations are related to
infinitesimal deformations.
40 CHAPTER 2. INFINITESIMAL STRAINS AND THEIR ACCUMULATIONS

2.4.1 Reference configuration


We have considered deformation by comparing the configurations that were represented by position
vectors ξ and x at the time 0 and t. However, the rate of deformation changes with time. The rate is
the infinitesimal deformation within a infinitesimal time interval dt, and the reference configuration
for the infinitesimal deformation is renewed every moment. Accordingly, consider the times t = 0, t
and τ, where |t − τ| = dt. Assuming the configuration at time t be the reference, we have

dx(τ) = Ft (τ) · dx(t),

where dx(τ) is the infinitesimal vector at the time τ and Ft (τ) is the and deformation gradient tensor
at the time τ with the reference configuration at the time t. Thus,
 
∂dx(τ)  ∂Ft (τ) 
 · dx(t) = Ḟt (t) · dx(t).
dv(t) = =
∂τ  τ=t ∂τ  τ=t
Note that
∂x(τ)i  ∂x(τ)i ∂x(t)k
F0 (τ)ij = = .
∂x(0)j ∂x(t)k ∂x(0)j
k

Namely, F0 (τ) = Ft (τ) · F0 (t), where F0 (t) is the deformation gradient tensor at time t with the
reference one at the time 0. Velocity gradient at time t is, therefore,
 
∂Ft (τ)  ∂ F0 (τ) · F0 (t)−1 

L(t) = Ḟt (t) =  =  = Ḟ0 (t) · F0 (t)−1 . (2.30)
∂τ  τ=t ∂τ 
τ=t

This is identical to Eq. (2.23), but its reference is indicated. Using the polar decomposition theorem
Ft (τ) = Rt (τ) · Ut (τ), the above equation becomes
  
∂Ft (τ) 
 = Rt (τ) · U̇t (τ) + Ṙt (τ) · Ut (τ)
L(t) = = Ṙt (t) + U̇t (t).
∂τ  τ=t τ=t

Similarly, we obtain L(t) = Ṙt (t) + V̇t (t). Comparing these and L = D + W,

D(t) = U̇t (t) = V̇t (t) (2.31)


W(t) = Ṙt (t). (2.32)

These equations relate stretching and spin tensors with the time rate of the tensors indicating finite
strain and rotation within a moment at time t.
However, it should be noted that the relations (Eqs. (2.31) and (2.31)) do not hold if the con-
figuration at time 0 is chosen for the reference. This is demonstrated as follows. Substituting
F0 (t) = R0 (t) · U0 (t) into Eq. (2.30),

L(t) = Ṙ0 (t) · U0 (t) + R0 (t) · U̇0 (t) · U0 (t)−1 · R0 (t)T
= Ṙ0 (t) · R0 (t)T + R0 (t) · U̇0 (t) · U0 (t) · R0 (t)T . (2.33)
2.4. PILE UP OF INFINITESIMAL DEFORMATIONS 41

Figure 2.2: Schematic cross-section of a pure shear rift, the central part of which is undergone pure
shear.

On the other hand, taking the material derivative of both sides of R · RT = I, we have

Ṙ0 (t) · R0 (t)T + R0 (t) · Ṙ0 (t)T = O. (2.34)

Combining Eqs. (2.33) and (2.34), we obtain the equations


1
D(t) = R0 (t) · U̇0 (t) · U0 (t)−1 + U0 (t)−1 · U̇0 (t) R0 (t)T , (2.35)
2
1
W(t) = Ṙ0 (t) · R0 (t)T + R0 (t) · U̇0 (t) · U0 (t)−1 − U0 (t)−1 · U̇0 (t) · R0 (t)T (2.36)
2
which indicates the relationship between instantaneous rates at time t (left-hand side) and finite
deformation and rotation (right-hand side) since the time 0. Note that W is no longer equal to
the material derivative of R indicated by Eq. (2.32). In the case of t → 0, both R0 (t) and U0 (t)
approaches I, hence, Eq. (2.36) reduces to Eq. (2.32).

2.4.2 Pure shear rift


Continental rifts have wide variations: there are, for example, wide and narrow rifts and symmetric
as well as asymmetric ones. If the lithospheric thinning is roughly approximated as a pure shear
on the vertical section across the rift zone, it is called a pure shear rift (Fig. 2.2). In contrast, if
extension occurs by displacement along a normal detachment fault that extends completely through
the crust, it is called a simple shear rift . The North Sea was a rift zone that was active before ocean
floor spreading began at the Mid Atlantic ridge.

Strain rate of pure shear In order for pure-shear rifts, we shall derive the expression for the strain
rate of coaxial deformation that has a deformation gradient in the form of Eq. (1.14). The velocity
gradient tensor is derived using Eq. (2.23):
⎛ ⎞⎛ ⎞ ⎛ ⎞
Ẋ 0 0 1/X 0 0 Ẋ/X 0 0
L = Ḟ · F−1 = ⎝ 0 Ẏ 0 ⎠ ⎝ 0 1/Y 0 ⎠=⎝ 0 Ẏ /Y 0 ⎠.
0 0 Ż 0 0 1/Z 0 0 Ż/Z
42 CHAPTER 2. INFINITESIMAL STRAINS AND THEIR ACCUMULATIONS

All the non-diagonal components are zero so that the spin tensor W vanishes for coaxial deforma-
tions3 . The stretching tensor, on the other hand, is identical to the volocity gradient tensor
⎛ ⎞
Ẋ/X 0 0
D=L=⎝ 0 Ẏ /Y 0 ⎠. (2.37)
0 0 Ż/Z

Coaxial, incompressible, plane strain is a pure shear. Taking the Y -axis perpendicular to the plane,
the plane strain condition results in Ẏ /Y = 0. Volume conservation is represented by trace L = 0.
Therefore, we have Ẋ/X = −Ż/Z for pure shear. In this case, the strain rate is
⎛ ⎞
Ẋ/X 0 0
Ė = D = ⎝ 0 0 0 ⎠. (2.38)
0 0 −Ẋ/X

The horizontal stretch of the lithosphere compared to the pre-rift stage is usually referred to as
the β factor in theoretical models of continental rifting. Taking the x and z axes in horizontal and
vertical directions (Fig. 2.2), Eq. (2.38) is rewritten to
⎛ ⎞
β̇/β 0 0
Ė = ⎝ 0 0 0 ⎠ (2.39)
0 0 −β̇/β.

Based on this equation, we are able to calculate the finite deformation of a pure shear rift from the
temporal variation of the rate of strain. It should be noted that if the rate is constant, the β factor
increases non-linearly (Exercise 2.3). Quantitative stratigraphy (§3.10) allows us to infer the history
of lithospheric thinning, from which the temporal variation of the rate of strain can be determined
(§3.15).

2.5 Rotation
Tectonic rotations are observable from geological objects including geologic structures and paleo-
magnetic deflections. A train of twisted en echelon veins (Fig. 1.13) is an example of the former.
Paleomagnetism is often used to study large-scale tectonic rotations. Accordingly, let us study how
rotation is related to a velocity field.

2.5.1 Angular velocity


Consider an angular velocity vector w = we, where w is the angular velocity and e is the unit vector
parallel to the rotation axis pointing as a right-handed screw. If the initial point of a position vector
a is on the axis of rotation, the radius of rotation of the end point is |a| sin θ, where θ is the angle
between e and a (Fig. 2.3(a)) The angle satisfies the equation, e × a = |e| |a| sin θ = |a| sin θ,
3A spin tensor characterizes rotational motion so that the velocity field of coaxial deformation is irrotational.
2.5. ROTATION 43

Figure 2.3: (a) Rotation corresponding to the angular velocity w = we, where e is the unit vector
parallel to the rotation axis. (b) Angular velocity is an infinitesimal rotation in an infinitesimal time
interval. Infinitesimal rotation can be resolved to the successive rotations around the coordinate axes.
The components of w indicate the rotation angles.

which equals the radius of rotation of the end point. In addition, the vector e × a points the direction
that the end point is orbiting. Since the angular velocity is w, the orbiting velocity of the point is
we × a = w × a. Similar to Eqs. (2.13) and (2.14), the spin tensor W and its axial vector w have the
following identities:

Wij = − ijk wk , (2.40)


W · a = w × a, (2.41)

where a is an arbitrary vector.


We have seen that for rigid-body rotation, an angular velocity vector represents both the rotation
axis and angular velocity. The components of w indicates the rotation around the ith coordinate axis
(Fig. 2.3(b)). It is discussed in the following subsection how a velocity field is related to the angular
velocity.

2.5.2 Physical interpretation of spin tensor

Vorticity Given a velocity field, the vector

 
∂v3 ∂v2 ∂v1 ∂v3 ∂v2 ∂v1
ω =∇×v = − , − , − T
(2.42)
∂x2 ∂x3 ∂x3 ∂x1 ∂x1 ∂x2
44 CHAPTER 2. INFINITESIMAL STRAINS AND THEIR ACCUMULATIONS

Figure 2.4: Physical interpretation of the additive decomposition L = D + W. An infinitesimal


deformation in an infinitesimal time interval can be resolved to D and W , indicating the coaxial
strain and rotation, respectively. The sides of this rectangular solid is parallel to the principal strain
axes. Di stands for the ith principal strain. W indicates the infinitesimal rotation, the angular
velocity of which is w.

is known as the vorticity. From Eqs. (2.25) and (2.42) the components of W are obviously
⎛    ⎞
0 1 ∂v1
− ∂v2 1 ∂v1
− ⎛
∂v3 ⎞
2 ∂x2 ∂x1 2 ∂x3 ∂x1 0 − 12 ω3 1
2 ω2
⎜ ⎟
⎜    ⎟ ⎜ ⎟
⎜ ∂v3 ⎟ ⎜ ⎟
W = ⎜ 12 ∂x
∂v2
− ∂v1
0 2 ∂x3 − ∂x2 ⎟
1 ∂v2
= ⎜ 1 ω3 0 − 12 ω1 ⎟ . (2.43)
⎜ 1 ∂x2 ⎟ ⎝ 2 ⎠
⎝     ⎠
− 12 ω2 1
2 ω1 0
2 ∂x1 − −
1 ∂v3 ∂v1 1 ∂v3 ∂v2
∂x3 2 ∂x2 ∂x3 0

Thus, the vorticity is twice the vector axial vector of W. Let w be the axial vector, then

ω = 2w. (2.44)

According to the discussion on the rotation tensor and rotation vector in §2.1.3, we see that w rep-
resents angular velocity. The vector points the direction of rotation axis like a right-hand screw, and
|w| indicates the scalar angular velocity. Since a deformation within an infinitesimal time interval is
also infinitesimal, strain and rotation are additive and commutative (Fig. 2.4).

Rigid-body rotation Let us demonstrate the correspondence between the vorticity and angular ve-
locity for the case of rigid-body rotation. Given a constant a, the velocity field v = (−ax2 , ax1 , 0)T
represents a rigid-body rotation about the third
 coordinate axis. This is proved by the observation
firstly that v·x = 0 and, secondly, that |v| = a2 x22 + a2 x21 ∝ |x|. The coefficient a is the angular ve-
locity. Using the identity ∂xi /∂xj = δij , we have the vorticity of this velocity field ω = (0, 0, 2a)T .
The magnitude is twice the angular velocity, so that the proof has been established.
2.5. ROTATION 45

2.5.3 Spin within shear zone


Physical interpretation of the additive decomposition of velocity gradient tensor L = D + W may
be obtained in an analysis of simple shear deformations. Rigid-body rotation within a shear zone is
considered assuming a simple shear deformation, where the velocity field is v = (2q̇x2 , 0, 0)T and
q̇ = const. The velocity gradient is
⎛ ⎞
  0 2q̇ 0
∂vi
L= = ⎝0 0 0 ⎠ ,
∂xj
0 0 0

which is decomposed into symmetric and antisymmetric parts:


⎛ ⎞ ⎛ ⎞
0 1 0 0 1 0
D = q̇ ⎝1 0 0⎠ , W = q̇ ⎝−1 0 0⎠ . (2.45)
0 0 0 0 0 0

D and W represent shape-change and rotation, both of which do not vanish but occur simultaneously
for simple shear (Eq. (2.45)). The characteristic equation of D is
 
−λ q̇ 0
 
 q̇ −λ 0  = −λ3 + q̇ 2 λ = −q̇λ(λ − q̇)(λ + q̇) = 0.
 
0 0 −λ

The eigenvalues are, therefore, −q̇, 0 and q̇. Ė has the same eigenvalues, because D = Ė.
Suppose a square that is deformed to a parallelogram by the instantaneous simple shear, the
time interval of which is Δt. Instantaneous strain is infinitesimal. The order of strain and rotation
are commutative for infinitesimal deformations, so that we consider the deformation as a two-step
operation, strain followed by rotation (Fig. 2.5). The strain in this velocity field has the principal
axes inclined at 45◦ to the flow direction (see §1.7). The rate of strain Ė has the eigenvalues q̇, 0
and −q̇, so that the diagonals of the square have undergone strain ±q̇Δt within the interval Δt. If the

square ABCD in Fig. 2.5 has sides with a length 2a, the diagonal OA and OB has the length 2a.
The square, first, becomesthe parallelogram A B C D . OA and OB are elongated to OA and OB ,

respectively. Both AA and BB have an equal length 2aq̇Δt. The inclination of the parallelogram
is represented by the angle between AB and A B . The angle is very small for Δt → 0, so that
the tangent of the angle is approximated by the angle itself. It is seen in Fig. 2.5(a) that the tangent
equals 2aq̇Δt/2a = q̇Δt. The inclined parallelogram A B C D is rotated by the angle. The result is a
parallelogram A B C D whose two sides are parallel to the flow direction. The angle of rigid-body
rotation is q̇Δt. It follows that the angular velocity is q̇.
The same quantity is obtained from the spin tensor in Eq. (2.45). We obtain the vorticity
ω = (0, 0, −2q̇)T , where the minus sign indicates that the direction of rotation in this case is the
counterclockwise relative to the third coordinate axis. The axial vector of W is w = (0, 0, −q̇)T ,
which is identical with the angular velocity of the parallelogram in Fig. 2.5. Therefore, the axial
46 CHAPTER 2. INFINITESIMAL STRAINS AND THEIR ACCUMULATIONS

Figure 2.5: Instantaneous simple shear of a square (gray) as the successive shape-change and rigid-
body rotation. (a) The instantaneous strain by the effect of D transforms the square ABCD into the
parallelogram shown by solid line A B C D . The principal axes of this strain are inclined at 45◦ to
the flow direction. The diagonals OA and OB of the square becomes OA and OB , respectively. (b)
Rigid-body rotation of the parallelogram A B C D to A B C D , completing the simple shear.

vector and spin tensor indicate instantaneous rotation in the simple shear, though both rotation and
shape change occur.
It should be noted that the three quantities W, w and ω indicate instantaneous spin within a
deforming material, whereas the orthogonal tensor R indicates the finite rotation that is the integral
of the spin. For the case of simple shear with the constant q̇, the spin tensor is a constant tensor
(Eq. (2.45)). Equation (2.32) seems to indicate that R is proportional to time for a constant spin.
However, this is not always true. In this case, Eq. (2.36) rather than Eq. (2.32). R converges to 90◦
rotation for the simple shear, even though the spin is constant.

2.6 Paleomagnetic rotation


2.6.1 Paleomagnetic rotation determined by vorticity
We have seen in the previous section that rotation in deforming media is correlated with vorticity. So,
the rotation of a tectonic plate is often thought to indicate the magnitude of vorticity in the viscous
layer beneath the plate. In that case, if small, rigid, tectonic blocks are floating on a viscous fluid that
is subject to shearing flow, the blocks can be used as indicators of the vorticity field in the fluid. This
further allows us to estimate the total displacement across the shear zone from the paleomagnetic
rotation of the blocks in the zone. Let h be the width of the shear zone in the viscous layer, and v0 be
2.6. PALEOMAGNETIC ROTATION 47

Figure 2.6: Rotating blocks in a deforming zone between microplates A and B. Plate B moves at a
velocity v0 with respect to microplate A, so that the width of the zone w(t) increases. Block rotation
within the zone is controlled by the tangential component vx .

the relative velocity across the zone. The displacement during a time interval t is d = v0 t. If simple
shear is proceeding in the zone, the velocity field is represented by v = (v0 y/h, 0, 0), provided that
the first coordinate axis is parallel to the movement and y is the length across the zone. Vorticity is,
therefore, |ω| = |∇ × v| = v0 /h, and the angular velocity of the blocks is v0 /2h. If h and v0 t/2h is
known from geological survey and paleomagnetism, respectively, we are able to estimate the total
displacement of the shear zone.
Consider a deformation zone between microplate A and B that is moving with a relative velocity
v0 (Fig. 2.6). The microplates are thought to be rigid. If the deformation is homogeneous and plane
strain on the horizontal plane, then the velocity field in the zone is
y 
v= v0 cos θ, v0 sin θ, 0 T ,
w

where the angle θ, width of zone w and coordinates O-xyz in Fig. 2.6 are used. The vorticity is
   
∂vz ∂vy ∂vx ∂vz ∂vy ∂vx v0 cos θ
ω= − , − , − T
= 0, 0, − T
.
∂y ∂z ∂z ∂x ∂x ∂y w

This indicates that ω does not depend on the velocity component normal to the zone vy . There-
fore, paleomagnetic rotation in the deforming zone only indicates the tangential movement. This is
convenient to estimate the tangential displacement.
We again assume a simple shear. If the viscous layer is made of a power-low fluid (see §10.8),
the width of the shear zone is inversely correlated to the power-law index of the fluid. Accordingly,
Sonder et al. [222] argue the rheology of the viscous layer under rotating tectonic blocks on the
basis of paleomagnetic declinations around a strike-slip fault. The width may be indicated by the
zone where the declination (therefore vorticity) shows an anomaly (Fig. 2.7).
48 CHAPTER 2. INFINITESIMAL STRAINS AND THEIR ACCUMULATIONS

Figure 2.7: Paleomagnetic rotation of surface blocks (thick arrows) by an underlying shear zone. y
and ω indicate the coordinate across the zone and vorticity in the viscous substratum, respectively.

Figure 2.8: Schematic picture showing bookshelf faulting.

2.6.2 Paleomagnetic rotation determined by vorticity and deformation


We have seen in the preveous subsection that vorticity in a viscous layer beneath a rigid tectonic
block controls the rotation of the blocks, where the vorticity does not depend on stretching tensor
D, instantaneous deformation within the fluid. The angular velocity of the blocks is related to the
vorticity by Eq. (2.44).
Secular tectonic rotation about a vertical axis is revealed by paleomagnetism and other tech-
niques4 . Accordingly, let us consider paleomagnetic rotations. The rotation is sometimes argued
in relation to Eq. (2.44) or to bookshelf faulting [139]. The activity of a fault system is said to be
bookshelf faulting if the activity is accompanied by the rotation of fault slices just like the rotation
of books when they fall in a bookshelf. If the angular velocity of fault slices, w, is given, the relative
velocity of the blocks A and B in Fig. 2.8 is written as v = (w sin θ, w cos θ)T .
Clockwise and counterclockwise rotation of the carapace of tectonic blocks are related to dextral
and sinistral shears, respectively, when they are interpreted either with the vorticity of underlying
viscous fluid or bookshelf faulting. The following argument shows that the correspondence between
4 Vertical-axis rotation is not easy to recognize for structural geology. However, it is sometimes possible [83]. Such

rotations were inferred from the disturbance in the trend of a Jurassic dike swarm in California [199]. The geometry of fault
systems, including bookshelf faulting, is also used to infer vertical-axis rotation.
2.6. PALEOMAGNETIC ROTATION 49

Figure 2.9: (a) Fault systems (solid lines) produced by sand-box experiments after Cobbold et al.
[41]. Broken gray lines are strain markers that initially made a grid pattern at the surface of sand 20
cm thick. The box was square, 20 by 20 cm, before strain. (b) Stretch and engineering shear strain
along the x-axis for the four cases in (a).

the senses of rotation and shear has exceptions [111].


The first example came from sand-box experiments [41]. Since the shallow levels of the crust
have fractures, and the fractures are thought to affect the mode of deformation of the crust. Therefore,
sand is often used for an analog medium for the crust.
Figure 2.9(a) shows fault systems produced in a sand box. The box was subject to various strains,
the parameters of which are shown in Fig. 2.9(b). The box was only extended along the x-axis by
about 50% for the case “A”, where a conjugate fault system was created. Marker lines showed that
the deformation was accompanied by little systematic rotation of the fault blocks. In experiment
“B”, an engineering shear strain by the small amount of 0.2 was added to the case of “A”. The result
was that both sinistral and dextral faults appeared with prominent bookshelf faulting in the central
area. The overall shear sense of the sand box is sinistral for case “B.” However, there are domains of
bookshelf faulting both with clockwise and counterclockwise vertical-axis rotations. The domains
encompass fault slices with sinistral and dextral movements, respectively, between the slices. A
larger sinistral shear was applied to the box in experiment “C”, resulting in the dominant dextral
faults associated with the clockwise rotation of fault blocks. A simple shear was applied to the box
in case “D”, where the schema of bookshelf faulting depicted in Fig. 2.8 applies.
The representative rate of simple shear and stretch for the sand-box experiments are indicated by
   
0 2u v 0
L =
(1)
, L =(2)
,
0 0 0 0
50 CHAPTER 2. INFINITESIMAL STRAINS AND THEIR ACCUMULATIONS

respectively. And, the representative velocity gradient is


 
v 2u
L = L(1) + L(2) = .
0 0

Therefore, we have
   
v u 0 u
D= , W= .
u 0 −u 0

Hence, the difference between the results of cases “C” and “D” in Fig. 2.9 indicates that D can affect
the rotation of fault blocks, even if W is constant.
It is beyond our present scope to present a physical explanation for the variety of rotations.
Instead, let us study the rotation of a single rigid body surrounded by moving viscous fluid, as it is
a intricate problem to deal with plural bodies in that situation [275]. For additional simplicity, we
assume Newtonian fluid (§9.1). In general, the rotation of such a rigid particle is affected by the rate
of translational movement, strain rate, and vorticity of the surrounding fluid. For example, a windmill
rotates by the translational motion of the atmosphere. Inversely, a propeller transforms rotation to
wind. However, it is known that translation and rotation are independent for axisymmetric bodies.
Propellers do not have the symmetry, so they are capable of transformation. Further, the rotation of
bodies with spherical symmetry, or a disk for the two-dimensional case, is independent from D of
the surrounding fluid, but depends only on W of the fluid.
To see this, suppose an ellipsoidal rigid body in a simple shearing viscous fluid (Fig. 2.10).
The body represents shapes without spherical symmetry. Simple shear within a very short time is
small, so that the instantaneous strain axes for the simple shearing fluid intersect at 45◦ with the
flowing orientation. In case of a spherical body, the pressing and stretching effects along the axes
have the opposite sense of torque on the body with equal magnitudes (Fig. 2.10(b)). The result is
that the instantaneous strain of the surrounding fluid, D, does not affect the rotation of the sphere, but
the component, W, controls the rotation. By contrast, torques exerted by the effects have different
magnitudes on the surface of the ellipsoidal body (Fig. 2.10(c)), so that both D and W affect the
rotation. Here, we neglect the interaction and deformation of blocks. Such a simple model does not
have reality, but gives a hint to consider the implication of the rotation of tectonic blocks.
For further simplicity, let us consider a two-dimensional problem. Suppose an orogenic belt with
parallel borders, and the rigid terrains on both sides of this zone are going away from each other with
a constant relative velocity. The orientation of their motion is indicated by the angle θ (Fig. 2.11).
Their motion gives rise to the dextral shear and stretching5 in the belt, which is affected by the
dextral and sinistral simple shear zones for θ = 0 and 180◦ , respectively. A rift zone without strike-
slip movement is represented by θ = 90◦ . Transtensional tectonics is represented by 0 < θ < 90◦
and 90 < θ < 180◦ . Transpressional tectonics is indicated by the angles in the other two quadrants,
−180 < θ < −90◦ and −90 < θ < 0◦ .
5 The term dextral “transtension” applies to such a case.
2.6. PALEOMAGNETIC ROTATION 51

Figure 2.10: Rotation of spherical and ellipsoidal bodies in a simple shearing viscous fluid. (a)
Flow pattern far from the bodies. (b) Instantaneous flow in the fluid (gray bold lines) observed from
the spherical body is a pure shear with the strain axes intersecting at 45◦ with the far-field simple
shearing motion of the fluid. The instantaneous pure shear causes torque (thin arrows) at the surface
of the body. (c) Torque by the same flow at the surface of an ellipsoidal body.

Suppose that there is a small, elliptical, rigid, tectonic block in the zone. The long and short
principal radii of the body are a and b, respectively, and the aspect ratio is k = b/a ≤ 1. Let φ be
the orientation of the long axis of the ellipse. The tectonic block is rotated by the deformation of the
surrounding rocks.
It is known that a circlular block (k = 1) rotates with the angular velocity that is equal to half the
vorticity (φ̇ = ω/2) of the orogenic belt (§2.6). In contrast, the angular velocity φ̇ of an elliptical
block oscillates [67, 111] with increasing φ as
  
1 − k2
φ̇ = −w (cos 2φ + tan θ sin 2φ) − 1 , (2.46)
1 + k2
where w is a reference angular velocity. Substituting k = 1 into Eq. (2.46), it is obvious that w
equals the angular velocity of a spherical block, determined solely by the vorticity. The content of
the brackets is rewritten as
   
1 − k 2 cos θ cos 2φ + sin θ sin 2φ 1 − k 2 cos(θ − 2φ)
−1= − 1. (2.47)
1 + k2 cos θ 1 + k2 cos θ
Therefore, the amplitude of the oscillation depends not only on the vorticity but also on k, shape of
the block, and on θ.
Equation (2.46) shows orbits in the phase space, O-φφ̇, representing the temporal change of
the orientation of the block. The oscillation occurs for blocks with k < 1. If k is lower than the
threshold, the block rotate oppositely, φ̇ < 0. For example, a fat block with a larger k rotates
counterclockwise in a sinistral transtensional zone with θ = 140◦ (Fig. 2.11). However, a thin
block with k smaller than the threshold rotates clockwise for a range bounded by the points that is
indicated by open circles in Fig. 2.11. The points represents stable and unstable singular points, and
the the orientation of the thin block approaches the value indicated by the left stable point. The result
52 CHAPTER 2. INFINITESIMAL STRAINS AND THEIR ACCUMULATIONS

Figure 2.11: Phase space O-φφ̇ for the rotation of an elliptical rigid block in a sinistral transtensional
zone with θ = 140◦ .

is that thin blocks stop rotation and are oriented at a constant angle to the deforming zone, though
the surrounding zone continues deformation.
This example shows that paleomagnetic rotations are not enough to determine the sense of shear
for a deforming zone. We have to know the shape of the blocks, from which paleomagnetic data are
obtained, and the strain rate normal to the zone. The shape may be understood by detailed geological
mapping of the blocks and by paleomagnetic data with spatially high resolution. The strain rate may
be determined by an analysis of the geological structures.
Based on these observations, let us look at the paleomagnetic rotation in Northeast Japan asso-
ciated with the opening of the Japan Sea backarc basin (Fig. 2.12). While the Northeast Japan arc
drifted southward from the eastern margin of Eurasia in the Early Miocene time, extensional tecton-
ics prevailed in the arc [263]. Figure 2.12(a) shows the paleomagnetic directions in NE Japan. It
is important that the rotation in the forearc ceased 1–2 million years earlier than that in the backarc
[84]. Figure 2.12(b) shows the movement of the arc inferred from the synthesis of onshore and off-
shore geological and geophysical data around Japan. Northeast Japan was separated from Shikhote
Alin, and drifted southward. Therefore, the overall sense of shear between Shikhote Alin and the
Northeast Japan foreac was dextral. Therefore, the model of bookshelf faulting suggest that clock-
wise rotations were dominant in the backarc. However, the observed paleomagnetic rotations have
mostly the opposite sense (Fig. 2.12(a)). The paleomagnetic directions are also opposite to those
predicted from the vorticity under the backarc associated with the dextral movement of the arc. This
discrepancy is probably due to the association of strike-slip movement and extensional deformations
(Fig. 2.12(c)). When the arc migrated, it was broken into blocks by transtensional tectonics [263].
2.7. DEFORMATIN OF ROCK MASS BY THE ACTIVITY OF MINOR FAULTS 53

2.7 Deformatin of rock mass by the activity of minor faults


A rock mass encompassing many small-scale faults deforms by the activity of the faults (Fig. 1.17).
Displacements by the movement of a fault are discontinuous at the fault surace. However, continuum
mechanics applies to the description of the gross deformation of the mass, as the fault displacements
are much smaller than the dimension of the mass and the number of faults is large. The deformation
caused by a single fault is very small, so that infinitesimal deformation is considered for the mass by
integrating the activity of the faults. Specifically, we are able to estimate E and X for the mass from
the data of the minor faults6 .
To see this, we assumed a rock mass with volume V has minor faults whose displacements are
very small compared to the size of the mass, . Let the vector n be the unit normal to a fault surface,
and b be a unit vector parallel to the fault movement (Fig. 2.13). The initial and terminal of the
position vector x is placed on the opposite sides of the fault. The displacement of the terminal
due to the fault movement is indicated by the vector u. The portion of the rock mass depicted by
dashed lines in Fig. 2.13 is broken into blocks by the fault. However, since we assume a very small
displacement of the fault, d, the deformation of the portion can be approximated by a simple shear
with the engineering shear strain γ. Therefore, we obtain d = |u| = γ (x · n), and the displacement
gradient is
∇u = ∇ [γ (x · n) b ] . (2.48)
Each fault has its own parameters b, n and γ. If these parameters are spatially constant, they can be
moved out of the operand of the differentiation in Eq. (2.48). Therefore, we have
 
∂ui ∂   ∂xk
= γbi xk nk = γbi nk
∂xj ∂xj ∂xj
k k

Note that x1 , x2 and x3 are free variables and independent of each other. Hence, ∂xk /∂xj = δkj and
∂ui /∂xj = γbi nj . The product bi nj is a second-order tensor, so that we write

Ma ≡ bn. (2.49)

This is called an asymmetric moment tensor. This bears the information on the orientation of the
fault plane and slip direction. Finally, we have the deformation gradient of the rock mass due to this
fault is
F = I + γMa . (2.50)
Our next task is to estimate γ from the parameters that we can obtain from outcrops. If the cross-
sectional area of the rock mass is S, we have V = S (x · n). Hence, we have γ = d/ (x · n) = Sd/V .
Note that S is the area of the fault surface. Let the geometric seismic moment of S and d of this
fault be
Mg ≡ Sd. (2.51)
6 See [155] for further reading.
54 CHAPTER 2. INFINITESIMAL STRAINS AND THEIR ACCUMULATIONS

Figure 2.12: (a) Paleomagnetic declinations [73, 79, 84, 85, 162, 165, 166, 242, 271, 273] and major
faults that were probably activated when the Japan Sea opened in the Early Miocene [271]. Note that
the forearc (Pacific side) of Northeast Japan ceased paleomagnetic rotation by 16–17 Ma, earlier than
the termination of the rotation in the backarc (Japan Sea side) [84]. Major faults in Northeast Japan
(FF, Futaba Fault; HI, Hachirogata–Ichinoseki Tectonic Line; HM, Honjo–Matsushima Tectonic
Line, NT, Nishikawa–Takarazawa Tectonic Line; NM, Nihonkoku–Miomote Tectonic Line) after
[100]. OF, Oisawa Fault; TSZ, Tanakura Shear Zone; KTL, Kanto Tectonic Line. (b) Paleo-positions
of Northeast Japan and Hokkaido. (c) A simple model to account for the rotation and transtensional
deformation in the drifting NE Japan arc.

Using this moment, Eq. (2.50) is rewritten as


Mg Ma
F=I+ . (2.52)
V
The second term in the right-hand side is composed of variables that can be determined from the
fault parameters and the size of the rock mass.
The faults are numbered by their ages. That is, the number 1 is the oldest and the number N
is the youngest. The total deformation of the mass by the fault activity is indicated by the product,
F = F(N) · F(N−1) · · · F(1) , where F(i) stands for the deformation gradient tensor for the ith fault.
2.7. DEFORMATIN OF ROCK MASS BY THE ACTIVITY OF MINOR FAULTS 55

Figure 2.13: Deformation of a rock body by the movement of a fault. Gray, thick, dotted line shows
the initial shape of the body. The resultant shape is shown by light gray rectangles. The latter shape
is approximated by a parallelopiped indicated by dotted line. The unit vectors n and b is the direction
of fault plane and slip, respectively. Displacement d is assumed to be very small campared to the
dimension of the rock body.

Substituting Eq. (2.52) into this equation, we obtain


(N) (N)
! (1) (1)
!
Mg Ma Mg Ma
F=F (N)
···F (1)
= I+ ··· I +
V V
1  (k) (k)
≈I+ Mg Ma . (2.53)
V
k

Based on Eq. (2.53), we have the strain and rotation of the rock mass

1  1  (k) " (k)  (k)  T #


E= δF + δFT = Mg Ma + Ma , (2.54)
2 V
k
1  (k) " (k)  (k)  T #
X= Mg Ma − Ma (2.55)
V
k

Since the deformation for each fault is very small, second-order terms are neglected in the last
approximation. The sum of tensors does not depend on their order. As a result, we do not need
to know the order of activity of the faults to estimate the total deformation of the rock mass. This
tolerance is important for structural geology, because it is difficult to determine the order7 .
If the tectonic plates were literally rigid, tectonic activity would have occurred only at plate
boundaries. However, there are extensive deformation zones in the world, especially continental
collision zones are usually wide. Applying the above theory, the rate of deformation in a zone are
estimated from the seismic moment of great earthquakes that are released within the zone during a
certain period of time [89].
7 Cross-cutting relationship among faults tells the relative chronology of the faults. However, we can only date a part of

fault assemblage.
56 CHAPTER 2. INFINITESIMAL STRAINS AND THEIR ACCUMULATIONS

The theory is applied to geological faults [133]. However, the accurate estimation of deformation
by this method is probably difficult. The reason is that exposed faults are only a part of the faults
in the rock mass. Large faults are often hidden by young sediments, though the faults contribute
significantly to the deformation of the rock mass. In addition, the area of the fault surface, S, is
difficult to estimate. The fractal structure of the size and frequency of faults is sometimes employed
to fill this gap, but it may not have enough precision.

2.8 Exercises
2.1 Derive the stretching and spin tensors of simple shearing with a stationary velocity field.

2.2 Derive Eq. (2.28).

2.3 The rate of strain of a pure shear rift is related to the β factor by Eq. (2.39). Show that β
increases exponentially with time if the rate of strain is constant. Determine the width of the rift
zone as the function of time for the two cases: a constant strain rate and exponentially decaying rate.

2.4 Show that vorticity vanishes for pure shear.

2.5 Finite rotation is observable as paleomagnetic deflection. Suppose a vector a that is fixed in
a sedimentary body and indicates due north when it deposited. A rotation with a constant angular
velocity vector w and fixed rotation axis has carried the vector a to b during a time interval t. Let
$ be the orthogonal tensor corresponding to the rotation (v = R
R $ · a). Derive a relationship between
$ $
R and W, the axial vector of which is w. Note that both R of this problem and R in Section 1.5 are
orthogonal tensors, they represent different phenomena. The present one indicates spining of a rock
mass, but that in Section 1.5 is the rotation of passive strain markers.

2.6 Many fault scarps have been found on Mercury. They are called lobate scarps because of
their meandering trace on the surface, suggesting that they are the surface expression of low-angle,
thrust faults [232]. Some impact craters are horizontally shortened (Fig. 2.14), supporting the thrust
hypothesis. If so, the fault activity must have decreased the surface area of the planet, though other
types of faults may exit. The reduction of the area is attributed to ancient global cooling that resulted
in a decrease of the radius and surface of Mercury. If the fault parameters used in the method of
geometric seismic moment were known for the mercurial thrust faults, we would be able to estimate
the deformation of the planet associated with the fault activity. Instead, a simple and very rough
estimate is possible.
2.8. EXERCISES 57

Figure 2.14: Photograph taken by the Mariner 10 spacecraft showing fault scarps on Mercury. North
is at the top of the image (FDS 27380). Two impact craters near the upper left corner of the pho-
tograph has been cut and shorterned at the scarp. The southern crater is 65 km in diameter, and its
northwestern rim has been offset by about 10 km [232]. Courtesy of NASA.

Let Li , hi and di be the surfacial length, vertical displacement and dip of the ith fault. Show that
the decrease of radius ΔR can be roughly estimated by the equation

1  Li hi
N
ΔR
= , (2.56)
R 8πR2 i=1 tan di

where R is the initial radius, and N is the number of fault.

2.7 The strata shown in Fig. 1.18(a) have been deformed by minor normal faults. The strain ellipse
in (e) was approximately drawn from visual estimation of (c) and (d). Quantify the deformation
using the asymmetric moment tensor of the faults. Assume that all the fault planes are perpendicular
to the surface of the outcrop and have the same area.
Chapter 3

Stress, Balance Equations, and


Isostasy

We define stress to consider forces acting at depths, and derive the equations of mass,
linear momentum, angular momentum, and energy conservation. Very slow tectonic
movements allow us to neglect inertia force and simplify the equation of motion to a
balance equation. Isostasy is the most basic but important application of force balance
equations to tectonics.

3.1 Definition of stress


Suppose that a square pole is composed of five rectangler solids (Fig. 3.1(a)). If the pole bears a
load at the top and base, the parts may push each other across the bounding planes. Parts A and B
may push each other with that load. However, parts C and D may push each other with much less
force. Force across a plane in a solid depends on the direction of the plane.
In order to consider a force exerted on a rock mass, suppose there is a surface element δS on
the mass, and a unit vector N normal to the surface and pointing outward indicates the attitude of
the surface element (Fig. 3.1(b)). In addition, let the vector δF be the traction force acting on the
element. The force may increase with the area δS so that we define a vector quantity t(N ) as the
force per unit area exerted there. N is in parentheses because the force depends on the attitude of the
surface element. The quantity is called a stress vector. Mathematically, the stress vector is defined
by
δF
t(N ) = lim . (3.1)
δS→0 δS

The stress vector and its components are expressed in the units Nm−2 = pascals (Pa)1 .
1 The same units are used to count pressure. 1 bar = 105 Pa = 1000 hPa = 0.1 MPa. Engineering literature often uses the

units kgf /cm2 which are approximately equal to 0.98 MPa. The unit dyn cm−2 is used in older literature. It is converted by
the factor 1 Pa = 10 dyn cm−2 .

59
60 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Figure 3.1: (a) A square pole loaded at the top and base is composed of five rectangular solids, A
through E. (b) A rock mass is pulled from the adjacent mass with the force δF on an element of the
surface δS. A unit vector N indicates the attitude of the surface element. (c) Stress vector t(N ) and
Cartesian coordinates O-123 whose third axis is parallel to the unit normal N.

Consider a surface element δS with the unit, outward normal N, to which the third axis of
Cartesian coordinates O-123 is adjusted (Fig. 3.1(b) and (c)). Let e(i) be the ith base vector of the
coordinates, then we write
t(N ) = S31 e(1) + S32 e(2) + S33 e(3) , (3.2)
where S31 , S32 and S33 are the components of the stress vector t(N ), that is, t(N ) = (S31 , S32 , S33 )T .
The stress vector depends on the direction of the surface on which it acts. The first subscript ‘3’ in-
dicates accordingly that the surface is normal to the third coordinate axis and its outward normal
points positive direction of the axis. The second subscript distinguishes the components of the stress
vector. A stress component normal to the surface on which it acts is known as the normal stress on
the surface, whereas the component tangent to the surface is known as shear stress. Generalizing
Eq. (3.2), the components of stress vectors acting on a surface elements normal to the coordinate
axes, we have
 
Sij = tj e(i) . (3.3)
We define the stress tensor from the components as
⎛ ⎞  
S11 S12 S13 · · · components of te(1) 
S = ⎝ S21 S22 S23 ⎠ · · · components of te(2)  (3.4)
S31 S32 S33 · · · components of t e(3) .
The limit in Eq. (3.1) indicates that stress vector is defined at each point on an infinitesimally small
surface. The diagonal components are the normal stresses acting on a surface element normal to the
coordinate axes, while the other components represents shear stresses acting on the surface elements.

Sign conventions Figure 3.2(a) shows the positive direction of those components. Since we have
defined the unit vector N as being the outward normal to the surface on which the stress vector is
3.1. DEFINITION OF STRESS 61

Figure 3.2: Sign convention in continuum mechanics (a, b) and solid earth science (c, d). Arrows
indicate the direction of the positive stress components acting on a cube whose surfaces are parallel
to the coordinate planes. (a) A stress component is positive when it acts in the positive direction
of the coordinate axis, and on a surface of the cube whose outward normal points in the one of the
positive coordinate directions. (b) A stress component is positive when its direction is opposite to
that of (a).  and ⊗ indicate vectors pointing to this and opposite sides of the page, respectively.

defined, the opposite faces of a cube have the unit vectors in the opposite directions. Accordingly,
the normal stress on the faces has the opposite positive directions, that is, the normal stresses are
positive in sign if the cube is under a tensional stress (Fig. 3.2(b)). Compressive stresses have
negative normal stresses.
We have seen that the strain tensor E has positive diagonal components if the length parallel to
coordinate axes become larger. Therefore, it is natural that positive stress components correspond
to positive strain components. However, compressive stresses are common in the crust due to over-
burden pressure. It is the sign convention in solid-earth science to define the sign of stress as being
positive for compression, opposite to the sign convention of continuum mechanics (Table 3.1). Ac-
cordingly, we will use the symbol r for the stress tensor with positive components for compression
to describe tectonic models. Positive directions of the components of r are shown in Fig. 3.2(c) and
62 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Table 3.1: Symbols and signs used in this book.

Stress Unit Normal Vector Strain


Continuum Mechanics S (tension = +) N (outward) E (elongation = +)
Solid-Earth Science r (compression = +) n (inward) e (shortening = + )

Figure 3.3: Direction of the unit normal vectors n and N at the surface of a rock mass.

(d). The stress tensor r should be used with the inward normal n (Fig. 3.3), where

S = −r, N = −n. (3.5)

Stress components in cylindrical coordinates Cylindrical coordinates O-rθz are sometimes con-
venient (Fig. 3.4). Transformation between the cylindrical and rectangular Cartesian coordinates
O-xyz is given by the equations,

x = r cos θ, y = r sin θ, z=z


−1
r =x +y ,
2 2 2
θ = tan (y/x).

Stress components in the coordinate systems are transformed by the two sets of equations

σrr = σxx cos2 θ + σyy sin2 θ + 2σxy sin θ cos θ, (3.6)


σθθ = σxx sin θ + σyy cos θ − 2σxy sin θ cos θ,
2 2
(3.7)
σrθ = (σyy − σxx ) sin θ cos θ + σxy (cos θ − sin θ).
2 2
(3.8)

3.2 Cauchy’s stress formula


  % (i)
It is seen from Eq. (3.3) that ti e(j) = Sji = k Sjk ek . Therefore, we have
 
t e(i) = ST · e(i) . (3.9)
3.2. CAUCHY’S STRESS FORMULA 63

Figure 3.4: Stress components in the cylindrical coordinates O-rθ.

Note that the choice of frame of reference, the base vectors of which are e(i) , is arbitrary. Equation
(3.9) is generalized as follows: the stress vector acting on a surface element with the unit outward
normal N is
t(N ) = ST · N, (3.10)
where N may be oblique to the coordinate axes2 . This is called Cauchy’s stress formula. For the
sign convention of solid-earth science we have the the equivalent formula

t(n) = rT · n. (3.11)

If t(N ) is the traction on a body at its surface element with the normal N, the adjacent body has the
outward normal −N at the surface element. The action-reaction law requires that

t(N ) = −t(−N) = −t(n), (3.12)

which is shown in Fig. 3.3.

Normal and shear stresses Normal and shear stresses are the components of t(n) parallel and
normal to n, respectively. Thus,

Normal stress: σN = t(n) · n, (3.13)



Shear stress: σS = |t(n)|2 − σN2 . (3.14)

Since the vector n points inward at the surface of a rock mass (Fig. 3.3), σN is positive if the mass is
pushed inward at the surface. By contrast, the normal stress defined by the equation

SN = t(N ) · N (3.15)

is positive for tension. The shear stress for this sign convention is

SS = |t(N )|2 − SN2 .
2 Equation (3.10) is demonstrated via the force balance equation [134], which is introduced in §3.3. However, axiomatic

continuum mechanics defines a stress tensor as the linear transformation from t(N ) to t(N ), rather than the force per unit
area on the surface elements parallel to coordinate planes. The interested reader will find more information in [243].
64 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

The vector σN (n) = σN n has a magnitude obviously equal to the normal stress σN and is parallel
to n. If σN is positive, σN (n) points inward at the surface of the rock in question. The vector σN (n)
is obtained by the orthogonal projection of t(n) onto the line parallel to n, so that the projector
P (n) = nn (Eq. (C.37)) is used to describe the vector:
 
σN (n) = P (n) · t(n) = nn · r · n = n · (r · n) n. (3.16)

On the other hand, σS (n) is defined as the vector that has a magnitude equal to the shear traction
and indicates the direction of the shear traction at the surface element of the mass. The elementary
orthogonal projector P⊥ (n) = I − nn (Eq. (C.38)) is useful:
 
σS (n) = P⊥ (n) · t(n) = I − nn · t(n) = r · n − n · (r · n) n = t(n) − tN (n). (3.17)

Hydrostatic pressure From Eq. (3.11), let us derive a useful equation for tectonics: the tensorial
formulation of hydrostatic state of stress. Consider the stress tensor defined by

r = p I, (3.18)

where I and p are the unit tensor and a scalar quantity. Coordinate rotation by Q transforms the stress
as r = Q · r · QT = Q · (p I) · QT = p Q · QT = p I = r, indicating that the stress defined by Eq.
(3.18) is isotropic, that is, independent from orientations. The state of stress is anisotropic if there
is the dependence.
In this case, the force acting on a surface element dS is t(n)dS = (r · n)dS = (pdS)n, which
is perpendicular to the surface and pointing inward with a constant magnitude of pdS. The stress
vector is parallel to n for hydrostatic stress state: shear stress always vanishes. Therefore, Eq. (3.18)
is known as the hydrostatic state of stress, and p is hydrostatic pressure. The equation S = −p I
expresses the same stress for the sign convention of continuum mechanics. Hydrostatic pressure
constricts a body: it shrinks with a similar figure with the original one if the body is homogeneous
inside. If the body is subject to an anisotropic stress, the shape may change.

3.3 Force balance


Tectonic motion to create a large-scale geologic structure is very slow. Forces acting on a rock mass
at depth may be thought of being balanced at every moment. Accordingly, we shall derive balance
equations for forces, and derive vertical stress at depth.
There are two categories of forces: surface force per unit area of the surface t and body force per
unit mass X. The former was discussed in the last section. The latter acts on all the internal elements
of a body from its exterior. Examples are gravity and electromagnetic forces. As an example, gravity
is written as the vector
X = (0, 0, g)T , (3.19)
where g is gravitational accerelation and the third coordinate axis points vertically downward. Body
force per unit volume is ρX.
3.3. FORCE BALANCE 65

3.3.1 Equation of motion


Consider a rock mass defined by a volume V and a closed surface S. Body and surface forces, X
and t(N ) act on every portion in the body and on the surface, respectively. If the forces are balanced,
we have  
0 = t(N ) dS + ρX dV. (3.20)
S V
&
Since V
ρv dV is the total linear momentum of the rock mass, the material derivative

D
ρv dV
Dt V
represents the inertia force. Therefore, the equation of motion of the body is
  
D
ρv dV = t(N ) dS + ρX dV. (3.21)
Dt V S V

The surface integral in the left-hand side of Eq. (3.21) is transformed into a volume integral with
Gauss’s divergence theorem (Eq. (C.63)). In addition, combining Eq. (3.10), we obtain
  
D
ρv dV = ∇ · ST dS + ρX dV. (3.22)
Dt V V V

In this case the order of D/Dt and integral is exchangable3 so that we have

 
ρv̇ − ∇ · ST − ρX dV = 0.
V

Since the choice of the volume is arbitrary, the integrand must vanish to satisfy this equation. Con-
sequently, we obtain the differential form of

ρv̇ = ∇ · ST + ρX or ρv̇ = −∇ · rT + ρX. (3.23)

If the forces are balanced, we have

0 = ∇ · ST + ρX or 0 = −∇ · rT + ρX. (3.24)

The indicial notation of Eq. (3.24) is


 ∂Sji  ∂σji
0= + ρXi or 0=− + ρXi . (3.25)
j
∂xj j
∂xj
3 If V and V are the initial and instantaneous volumes that the rock mass occupies, respectively, at time t = 0 and t,
0
the exchangeability is demonstrated as follows. The initial volume V0 does not depend on time, so that the operator D/Dt
can be included in the integral over V0 . In addition, because of V = J V0 where J is the Jacobian, the initial and temporal
densities, ρ0 and ρ, are related to each other as ρ = ρ0 /J . The initial one does not depend on time, either. That is,
Dρ0 /Dt = D(ρJ )/Dt = 0. Therefore,
      
D D D(ρJ ) Dv Dv Dv
ρv dV = ρvJ dV0 = v· + ρJ dV0 = ρ J dV0 = ρ dV.
Dt V Dt V0 V0 Dt Dt V0 Dt V Dt
66 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Based on Eq. (3.24), we shall derive a simple but important quantity for tectonics. Suppose
that the ground surface is flat and there is no horizontal density variation. Rocks are subject to the
gravitational force shown in Eq. (3.19). Defining z-axis vertically downward from sea level, the
force balance in this direction is obtained from Eq. (3.24) as

∂σxz ∂σyz ∂σzz


+ + = ρ(z)g,
∂x ∂y ∂z

where density, ρ(z), is the function of depth, z. Since there is no horizontal variation at all, the first
and second terms in the above equation vanish. We therefore have

∂σzz
= ρ(z)g.
∂z

Integrating both sides, we obtain


z
σzz = ρ(ζ)g dζ, (3.26)
h

where h is altitude. This equation enables us to calculate the vertical stress σzz from the density
&z
profile ρ(z). If g is constant over the depth range in question, Eq. (3.26) becomes σzz = g h ρ(ζ) dζ.
The integral in the right-hand side of this equation is the mass of the vertical column with the unit
cross-section of the height from the surface to the depth z. Therefore, σzz is the stress at the base of
the column due to its own weight4 . The vertical stress by the weight is called overburden stress. If
density can be assumed to be constant, the overburden at the depth z is simply

pL = ρgz. (3.27)

We have found how the vertical stress σzz is determined. What are the other stress components?
Areas of compressional and extensional tectonics may have different stress states. Stress may not
only have spatial but also tempral variations, which are represented by components other than the
vertical one. However, it is convenient to define a reference stress state for later discussions. Al-
though choice of the reference is arbitrary, the isotropic stress

r = pL I (3.28)

is often used for the reference in the models of tectonics, where


z
pL = ρ(ζ)g dζ (3.29)
h

is the overburden at depth z. This is the overburden with the same dimension with pressure. Equation
(3.28) has the same form as Eq. (3.18), so that the stress indicated by Eq. (3.28) is known as
lithostatic stress, and pL is called lithostatic pressure.
3.4. SYMMETRY OF STRESS TENSOR 67

Figure 3.5: Force F at the point P with a position vector x yields a moment of that force M about
the origin O. The magnitude of M equals |x||F | sin θ, where θ is the angle between the vectors.

3.4 Symmetry of stress tensor


Both surface and body forces acting on a rock mass exert moments (torques) on the mass. If a force
F acts at a point P whose position vector is x (Fig. 3.5), the moment of the force about the origin is
&
M = x × F . Likewise, the moment about the origin by surface force is S x × t(N ) dS, where x is
the position vector of the point where the surface force t(N )dS acts on the rock mass. The moment
&
by body force is V x × ρX dV , where x is the position vector of a particle in the mass where the
body force acts. Therefore, in case that the total moment is in equilibrium to conserve the angular
momentum, we have the balance equation of momentum as
 
0 = x × t(N ) dS + x × ρX dV. (3.30)
S V

If they are imbalanced, the residual moment causes the acceleration in the angular velocity of the
body:   
D
(x × ρv) dV = (x × ρX) dV + (x × t) dS. (3.31)
Dt V V S
The choice of position of the origin is arbitrary. Hhowever, Eq. (3.31) holds wherever the origin is
defined5 .
4 In-situ stress measurements have shown that this picture gives a close approximation to observed σ
zz [3]. However,
deviations from predicted σzz from the density profile are rarely observed [30, 75].
5 Let us calculate the total moment about a position with the position vector o, from which position vector x is defined.

That is, x = x + o. By this additive decomposition of x, Eq. (3.31) becomes


   
Left-hand side = (x + o) × ρv̇ dV = x × ρv̇ dV + o × ρv̇ dV,
V V V
 
Right-hand side = (x + o) × ρX dV + (x + o) × t dS
V S
    
= x × ρX dV + x × t dS +

o × ρX dV + o × t dS.
V S V S

Combining the terms in and outside the square brackets in the above equations, we have the equations
  
x × ρv̇ dV = x × ρX dV + x × t dS,
V V S

which is identical to Eq. (3.31) provided that x is replaced by x and


    
o× ρv̇ dV = o × ρX dV + t dS . ()
V V S
68 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

It is important for discussions in the rest of this book that the stress tensor is symmetric:

S = ST , r = rT . (3.32)

These formulas hold for both static and dynamical problems unless there is neither “couple stress”
nor “body torque”. The former is the surface force acting on a body to rotate it about the axis
perpendicular to the object surface, and the latter is the distributed moment in the body6 . Equation
(3.32) is derived from the conservation of angular momentum7 .

Plane Stress It is a basic theorem in linear algebra that all the eigenvalues are of real for a real,
symmetric matrix. In addition, the eigenvectors are perpendicular to each other. If one and only one
of the eigenvalues is zero, a state of plane stress is said to exist. In that case, taking coordinate axes
parallel to the eigenvectors, the stress tensor is written as
⎛ ⎞
S1 0 0
S = ⎝ 0 S2 0⎠ ,
0 0 0
where S1 and S2 are non-zero eigenvalues of S. The physical interpretation of the stress state is
that traction always vanishes on a plane perpendicular to the eigenvector corresponding to the zero
eigenvalue. This is demonstrated by substituting N = (0, 0, 1)T into t(N ) = ST · N. It should be
noted that plane strain and plane stress do not go together for most cases (§7.1.3).

3.5 Conservation of energy


Let us derive an equation governing temperature changes from the conservation of energy. Multi-
plying both sides of Eq. (3.23) by v, we have
  
 
ρv · a dV = v · ∇ · ST dV + ρv · X dV, (3.33)
V V V

where V stands for the volume that the rock mass in question occupies. The left-hand side of this
equation is rewritten as
   
D v·v D ρ|v|2
ρv · a dV = ρv · v̇ dV = ρ dV = dV = K̇.
V V Dt V 2 Dt V 2
The symbol K stands for the total kinetic energy, so that K̇ is the rate of its increase. On the other
hand, the integrand v ·(∇·ST ) in the first term in the right-hand side of Eq. (3.33) has the components
∂Sji ∂(vi Sji ) ∂vi ∂(vi Sji )
vi = − Sji = − (Dij + ωij )Sji ,
∂xj ∂xj ∂xj ∂xj
This is identical to the cross-product of the constant vector o and Eq. (3.22), so that Eq. () is valid for any o. Therefore, Eq.
(3.31) holds wherever the origin is chosen.
6 See advance continuum mechanics books such as [132].
7 The derivation needs a tricky calculation. See, for example, [134].
3.5. CONSERVATION OF ENERGY 69

where Lij and ωij are stretching and spin tensors, respectively. Therefore, Eq. (3.33) is rewritten as
   
K̇ + D : S dV = ∇(v · S) dV − W : S dV + ρv · X dV. (3.34)
V V V V

Since W = (ωij ) is antisymmetric we have W : S = 0. Combining this and Eq. (3.34), we obtain
  
K̇ + D : S dV = ∇(v · S) dV + ρv · X dV. (3.35)
V V V

Converting the first term on the right-hand side of this equation into a surface integral by Gauss’s
divergence theorem, we obtain the energy conservation equation
  
K̇ + D : S dV = [v · t(N )] dS + ρv · X dV. (3.36)
V S V

Energy equals the product of force and distance, so that the rate of energy change equals the product
of force and velocity. The right-hand side of Eq. (3.36) indicates the energy that the rock mass
accepts from outside. The second term on the left-hand side is the energy dissipation due to the
deformation of the mass against the stress S. The dissipated energy is transformed to thermal or
internal energy. Let e be the internal energy per unit mass, then the internal energy per unit volume
is equal to ρe, and
 
ρė dV = D : S dV. (3.37)
V V

Therefore, the energy conservation equation is rewritten as


    
D v2
ρ + e dV = v · t(N ) dS + ρv · X dV. (3.38)
Dt V 2 S V

The left-hand side of this equation represents the energy increase of the rock mass, and the right-hand
side indicates the work done by its outside.
Equation (3.38) was derived from the equation of motion multiplied by velocity. That is, only the
balance of kinetic energy was taken into account. However, vertical motion of a rock mass causes
cooling or heating of the mass. Temperature is an important factor in controlling the mechanical
properties of rocks. Accordingly, heat transfer is important in understanding the mechanical aspect
of tectonics.
Heat flux is defined by the amount of heat that passes through a unit area in a unit of time. The
heat flux is a vector quantity, so let us use the symbol q for the quantity. The heat energy passing
through the surface dS of a rock mass in a unit time equals N ·q. Substituting this into the right-hand
side of Eq. (3.38), we have
     
D v2
ρ + e dV = v · t(N ) dS + ρv · X dV − N · q dS. (3.39)
Dt V 2 S V S
70 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Using Gauss’s divergence theorem again, the last term in this equation is converted to a volume
&
integral, and we combine all terms into the form V (· · · ) dV = 0. Since the volume V is arbitrary,
this equation is satisfied only if
 
D v2  
+ e = ∇ · S · v − q + ρX · v. (3.40)
Dt 2

This is called the energy equation. Combining Eqs. (3.40) and (3.33), we have

ė = −∇ · q + ∇ · (S · v) − v · (∇ · S).

Components of the first and second terms in the right-hand side are
  ∂(Sij vj )  
∂Sij ∂vi
− vi = Sij = S : (∇v).
i,j
∂x i ∂x j i,j
∂x j

Therefore, we finally obtain the equation

ρė = S : (∇v) − ∇ · q. (3.41)

3.6 Self gravitational, spherically symmetric body


The direct objects of geological observations occupy shallow levels in the Earth. However, those
objects sometimes provide constraints on ancient global changes including planetary differentiation.
The differentiation occurred at the beginning of planetary evolution, so that later geological processes
have erased all traces of the event in the Earth. However, satellites and terrestrial planets smaller than
the Earth have geological structures that evidence the early history of the planetary bodies.
For example, wide rift zones on the Jovian satellite Ganymede are thought to be evidence of
planetary differentiation (Fig. 3.6). An icy mantle ocupies one third of the radius, and 60% the mass
of the satellite. The surface of Ganymede has high and low albedo regions which appear as light
gray and dark terrains, respectively, in Fig. 3.6. Number density of impact craters indicates that the
dark terrains are older than the lighter ones. White spots in this photograph are ejecta blankets from
young impact craters. The light, young terrains with stripes are called sulci. These are thought to be
the surface manifestation of normal faults, and the sulci are wide rift zones [171]. The formation of
those normal normal faults is modeled in Section 12.4.
Ganymede was initially wrapped up in a dark icy layer, but global expansion broke the layer into
dark terrains. Young, light gray ice spread between the terrains. The global expansion was due to
core formation in the satellite; initially deep-seated ice was raised to shallow levels in the moon by
sinking rocks and metalic materials. The phase transformation of the ice by decompression caused
the global expansion.
We have seen in the previous section (§3.3.1) that the vertical stress σzz is determined by gravity,
where the gravitational acceleration g is assumed to be constant. However, g may vary with depth
3.6. SELF GRAVITATIONAL, SPHERICALLY SYMMETRIC BODY 71

Figure 3.6: Surface of Ganymede, the largest Jovian satellite with a radius of 2640 km. The circular
region labeled “E” is the palimpsest Epigeous, an old and degraded large impact basin 350 km in
diameter. Voyager image, 0370J2-001. Courtesy of NASA.

in the real Earth. Accordingly, in this section we study the depth-dependence of g and lithostatic
pressure.

Firstly, let us assume a density distribution with spherical symmetry, and M (r) be the mass
72 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

within the distance r from the center. The gravitational acceleration at the radius is

GM (r)
g(r) = , (3.42)
r2

where G = 6.67 × 10−11 Nm2 kg−2 is the gravitational constant. If density at the radius is ρ(r), we
have  r
M (r) = 4πr  2 ρ(r ) dr . (3.43)
0

Combining Eqs. (3.42) and (3.43), we obtain the gravitational acceleration as a function of r as

G r
g(r) = 2 4πr  2 ρ(r ) dr . (3.44)
r 0

Scondly, pressure p at r is obtained from the force balance equation (Eq. (3.25)). As the force
in this case has merely originated from the self-gravity of the spherically symmetric body, there is
no horizontal variation in the stress field. Equation (3.25) is rewritten for the spherical coordinates
O-rθφ as
∂σrr 1 ∂σθr 1 ∂σφr
0=− − − − ρg, (3.45)
∂r r ∂θ r sin θ ∂φ
where both ρ and g are the functions r and the latter is given by Eq. (3.44). The gravity term ρg
in Eq. (3.45) is negative in sign, because gravity points downward, i.e., in the direction of −r. If
the spherical body is composed of viscous fluid at rest, the state of stress in the body is hydrostatic.
In this case, shear stresses σθr and σφr vanish, and in Eq. (3.45) σrr should be replaced by the
hydrostatic pressure p. Consequently, we have

dp
= −ρg. (3.46)
dr
As ρ and g are always positive in sign, Eq. (3.46) indicates that p should decrease with r.
Let us calculate g and p as the functions of r for the two cases in Fig. 3.7. Density in the layers
is assumed to be constant. The former is a simple model for the Moon. Our Moon had its own
magnetic field a few billion years ago, suggesting that there was a melted metalic core at the center.
However, the core is so small that it is difficult to estimate its dimension. Accordingly, we assume
the Moon to be a rocky spherical body with a constant density ρ = 3.3 × 103 kg m−3 to roughly
estimate the gravitational acceleration and pressure at depths8 . In the case of a constant density ρ,
the integral in Eq. (3.43) results in M (r) = 4πρr 3 /3, and we obtain


g(r) = Gρr. (3.47)
3
8 The Moon has a less dense crust than the mantle. However, the mean thickness of the crust is less than 100 km, two

orders of magnitude smaller than the radius of the Moon, which is about 1740 km. Accordingly, the crust is neglected in this
estimation.
3.6. SELF GRAVITATIONAL, SPHERICALLY SYMMETRIC BODY 73

Figure 3.7: Pressure p and gravitational acceleration g versus distance r from the center of a spher-
ically symmetric body. (a) Single-layer model Moon with a constant density with the parameters
ρ = 3.3×103 kg m−3 and R = 1740 km. (b) Two-layer model Earth with the parameters ρm = 4×103
kg m−3 , ρc = 10 × 103 kg m−3 , Rc = 3490 km, and R = 6380 km.

It follows that gravitational acceleration for a constant density, spherically symmetric body is pro-
portional to the distance from the center (Fig. 3.7(a)). Substituting Eq. (3.47) into (3.46), we
have
dp 4πGρ2
=− r.
dr 3
Using the boundary condition p = 0 at r = R, where R stands for the radius of the Moon, we obtain
the pressure distribution
2π 2
p(r) = ρ G(R2 − r2 ). (3.48)
3
The density distribution in the Earth is roughly simulated by a two-layer model, as about half
the radius of the Earth is occupied by a high density, metalic core at the center and the crust is
negligibly thin compared to the core and mantle. We do not take into account the tiny density
difference between the inner and outer cores. Let R and RC be the radius of the Earth and the core,
respectively, then the density is
'
ρC (r ≤ RC )
ρ= (3.49)
ρm (RC < r ≤ R).

Gravitational acceleration in the core is the same as Eq. (3.47) and we have


g(r) = GρC r (r ≤ RC ),
3
74 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

where ρC is the density of the core. If that of the mantle is ρm , gravitational acceleration in the mantle
is obtained by using Eqs. (3.44) and (3.49) as
!
4πG RC3
g(r) = ρm r + (ρC − ρm ) (RC < r ≤ R).
3 r2

Figure 3.7(b) shows g(r) for a double-layered model Earth. It is seen in this figure that g is roughly
constant in the upper mantle. The models of tectonics in the following sections of this book deal
with the crust and the upper mantle, so that we will always assume that gravitational acceleration is
constant.
Pressure distribution in the model Earth is given by integrating Eq. (3.46) as
 
4π 1 1
p(r) = Gρm (ρC − ρm ) RC3 −
3 r R
2π   2π  
+ Gρ2m R2 − RC2 + Gρ2C RC2 − r2 (r < RC ) ,
3   3
4π 1 1 2π  
p(r) = Gρm Rc (ρC − ρm )
3
− + Gρ2m R2 − r2 (RC < r < R) .
3 Rc R 3

The gravitational acceleration and pressure in the model Earth are calculated with the values, ρm =
4.0 × 103 kg m−3 , ρcore = 10 × 103 kg m−3 , Rc = 3490 km, R = 6380 km, and G = 6.7 × 10−11
Nm2 kg−2 , and are shown in Fig. 3.7. It is seen that gravitational acceleration is more or less constant
in the upper mantle. Therefore, the acceleration is always assumed to be constant in the following
sections of this book.

3.7 Isostasy
Buoyancy is an important force to drive tectonics. The crust has smaller density than the underlying
mantle so that the crust is subject to buoyancy forces from the mantle and is floating on it. The
tectonic thinning and thickening of the crust cause the subsidence and uplift of the surface. We will
derive Archimedes’ principle from the force balance equation.
Suppose a solid body is immersed in a fluid with a density of ρ (Fig. 3.8). The body and fluid are
assumed to stand still so that all forces acting on the body are balanced. The force balance equation
(Eq. (3.20)) in this case is
 
0 = t(n) dS + X dV, (3.50)
S V

where V and S are the volume and surface of the body. The first and second terms on the right-hand
side of this equation represent surface force from the fluid to the body and gravitational force (Eq.
(3.19)), respectively. The buoyancy force exerting on the body is the former. The traction at the
surface is pressure, p. That is, t(n) = pn. Using Gauss’s divergence theorem, the surface integral is
3.7. ISOSTASY 75

Figure 3.8: Schematic illustration to explain Archimedes’ principle.

converted to a volume integral. Therefore, the buoyancy term is rewritten as


 
pn dS = − ∇p dV. (3.51)
S V

A minus sign is placed on the right-hand side of this equation because the unit normal n is defined
inward, unlike the case of Eq. (C.63). It should be noted that the integration on the right-hand side of
Eq. (3.51) is taken inside the body, although p is not the pressure in the body but fluid pressure at the
surface. Since the fluid is at rest, the pressure depends only on the depth z. The pressure gradient,
∇p, and the gravity term in Eq. (3.50) have no horizontal component. Therefore, an equation of
scalar quantities is enough to consider the vertical force balance of this system. Let Fb be buoyancy,
and ρ be constant, then Eq. (3.50) is converted to the equation

Fb = −ρ dV = −ρV,
V

where the negative signs came from the definition of the force Fb that is positive upward. This
equation stands for Archimedes’ principle: an object completely immersed in a fluid experiences an
upward buoyant force equal in magnitude to the weight of the fluid displaced by the object.
The concept of isostasy is derived from that principle. Suppose the structure shown in Fig. 3.9,
where a continental crust with a thickness of tc is floating on a mantle. According to Archimedes’
principle, the buoyancy force exerted at the base of the crust is the weight of the mantle displaced
by the part of the crust that is immersed in the mantle. The weight is ρm ghc S, where ρm the mantle
density and S the area of the base. Other symbols are shown in Fig. 3.9. The gravitational force
that the crustal block experiences is ρc gtc S, which should be balanced with the buoyancy force.
Therefore, we have
ρc gtc = ρm ghc . (3.52)
76 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Figure 3.9: A simple model for a continental crust floating on a mantle.

This equation says that the lithostatic pressure at the level of the base is spatially constant. If density
variations in the upper mantle are not negligible, this level should be taken to be deeper. The litho-
spheric plate that is composed of a crust and upper mantle is often envisaged as floating on a fluid
asthenosphere. Rocks behave as fluid in a geological timescale, if they have high temperature. On
the other hand, if the lower mantle is heated to become fluid, upper crustal blocks with smaller den-
sities float on it. The level is placed in the middle of the crust. The level under which rocks behave
as fluid and they cannot support a lithostatic pressure difference is called the depth of compensation.
A simple model of isostasy asserts that overburden is constant at a depth of compensation. The
following equation represents this concept:
 dc
ρg dz = constant, (3.53)
surface

where dc is the depth of compensation, and the z-axis is defined downward with the sea level at z = 0.
If an offshore region is considered, the lower bound of this integral is taken to be at sea level to take
into account the load of the ocean water. The integral is spatially or temporarily constant. When it
is thought to be temporarily constant, we can calculate subsidence and uplift accompanied by the
temporal variation of density distribution at depths, i.e., vertical movements that are recognized from
stratigraphy give constraints to the variation.
Equation (3.53) represents a one-dimensional model of isostasy, where density variations only
along the z direction are taken into account. This idea is sometimes referred to as local isostasy
compared to regional isostasy that is introduced in Section 8.1. In the latter model, horizontal density
variations are explicitly included in its equation.
In the rest of this chapter we will use a simple isostatic model to calculate topography and vertical
tectonic movements.

3.8 Balance between ocean and continent


There are two tectonic classes of regions, continents and oceans. This division is reflected in the
frequency distribution of the height or water depth of the surface of the solid Earth (Fig. 3.10).
3.8. BALANCE BETWEEN OCEAN AND CONTINENT 77

Figure 3.10: Cumulative frequency distribution of the altitude of the surface of the solid Earth.

There are apparently two modes at about 1 km above and 3–4 km below sea level. They correspond
to the average altitude of the continents at 0.84 km and the average depth of oceans at 3.4 km.
Let us see that the two levels are in isostatic equilibrium [78]. Assuming g = 10 ms−1 and
the depths and densities shown in Fig. 3.11(a) for representative continent and ocean, we have the
overburden at the Moho depth under the continent

pL = ρc gto = 0.98 GPa. (3.54)

The overburden under the ocean basin at the same level is

pL = ρw gb + ρo gto + ρm g (tc − h − b − to ) ≈ 1.006 GPa, (3.55)

which is about the same as that under the continent (Eq. (3.54)). Therefore, continents and oceans
are roughly in isostatic equilibrium.
Seawater pushes the oceanic crust downward by its weight, and pushes the continents upward.
What if the Earth loses oceans by extreme global warming? The average continental altitude from
the average oceanic surface, x in Fig. 3.11(b), would have been decreased. Neglecting other factors
including thermal expansion of rocks, the overburden at the level of continental Moho is

pL = ρc gto + ρm g (tc − to − x) .

Equating this pressure and that in Eq. (3.54), we obtain x = (tc − to ) (ρm − ρc )/ρm ≈ 3.3 km.
The frequency distribution of global topographic height, hypsometry, is used to investigate the
tectonics of terrestrial planets and moons. If impact cratering was the most significant process for
shaping the surface of a planetary body, the resultant frequency distribution would be something
like a normal distribution. In spite of the existence of numberless impact craters, the Moon has two
modes in its distribution. They correspond to highlands and maria, suggesting that this distinction
was formed through some global processes.
78 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Figure 3.11: (a) Representative large-scale structure of the shallow part of the Earth. (b) A structure
without seawater.

The bimodal distribution for the case of the Earth reflects the density contrast of the continen-
tal crust to the oceanic crust and to the mantle. The density distribution with spherical symmetry
is obviously the most stable structure in consideration of potential energy. However, the Earth has
the dichotomy of continents and ocean basins. Plate tectonics is an important action to keep the
dichotomy. Surface processes, including erosion and sedimentation, are also important for redis-
tributing rocks and sediments. The strength of the lithosphere also affects the distribution. The
continental lithosphere has mechanical strength so that plate convergence causes continental thick-
ening with a limiting average altitude. As the average altitude exceeds it, increased gravitational
force flattens the crust to lower the surface (§3.11).

3.9 Sediment load


A thick sedimentary pile pushes its basement downward by its weight. Let us consider the effect of
sediment loading, using the thick sedimentary sequence resting on the Atlantic passive continental
margin (Fig. 3.12) as an example. Sedimentary layers that lie beneath the continental shelf ac-
cumulated in sublittoral paleoenvironments9 . The total thickness of the sediments reaches several
kilometers. The paleo water depth, which is called paleobathymetry, and the thickness indicate that
the basement has subsided by the same amount as the thickness. Some old literature explained the
subsidence by the increasing sediment load. Sediments have greater densities than water, and sedi-
mentation replaces water by sediments. Therefore, sedimentation increases loading to the basement.
Namely, sedimentation created the sedimentary basin.
Now we are ready to deny this idea for the case of passive margin subsidence. Consider, for
9 Sublittoral zone lies immediately below the intertidal zone and extending to a depth of about 200 m or to the edge of the

contiental shelf [2].


3.9. SEDIMENT LOAD 79

Figure 3.12: Crustal structure of the southern Baltimore Canyon Trough, Atlantic margin of North
America. LJ, Lower Jurassic; MJ, Middle Jurassic; UJ, Upper Jurassic; LK, Lower Cretaceous; UK,
Upper Cretaceous; T, Tertiary. Simplified from [101].

Figure 3.13: Sedimentary basin subsidence by sediment load.

simplicity, a constant density for sediments ρs . If an ocean basin with an initial depth of b is filled
up with sediments (Fig. 3.13), the final thickness of the sedimentary pile x is given by the isostasy
equation
ρw gb + ρc gtc + ρm g (x − b) = ρs gx + ρc gtc .

The left- and right-hand sides of this equation correspond to the left and right columns, respectively,
in Fig. 3.13. Therefore, we obtain
 
ρm − ρw
x= b. (3.56)
ρm − ρ s
The density of sediment depends on the lithology. Using a representative value ρs = 2.3 × 103
kg m−3 , we have x = 2.3b. If excessive sediments are supplied from the hinterland, the sediments
would be exported to the lower reaches. For the case of the thick shallow marine sequence on the
passive continental margin, b ≤ 0.2 km results in x ≈ 0.5 km. Therefore, the basement subsided
actively, not passively, because of the sediment load (§3.14).
It is necessary to grasp subsidence history, and not only the initial water depth and final sediment
80 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

thickness, to understand the mechanism of subsidence. Quantitative stratigraphy provides a way to


reconstuct the history.

3.10 Quantitative stratigraphy


The subsidence history of a sedimentary basin is an observable function of time. Quantitative stratig-
raphy or geohistory analysis is the technique used to reconstruct the history from formations, where
isostatic balance plays an important role. The result of the analysis designates the vertical component
of the tectonic movements.
For this purpose, the necessary data are the present depth, lithology, porosity, sedimentation age,
the paleobathymetry of formations, and the eustatic sea level curve for the period covering the ages
of the formations. The paleobathymetry is estimated from sedimentary structures characteristic for a
certain depth such as hammocky cross-stratification, which indicates the depths were so shallow that
the sedimentation was affected by surface waves. Some kinds of molluscs and benthic foraminifers
live on the ocean bottom at certain water depths so that benthic fossils are another indicator of
paleobathymetry. Figure 3.14 shows the paleo topography inferred from those kinds of data around
Japan. The Japanese island arcs experienced vertical movements on the order of kilometers in the
Miocene [263, 272].
The basic idea of geohistory analysis is that the sum of the formation thickness and paleo-
bathymetry gives the depth to the basement from the present sea level, and applies several cor-
rections. There are three key factors for the accumulation of thick sedimentary pile. Firstly, tectonic
subsidence is necessary to form a container of sediments. Secondly, sediment supply is impor-
tant. If the supply is scarce, a sediment-starved basin is the result. Thirdly, eustasy affects not
only paleobathymetry but also sedimentation. A eustatic rise is equivalent to tectonic subsidence
for sedimentation. We can abstract vertical tectonic movements from a sedimentary basin by taking
into account these and additional factors. The mathematical procedure employed for this purpose is
comparable to the stripping of formations from the top to lowest stratigraphic horizons (Fig. 3.15),
so that it is known as the backstripping technique.
The density of rocks is important for isostasy. Sediments lose their porosity and get a larger
density while they are buried. Pores between sedimentary particles are usually filled with forma-
tion fluids that are mostly water. For example, mudstone has a porosity of some 50% shortly after
deposition, However, increasing compactness of the particles expels the fluids during burial, i.e.,
compaction occurs. The porosity φ is the volume fraction of a pore. The porosity of mudstone is
less than 10% at depths of a few kilometers. The total volume of sedimentary particles in a unit
volume is 1 − φ. It is known that porosity of sedimentary rocks decreases roughly exponentially
with depth,
φ = φ0 e−Cζ , (3.57)

where ζ is depth from the ocean bottom. The parameters φ0 and C depend on lithology. Accordingly,
it is assumed that the parameters are determined for rock types at depths, e.g., in boreholes.
3.10. QUANTITATIVE STRATIGRAPHY 81

Figure 3.14: Paleo topography around Japanese islands from 23 to 15 Ma inferred from sedimentary
environments. Benthic fossils and depositional facies are the keys for estimating the environment.
The Japan arc drifted to form the Japan Sea backarc basin in the Early Miocene. The paleo posi-
tion of the arc before 15 Ma, when the opening was completed, is controversial so that the paleo
environments are indicated on the present map of this region.

If a formation has a horizonally infinite extension, the porosity reduction results in the thinning of
the formation. Consider a formation with an initial thickness of δd0 decreasing to δdN by increasing
burial depth from d0 to dN . A column of sedimentary rocks with a unit sectional area has the total
& d +δd
volume of sedimentary particles d00 0 (1 − φ) dζ between the depths d0 and d0 + δd0 . Pore fluid
is expelled from this part of the column by burial, but the particles are not. The conservation of
82 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Figure 3.15: Schematic illustration to explain the backstripping technique. The left column shows
the present state. The central and right columns indicate the situation just after the ith and (i + 1)
layers deposited, respectively.

particles is indicated by the equation


 d0 +δd0  dN +δdN
(1 − φ) dζ = (1 − φ) dζ.
d0 dN

Substituting Eq. (3.57), we have


φ0 Cd0  −Cδd0  φ0 CdN  −CδdN 
δd0 + e e − 1 = δdN + e e −1 . (3.58)
C C
The parameters C and φ0 are known for specific rock types. Therefore, if the thickness of a formation
δdN when the depth of the formation was dN is given, we can calculate the thickness δd0 at a given
depth of d0 through Eq. (3.58). That is, this equation corrects the effect of compaction on the
thickness of a formation. In addition, it is easy for us to grasp the present depth and thickness of a
formation. The ancient depth of the formation is obtained as the total thicknesses of the overlying
formations. Therefore, the right-hand side of this equation is known for the youngest formation that
occupies the top of the sedimentary pile. Accordingly, Eq. (3.58) works as a recursive function
to calculate ancient thicknesses which are determined successively from the top to the base of the
layers. This procedure is called decompaction, which increases layer thickness.
The next task is the correction of sediment loading. Let us assign a number for each layer from
the top (Fig. 3.15). The mass of the ith layer includes those of pore water φi ρw and sedimentary
grains (1 − φi )ρg , where φi is the porosity of the layer and ρg is the average density of the grains. If
3.11. TECTONIC FORCE CAUSED BY HORIZONTAL DENSITY VARIATIONS 83

δdi is the thickness of this layer just after it was deposited, we obtain δdi using the recursive equation
Eq. (3.58). The mass of the layer per unit basal area is
 
mi = φi ρw + (1 − φi ) ρg δdi . (3.59)

If the lowermost layer has the ordinal number n, the sedimentary column with a unit sectional area
has the mass of
Mi = mi + mi+1 + · · · + mn−1 + mn (3.60)
when the ith layer was deposited. Consider that the Moho is displaced downward by the weight of
the ocean water and the ith layer. If the displacement is δhi , we have the equation

ρw (bi − ei ) + Mi = ρw (bi+1 − ei+1 ) + Mi+1 + ρm δhc (3.61)

from isostasy10 , where bi is the water depth and ei is the level of the ocean surface relative to the
present one when the ith layer was deposited. That is, the sequence {en , en−1 , . . . , e0 } represents
the eustatic sea level changes through time. They are estimated from ancient marine transgression
and regression observed around stable continents11 . Combining Eqs. (3.58)–(3.60), we obtain Mi .
Therefore, δhc is calculated through Eq. (3.61).
We are able to estimate the ancient depth of burial for every formation from a sedimentary record
using this backstripping technique, and to quantify the history of tectonic subsidence. Quantitative
stratigraphy has a precision of ∼100 m for a sedimentary sequence that was deposited near sea level.
Not only subsidence, but also uplift can be evaluated with this method. A sedimentary pile
records the uplift as a decrease of paleo water depth or regression. We are able to quantify these
phenomena, provided that sedimentation lasted during the uplift. If vertical movement occurred
above sea level, fission track thermochronology is useful to evaluate uplifts of the order of kilo
meters. Recently, a geochemical signature of continental carbonates was used to estimate paleo
altitudes [66].
Figure 3.16 shows the subsidence history of the continental shelf off the Grand Banks of Canada
determined by the backstripping technique. The vertical movements of the basement is illustrated
by a subsidence curve. The subsidence of passive continental margins is modeled as the surface
manifestation of post-rift lithospheric cooling (§3.14). Kominz et al. [103] exhibited a eustatic sea
level curve from the subsidence curves of passive margins combined with the cooling model.

3.11 Tectonic force caused by horizontal density variations


Assuming the lithostatic state of stress, horizontal forces in the lithosphere are discussed in this
section. Under this condition, lithostatic pressure is proportional to depth z, and the constant of
10 We asssume that the crustal thickness except for the sedimentary pile does not change so that the thickness does not

appear in Eq. (3.61). Local isostasy is assumed here, but a technique called flexural backstripping has been developed to
account for the strength of the lithosphere [74].
11 Epeirogeny (§9.5) is an important but serious problem for this purpose.
84 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Figure 3.16: Subsidence history of the continental shelf off the Grand Banks, Canada [1]. The
present sea level is the origin of the vertical axis. Ages on the right side of this graph depict the
depositional age of selected stratigraphic horizons.

Figure 3.17: Diagram showing the lithostatic pressure versus depth for high and low density rocks.

proportionality depends linearly on density (Fig. 3.17). We have an interesting consequence from
this simple model if there is a horizontal density variation.
Continents and oceans are isostatically compensated. If a depth of compensation is placed at
the continental Moho, the depth dependence of the lithostatic pressures under a continent and an
ocean basin are illustrated in Fig. 3.18(a), which is drawn with the same structure as shown in Fig.
3.11(a). Namely, the pressure is greater under the continent than under the ocean around the sea level
because of the topographic bulge of the continent and of the density difference of the continental
crust and seawater. However, the lithostatic pressure under the ocean catches up with that under the
continent at the compensation depth. Pressure difference generally drives deformation. Accordingly,
the difference between the lithostatic pressures produces tectonic force, which horizontally extends
continents and constricts ocean basins (Fig. 3.18(c)).
3.11. TECTONIC FORCE CAUSED BY HORIZONTAL DENSITY VARIATIONS 85

Figure 3.18: (a) Lithostatic pressure under a continent and an ocean basin that are isostatically
compensated. It is assumed that there is no density variation in the crust and in the mantle. (b)
Vertical profile of density contrast between the regions. (c) Tectonic flow driven by the pressure
difference.

A continent pushes an ocean basin with a force that is designated by the gap between the line
graphs of the lithostatic pressures under the regions. The gap depends on the profile of the density
difference δρ, which is schematically shown in Fig. 3.18(b). The density profile exhibits deviations
with the opposite signs at two levels. The dipole moment is the magnitude of positive and negative
anomalies multiplied by the distance between the anomalies. Accordingly, the gray area in Fig.
3.18(a) increases with the increasing dipole moment of the density profile. As the continental crust
is thickened, both the average altitude and the Moho depth increase. They expand the gap, and
further strengthen the horizontal force (Fig. 3.18(a)). Consequently, a continent pushes oceans with
a force of 1.6 × 1012 Nm−1 = 1.6 TNm−1 per unit length of the continent-ocean boundary, if the
structure shown in Fig. 3.11(a) is assumed. This value is obtained by the downward integration of
the gap between the pressures. The force is as great as other tectonic forces. That is, large-scale
variation of topography can drive tectonic deformation (§12.2).
The same argument also applies to the horizontal force accompanied by a large-scale variation
in altitude. That is, highlands tend to extend horizontally and pushes the neighboring lowland.
The extensional force by this effect builds up with an uplift of the highland. The result is that the
altitude of an extensive highland cannot exceed a certain limit. Gravitational collapse occurs when
86 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Figure 3.19: Viscous fluid has a level surface at rest. However, a moving belt installed on the bottom
of a container produces a slope at the surface.

the highland is uplifted above the limiting altitude.


The Tibetan Plateau has been uplifted to an altitude of ∼5 km by N–S cotinental convergence, and
the crustal thickness reaches ∼70 km. E–W trending thrust faults are dominant along the Himalayas,
but there are N–S or NNE–SSW trending normal faults and strike-slip faults oblique to the normal
faults in the plateau [12, 240]. The crustal thickness is attenuated rather than increased in the plateau,
and the normal and strike-slip faulting transfers crustal blocks eastward, suggesting that the plateau
has reached the limit12 .
The limiting altitude depends on the strength of the continental lithosphere, which has temporal
and spatial variations13 . To see this, consider viscous fluid like honey in a container that has a
moving belt at its bottom. This thought experiment models a convergent orogen, where the fluid
simulates the continental crust. The belt drags the fluid to form a slope at the surface of the fluid
(Fig. 3.19). The slope is dynamically maintained by the competing factors, the shear traction and
gravitational spreading caused by the slope. The driving force of the latter results from the horizontal
density difference of the fluid and air across the slope. The inclination of the surface is enhanced
by the shear stress caused by the belt, and the shear stress increases with increasing viscosity. The
ratio of the gravitational and viscous forces is called the Argand number [56] and indicates the
tendency of gravitational spreading of large-scale positive undulations. Several factors including
lithology, thermal regime and pore pressure, control the effective viscosity of rocks (§10.8). Since
these factors may have temporal and spatial variations, the limiting altitude has these variations, also.
If the lithosphere is soft or softened, topographic bulges on the lithosphere collapse.

3.12 Thermal isostasy


In this section we consider ocean-floor subsidence (Fig. 3.20) by combining the isostasy and thermal
evolution of the oceanic lithosphere.
The solid Earth discharges energy mostly through the cooling oceanic plates. Thermal conduc-
tion is the primary mechanism of heat transfer in the lithosphere14 . As rocks have low thermal
12 The Himalayas have higher altitudes than the plateau. The peaks are supported not only by the buoyancy force associated

with isostatic compensation but also by the flexural strength of the lithosphere.
13 See Chapter 12.
14 The Galilean Satelite, Io, is a remarkable extra terrestrial body. The moon is as the size of the Moon, but discharges with
3.12. THERMAL ISOSTASY 87

Figure 3.20: Depth–age relationship of ocean basins [228]. The vertical axis shows the moving
average with a window of 5 m.y. of the depth that is corrected for the load and thickness of sedimen-
tary cover. GDH1, PSM and HS are synthesized depth-age curves based on different models. PMS
assumes the cooling oceanic plate with the constant thickness and constant basal temperature [174].
HS is the curve calculated by the cooling half-space with the thermal properties derived from PMS.
GDH1 is the compromised model that assumes a cooling half-space for the lithosphere younger than
20 m.y. and the plate model for the older one [228].

conductivity, the conduction is associated with steeper temperature gradients in the lithosphere than
in the asthenosphere where heat is transferred by convection. That is, the lithosphere is the thermal
boundary layer of the convecting mantle. Rocks lose their ductile strength with increasing tempera-
ture. Therefore, the lithosphere is the cold and hard lid of the convecting cells.
The thickness of the thermal boundary layer changes with cooling and heating. The density
of rocks depends on temperature so that a change in the temperature field affects isostatic balance.
This is known as thermal isostasy. Therefore, the thermal regime of the lithosphere affects the
vertical movement of ocean basins. The cooling lithosphere causes subsidence. To understand this
phenomenon, we firstly consider the heat conduction equation.

Conduction of heat

A large rock mass has a large thermal capacity so that temperature change takes a long time in the
mass. Consider a mass with the diameter L at a standstill, The temperature change is described by

the same great energy as the Earth and most of the energy is emitted from its hotspot volcanoes [135].
88 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Figure 3.21: Diagram showing the model of cooling oceanic lithosphere.

the heat conduction equation


∂T
= κ∇2 T, (3.62)
∂t
where T is the temperature that is a function of time t and position, and κ is the thermal diffusivity
of the rock mass. If heat is conducted only in the x direction, this equation reduces to

∂T ∂2 T
=κ 2. (3.63)
∂t ∂x

The representative value for the thermal diffusivity of rocks is 1 m2 s−1 . This unit designates that
κ/L2 has the dimension of time. Accordingly, we define the dimensionless length x̄ and time t̄ by
dividing L and L2 /κ and the dimensionless temperature T by dividing by a reference temperature
T0 . Namely,
x̄ = x/L, t̄ = κt/L2 , T = T/T0 ,
(3.64)
dx = Ldx̄, dt = L dt̄/κ, dT = T0 dT .
2

Using these dx, dt dT , we rewrite Eq. (3.62) in the form

∂T ∂2 T
= (t̄ > 0, 0 < x̄ < 1) .
∂t̄ ∂x̄2
This equation does not include material constants such as κ, so that its solution depends only on the
initial and boundary conditions. The solution can be applied to specific problems using Eq. (3.64).
The time representing the cooling speed of the rock mass is evaluated by the ratio, τ = L2 /κ.
This is the time constant for thermal conduction. This equation demonstrates that the cooling of a
large body takes a long time.

Subsidence of the oceanic basins

Consider an oceanic lithosphere that is moving with a constant velocity of v away from the oceanic
ridge. We define the Eulerian coordinates O-xz with the origin at the spreading center and the
Laglangian coordinate ξ that travels with the oceanic plate (Fig. 3.21). The origin of ξ-coordinate
3.12. THERMAL ISOSTASY 89

is placed at z = 0, i.e., z = ξ. For simplicity, we assume that heat is transferred only by conduction.
To calculate temperature change, the thermal conduction equation,

∂2 T ∂2 T
= κ
∂t2 ∂ξ 2
is solved with the initial and boundary conditions

T = Ta at ξ ≥ 0, t = 0,
T = T0 at ξ = 0, t > 0,
T → Ta as ξ → ∞, t > 0.

That is, we assume a cooling half-space with the constant temperatures T0 and Ta at the ocean floor
and at the deep mantle, respectively. As the lithosphere ages, the ocean floor subsides and departs
from z = 0 by a few kilometers. Therefore, ξ = 0 is not the level of the floor, but this departure is
neglected in the above conditions because it is negligible compared to the dimension of the mantle.
After all, we have the solution of this problem,
 
T − T0 ξ
= Erf √ , (3.65)
Ta − T0 2 κt
where Erf( ) is the error function (Fig. 3.22) defined by the equation

2 x −η2
Erf(x) = √ e dη (x ≥ 0). (3.66)
π 0
The graph of the error function designates that the shallow and deep parts of the suboceanic mantle
have steep and gentle geothermal gradients (Fig. 3.22). The turning point gradually goes down

in the mantle with the denominator, κt, in the operand of the error function in Eq. (3.65). The

denominator has the dimensions of length so that κt is called the diffusion length. Obviously, an

isotherm descends with age as ξ ∝ κt.
The error function has an inverse function. Therefore, the depth of the isotherm with a specific
temperature T is given by Eq. (3.65), where the left-hand side is known. Substituting t = x/v and
ξ = x into the equation, we have
 ( 
T − T0 z v
= Erf . (3.67)
Ta − T0 2 κx

The thermal evolution has been investigated in the above argument. Now we can calculate the
subsidence. It is know that the density of rock at the temperature ΔT , which is measured from the
reference temperature T0 , changes as

ρ = ρ0 (1 − αΔT ), (3.68)

where ρ0 is the density at the reference temperature and α is the thermal expansion coefficient. Rocks
have a representative value for the coefficient at about 3 × 10−5 K−1 . We use ρa = ρm0 (1 − αTa ) for
90 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Figure 3.22: Graphs of error function Erf(x) and complementary error function Erfc(x).

the density of the asthenosphere. If b is the depth of the sea floor from the level ξ = 0, the isostasy
equation gives
 dc
ρw b + ρ dξ = const. (3.69)
0
The value of this equation is evaluated at the spreading center, where the asthenosphere comes
immediately below the ridge axis. Therefore, we obtain the value of the above equation there,
ρa dc + ρw bR , where bR is the water depth of the axis. Accordingly, Eq. (3.69) is rewritten as
 dc
ρw b + ρ dξ = ρa dc + ρw bR . (3.70)
0

We assume the half-space so that the compensation depth can be treated as dc = ∞. Combining Eqs.
(3.70), (3.67) and (3.68), we obtain
∞  ( !
z v
b(ρa − ρw ) = ρm0 α(Ta − T0 ) 1 − Erf dz
0 2 κx
∞  ( 
z v
= ρm0 α(Ta − T0 ) Erfc dz
0 2 κx

Erfc(x) = 1 − Erf(x) is the complementary error function (Fig. 3.22). This function satisfies
∞
1
Erfc(x) dx = √
0 π,
therefore, the depth of the sea floor at a distance of x from the spreading axis is given by the equation
(
2ρa α(Ta − T0 ) κx
b= + bR . (3.71)
ρa − ρ w πv
3.13. HORIZONAL STRESS RESULTING FROM TOPOGRAPHY 91

The depth increases with the square root of the distance, as we have assumed a constant spreading
rate at the ridge center.
We are able to estimate the effective values for the thermal diffusivity and thermal expansion
coefficient of the oceanic mantle by comparing the above model with observations. However, it is
important that old sea floors are shallower than the depth predicted by the model (Fig. 3.20) and the
ocean basins have different deviations. Some factors other than the cooling are responsible for this
flattening. It is pointed out that reheating by hotspots cannot account for the deviations [198].

3.13 Horizonal stress resulting from topography


Large-scale topographic undulation is one of the sources of tectonic stress. Ocean floors show the
difference in their depth up to a few kilometers (Fig. 3.20) so that great stress may be generated.
Following Dahlen [47], let us consider how large is the stress level associated with the variation of
the depth of the ocean floor. In conclusion, the ridge push is the force resulting from the topography.
The force is important among other origins of tectonic stresses including the colliding force of plates.
The asthenosphere comes immediately below the spreading axis, therefore we assume a constant
density ρ0 under the axis. The coordinates O-xz shown in Fig. 3.21 are used in this section. Consider
the density distribution
ρ(x, z) = ρ0 + ρ$(x, z),

where the last term is the density anomaly from the reference density ρ0 . According to Eq. (3.65),
the temperature and the resultant density anomalies were derived in the previous section as
 
z
T (z, t) = Ta Erf √ ,
2 κt
 
z
∴ ρ̃(z, t) = ρm 0 αTa Erfc √ . (3.72)
2 κt

The force FR that a unit length of the ridge axis is determined by the dipole moment of vertical
density distribution (§3.11). The moment is obtained from Eq. (3.72) so that
∞
 
FR = ρ$gz dz = ρm 0 gαTa κ t.
0

Consequently, the force is proportional to the age of the oceanic lithosphere15 . The constant of
proportionality is about 0.035 TNm−1 /m.y. For example, the ocean floor with an age of 80 m.y.
has FR = 2.8 TNm−1 . The force resulting from the ocean-continent topographic contrast has a
magnitude of about 1.6 TNm−1 (§3.11). Therefore, the ridge push is as great as the force by the
contrast.
15 The integral is taken over the range [0, ∞]. However, the contribution from the asthenosphere is negligible because of

Erfc(x)  1 for x > 2 (Fig. 3.22).


92 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

In the above argument, we used the density profile under the spreading axis as the reference.
It is the next problem where the proper area that provides the reference is sought. If the density
distribution was under the summit of Mount Everest, the entire world would have been under a
compressional stress field. However, this is not the case. Accordingly, Coblentz and others excluded
the arbitrariness in the choice of reference by taking global topography into account [42].
Let the levels of the top and base of the lithosphere be z = h and z = L, respectively. Note
that the vertial force per unit area is σzz × (1 m2 ) and that the product of force and length has the
dimensions of energy. Then the quantity
L  L h
U= σzz dz = ρ(ξ)g dξdz
h h ξ

is the potential energy of the lithospheric column with the unit sectional area. A region with a large
U tends to be subject to extensional tectonics. That with a small U is ready to be horizontally
contracted. The global average is estimated at 2.4 TNm−1 [42]. The threshold altitude for the
continent is ±200 m, and the threshold depth of the ocean is about 4.3 km. Spreading centers have
depths actually shallower than this threshold. However, the topography is not the only cause of
lithospheric stress. Therefore, an extensional stress field rifts continents and eventually breaks up
the continental lithosphere.

3.14 Thermal subsidence of continental rifts


It was shown in the previous chapter that the subsidence history of a passive continental margin is
revealed from stratigraphic records. Here we consider Mckenzie’s simple stretching model to link
the subsidence and rifting together.

Post-rift subsidence

Passive margin subsidence after continental breakup is generally associated with very weak or no
fault activity. Extensional tectonics does not account for the subsidence. In contrast, the continental
lithosphere is thinned and horizontally extended during the rift phase. The fact is that the rifting
accounts for the post-rift subsidence.
Consider the thermally equilibrium state of the lithosphere illustrated in Figs. 3.23(a) and (c),
i.e., the temperature at the base of the lithosphere is Ta at a depth of z = tL . The temperature
distribution before and long after rifting is assumed to be in equilibrium. The continental crust is
thinned with the lithospheric mantle. The thinning of the crust results in subsidence. Let tc and tL be
the initial thickness of the crust and the lithosphere, respectively (Fig. 3.23(a)). It is also assumed
that the temperature at z = tL is kept at Ta . The top of the lithosphere was at sea level before rifting.
Here, the lithosphere is thought to be subject to homogeneous deformation in the rift phase. If
the lithosphere is horizontally extended by a factor of β, the thickness of the lithosphere and crust
become tc /β and tL /β, respectively. The rift-stage subsidence is affected not only by the thinning
3.14. THERMAL SUBSIDENCE OF CONTINENTAL RIFTS 93

Figure 3.23: Diagrams showing the simple stretching model [138]. (a) Structure and temperature
distribution in the lithosphere in the pre-rift stage, (b) at the end of rifting, and (c) in post-rift stage.

but also by the temporal variation of the thermal regime, because rock density changes with the
variation as
ρ = ρ0 (1 − αT ), (3.73)

where ρ0 is the reference density at T = 0◦ C. If the thinning of the lithosphere is quick compared to
the time constant of heat conduction, isotherms in the lithosphere are uplifted adiabatically16 (Fig.
3.23(b)). If we assume isostatic compensation throughout the rifting process, we obtain the syn-rift
subsidence  
tL ρa − tc ρc − (tL − tc )ρm 1
Si = 1− , (3.74)
ρa − ρ w β
where ρc , ρm and ρa are the representative densities of the crust, mantle lithosphere, and astheno-
sphere, respectively, and are given by
   
αTa tc αTa αTa tc
ρa = ρm0 (1 − αTa ), ρc = ρc0 1 − , ρm = ρm0 1 − − .
2 tL 2 2 tL
16 This assumption is valid for the rifting with a duration shorter than ∼25 m.y. [92].
94 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Table 3.2: Parameters for the simple stretching model.

ρm0 = 3.3 × 103 kg m−3 α = 3.5×10−5 K−1 tL = 100 km


ρc0 = 2.8 × 103 kg m−3 κ = 1.0×10−6 m2 s−1 tc = 35 km
ρw = 1.0 × 103 kg m−3 Ta = 1333◦ C

The coefficients, ρc0 and ρm0 are the reference densities of the crust and mantle. Si is called the
initial subsidence in the passive margin formation. Substituting the representative densities into Eq.
(3.74), we obtain the magnitude of the initial subsidence
 
1
Si = Si0 1 − , (3.75)
β
   
αTa tc tc αTa
(ρm0 − ρc0 ) 1 − − ρm0 tL
2 tL tL 2
Si0 = . (3.76)
ρm0 (1 − αTa ) − ρw
The latter quantity, (Si0 ) can have positive and negative signs depending on the initial thickness of
the crust. Since the denominator in the right-hand side of Eq. (3.76) is positive, the sign depends on
the numerator. Solving the equation, (numerator) = 0, we obtain the solution,

tc ρm0 − ρc0 4(ρm0 − ρc0 )2 + 4α 2 (ρm0 − ρc0 )ρm0 Ta2
= ± .
tL α(ρm0 − ρc0 )Ta α(ρm0 − ρc0 )Ta
Using the parameters shown in Table 3.2, we obtain the ratio tc /tL = 0.154 and 42.7. However, this
ratio should be less than unity so that the latter is a meaningless solution. The former indicates a
threshold initial thickness of tc ≈ 15 km. The lithosphere with a continental crust initially thicker
than ∼15 km subsides through the lithospheric thinning, but the surface is uplifted by rifting if the
crust is initially thinner than this value. The uplift is due to a steepening of the geothermal gradient.
That is, the attenuation of the lithospheric mantle plays an opposite role to that of the continental
crust. Island arcs have volcanism so that their lithosphere is thinner than the ordinary continental
lithosphere. If an island arc has a lithosphere with a pre-rift thickness of tL = 50 km, then the arc
has the threshold at about 7 km. The geohistory analysis of syn-rift sediments can constrain β with
Eq. (3.75). This parameter is called the stretching factor.
The thermal regime is altered by the rifting. Isotherms between 0◦ and Ta are raised by the rifting
(Fig. 3.23(b)). After the rifting, the isostherms descend to the equilibrium level in the long run (Fig.
3.23(c)). This post-rift cooling causes a gradual subsidence of the passive margin. This is called the
thermal subsidence of the lithosphere. The simple stretching model (Fig. 3.23) allows us to estimate
the rate and amount of this subsidence [138].
In order to calculate the heat conduction after the termination of rifting, the initial subsidence is
neglected because Si  tL . Therefore, the thermal conduction equation ∂T/∂t = κ∂2 T/∂z2 is solved
with the boundary condition T = 0 at z = 0. In addition, the temperature at z = tL is assumed to be
3.15. KINEMATIC MODEL OF SYN-RIFT SUBSIDENCE 95

kept at T = Ta . The time t is measured from the termination. The initial condition is shown by the
graph in Fig. 3.23(b), which is the temperature distribution at the end of rifting. Solving this system,
we obtain the thermal evolution
 
2  (−1)n+1 β

T z nπ nπz −n2 t/τ
= 1− + sin sin e , (3.77)
Ta tL π n nπ β tL
n=1
τ= π 2 t2L /κ. (3.78)

The parameters shown in Table 3.2 give the time constant τ = 3.2 m.y.
Combining Eqs. (3.73) and (3.77), we obtain the evolution of the density structure ρ(z, t) .
Isostatic compensation through time is expressed as
 tL  tL
ρ(z, t) dz + ρw St = ρ(z, 0) dz,
0 0

where St is the thermal subsidence,


 
4αρm0 Ta tL β π 
St ≈ sin 1 − e−t/τ . (3.79)
π (ρm0 − ρc0 ) π
2 β

Higher-order terms in Eq. (3.77) are neglected in deriving Eq. (3.79).


Equation (3.79) shows that the rate of thermal subsidence decays as the exponential function.
The subsidence curve shown in Fig. 3.16 has this shape. Fitting the curve given by Eq. (3.79), we
obtain the stretching factor for the area the curve was determined.

3.15 Kinematic model of syn-rift subsidence


Post-rift subsidence results from the cooling of the lithosphere, whereas syn-rift subsidence is con-
trolled by not only the thermal but also the mechanical properties of the lithosphere. The latter
affects the characteristics of rift zones such as their widths and depths. Decompression melting of
the asthenosphere under rift zones causes uplift through the buoyancy of magma. Therefore, the
observation of syn-rift subsidence sheds light on the deep seated factors and on the secular changes
of these factors.
The simple stretching model (§3.14) assumed an instantaneous thinning of the lithosphere and
neglected the details of the rifting processes. In this section, we consider a mathematical inverse
method to constrain the rift-phase strain rate of the lithosphere from an observed rift-phase subsi-
dence curve (Fig. 3.24) [259].
Consider a pure shear rift with horizontal and vertical strain rates of Ėxx and Ėzz , respectively.
The lithosphere is extending horizontally so that we have Ėzz < 0 < Ėxx . The pre-rift levels of the
top and base of the lithosphere is z = 0 and z = a. The lithosphere is subject to a pure shear, so that
we have
−ε̇zz = ε̇xx = G(t),
96 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

where the function G(t) is defined to describe the temporal variation of the strain rates and is linked
to the rate of subsidence in the following discussion.
Consider that the horizontal velocity field is symmetric with respect to the plane x = 0. The
lithosphere is assumed to undergo homogeneous strain so that the horizontal velocity u(x, t) is pro-
portional to x. That is, we have
u(x, t) = G(t) x. (3.80)
Syn-rift subsidence reaches several kilometers, but is negligible compared to the representative
thickness of continental lithosphere. Therefore, we assume the top of the lithosphere is kept at z = 0
during the calculation of the lithospheric strain. This assumption allows us to write the descending
velocity of rocks at depth z as
v(x, t) = −G(t) z. (3.81)
Note that this is the velocity in the +z direction.
Subsurface temperature distribution affects the subsidence of this rift. The thermal conduction
equation to be solved should include velocity, because heat is transported not only by conduction
but also by travelling rocks. Hence, the left-hand side of the thermal conduction equation, ∂T/∂t =
κ∇2 T , should be replaced by the material derivative (Eq. (2.20)). That is, the equation
    2 
∂T u ∂ T ∂2 T
+ ∇· T =κ + (3.82)
∂t v ∂x2 ∂z2
governs the thermal evolution. From Eqs. (3.81) and (3.80), we have
  
u ∂T ∂T
∇· T = G(t)x − G(t)z . (3.83)
v ∂x ∂z
If there was no horizontal variation in the initial temperature distribution, the pure shear would not
raise horizontal temperature variation. Therefore, we have
∂T ∂2 T
= = 0.
∂x ∂x2
Combining this and Eqs. (3.82) and (3.83), we obtain
∂T ∂T ∂2 T
+ Gz =κ 2 . (3.84)
∂t ∂z ∂z
The boundary conditions that we assume for the temperature evolution are T = 0 at z = 0 and
T = Ta at z = a. The initial condition and the equilibrium temperature distribution is the constant
geothermal gradient to a depth of z = a.
The subsidence of this rift is the sum of the subsidence due to the thinning of the lithosphere
and to the thermal perturbation. The former is described by Eq. (3.75). The equation describing the
latter factor should have the form
 a 
Q(t) = B T (z, t) − T (z, ∞) dz, (3.85)
0
3.15. KINEMATIC MODEL OF SYN-RIFT SUBSIDENCE 97

Figure 3.24: (a) Strain rate history of a Triassic rift in northern Italy determined from a subsidence
curve of the rift (b). The 95% confidence region is shown in (a) by white [259].

where B = αρm /(ρa − ρw ) is the factor to account for the buoyancy of the crust. Therefore, we have
the total subsidence  
1
S(t) = A 1 − − BQ(t), (3.86)
β
where A = tc (ρm − ρc )/(ρa − ρw ). Crustal thinning has a positive effect for the total subsidence, but
the thinning of the lithosphere increases the geothermal gradient which has a negative effect for the
subsidence. Rearranging Eq. (3.86), we have

A
β= . (3.87)
A − S − BQ(t)
Equation (2.39) links the stretching factor β and G(t) as
 tR 
β = exp G(t) dt , (3.88)
0

where tR is the time ellapsed since the rifting began. Accordingly, combining Eqs. (3.87) and (3.88),
we obtain the strain-rate history of the rift
)  *
d A
G(t) = ln . (3.89)
dt A − S(t) − BQ(t)

The rift-phase strain-rate history can be evaluated from observations for a pure shear rift. The
right-hand side of Eq. (3.89) has S(t) that is determined from the quantitative stratigraphy of the
rift. However, Q(t) depends on the unknown G(t) through Eq. (3.84). Therefore, G(t) is determined
by a mathematical inversion17 from the observed subsidence history S(t). Figure 3.24(a) shows the
17 See [259] for detail.
98 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

optimal strain-rate history for the subsidence history of a Triassic rift zone in northern Italy (Fig.
3.24(b)). The subsidence curve of the Triassic rift has an inflection point at about 220 Ma (Fig.
3.24(b)). After the point, the subsidence curve is similar to an exponential curve, suggesting thermal
subsidence. The point designates a rejuvenated rifting, and corresponding to the second peak in the
strain-rate history.
A sharp decline in the strain rate by several orders of magnitude indicates the termination of rift-
ing; the rift was aborted or the area became a passive margin. Timing of the termination determined
by the inverse method for the Triassic rift has a large uncertainty which is illustrated by the 90%
confidence region between 100 and 220 Ma in Fig. 3.24(a). The umbiguity stems from the difficulty
to detect very slow rifting if it overlaps simultaneous thermal subsidence. More definite constrains
come from geological structures. The termination of rifting is marked by the cessation or distinct
decline of fault activity. For example, the cross-section in Fig. 3.12 shows that the subsidence of
wedge shaped grabens had ceased by the early Jurassic. If rifting was successful to form a passive
continental margin, the magnetic anomaly near the continent-ocean boundary provides a constraint
for the timing. Some passive margins were remarkably uplifted when continents were broken off the
margins. Such a unconformity is known as a breakup unconformity [59].

3.16 Exercises
3.1 Assume a state of stress with the principal stresses σ1 = 4 MPa, σ2 = 5 MPa and σ3 = 8 MPa,
and vertical σ1 -axis and horizontal σ3 -axis oriented due east. Determine the magnitude of the normal
and shear stresses acting on the block beneath a fracture plane of which trend and dip are N030◦ E
and 60◦ SE, respectively. How much is the rake of the maximum shear stress on the plane?

3.2 Explain qualitatively why the function g(r) shown in Fig. 3.7(b) is convex downward and why
the function decreases with increasing r in the lower mantle.

3.3 What is the magnitude of the lithostatic pressure at the Moho interface. Assume the thickness
and density of the crust at 35 km and 2.8 × 103 kg m−3 , respectively.

3.4 Derive Eq. (3.46).

3.5 Estimate how much is the force per unit boundary length with which the Tibetan Plateau pushes
the Hindustan Plain due solely to the structure illustrated in Fig. 3.25.
3.16. EXERCISES 99

Figure 3.25: Lithostatic pressure under the neighboring plateau and lowland.
Chapter 4

Principal Stresses

Anisotropic stress causes tectonic deformations. Tectonic stress is defined in


an early part of this chapter, and natural examples of tectonic stresses are also
presented.

4.1 Principal stresses and principal stress axes


Because of symmetry (Eq. (3.32)), the stress tensor r has real eigenvalues and mutually perpendicu-
lar eigenvectors. The eigenvalues are called the principal stresses of the stress. The principal stresses
are numbered conventionally in descending order of magnitude, σ1 ≥ σ2 ≥ σ3 , so that they are the
maximum, intermediate and minimum principal stresses. Note that these principal stresses indicate
the magnitudes of compressional stress. On the other hand, the three quantities S1 ≥ S2 ≥ S3
are the principal stresses of S, so that the quantities indicate the magnitudes of tensile stress. The
orientations defined by the eigenvectors are called the principal axes of stress or simply stress axes,
and the orientation corresponding to the principal stress, e.g., σ1 is called the σ1 -axis (Fig. 4.1).
Principal planes of stress are the planes parallel to two of the stress axes, or perpendicular to one of
the stress axes.
Deformation is driven by the anisotropic state of stress with a large difference of the principal
stresses. Accordingly, the differential stress
ΔS = S1 − S3 , Δσ = σ1 − σ3
is defined to indicate the difference1 .
Taking the axes of the rectangular Cartesian coordinates O-123 parallel to the stress axes, a stress
tensor is expressed by a diagonal matrix. Therefore, Cauchy’s stress (Eq. (3.10)) formula is
⎛ ⎞ ⎛ ⎞⎛ ⎞
t1 (n) σ1 0 0 n1
t(n) = ⎝t2 (n)⎠ = ⎝ 0 σ2 0 ⎠ ⎝n2 ⎠ .
t3 (n) 0 0 σ3 n3
1 ΔS is sometimes called the stress difference [2]. However, we keep this term for another quantity (Eq. (11.19)).

101
102 CHAPTER 4. PRINCIPAL STRESSES

Figure 4.1: Stress ellipsoid and principal stress axes. White lines indicate the principal planes of
stress.

∴ t1 (n) = σ1 n1 , t2 (n) = σ2 n2 , t3 (n) = σ3 n3 . (4.1)


The planes perpendicular to the stress axes have unit normals, n = (1 0 0)T , (0 1 0)T , (0 0 1)T .
Hence, the traction vectors upon the planes are
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
σ1 0 0
t(n) = ⎝ 0 ⎠ , ⎝σ2 ⎠ , ⎝ 0 ⎠ .
0 0 σ3

Therefore, the principal stresses equal the normal stresses working on the principal stress planes.
Shear stress vanishes on the planes. The normal and shear stresses are written in this coordinates as

σN = t(n) · n = σ1 n21 + σ2 n22 + σ3 n23 (4.2)


1/2  2 2  2 1/2
σS = |t(n)|2 − |σN |2 = σ1 n1 + σ22 n22 + σ3 n23 − σ1 n21 + σ2 n22 + σ3 n23 . (4.3)

If there is a plane on which traction vanishes, the state of stress is said to be plane stress (§3.4).
The stress tensor corresponding to this state has a null principal stress. If σ3 = 0, traction on the
σ1 σ2 -plane vanishes. Combining Eq. (4.1) and n21 + n22 + n23 = 1, we have
 2  2  2
t1 (n) t2 (n) t3 (n)
+ + = 1. (4.4)
σ1 σ2 σ3

This is an ellipsoid equation whose principal axes coincide with the stress axes and radii equals |σ1 |,
|σ2 | and |σ3 | (Fig. 4.1). This is called Lamé’s stress ellipsoid. Regarding t(n) as a position vector
with its initial point at the coordinate origin, the end point of the vector is on the ellipsoid, provided
that 0 < σ3 . The ellipsoid becomes a sphere for the hydrostatic state of stress and its radius equals
the hydrostatic stress. The stress ellipsoid is given uniquely from the stress tensor regardless of the
orientation of coordinate axes, although Eq. (4.1) was derived for a specific orientation of the axes.
The basic invariants (§C.6) of the stress tensor r are σI , σII and σIII . Those of the tensor S are
SI , SII and SIII . These quantities designate the “absolute values” of the stress tensor, independent
from the orientation of rectangular Cartesian coordinates. We can define variables with the same
4.2. TECTONIC STRESS 103

independence from the basic invariants. Among them is the stress ratio
σ2 − σ3
Φ= . (4.5)
σ1 − σ3
It is seen that the ratio is in the range of 0 ≤ Φ ≤ 1. When the state of stress is hydrostatic, the stress
ratio is not defined. For anisotropic stress states, the stress ratio designates the shape of the stress
elliosoid. If Φ = 0, σ3 = σ2 < σ1 , then, the ellipsoid is prolate. Φ = 1 corresponds to an oblate
ellipsoid, σ3 < σ2 = σ1 . The stress ellipsoid has axial symmetry for the extreme cases Φ = 0 and 1.
In contrast, intermediate Φ indicates triaxial stresses σ3 = σ2 = σ1 . In this case, the stress ellipsoid
and stress tensor have orthorhombic symmetry.

4.2 Tectonic stress


The lithostatic state of stress causes isotropic contraction for a homogeneous material. Deviation
from this state results in strain. Tectonic stress is generally defined as this deviation. The lithostatic
state of stress is used as the reference. Other states of stress can also be taken as the reference (§7.2).
The average

σ1 + σ2 + σ3 trace r σI S1 + S2 + S3 trace S SI
σ0 ≡ = = or S0 ≡ = = (4.6)
3 3 3 3 3 3
is said to be mean stress.
For lithostatic stresses, mean stress coincides with pressure. Accordingly, it is convenient to
define deviatoric stress
s = r − σ0 I (4.7)
as the deviation from the lithostatic state of stress. This is the deviatoric tensor introduced in Section
C.6. Let us define the symbol T as
T = S − S0 I. (4.8)
It is obvious that a deviatoric tensor is symmetric, so that the tensor has its principal orientations
parallel to the stress axes (§C.5). If T1 , T2 and T3 are the principal deviatoric stresses, they are
related to the principal stresses as
Ti = S i − S 0 . (4.9)
Combining Eqs. (4.6), we have

2S1 − S2 − S3 2S2 − S3 − S1 2S3 − S1 − S2


T1 = , T2 = , T3 = . (4.10)
3 3 3
Using the stress ratio, these are rewritten as
     
2−Φ 2Φ − 1 Φ+1
T1 = ΔS, T2 = ΔS, T3 = − ΔS. (4.11)
3 3 3
104 CHAPTER 4. PRINCIPAL STRESSES

Figure 4.2: Diagram for the explanation of the Hubbert-Rubey model.

Hubbert-Rubey model

In the previous chapter, we considered only the lithostatic state of stress. Our first application of the
deviatoric stress tensor is the statics of a thrust sheet. There are thrust sheets that travel horizontally
for more than 100 km. The Hubbert-Rubey model explains why such long-distance displacement is
possible or what it implies. The very low friction of the fault surface is the key [86, 87].
Consider a thrust sheet that has a thickness of t and is advancing in the x direction (Fig. 4.2).
For simplicity, the topographic surface coincides with the top of this sheet. We pay attention to a
part of the sheet with length L in the x direction and unit length in the y direction. The tectonic
forces pushing at the rear and front of this portion are τxx t in the +x direction and −(1 + Δ)τxx t
in the −x direction. Therefore, the net force to push the portion of the sheet in the +x direction is
−Δτxx t. Frictional resistance is the one significant force for the sheet. The resistance per unit basal
area is σzx , and the law of friction gives σzx = μf σzz = μf ρgt, where μf is the coefficient of friction
and ρ is the mean density of the sheet. Therefore, a portion feels the resistance, −μf ρgLt, in the +x
direction. The motion of the sheet is so slow that inertia is negligible. Hence, all forces acting on the
portion is balanced: −Δτxx t = −μf ρgLt. Therefore, we obtain Δτxx = ρμf gL. In order to evaluate
the coefficient of friction, we substitute Δτxx = 100 MPa as the representative magnitude of tectonic
stresses and ρ = 2.75 × 103 kg m−3 in this equation. We consider a thrust sheet that travels hundreds
of kilometers so that we assume L = 100 km. The result is that μf = 0.036. The friction at the base
of those giant thrust sheets is very low compared to the coefficient of the friction of ordinary rocks.
Pore fluid lowers the friction (§6.5).

Octahedral shear stress

The deviatoric stress tensor has the characteristic equation shown in Eq. (C.48), and its first basic
invariant is

TI = trace T = T1 + T2 + T3 = 0. (4.12)
4.2. TECTONIC STRESS 105

Figure 4.3: Basic invariants of deviatoric stress tensor versus stress ratio.

The second basic invariant satisfies


1 1
TII = T:T= Tij Tij
2 2 i,j
1 2 
= T11 + T22
2
+ T332
+ T12
2
+ T23
2
+ T31
2
(4.13)
2
1 
= (S11 − S22 )2 + (S22 − S33 )2 + (S33 − S11 )2 + S12
2
+ S23
2
+ S31
2
6
1 
= (S1 − S2 )2 + (S2 − S3 )2 + (S3 − S1 )2 ≥ 0.
6
It is obvious in the last line that τII ≥ 0. TII is related to the principal deviatoric stresses as

1 2  1 
TII = T1 + T22 + T32 = (T1 + T2 + T3 )2 − 2T1 T2 − 2T2 T3 − 2T3 T1
2 2
= −T1 T2 − T2 T3 − T3 T1 . (4.14)

The transformation at the second equal sign is based on Eq. (4.12). Substituting Eq. (4.11), we
obtain
Φ2 − Φ + 1
TII = (ΔS)2 . (4.15)
3
Therefore, TII ranges from 41 (ΔS)2 at Φ = 12 to 13 (ΔS)2 at Φ = 0 and 1.
The third basic invariant is TIII = T1 T2 T3 . Combining this equation and Eq. (4.11), we have

(2 − Φ)(2Φ − 1)(Φ + 1)
TIII = − (ΔS)3 .
27

Therefore, TIII ranges from − 27


2
(ΔS)3 ≈ −0.074(ΔS)3 at Φ = 1 to 27 2
(ΔS)3 ≈ 0.074(ΔS)3 at
Φ = 0. When Φ = 2 , we have TIII = 0. Figure 4.3 shows the graphs of TII and TIII .
1
106 CHAPTER 4. PRINCIPAL STRESSES

Figure 4.4: An octahedral plane and its unit normal.

Octahedral shear stress



Taking coordinate axes parallel to the stress axes, the eight unit vectors e⊥ = (±1, ±1, ±1)T / 3
are the unit normals for the faces of a regular octahedron. The octahedral plane is the plane which
makes equal angles with the principal stress axes (Fig. 4.4). The normal and shear stresses upon the
planes are called the octahedral normal stress and octahedral shear stress. The former is given by
the equation
SNoct = e⊥ · S · e⊥ = S : e⊥ = S0 , (4.16)
 
where e⊥ ≡ e⊥ e⊥ is the projector onto the vector e⊥ . Equation (4.16) designates that the octahe-
dral normal stress equals the mean stress.
The octahedral shear stress is used in the theory of plasticity (§10.2), and is proportional to the
second basic invariant of the deviatoric stress tensor
 oct 2   2 1   2
SS =  S : I − e⊥  = (T1 − T2 )2 + (T2 − T3 )2 + (T3 − T1 )2 = TII ,
9 3
or, equivalently,
(
2 1
SSoct = TII = (T1 − T2 )2 + (T2 − T3 )2 + (T3 − T1 )2 . (4.17)
3 3

For the convention that compression is positive stress, we use the symbol σSoct for the octahedral
shear stress.

4.3 Mohr diagram


The Mohr diagram is useful to illustrate the state of stress and the combinations of normal and
shear stresses for various directions of surface elements. These combinations are important when
we consider friction on a fault plane. Traction on a surface element depends on the direction of its
normal n, and the normal and shear components also changes with n. The admissible combinations
4.3. MOHR DIAGRAM 107

Figure 4.5: Mohr’s circles illustrate the admissible combinations of normal and shear stresses, σN
and σS , for a given set of principal stresses σ1 , σ2 and σ3 which are indicated by shaded  area. The
sign σS indicate the sense of shear. The centers of the circles are located at σ2 +σ 3
, 0 , σ1 +σ3
, 0
  2 2
and σ1 +σ
2
2
, 0 . The maximum shear stress equals Δσ/2.

for a given state of stress are represented by areas bounded by three circles C1 , C2 and C3 on O-σN σS
plane (Fig. 4.5). They are called Mohr’s circles. The horizontal axis O-σN passes the centers of the
circles of which the diameters are σ1 − σ3 , σ1 − σ2 and σ2 − σ3 . The first one equals the differential
stress Δσ. The circles are symmetric with respect to the horizontal axis O-σN , so that the upper half
of Mohr’s stress diagram, or Mohr diagram can illustrate the state of stress. No matter how great or
small the magnitude of σ2 is, the maximum shear stress for a given state of stress is determined only
by σ1 and σ3 as
σ1 − σ3
Maximum shear stress = . (4.18)
2
This is equal to the diameter of the outermost circle C2 (Fig. 4.5).
Given the principal stresses and the unit normal of a surface element, the normal and shear
stresses on the element are represented in Fig. 4.5 as the shaded area bounded by the Mohr’s
circles. The relationship between the position of the point in the diagram and the direction of
n = (cos φ, cos β, cos θ)T is established by taking Fig. 4.6 into consideration. These components
are the direction cosines of n. Since a stress tensor has orthorhombic symmetry, Fig. 4.6(a) shows
the first octant of a sphere on which a point Q represents the direction of n. The line segment QP is
parallel to n, where P is the center of the sphere. It is shown (Exercise 4.1) that the octant is mapped
onto the shaded area in Fig. 4.6(b), and that the points “A” through “H” on the octant correspond to
the points “a” through “h” in the figure. The point “Q” that represents the direction of n corresponds
108 CHAPTER 4. PRINCIPAL STRESSES

Figure 4.6: Correspondence between the direction of n and the point in Mohr’s stress diagram. (a)
The direction is represented by a point “Q” on an octant of a sphere. The points “K”, “Q” and “D”
indicate the directions that meet the σ1 axis at the common angle φ. (b) The points “a”, “b” and “c”
are plotted at the points (σ1 , 0), (σ2 , 0) and (σ3 , 0), respectively. The points “A” through “H” and
“Q” on the sphere correspond to the points “a” through “h” and “q” in the diagram. The small circle
“KQD” corresponds to the circle “kqd”.

to the point “q” in Mohr’s stress diagram.


To find the point “q” in the diagram for a given n = (cos φ, cos β, cos θ)T , one has to plot
firstly the point “h” that is rotated clockwise along the circie C1 from the point “c” by the angle 2θ.
Secondly, the point “d” is plotted on the circle C3 . The point is rotated counterclockwise from the
point “a” by the angle 2φ. Thirdly, one draws arcs “hf” and “dk” that are concentric with the circles
C3 and C1 , respectively. The intersection of the arcs is the point “q”.
The maximum shear stress is indicated by Eq. (4.18). Obviously, shear stress becomes the
maximum at the point “m” in Fig. 4.6(b), where σN = (σ1 + σ3 )/2 and σS = (σ1 − σ3 )/2. The point
“m” corresponds to the point “M” in Fig. 4.6(a), and ∠APM = ∠CPM = 45◦ , i.e., shear stress is
the maximum on the surface that is parallel to the σ2 axis and that makes an angle of 45◦ with the σ1
and σ3 axes.

4.3.1 Classes of stress


The state of stress at a point can be classified as shown in Fig. 4.7.

Lithostatic or hydrostatic pressure (σ3 = σ2 = σ1 = p): The normal stress across all planes is equal
to pressure p, but there are no shearing stresses. In this case, the state of stress is symmetric,
r = pI. The stress plots on the Mohr diagram degenerate into a point on the abscissa.
4.3. MOHR DIAGRAM 109

Uniaxial compression (0 = σ3 = σ2 < σ1 ): The only stress applied is a compressive stress in one
direction. The state of stress has axial symmetry about the σ1 axis. The Mohr diagram for this
case is a single circle tangent to the ordinate at the origin.

Uniaxial tension (σ3 < σ2 = σ1 = 0): The only stress applied is a tension in one direction. The
state of stress has axial symmetry about the σ3 axis. The Mohr diagram for this case is a single
circle tangent to the ordinate at the origin.

Axial compression (σ3 = σ2 < σ1 ): The state of stress has symmetric about σ1 axis. The Mohr
diagram degenerates into a single circle.

Axial tension (σ3 < σ2 = σ1 ): Some authors avoid using this term, because “tension” seems to
indicate negative principal stresses. The distinction should always be made clear [248]. The
state of stress has symmetry about the σ3 axis. The Mohr diagram degenerates into a single
circle for this case.

Plane stress (0 = σ3 < σ2 ≤ σ1 ): If a principal stress is zero, plane stress is said to exist. This is
synonymous with biaxial stress [91]. The state of stress has orthorhombic symmetry.

Pure shear stress (σ3 = −σ1 , σ2 = 0): The normal stress on planes of maximum shear stress
is zero. The state of stress has orthorhombic symmetry. The Mohr diagram for this case is
symmetric with respect to the ordinate and to the origin.

Triaxial stress (σ3 < σ2 < σ1 ): All the principal stresses are different. The state of stress has
orthorhombic symmetry. Three distinct circles compose the Mohr diagram.

If there are planes on which traction vanishes (σN = σS = 0), they are called free surfaces or free
boundaries. Uniaxial and plane stresses have those planes. If O-3 axis is chosen tobe perpendicular
to the surface, then σ33 = σN = 0 and the shear stress on the surface is |σS | = σ31 2
+ σ32
2
= 0.
Therefore σ31 = σ32 = 0. Consequently, the stress tensor for the stress state has the form
⎛ ⎞
• • 0
r = ⎝• • 0⎠ , (4.19)
0 0 0

where the black dots represent the components that are not determined by the boundary condition.

4.3.2 Mohr diagram for two-dimensional problems


Axial stresses can be treated as two-dimensional problems, where there are two principal stresses,
σ1 and σ2 , accompanied by two principal axes. One of the axes is parallel to the axis of symmetry
for the state of stress. Even for triaxial stresses, such a problem can be considered to estimate the
normal and shear stresses to investigate faulting (Chapter 6), as σ2 has a limited effect on faulting
(Chapter 10).
110 CHAPTER 4. PRINCIPAL STRESSES

Figure 4.7: Mohr diagrams for special states of stress.

The admissible combinations of normal and shear stresses for a given state of stress are illustrated
by a Mohr diagram that consists of a single circle. To see this, suppose a stress tensor
 
σ1 0
r= ,
0 σ2

for which the rectangular Cartesian coordinates O-12 are taken as parallel to the principal stress axes
(Fig. 4.8). Other coordinates O-1 2 are defined to have the same origin O with the former coordi-
nate system. The coordinate axes meet at an angle of θ. The orthogonal tensor for this coordinate
transformation is given by Eq. (C.8). Thus, stress components are transformed from O-12 to O-1 2
as
  
    
σ11 σ12 cos θ sin θ σ1 0 cos θ − sin θ
  =
σ21 σ22 − sin θ cos θ 0 σ2 sin θ cos θ
 
σ1 cos2 θ + σ2 sin2 θ −(σ1 − σ2 ) sin θ cos θ
= . (4.20)
−(σ1 − σ2 ) sin θ cos θ σ1 cos2 θ + σ2 sin2 θ

 
It is obvious in Fig. 4.8(a) that σN and σS are equivalent to σ11 and σ12 , respectively. These stress
4.3. MOHR DIAGRAM 111

Figure 4.8: Mohr diagrams for two-dimensional problems. The axes O-12 and O-1 2 represent two
rectangular Cartesian coordinate systems with a common origin O. The former axes are taken to be
parallel to the principal stress axes. O-1 axis is taken to be parallel to N that is the outward normal
to the surface element on which the traction is considered. This surface element has positive and
negative directions for N and n, respectively, leading to the opposite sign of the shear stress, SS and
σS . See Fig. 3.2 for the relationship between positive shear directions. (b, c) Mohr’s circles for each
of the sign conventions.

components are shown in Eq. (4.20). Namely,


σN = σ11 = σ1 cos2 θ + σ2 sin2 θ,

σS = σ12 = −(σ1 − σ2 ) sin θ cos θ.

Using the formulas, sin2 θ = (1 − cos 2θ)/2, cos2 θ = (1 + cos 2θ)/2 and sin θ cos θ = 1
2 sin 2θ, we
obtain

1 1
σN = (σ1 + σ2 ) + (σ1 − σ2 ) cos 2θ (4.21)
2 2
1
σS = − (σ1 − σ2 ) sin 2θ. (4.22)
2
112 CHAPTER 4. PRINCIPAL STRESSES

Figure 4.9: (a) Planes perpendicular to the O-3 axis are assumed to be the free surfaces. (b) Mohr
diagram for the plane stress.

Plane stress Suppose a state of plane stress where the free surface is parallel to the O-12 plane
(Fig. 4.8(a)). That is, the stress tensor has the components
⎛ ⎞
S11 S12 0
S = ⎝S21 S22 0⎠ .
0 0 0

The Mohr’s circle for this stress is drawn as the following procedure. The center of the circle B in
Fig. 4.9(b) is located at SN = (S11 + S22 )/2, because the trace of a tensor is invariant to coordinate
rotations. The points A and D on the circle represent the traction on the right-hand and top faces,
respectively, of the rectangular parallelepiped in Fig. 4.8(a). The line segment AD is the diameter
of the circle, as the faces make a right angle. The phase angle in a Mohr diagram is twice as much
as an angle in physical space. The line segment AB in Fig. 4.8(b) is the radius of the circle, and

2
AB = (S11 − S22 )/2 + S12
2
.

4.4 Boundary conditions of stress


In the shallow part of the solid Earth, one of the stress axes is perpendicular to the surface. To see
this, consider
 two  materials I and
 II that
 are firmly fixed to each other at the O-12 plane (Fig. 4.10a).
(I) (II)
Let r = σij and r = σij be the stress tensors in materials I and II, respectively, where
(I) (II)

the coordinates O-123 are defined in Fig. 4.10. The traction t(I) is exerted on material I at the point
P on the interface. Likewise, material II feels the traction t(II) at the same point. If all forces are
balanced and the materials are at rest, we have

t(I) + t(II) = 0. (4.23)


4.4. BOUNDARY CONDITIONS OF STRESS 113

Figure 4.10: Schematic illustrations for the explanation of stress boundary coditions. (a, b) Materials
I and II are firmly fixed at the O-12 plane. O-2 axis is perpendicular to the page. (c, d) Materials I
and II with a lubricated interface.

Given the unit normal to the interface n = (0, 0, −1)T , the first traction vector is written as
⎛ (I) (I) (I) ⎞ ⎛ ⎞ ⎛ (I) ⎞
σ11 σ12 σ13 0 σ13
t(I) = r(I) · n = ⎝σ21 σ22 σ23 ⎠ ⎝ 0 ⎠ = − ⎝σ23 ⎠ .
(I) (I) (I) (I)
(I) (I)
σ31 σ32 σ33
(I) −1 (I)
σ33
(II) (II) (II)
The unit normal for material II is −n = (0, 0, 1)T , and we have t(II) = (σ13 , σ23 , σ33 )T . Substitut-
ing these traction vectors into Eq. (4.23), we obtain
   
(I) (I) (I) (II) (II) (II)
σ13 , σ23 , σ33 T = σ13 , σ23 , σ33 T .

Since the stress tensors are symmetric, we obtain


⎛ (I) ⎞ ⎛ (II) ⎞
• • σ13 • • σ13
σ23 ⎠
⎝ • (I) ⎠ ⎝
• σ23 = • •
(II)
(I) (I) (I) (II) (II) (II)
σ13 σ23 σ33 σ13 σ23 σ33

as the boundary condition of the stress field at the interface2 . The black dots in these matrices
represent the stress components that are not constrained by the boundary condition.
If the interface is not fixed but lubricated, the shear stress vanishes on the interface. That is, we
have that the boundary condition for this case is
⎛ ⎞
 • • 0

r = ⎝• • 0 ⎠ . (4.24)
interface 0 0 σ33
2 This is a special case of the jump condition of physical properties at a surface. If the surface and the materials on its sides

move with different velocities, the impluse by the differential motions should be taken into account [38].
114 CHAPTER 4. PRINCIPAL STRESSES

Namely, only the normal stress σ33 is transmitted across the interface, but the shear stress is not.
Values of the components σ11 , σ22 and σ12 may jump across the interface. The σH axes are oriented
NNW–SSW far from the San Andreas Fault. However, those are nearly perpendicular to the fault
trace near the fault, suggesting a very low friction on the fault surface3 [277].

Vertical and horizontal stresses The stress tensor in Eq. (4.24) is diagonalized for the third
column and row. Therefore, σ33 is a principal stress and the third coordinate axis O-3 is the stress
axis corresponding to this principal stress. That is, one of the stress axes is perpendicular to a
lubricated surface. The other stress axes lie on the surface.
The solid Earth is covered by the atmosphere or seawater. These fluids have much less viscosity
than rocks. We regard the surface of the Earth as a lubricated interface which is covered by a fluid.
Therefore, Eq. (4.24) holds at the surface of the Earth. The solid Earth is exerted a shear stress
by winds and ocean bottom currents, but the shear stress is much less than tectonic stress. For this
reason, we neglect those drag forces in the discussion of tectonics. The result is that one of the stress
axes is perpendicular to the surface at the shallow levels in the Earth. In addition, if we assume that
the Earth has a level surface, the stress axis is largely vertical.
The component σ33 in Eq. (4.24) is, in these cases, the atmospheric or hydrostatic pressure at the
surface of the solid Earth. The atmospheric pressure is about 0.1 MPa, which is negligible compared
to tectonic stress. The pressure at the bottom of ocean is usually negligible. Therefore, we have the
stress boundary condition at the surface,
⎛ ⎞
 • • 0

r
 = ⎝• • 0⎠ .
surface 0 0 0

This is, of course, applicable to the surface of airless planets and satellites. However, hydrostatic
pressure at the bottom of ocean trenches may not be negligible, as the pressure reaches 100 MPa.
If one of the stress axes is vertical, stress in the shallow part of the Earth is expressed as
⎛ ⎞
σH 0 0
r=⎝0 σh 0 ⎠,
0 0 σv

where σH and σh are the maximum and minimum horizontal stresses, and σv is the vertical stress
that is usually approximated by the overburden stress (Eq. (3.29)). The σH - and σh -axes are the
corresponding horizontal principal axes. Some researchers use the notation, σHmax and σHmin , for the
horizontal axes.

3 Some researchers argue against this interpretation [207].


4.5. IN-SITU STRESS MEASUREMENTS 115

Figure 4.11: Deformed borehole determined by well logging at ∼3 km below the surface [180].

4.5 In-situ stress measurements

The present state of stress is estimated by several methods. Among them, two methods of in-situ
stress measurement in boreholes are introduced in this section4 .
If a deep borehole is filled with a fluid that has a smaller density than the surrounding rocks,
there is a great gap between the pressure in the fluid and in the rocks. The pressure difference
sometimes deforms or even collapses the borehole. Figure 4.11a shows the geometry of a borehole
no longer smooth or round. A zone along the wall of a well which has failed is called borehole
breakout. There are a few types of breakouts reflecting the state of stress in the surrounding rocks.
Detailed observation of the borehole geometry allows us to infer the principal orientations of stress
perpendicular to the hole.
Hydraulic fracture stress measurement is an active measurement of in-situ stress. Hydraulic
fracture or hydrofracturing is a process of breaking up the rocks immediately out of the wall of
a borehole under pressure by pumping water into an interval of the hole of which both ends are
sealed. Hydraulic fracture experiments are usually done to increase permeability in hydrocarbons
and geothermal reservoirs. The pressure opens the pre-existing discontinuity surface including joints
and bedding planes. The opened surfaces tend to be oriented parallel to the far field σH orientation.
Therefore, the borehole geometry again indicates the stress orientations. The experiment not only
determine the orientations but also the stress magnitudes.

4 See [54] for further reading.


116 CHAPTER 4. PRINCIPAL STRESSES

Figure 4.12: A basaltic dike intruding Oligocene basaltic lavas and tuffs cropping out on the Japan
Sea coast of Northeast Japan. Dotted lines outline the dike. Note that the dike expands its width at
the level of the tuff layer. The dike looks like the head of a cobra. A close-up photograph of the
lower left part of the cliff (indicated by angles) is shown in 4.13.

4.6 Dikes as paleostress indicators


Dikes, tabular intrusions of magma, are the results of the natural analog of hydraulic fracturing.
Magma pushed its way through the country rock via its pressure for intrusion. This is demonstrated
by the dike in the outcrop shown in Fig. 4.12. The dike vertically penetrates basaltic volcanic rocks
and tuffs, and the width of the dike increases abruptly at the horizon of a tuff. Figure 4.13 is a close-
up photograph of the tuff to the left of the dike. The tuff and overlying lava are intruded by an old
dike that is cut by a fault parallel to the bedding in the tuff, and the upper block is transported away
from the young dike shown in Fig. 4.12. The offset of the old dike is as large as the expansion of
the young dike. These observations indicate that the young dike pushed its country rocks when it
intruded.
4.6. DIKES AS PALEOSTRESS INDICATORS 117

Figure 4.13: A dike (D) broken into three parts by bedding parallel faults (solid lines) in the tuff
layer neighboring the dike shown in Fig. 4.12.

The Mohr diagram in Fig. 4.14(a) shows the admissible orientations of dike intrusion relative to
the principal axes of stress in the country rock. Consider a rock mass in which magma is intruding
with a pressure pm , which has pre-existing cracks with various attitudes. The hatched area is bounded
by the vertical line σN = pm , and the normal stress on the crack is smaller than the magma pressure
in the area. Magma pressure has to overcome the normal stress for intrusion. The state of stress
in the country rock is illustrated by the Mohr’s circles in Fig. 4.14(a), and the dark gray portion
of the Mohr diagram shows the condition under which the magma pressure overcomes the normal
stress. This portion designates the traction on the surfaces whose orientations are shown by the dark
gray portion on the sphere in Fig. 4.14(b). Magma can intrude cracks with orientations that are
indicated by the dark gray part of the sphere. As the magma pressure increases, if pm is smaller than
σ3 , no magma intrusion occurs. When the pressure is slightly above the minimum principal stress
(pm  σ3 ), magmas can only intrude cracks nearly perpendicular to the σ3 -axis. As the pressure
increases, the admissible orientations of dike intrusion expands. Finally, magma with pm > σ1
intrudes any crack, resulting in a network of intrusions.
The above argument demonstrate that a dike does not necessarily indicate the orientation of stress
118 CHAPTER 4. PRINCIPAL STRESSES

Figure 4.14: (a) Mohr diagram showing the magnitude of normal stress on surfaces. (b) Diagram
showing the attitude of the surfaces relative to the stress axes of country rocks.

axes when it intruded. If the magma pressure was appropriate, that is, slightly greater than σ3 , dikes
are paleostress indicators perpendicular to the σ3 -axis5 . If a dike swarm, parallel dikes, occur in a
large regional area, they are evidence for paleo σ3 orientation.
One good point of dike swarms in determining paleostress is that the age of the paleostress is
known from the radiometric age of the dikes. However, it should be determined from other types of
data whether a horizontally extensional or compressional stress field affected the area.
Magma pressure may decrease with the distance of a magma chamber. Dikes accompanied by
the Otoge cauldron shown in Fig. 4.15 illustrate this idea. The magmatism occurred at ∼15 Ma.
NNE–SSW to N–S trending dikes penetrates a Mesozoic basement in this area, but there are ring
dikes in the vicinity of the cauldron. The dike swarm is perpendicular to the present σH orientation,
but is parallel to the Median Tectonic Line, which is the most significant fault in Southwest Japan.
The line is parallel to the arc, and juxtaposes the high T/low P metamorphic belt and the low T/high P
belt. There are cauldrons with the same age along the Median Tectonic Line, and one other cauldron
in the central part of the arc has a dike swarm parallel to the line. The swarm was associated with
normal faulting, therefore, the dikes indicate an arc perpendicular extensional stress field at ∼15 Ma
[102, 108].

4.7 First-order regional stress field


The state of stress is measured through out the world using various methods, and the gross pattern
of the present stress field in the shallow levels in the Earth has been understood [279]. Figure 4.16
shows the results published in 1992, indicating several important points.
5 It was attempted to estimate magma pressure and intermediate stress from the variation of dike orientations in [14].
4.7. FIRST-ORDER REGIONAL STRESS FIELD 119

Figure 4.15: (a) Dike swarm and (b) ring dikes accompanied by the Otoge cauldron in Central Japan
[235]. The cauldron is the result of extensive volcanism in the Southwest Japan arc at ∼15 Ma. There
are too many dikes in this area so that their density, i.e., number per a length of 1 km across the trend
of the dikes, is shown. Note the difference of scale.

First of all, it was shown that there are vast regions where the principal orientations of stress are
roughly uniform. This was doubted because there are a variety of stress sources in the lithosphere
including topographic and lithological variations. Lithosphere thickness is a reference to argue how
the stress field is uniform. The thickness of the mechanically strong part in the lithosphere has spatial
and temporal variations, but is several tens of kilometers. There are regions where σH orientation
is uniform over the dimension 20–200 times greater than the thickness of the lithosphere beneath
the regions. The central to northwestern North America is one of them, where σH axis is oriented
ENE–WSW. In the Far East, the axis is oriented E–W between Japan and northern China.
The first-order stress field in each tectonic plate is believed to be controlled by collisional forces
at plate boundaries. If this is the case, the intraplate stress field is determined by the forces and the
shape of the boundary. In the northern Indo-Australia plate, σH -axis is parallel to the velocity of the
plate relative to the Eurasia plate. The Indo-Australia plate is bordered by a mid-oceanic ridge. The
axis is oriented perpendicular to the boundary, due to the ridge push force.
The second point is that the stress gradients within a plate are smaller than expected. The small
gradients suggest that the viscous coupling of the lithosphere and the underlying mantle is weak. If
the plate had been fixed upon the stationary or horizontally travelling deep mantle, plate boundary
forces caused by differential movement of neighboring plates would have resulted in some gradients
120 CHAPTER 4. PRINCIPAL STRESSES

Figure 4.16: σH orientations in the shallow levels of the crust [279]. Altitude and ocean depth are
indicated by a gray scale.

in the intra plate stress field. The observed first-order stress field is not consistent with absolute
motion, which also demonstrates little coupling.

The third point is that plateaus coincide with the region of extensional tectonics. The Basin and
Range Province in the western United States is an example where the surface has an altitude of ∼2
km and the vigorous extensional tectonics has been going on since the late Tertiary. However, it
should be noted that topographic height is not the only source of intra plate stress. This is demon-
strated by the existence of continental rifts such as the Suez, where rifting occurs near and under the
sea level.
4.8. EXERCISES 121

Figure 4.17: Mohr diagram showing a triaxial stress. The points P, Q and R designate the centers of
Mohr’s circles.

4.8 Exercises
4.1 Explain mathematically why the admissible combinations of normal and shear stresses are
indicated by Mohr’s stress diagram. Show the correspondence of the points with the labels of upper-
and lower-case letters in Fig. 4.6.

4.2 A stress state is designated by the Mohr diagram in Fig. 4.17. (a) What are the orientations of
the planes corresponding to the closed circles A, B, C and D in the diagram? (2) Which is the point
in the diagram corresponding to an octahedral plane?

4.3 Suppose a hot diapir with a diameter of L = 1 km ascending through the crust. The diapir
loses its heat during the ascent, so that it has to rise rapidly to cause contact metamorphism in the
country rocks when it reaches a shallow level of the crust. Estimate the lower bound of the ascending
velocity of the diapir for the metamorphism to occur. Use the thermal diffusivity κ = 10−6 m2 s−1
and a distance of 10 km for the diapir to ascend.
Chapter 5

Constitutive Equations

We have studied stress and strain in the first part of this book. This is the introductory
chapter for the second part where stress and strain are related to constructing models for
tectonics. This chapter is the introduction to the stress-strain relationships of materials
used in the following chapters.

5.1 Stress-strain diagram


Tectonic deformation proceeds under high temperature and high pressure at depths. Laboratory
experiments have revealed the mechanical properties of rocks under those conditions [176, 247].
One popular type of apparatus loads a cylindrical rock sample with axial stresses, σ1 > σ2 = σ3 ,
to observe the stress-strain relation of the sample. The maximum stress, σ1 , is loaded along the
longitudinal axis of the sample. The load is laterally isotropic, σ2 = σ3 , which is called confining
pressure. The apparatus controls the longitudinal strain ε and confining pressure to measure the
change of longitudinal stress σ. Figure 5.1 shows the stress-strain diagrams of a hypothetical material
obtained using such apparatus. Given a confining pressure, the longitudinal strain ε and stress σ are
the input and output variables, respectively.

Basic Concepts
Deformation is called elastic if it is reversible. While applied strain is small, stress is proportional
to strain. However, the stress-strain curve becomes less steep beyond a limit called the proportional
limit (Fig. 5.1). Samples have their own limit beyond which deformation becomes irreversible.
Once the load under the limit is removed, the sample recovers its original, initial shape. The limit
is known as elastic limit or yield point, and the stress at which the sample yields is known as yield
stress, σY . If we use the convention that tension is positive stress, the yield stress is expressed as SY .
Proportional and elastic limits are indistinguishable for most rocks. Linear elasticity is the elastic

125
126 CHAPTER 5. CONSTITUTIVE EQUATIONS

Figure 5.1: Stress-strain curve of a hypothetical material. Experimental apparatus controls the in-
crease of longitudinal strain ε of a cylindrical sample to observe longitudinal stress σ at the ends
of the sample. For small strains, the sample behaves as an elastic material of which deformation is
reversible. However, the deformation become irreversible at the elastic limit of the sample, beyond
which brittle or plastic deformation occurs. Brittle failure of the sample abruptly relieves the stress
so that the breakdown of the sample is indicated by a vertical line in the stress-strain diagram.

behavior where stress and strain is proportional to each other1 . If a load over the yield stress is
applied, the sample may break or deform like a piece of plastic clay. They are brittle and plastic
deformations. Once the load is removed, stress and strain become smaller along the line parallel to
the line OA, remaining a permanent strain. Irreversible deformation over the yield point is known
as inelastic deformation.
Overburden causes a high confining pressure for a rock mass at depth so that broken pieces
slide on their interface rather, i.e., faulting is the important consequence of brittle deformation for
tectonics.

Simplification
Actual materials show a wide variation in behavior when subjected to loading. Their stress-strain
diagrams are affected by various factors including rock type, confining pressure, tempearture, strain
rate, chemical atmosphere, etc. Real materials respond in an extremely complicated fashion under
various loadings. Rocks tend to be plastic rather than brittle material for increasing temperature, as is
the case of a taffy. Deformation may dissipate kinetic energy to increase temperature that decreases
the strength of the deforming material. Increasing temperature may cause phase transition that may
affect the mechanical properties, and may cause an additional strain by thermal expansion. Thermal
1 Linear elastic body hehaves equally for both compressional and tensile stresses. However, actual rocks do not. Rocks
have numerous cracks that decrease the rigidity of the rocks. Cracks open in tensile stresses and close in compression.
Laboratory experiments show that G in extension is in some cases tens of % smaller than that in compression [91]. The effect
of this asymmetry has been found to a mappable scale [178]. However, we neglect the nonlinearity in the following chapters.
5.1. STRESS-STRAIN DIAGRAM 127

Figure 5.2: Stress-strain curve and simple mechanical model for ideal materials. Displacement of
the mass, M, and force, F , depict the strain and stress, respectively. The springs represent elasticity.
Slidng on the rough surface indicates plasticity. Yield stress is represented by σY .

and mechanical phenomena are often coupled. Incremental deformation alters the internal structure
of a rock, i.e., the pile-up of microfractures, which in turn affects the strength of the rock against the
incremental deformation. A rock mass may consist of various types of rocks, each of which has its
own behavior.
Even a single crystal has its own complexity: its strength depends on direction with respect
to crystal axes. A material is called anistropic if some property depends on direction. Isotropic
materials show no such dependency. Although a rock mass is composed of various kinds of minerals
with anisotropy, most rocks have constituent minerals with random crysatalographic orientations,
vanishing anisotropy as a whole. Accordingly, details are often neglected but their average properties
are important for tectonics.
It is quite difficult to take into account all the observed phenomena under various conditions.
Instead, we need to define certain ideal materials such as ideal elastic solids. Such idealized materials
are useful in that they portray reasonably well over a definite range of loads and temperatures the
behavior of rocks. It is permitted to neglect the interaction of mechanical and thermal processes in
many situations.
Theories of tectonics utilize the phenomenological models of idealized materials. The stress-
strain diagrams of common, idealized materials are shown in Fig. 5.2. A linear elastic body shows
a linear stress-strain relation. A rigid-perfectly-plastic body has no elastic behavior. The adverb
“perfectly” carries the property that stress is constant at σY for incremental strain. This model may
represent not only plastic behavior but also a frictional sliding of faults.
128 CHAPTER 5. CONSTITUTIVE EQUATIONS

5.2 Representation theorem


A constitutive equation indicates a mathematical relationship among the statical, kinematic, and
thermal variables, which will describe the behavior of the material when subjected to applied me-
chanical or thermal loads. The equation relates stress S with strain F for statical equillibria or with L
for kinematic problems. The stress caused by the deformation F or rate of deformation L is calculated
by the equation.
Whatever the material in question is, its constitutive equation must follow the principle of ma-
terial frame-indefference if the stress in the material is caused by purely mechanical phenomena
[132, 244]. Namely, constitutive equations must be invariant under changes in the frame of refer-
ence. Namely, two observers, even if in relative motion with respect to each other, observe the same
stress in the body. This is also called the principle of material objectivity. The principle provides a
tight constraint to mathematical formulation of the constitutive equations.
The change of reference is given by

x = o(t) + Q(t) · x, (5.1)

where o(t) is a vector and Q(t) is an orthogonal tensor representing the translation and rotation of
the coordinate system, respectively. Since the stress tensor S defines a linear transformation (Eq.
(3.10)) between the two vectors N and t(N ), the tensorial representations of the stress for the two
reference systems are related to each other by the equation S = Q · S · QT (Exercise 5.1). However,
the deformation gradient F is transformed in a different way. To show this, suppose that the frames
coincide at time t0 so that ξ = ξ  at t = t0 . Note that the spatial coordinates x and x are dependent
on time t, but the material coordinates ξ and ξ  are not. From Eq. (5.1) we obtain
 
dx = Q(t) · dx = Q(t) · F · dξ = Q(t) · F · dξ. (5.2)

In this case, dξ = dξ  , so that dx = F dξ  = F dξ. Combining Eq. (5.2), we have



F − Q(t) · F · dξ = 0

for any ξ. Therefore, we obtain


F = Q(t) · F, (5.3)

showing that the deformation gradient tensor transforms like a vector (Eq. (C.11)).
 
Now let S = F F be a constitutive equation, where F ( ) represents a tensor-valued function of
a second-order tensor. Stress tensor S is transformed by the equation S = Q · S · QT , and Eq. (5.3)
transforms the deformation gradient. The function F indicates the material property of a subject
material, so that the function does not depend on the choice of reference frames. Therefore, the
 
function also applies to the primed tensors, that is, S = F F . Accordingly, we have
   
Q · F F · QT = F Q · F .
5.3. EXERCISES 129

We are able to constrain the equation further for isotropic materials. The rotation of reference
   
frames should not affect the function F for such materials, so that we have F F = F Q · F . As
the deformation gradient can always be decomposed as F = R · U = V · R, we have
     
F F = F R · U = F (Q · R) · U .

In this equation, (Q · R) is an orthogonal tensor, therefore, the function has only the argument U or
V. The constitutive equation of isotropic material is, therefore,
 
S=F V . (5.4)

If an isotropic material is extended, tensile stress may be induced in the extensional direction,
that is, the principal axes of V and S may be parallel to each other (Exercise 5.2). Therefore, the
function F does not vary principal axes, so that the Cayley-Hamilton theorem applies to this case.
Here, Eq. (C.46) works as the recursive relation to generate Vn from Vn−1 , Vn−2 and Vn−3 , so that Vn
can be expressed by the linear combination of V2 , V and I. Namely,

V3 = VI V2 − VII V + VIII I,
V4 = VI V3 − VII V2 + VIII V = (VI2 − VII )V2 + (VIII − VI VII )V + VI VIII I

and so on, where VI , VII and VIII are the basic invariants of V. The function F is assumed not to
change eigenvectors, so that Eq. (C.27) holds for V.
The Taylor series of V have terms of Vn , which can be rewritten with V2 , V and I. Thus

S = F (V) = φ0 I + φ1 V + φ2 V2 , (5.5)

where the coefficients φ1 , φ2 and φ3 are the scalar-valued functions of the basic invariants2 . The
concrete forms of φ1 , φ2 and φ3 depend on F. We have seen that isotropic materials have the simple
constitutive equation (Eq. (5.5)). This is known as the representation theorem.

5.3 Exercises
5.1 Demonstrate that the stress tensor is transformed as S = Q · S · QT .

5.2 Consider what if an anisotropic material undergoes uniaxial compression. Are the principal
strain axes parallel to the principal stress axes?

2 Equation (5.5) seems to be a quadratic equation of V, but the coefficients depend on the higher-order terms of the
invariants.
Chapter 6

Faulting and Brittle Strength

In this chapter, faulting is firstly considered as brittle failure. It is shown that tectonic
stress is limited in its magnitude by the brittle strength of faults in the shallow levels in
the lithosphere.

6.1 Primary fractures


Faulting generates earthquakes that cause serious natural disasters so that great efforts have been
made to understand it. Faults appear as shear fracturing in intact rock. In addition, faulting occurs
along a plane of pre-existing weakness such as fractures and bedding planes. It is known that the
Coulomb-Naiver criterion of brittle failure provides a basis for simple models for tectonic faulting.
The criterion is a yield condition for intact rocks.
The results of experimental rock mechanics with triaxial testing apparatus were synthesized to
propose the criterion. If the orientation of a fracture plane relative to σ1 axis is measured with the
axial load and confining pressure when fracturing occurred, the normal and shear stresses exerted
on the plane just before fracturing can be calculated. Axial stress σ1 > σ2 = σ3 is applied to
the test sample, therefore, Eqs. (4.21) and (4.22) hold. The axial load and confining pressure are
known variables and are equal to σ1 and σ2 , respectively. Thus, a Mohr circle can be drawn for
each experiment. The fracture plane angle θ is the angle between the σ1 -axis and the fracture plane.
This angle is equal to the orientation perpendicular to the fracture plane from σ3 axis, so that the
normal and shear stresses are indicated by the point that makes an angle of 2θ from (σ2 , 0) along
the Mohr circle (Fig. 6.1). The point represents the critical condition for shear fracturing. A series
of experiments conducted for a specific type of rock with various confining pressures provide a
series of such critical points in the Mohr diagram. Connecting those points, we obtain the failure
envelope or Mohr envelope for the sample material. The envelope is symmetric with respect to
the horizontal axis of the Mohr diagram, because the specimens and the applied stress have axial
symmetry. Accordingly, the envelope is expressed by a one-valued function σS = f (σN ). Failure
takes place when the Mohr circle expands with increasing differential stress so that it is just tangent

131
132 CHAPTER 6. FAULTING AND BRITTLE STRENGTH

Figure 6.1: Schematic diagram showing the normal and shear stresses at which the cylindrical spec-
imens of the same rock are fractured by axial stresses (σ3 = σ2 < σ1 ).

to the envelope. According to experiments on rock failure, the shear stress needed to produce failure
increases as the confining pressure increases. On the other hand, rocks cannot bear very large tensile
stresses. Therefore, the envelope has a V-shape, as shown in Fig. 6.1.
The envelope divides the O-σN σS plane into two regions. The region on the right side of the
envelope represents the combinations of normal and shear stresses that the type of rocks, for which
the envelope was drawn, can support the stresses and does not fail. The rocks cannot support the
stresses designated by the other region. The envelope therefore represents the fracture strength of
the rocks.
Experiments shows that the slope of the envelope decreases for large confining pressures, and
eventually vanishes. Then, a rock yields to the applied stress. Figure 6.2 shows the brittle strength
of various types of rocks. Carbonate rocks and halite have lower yield strength than plutonic rocks.
Any solid material breaks under very large uniaxial tension (σ1 = σ2 = 0 and σ3 < 0). Hence,
the Mohr envelopes above and under the σN -axis are connected on the horizontal axis on the right of
the σS -axis. The normal stress designated by the intercept is called tensile strength, and is designated
6.1. PRIMARY FRACTURES 133

Figure 6.2: Brittle strength of various types of rocks at room temperature [125]. The strength pre-
sented in Eq. (6.7) is also shown. Note that the vertical axes of the five panels have different scales.

by the symbol σ T . When the uniaxial stress with σN < σ T , tension fracture occurs1 . The Mohr circle
is tangent to the envelope at the intercept (6.1). That is, a specimen tends to be split by a surface
perpendicular to the extensional orientation.
Longitudinal splitting tends to occur under conditions of uniaxial compression (σ1 > 0 and
1 Note that tensional stress is indicated by σN < 0.
134 CHAPTER 6. FAULTING AND BRITTLE STRENGTH

Figure 6.3: Compressive and tensile strengths of intact rocks at room temperature [116, 125]. Bar,
minimum and maximum values; Box, ±25% intervals about average.

σ2 = σ3 = 0). The compressive strength is the minimum σ1 to cause brittle failure for the conditions
and is designated by σ C . Figure 6.3 shows the tensile and compressive strengths of various types of
rocks.
The longitudinal splitting causes expansion perpendicular to the compressive axis. Hence, posi-
tive confining pressures suppress the splitting, resulting in the formation of shear fractures.
Rocks have numberless micro cracks with various orientations. The growth and coalescence of
the cracks control macroscopic failure. Under conditions of uniaxial tension, the applied stress builds
up the stresses immediately adjacent to cracks (Fig. 6.4(a)), i.e., stress concentration occurs at the
tips of those cracks which are grown by the concentrated stresses. The crack propagation enhances
the concentration. This feedback cycle is formed most favorably from the cracks perpendicular to
the applied tensile stress, resulting in tension fracture.
Shear fracture is a more complicated phenomenon [118]. Frictional sliding on closed surfaces
oblique to the σ1 -axis gives rise to crack propagation at the tips to the σ1 orientation (Fig. 6.4(b)).
Unlike the formation of tensile fractures, the crack propagation does not directly make a feedback
cycle but mitigates the stress concentration at the tips. Therefore, compressive strength is much
greater than tensile stress.
The Griffith theory of fracture formulates the behavior of those micro cracks as the equation for
failure envelope:
σS2
σN = − |σ T |. (6.1)
4|σ T |
This is a parabola with the apex at the point (σ T , 0) and (0, ±2σ T ). The failure envelopes determined
by triaxial deformation tests2 are roughly similar to the parabola (Fig. 6.2).

2 Triaxial testing apparatus applies axial stresses to specimens. The Griffith-Murrell failure criterion, (τ oct ) 2 = |8σ σ |, is
T 0
proposed for conditions of triaxial stress [91].
6.2. COULOMB-NAVIER CRITERION 135

Figure 6.4: (a) Stress concentration by tensile stress. (b) Formation of shear fracture by confined
compression.

6.2 Coulomb-Navier criterion


If σS = ±f (σN ) represents the failure envelope, its linear approximation to this equation,

|σS | = τ0 + μσN , (6.2)

is often used, i.e., brittle strength is linearly related with normal stress (Fig. 6.5(a)). This is known
as the Coulomb-Navier criterion3 . In relation to the coefficient of friction, μ is called the coefficient
of internal friction. The angle φ that satisfies

μ = tan φ (6.3)

is known as the angle of internal friction and is designated by ∠CAB in Fig. 6.5(a). The intercept
τ0 is known as cohesion. These are material constants depending on lithology and other factors
including temperature and strain rate.
This criterion is often used to link faulting to stress. As differential stress increases, the Mohr
circle between the least and greatest principal stresses expand. When the circle eventually touches
the failure envelope, a shear fracture is formed, i.e., a fault is generated. Note that the magnitude of
the intermediate principal stress has nothing to do with the failure criterion. However, the attitude of
the fault plane predicted by this criterion is parallel to the σ2 -axis, because the points on the Mohr
circle that links the σ1 and σ3 represent the surfaces that contain the σ2 -axis (Fig. 6.5(b)). Among
those surfaces, one in the pair represented by the points C and D in Fig. 6.5(a) is chosen to form a
fault. The two surfaces are called the conjugate shear planes or conjugate faults. The half line BC
or BC makes an angle 2θ with the half line BE (Fig. 6.5(a)). The point E designates the normal
stress σ3 . Accordingly, the point C or D indicates the angle θ between the σ3 -axis and the line
perpendicular to the fault surface in physical space. The latter equals the angle between the surface
and the σ1 -axis (Fig. 6.5(b)). The acute angle between the conjugate faults equals 2θ. The angle θ
is sometimes called the angle of shear.
3 Various names are used for this criterion, such as the Coulomb, Navier-Coulomb, and the Coulomb-Mohr yield criterion.
136 CHAPTER 6. FAULTING AND BRITTLE STRENGTH

Figure 6.5: (a) Mohr diagram showing the Coulomb-Navier criterion. (b) Diagram showing the
relationship among the stress axes, shear angle and the sense of sear. Note that the σ2 -axis is per-
pendicular to the page.

The point B is the center of the Mohr circle, so that ∠BCA = 90◦ . Therefore, we have

2θ + φ = 90◦ , (6.4)

where ∠ABC = 2θ. Rearranging this equation, we have μ = tan φ = tan(90◦ − 2θ) = cot 2θ. That
is, we obtain the relationship between the coefficient of internal friction and the shear angle,

tan 2θ = 1/μ. (6.5)

This equation shows that μ is inversely correlated with the angle of shear. The angle has a minimum
θ = 0 and a maximum θ = 45◦ for μ = ∞ and 0, respectively. Laboratory experiments show that θ
is roughly equal to 30◦ for various rock types.
The criterion is simple and, therefore, convenient. However, laboratory experiments using ap-
paratus that produces true triaxial stresses σ3 < σ2 < σ1 demonstrated that the magnitude of σ2
affects the brittle strength of rocks. The strength predicted by the criterion has errors of up to ∼20%
(§10.3). In addition, the attitude of fault surfaces coincide with the criterion only for limited cases
(§11.7).

Strength of ice The Coulomb-Navier criterion is used to describe the brittle strength not only of
rocks but also of ice. Many moons in the outer planets have thick icy crusts, so that they are called
icy satellites. Large icy satellites had tectonic activities that left tectonic features on the surface.
Ganymede is one of them (Fig. 3.6). Therefore, the strength of ice is important to understand the
activities. The surface temperature of the icy satellites is lower than ∼100 K. Laboratory experiments
have presented the values μ = 0.55 and τ0 = 1 MPa at low confining pressures σ3 < 500 MPa and
at temperatures between 77 and 115 K and pressures [172]. The compressive and tensile strength of
6.3. ANDERSON’S THEORY OF FAULTING 137

Figure 6.6: Types of faulting predicted by Anderson’s theory of faulting.

ice are estimated at 3.4 and –1.2 MPa, respectively, by Eq. (D.22). Both are one orders of magnitude
smaller than those of rocks (Fig. 6.3). That is, ice is weaker than rocks.

6.3 Anderson’s theory of faulting


One of stress axes stands upright near the surface if the surface is largely horizontal (§4.4). Com-
bining this and Coulomb-Navier criterion, we obtain an important relationship between the type of
faulting and stress. This is known as Anderson’s theory of faulting [4].
The criterion predicts that fault planes are parallel to σ2 -axis, and that the planes make angles
−1
1
2 tan (1/μ), which is less than 45◦ with the σ1 axis. Therefore, if the σ3 -axis is vertical, the
stress gives rise to reverse faulting with the dip angle of the faults less than 45◦ (Fig. 6.6(a)).
Namely, thrusts are formed. If the σ1 - or σ2 -axis is vertical, normal or strike-slip faults are formed,
respectively (Figs. 6.6(b) and (c)). The latter should have vertical fault planes.
Anderson’s theory of faulting is convenient for infering the state of stress from faults and vice
verse so that it is often used if no other evidence is available. However, it is now known that the
theory is not always correct. It cannot account for the abundance of oblique-slip faults in the field
[20]. In addition, the fault activity predicted by this theory results in a plane strain4 . However, strains
should be generally three-dimensional even if they are caused by brittle faulting (§11.7).
4 See Exercise 11.1 and its answer.
138 CHAPTER 6. FAULTING AND BRITTLE STRENGTH

Anderson’s scheme is currently used for the classification of stress regimes in the lithosphere5 ,
involving the normal fault regime (σv > σH ), reverse fault regime (σH > σh > σv ), and strike-slip
regime (σH > σv > σh ).

6.4 Frictional strength of fault


A rock mass has discontinuous surfaces such as joints, fault planes and bedding planes. Those
surfaces can be activated as faults by applied forces. Frictional sliding occurs on the planes. The
frictional resistance or frictional strength is represented by the shear stress on the surface, and is
given by
τf = μf σN , (6.6)

where σN is the normal stress acting on the surface and μf is the coefficient of friction. In the case
of shear stresses that are less than this strength, the surface is locked. When the applied shear stress
overcomes this, sliding occurs on the surface.
The frictional strength of rocks is important for the investigation of earthquakes and landslides
so that the coefficient of friction of various rock types are experimentally estimated in the range
0.6–1.0. The strength roughly obeys the piecewise linear relationship6 [33],
)
0.85σN (σN < 200 MPa)
μf = (6.7)
0.6σN + 50 MPa (0.2 < σN < 1700 MPa).

An overburden stress of 200 MPa corresponds to the depth at about 6 km. This relationship is
sometimes referred to as Byerlee’s law. The continuous function of σN ,

τf = σN0.94 (σN < 1700 MPa), (6.8)

is sometimes convenient [125]. The coefficient of the friction of ice is estimated at μf = 0.69 [19].
The coefficient of friction introduced above was determined by laboratory experiments. How-
ever, field observations suggest that great faults have low friction [197]. The σH orientations near the
San Andreas Fault is consistent with this view. Heat-flow measurements around this fault suggests
friction as low as μf < 0.2 [115].
Phyllosillicate minerals such as clay, serpentine, and mica show lower coefficients of friction.
Since fault surfaces have those minerals, their strengths are important. Montmorillonite, a clay
mineral stable at low temperatures, has μf ≈ 0.2 [158]. Fault gauges and clay sampled from the San
Andreas Fault have the values 0.15–0.55 [158]. It is a matter of debate how these materials lower
the frictional strength of faults. It is argued that clay minerals generally show strain hardening, so
that highly deformed fault clay is as hard as ordinary rocks. In addition, frictional sliding has similar
5 Theregimes are also known as the normal faulting, reverse faulting, and strike-slip faulting regimes of stress.
6 It
is reported that surface roughness affects the strength in the shallow crust where the confining pressure lower than
∼300 MPa.
6.5. EFFECT OF PORE FLUID PRESSURE 139

Figure 6.7: Force to move an air-cushion vehicle.

microscopic processes with brittle failure at high confining pressure. Therefore, clay minerals may
not lower the strength of the faults.
Frictional sliding and brittle failure have similar microscopic phenomena. Surfaces on which
frictional sliding occurs have roughness so that brittle failure or ductile deformation of bumps on
the surfaces are needed for the sliding. In other cases, bumps on one side of the fault plane should
overcome the normal stress to climb over the others on the other side of the plane. The normal stress
becomes large with increasing confining pressure so that the microscopic brittle failure and yielding
control the frictional strength at high confining pressures.

6.5 Effect of pore fluid pressure


Rocks at depth have pore water to a greater or lesser extent. Pore fluids have strong control over
the brittle and frictional strengths of rocks so that they are important when considering landslides,
earthquake generation, and the strength of the lithosphere.
Consider an air-cushion vehicle with the basal area of S and the weight M. The horizontal force
needed to overcome the basal friction and to move the vehicle is (μf Mg/S  )S  = μf Mg, where S  is
the area on which the vehicle rests on the ground (Fig. 6.7). The air pressure p in its skirt lowers this
force to μf (Mg − pS), S is the basal area of the skirt. The pS is the force due to the pressure p to
float the vehicle. When the pressure reaches Mg/S, the vehicle floats on the surface and eventually
moves.
The air pressure is comparable with pore fluid pressure7 pf to lower the frictional and brittle
strength of rocks. Without pore fluid, normal stress in a rock should be supported by a framework of
solid grains. Pore fluid pressure, if it exists, takes a part of the forces in the rock to lower the normal
stress. The decreased normal stress reduces the brittle and frictional strengths. Pore pressure does
not affect shear stress but lowers normal stresses so that this effect is described by the equation

re = r − pf I, (6.9)

where re is the effective stress. The frictional strength is reduced by the pore pressure as

τf = μf (σN − pf ), (6.10)
7 This is simply called pore pressure.
140 CHAPTER 6. FAULTING AND BRITTLE STRENGTH

Figure 6.8: Diagram showing the effect of pore-fluid pressure on brittle strength.

where (σN − pf ) is the normal stress due to re . It was experimentally confirmed [71] that the brittle
strength predicted by the Coulomb-Navier criterion reduces to

|σS | = τ0 + μ(σN − pf ). (6.11)

Instead of pore fluid pressure, the pore fluid pressure ratio

λf = pf /pL (6.12)

is often used to discuss the strength of the lithosphere.


Note that re is different from r only through the hydrostatic pressure pf I (Eq. (6.9)). Therefore,
the two tensors have common principal orientations (§C.5). The tensors have a common shear stress
acting on a particular plane. Therefore, the difference is graphically expressed by the leftward shift
of the Mohr circles (Fig. 6.8). If r is constant, increased pore fluid pressure shifts the circles
to traverse the failure envelope to activate a shear fracture. Consequently, pore pressure reduces
the brittle strength of rocks. Increased ground-water pressure through heavy rain can weaken the
frictional resistance of a pre-existing slip surface to cause a landslide. High pore pressure is essential
for the formation of an extensive decollement surface (§4.2).
There are several origins of high pore-fluid pressures. Dehydration associated with metamor-
phism can build up the pressure. High pore pressure zones are often met in hydrocarbon exploration
in sedimentary basins. As formations are deeply buried, increased overburden stress encourages de-
watering of the formations. However, the formations are sealed by an overlying impermeable layer
such as a claystone bed. The formations are hindered from dewatering, resulting in overpressured
formations. Without this type of cap rock, increased compaction impede dewatering (Fig. 6.9(a)).
There are other origins of overpressured zones [234]. Tectonic stress can control pore-fluid pressure
[70, 251].
An overpressured zone is defined as the zone where the pore-fluid pressure is higher than the
hydrostatic pressure of the same depth (Fig. 6.9(b)). The existence of overpressured formations is
crucial for petroleum industry, because pore-fluid pressure in a hydrocarbon reservoir affects efficient
drain-up of the resource. In addition, pore-pressure affects efficient borehole drilling, because high
6.6. REACTIVATION OF FAULT 141

Figure 6.9: Schematic pictures showing overpressured formation in a sedimentary basin. (a) Pore
fluid pressure pf help the framework of solid particles to support normal stresses. (b) Pore spaces
between sedimentary particles are linked to each other in the shallow part of the basin, but the
continuity is reduced by compaction. Therefore, overpressured formation appears at depths.

pore-pressure weakens the wall of a borehole. Therefore, pore-fluid pressure is routinely monitored
in hydrocarbon exploration. Overpressured formations appear at a few kilometers below the surface
or deeper [54].

6.6 Reactivation of fault


The Coulomb-Navier criterion predicts that the formation of a fault in an intact rock needs a differ-
ential stress as great as the cohesion τ0 . By contrast, frictional sliding is much easier, as the sliding
has τ0 = 0. Shear stress needed for frictional sliding is smaller than that for forming a shear frac-
ture in an intact rock. Therefore, pre-existing fractures are easily activated as faults, if they have
not been cemented. Accordingly, a fault can be repeatedly activated. Many geological faults have
displacements much larger than the representative displacement of one earthquake faulting, which
is less than ∼10 m. Therefore, they demonstrate the reactivation, and active faults that moved in the
Late Quaternary are dangerous.

Example

Old large faults are often reactivated in different stress regimes. Normal faults in a passive continen-
tal margin are sometimes reactivated as reverse or thrust faults when the ocean basin in front of the
margin is consumed. This is known as tectonic inversion or basin inversion. Figure 6.10 shows an
example. Old reverse faults are sometimes reactivated as normal faults.
The timing of inversion is constrained by geological structures and the ages of involved forma-
tions. Triassic to Jurassic rifting in the North Sea created normal faults before the earliest Cretaceous
142 CHAPTER 6. FAULTING AND BRITTLE STRENGTH

Figure 6.10: Tectonic inversion in the North Sea [13]. The upper panel shows the seismic profile in
the basin. The middle panel shows the line drawing for the structures interpreted from the profile.
The lower panel shows the inferred geological structure when the horizon A/C deposited. Significant
horizons are labeled as IUC = a reflecting horizon in the Upper Cretaceous, IC = a horizon in
the Upper Cretaceous Chalk Group, A/C = Aptian-Cenomanian (Lower and Upper Cretaceous)
boundary, BC = base Cretaceous, TR = Triassic, Z = top Zechstein Supergroups (Upper Permian),
R = top Rotliegend Group (Lower Permian). Note that normal faults that bounded wedge-shaped
half grabens (lower panel) were reactivated as reverse faults (middle panel). Arrows indicate onlap,
offlap or erosional truncation of sedimentary layers.

in this area, and some of these faults have been reactivated by compression since the Late Cretaceous.
The angular unconformity at the horizon B/C in the lower panel in Fig. 6.10 demonstrates the
termination of rifting before the middle of the Cretaceous. Arrows pointing to the horizon BC from
below show that a sedimentary sequence older than the horizon is tilted and truncated by erosion. The
formation between Z and BC shows lateral variations in thickness. The formation thickens toward
normal faults, indicating syndepositional normal faulting. The thickness of the Lower Cretaceous,
which lies between BC and A/C, is relatively constant, suggesting that the differential subsidence
accompanied by normal faulting had ceased before the accumulation of the Lower Cretaceous.
6.6. REACTIVATION OF FAULT 143

Figure 6.11: Reactivation condition of faults. (a) Two stress states with the same differential stress.
The fault orientations, allowed to move under the stresses, which are indicated by gray fans in
the Mohr circles, become narrower with increasing mean stress. (b) Since the vertical stress σv is
approximately equal to the overburden pressure, faults are more easily reactivated by the normal
faulting regime of stress than the reverse faulting one if the two regimes have the same differential
stress and the same overburden pressure.

The timing of inversion is known from the structure and age of the overlying strata. The Upper
Cretaceous formations make asymmetric folds with the same intervals as the normal faults. The for-
mations onlap the anticlines (arrows with number 1) so that the formations are thick in the synclines
and thin in the anticlines. These observations demonstrate that the folds grew with the accumulation
of the Late Cretaceous. Therefore, compressional tectonics began sometime between A/C and IUC.
The compressional tectonics had ceased at around the horizon IUC, because the thickness of the
strata younger than this horizon is hardly affected by the folds. The young formations accumulated
through the widespread uniform subsidence of this region.

Reactivation condition

There are conditions for the reactivation of a fault. Theis reactivation would be difficult for the fault
surfaces that have been firmly fixed by the lithification of fault materials8 . In addition, the new
state of stress should have favorable principal orientations for the faults. For example, fault surfaces
parallel to a principal plane of stress could not move because shear stress on the surface vanishes.
The state of stress controls the mobility of pre-existing fractures. Figure 6.11(a) shows that the
orientations of faults that can be activated by the same differential stress expand with decreasing σ3 ,
because the brittle and frictional strengths increase with σN . Therefore, the orientations are more
limited with increasing depth and pore pressure. Great faults appear to have low friction (§6.4). If
so, the admissible orientations are widen, also. Accordingly, great faults are easily reactivated.
More importantly, the normal fault regime of stress enables the activation or reactivation of faults
8 If the lithified material has elastic constants different from those of country rocks, stress concentration at the fault surface

would encourage reactivation.


144 CHAPTER 6. FAULTING AND BRITTLE STRENGTH

more easily than the reverse fault regime if the differential stress and the burial depth are the same.
This difference results again from the positive gradient of the strengths, μ > 0 and μf > 0 (Fig.
&z
6.11(b)). Vertical stress obeys the relationship, σv ≈ 0 ρ(z)gdz ≈ ρgz. In the last approximation,
the density is assumed to be constant, because common rocks9 have densities around 3×103 kg m−3 .
Therefore, σv is determined by depth. If a rock mass is in the normal faulting regime of stress, we
have σv = σ1 and σh = σ3 = σv − Δσ < σv . On the other hand, if it is in the reverse faulting regime,
we have σv = σ3 = σ1 − Δσ < σH and σH = σ1 . The Mohr circles of the former regime are nearer
to the strength line than those of the latter (Fig. 6.11(b)). Therefore, faulting occurs in a normal
faulting regime more easily than in a reverse faulting regime.

6.7 Depth dependence of brittle strength


Rocks support differential stresses less than their strength. When differential stress reaches rock
strength, brittle failure occurs to mitigate the stress. Consequently, the brittle strength determines
the maximum differential stress in the shallow part of the Earth, where brittle failure is the most
significant deformation mechanism. Let us derive the critical stress at depth z for normal and reverse
fault regimes from the friction law (Eq. (6.6)). Pre-existing fractures with various orientations are
assumed to exist so that the law is employed here. Since we use the frictional strength (Eq. (6.6))
that is independent from the magnitude of σ2 , only the σ1 and σ3 orientations are considered. The
coordinate system O-xz is defined as shown in Fig. 6.12(a).
In order to distinguish the regimes, we define Δ = σH − σv . The sign of this quantity indicates
one of the regimes (Figs. 6.12(b) and (c)). This is equivalent to differential stress for a reverse fault
regime. However, −Δ equals differential stress for a normal fault regime.
Secondly, we have a uniform density ρ and a level surface, so that stress axes are parallel to the
x- and z-axes. The principal stresses are

σz = ρgz and σx = σz + Δ. (6.13)

We investigate whether a pre-existing fracture surface dipping at an angle (π/2 − θ) is activated


as a fault (Fig. 6.12(a)). Since we are considering a two-dimensional problem, Eqs. (4.21) and
(4.22) apply to our problem.
In the case of the reverse fault regime, σ1 and σ2 in Eqs. (4.21) and (4.22) should be replaced by
σx and σz , i.e., we have the normal and shear stresses acting on the footwall
1 1 σx + σz Δ
σN = (σx + σz ) + (σx − σz ) cos 2θ = + cos 2θ (6.14)
2 2 2 2
1 Δ
σS = (σx − σz ) sin 2θ = sin 2θ. (6.15)
2 2
Note that the definition of θ in this problem is different from that in Fig. 4.8(a) in its sign. Therefore,
the right-hand side of Eq. (6.15) lacks the negative sign that is attached to Eq. (4.22).
9 Here, materials that are abundant and have low densities (e.g., water and rock salt) are neglected.
6.7. DEPTH DEPENDENCE OF BRITTLE STRENGTH 145

Figure 6.12: Schematic illustration to explain the activation condition of a pre-existing fracture
surface as a fault. (a) The dip of the surface is π/2 − θ. Unit vector n points inward to the footwall.
(b, c) The relationship of the sign of Δ with stress regimes.

In the case of the normal fault regime (Fig. 6.12(c)), we have the correspondence σ1 = σz and
σ2 = σx . Comparing Figs. 4.8(a) and 6.12(a), we replace θ by π/2 − θ in Eqs. (6.14) and (6.15).
By means of formulas cos(π − 2θ) = − cos 2θ and sin(π − 2θ) = sin 2θ, we obtain
1 1 1 Δ
σN = (σz + σx ) − (σz − σx ) cos 2θ = (σz + σx ) + cos 2θ
2 2 2 2
1 Δ
σS = (σz − σx ) sin 2θ = − sin 2θ.
2 2
Consequently, the results for the two cases are different only in the signs of the shear stress.
Therefore, combining Eqs. (6.13), the normal and shear stresses are expressed by the equations
Δ
σN = ρgz + (1 + cos 2θ), (6.16)
2
Δ
σS = ± sin 2θ, (6.17)
2
where the upper and lower signs in the right-hand side of Eq. (6.17) indicate the reverse and normal
fault regimes, respectively.
The critical condition for this surface is given by Eq. (6.10), i.e., σS = μf (σN − pf ). Substituting
Eqs. (6.16) and (6.17), we obtain
 
Δ Δ
± sin 2θ = μf ρgz + (1 + cos 2θ) − pf ,
2 2
Δ 
∴ ± sin 2θ − μf (cos 2θ + 1) = μf (ρgz − pf ).
2
Rearranging this equation, we obtain the critical condition
2μf (ρgz − pf )
Δ= . (6.18)
± sin 2θ − μf (cos 2θ + 1)
146 CHAPTER 6. FAULTING AND BRITTLE STRENGTH

Figure 6.13: Graphs of F (θ) and G(θ) for the reverse and normal fault regimes, respectively. Open
circles designate their peaks. The gray region indicates the domain of the function.

Hence, the critical stress increases with depth z, but pf can cancel this effect.
The continental crust has experienced a long deformation history that has created the planes
of weakness such as old faults, joints, metamorphic foliations, and bedding planes. Therefore, we
assume that there are pre-existing planes of weakness with various orientations. If so, the surfaces
with the most favorable orientations are activated as faults. The optimal orientation is given by
the minimum differential stress |Δ|. In order to determine the optimal orientation, we seek θ that
maximizes the denominator in the left-hand side of Eq. (6.18).
In the case of the reverse fault regime, the function F (θ) = sin 2θ − μf (cos 2θ + 1) should be
maximized for the domain 0 ≤ θ ≤ π/2. By way of the formula sin a cos b + cos a sin b = sin(a + b),
we have
⎛ ⎞

⎜ 1 μf ⎟
F (θ) = sin 2θ − μf cos 2θ − μf = 1 + μ2f ⎝  sin 2θ −  cos 2θ ⎠ − μf
1 + μf2
1 + μf
2


= 1 + μ2f sin(2θ + α) − μf , (6.19)

where
1 μf
cos α =  , sin α = −  , tan α = −μf , (−π/4 ≤ α ≤ 0).
1+ μ2f 1+ μ2f
Since the coefficient of friction ranges between 0 and 1, the phase angle α is in the range −π/4 <
α < 0. Equation (6.19) represents the sine curve with a period of π. Therefore, F (θ) have the
maximum 
F (θ) = 1 + μ2f − μf
within the range 0 ≤ θ ≤ π/4 (Fig. 6.13).
For the case of the normal fault regime, the function

G(θ) = sin 2θ + μf cos 2θ + μf = 1 + μ2f sin(2θ + α) + μf
6.8. JOINT 147

must be maximized, where


1 μf
cos α =  , sin α =  , tan α = μf (0 ≤ α ≤ π/4).
1+ μ2f 1+ μ2f

The function G(θ) has the maximum



G(θ) = 1 + μ2f + μf

within the range 0 ≤ θ ≤ π/4 (Fig. 6.13).


Substituting the maximum denominator into Eq. (6.18), we obtain the critical condition
2μf (ρgz − pf )
Δ=  , (6.20)
1 + μ2f ∓ μf

where the upper and lower signs in the denominator of the right-hand side indicate the reverse and
normal fault regimes, respectively. Using the pore-fluid pressure ratio (Eq. (6.12)), this is rewritten
as (Fig. 6.14) ⎡ ⎤
⎢ 2μf ⎥
Δ = ± ⎣ ⎦ (1 − λf ) ρgz. (6.21)
1 + μ2f ∓ μf
The double signs correspond to the reverse (upper sign) and normal (lower sign) fault regimes.
Equation (6.21) designates that the critical differential stress is proportional to depth z.
Figure 6.14(a) shows a graph for the depth-brittle strength relationship given by Eq. (6.21).
Withing this context, the absolute value of the critical Δ is the strength. The figure shows that the
strength is smaller in the normal fault regime than in the reverse fault regime. In addition, pore
pressure has a strong control over the brittle strength. This difference results from the fact illustrated
in Fig. 6.11(b). The dependence of the strength on μf is indicated by the factor within the brackets
in Eq. (6.21). Figure 6.14(b) shows that the brittle strength has a larger μf dependence in the reverse
fault regime than in normal fault regime. It should be noted that the differential stresses shown in
Fig. 6.14(a) are the limits, and those may be much smaller than the critical values in tectonically
quiet regions.
Large faults may have a small coefficient of friction. In-situ stress measurement is a clue to
estimate the frictional resistance of large faults. For this purpose, stress states were measured at
various depths in tectonically active regions where the stress state is at or near the critical condition,
therefore, Eq. (6.21) is applicable. The results indicate that the effective coefficient of friction is
about the same as that determined in the laboratory [3, 21, 278].

6.8 Joint
Fractures of rocks that show little displacement parallel to their surfaces but with very small openings
normal to them are called joints. They are basically extension fractures. The lack of displacement
148 CHAPTER 6. FAULTING AND BRITTLE STRENGTH

Figure 6.14: (a) Brittle strength (Δ) as a function of depth z. Parameters, ρ = 2.7 × 103 kg m−3 ,
g = 10 ms−2 and μf = 0.75 are used. (b) Dependence of the brittle strength on the pore-fluid
pressure ratio.

for a crack is verified by the feather-like pattern of subtle ridges and grooves on the fracture surface
and the lack of fault striation (Fig. 11.2). Most ourcrops contain one or more sets of joints. Each
joint set has a characteristic orientation and spacing. The photograph in Fig. 6.15 shows two joint
sets in Lower Miocene sedimentary layers. One set has a spacing of several centimeters and the
other has that of a few tens of centimeters. Joint sets in the same rock mass is called a joint system.
The abundance of joints designates that there are numberless cracks in the shallow part of the crust,
so that they are important for the transfer of formation fluids and for the stability of a rock mass
containing fractures.
The relative age of joint sets is determined at outcrops by the abutting relationship, i.e., an old
joint surface acts as a free surface so that younger crack propagation is stopped at the surface10 .
Joints have multiple origins including the contraction of rocks due to cooling, unloading of over-
lying rocks by erosion, or tectonism11 . Joints can form in sediments before they have consolidated
into rock. For example, extension fracturing of unconsolidated sediment in the presense of high
pore pressure can result in the formation of clastic dikes. As evidence, Fig. 6.16 shows a clastic
dike which penetrates a tuff breccia. The dike is tapped from the thin pumice tuff layer beneath the
breccia. The crack is filled with the pumice tuff which covers the laminated siliceous tuff and thick
fine tuff. The substrata of the breccia probably deposited in a swamp. The negative pressure pro-
duced by the extension fracturing in the crack with respect to the overpressured substrata not only

10 See [248, §3.4] for further reading.


11 See [54] for further reading about the mechanics of joint formation.
6.8. JOINT 149

Figure 6.15: Joint system in alternating beds of sand- and siltstones, the Lower Miocene Oya For-
mation, northwestern Kyushu, Japan. Bedding planes are gently inclined to this side. There are two
sets of joints here; fractures trending toward the sea with smaller spaces, and those with larger spaces
perpendicular to the grip of the hammer.

sucked up the pumice tuff but also broke the laminated layer into slices and gathered them under the
dike. Consequently, the slices made an antiformal stuck, a type of duplex structure. Formation of
clastic dikes are comparable with hydraulic fracturing. They were formed in the shallow levels of
sedimentary basins.
Sediments gradually gain stiffness during burial. Consider a horizontally extensive layer of sed-
iments that is assumed to be an elastic plate. Lithification of the sediments may be modeled by
increasing the tensile strength ST = −σT . The incremental depth of a horizontally lying elastic slab
causes a normal fault regime of stress within the slab, i.e., horizontally extensional stress (§7.2).
Therefore, clastic dikes are formed vertically. However, stress levels there are very small, so that
multiple factors give rise to disturbance in the stress field. Consequently, clastic dikes are not so
parallel as swarms of igneous dikes. By contrast, joints surprisingly make parallel sets.
Figure 6.17 shows the orientation of joints in the Upper Miocene formations in the Boso penin-
sula of central Japan. The joints are nearly vertical and largely trend NE–SW. This orientation is
150 CHAPTER 6. FAULTING AND BRITTLE STRENGTH

Figure 6.16: Photographs showing the source of a clastic dike. Upper Oligocene Shioseno-misaki
Conglomerate, northwestern Honshu, Japan. (a) A hammer is placed at the root of the dike. (b)
Close-up of the root. The slices of a thin layer of laminated siliceous tuff makes an antiformal stack,
a type of duplex structure, at the base of the dike. The coin is about 2 cm in diameter.
6.9. EXERCISES 151

Figure 6.17: Orientation of joints in the Upper Miocene Amatsu and Anno Formations, Boso penin-
sula, central Japan. Each solid square indicates the pole to a joint surface that the correction for the
dip of strata has been carried out. The open circle designates the orientation of fold hinge lines in
this area. Arrows indicate the trend of maximum shortening perpendicular to the hinge line.

perpendicular to the hinge lines12 of folds in this area. The folds suggest Late Miocene to Early
Quaternary shortening in NE-SW trend. Therefore, the majority of the joint were perpendicular to
the σh orientation at that time, suggesting the coeval formation of the joints and folds under NE–SW
compression. The peninsula is located near a triple-trench-type plate junction, and has experienced
polyphase tectonics since the Tertiary. This joint set was probably reactivated as normal and oblique-
normal faults in the mid Quaternary (§11.5).

6.9 Exercises
6.1 Eliminate σN , σS and τ0 from Eq. (6.2), and rewrite the equation to relate σ1 , σ3 , σ T and σ C .

6.2 Pore fluid pressure is usually limited by overburden stress pL . Therefore, intact rocks are stable
if the differential stress Δσ is smaller than their tensile strength |σT | regardless of pore pressure (Fig.
6.18). Common rocks have tensile strengths of the order of 100 –101 MPa. Converting these values
with a constant density of 2.8 × 103 kg m−3 , we have a depth of 102 –103 m. Therefore, intact rocks
shallower than this can support their overburden. However, sediments before consolidation have so
small effective values of tensile strength that they do not allow extension fracturing.
By means of the failure envelope predicted by the Griffith theory (Eq. (6.1)), determine the
maximum depth to which cracks are allowed to propagate in unconsolidated sediments to form
12 The hinge line of a fold is the line in the folded surface along which the curvature is a maximum.
152 CHAPTER 6. FAULTING AND BRITTLE STRENGTH

Figure 6.18: Stability of intact rocks. If pore fluid pressure pf equals overburden stress pL , the Mohr
circle of the effective stress passes the the origin O of this diagram.

joints. Estimate the order of magnitude for this depth using the values ρ = 2.3 × 103 kg m−3 and
|σT | = 0.1 MPa = 1 bar.
Chapter 7

Elasticity

Elastic strain of rocks is very small, so that it is usually difficult to detect the strain
from geologic structures. However, it is possible in some cases, allowing us to infer
mechanical properties of rocks at depths. A simple model, linear elasticity, is introduced
in this chapter to discuss the states of stress. It should be noted that models in this
chapter are ideal ones, because elastic strain energy tends to be released, over millions
of years, in shallow levels in the Earth by weathering, diagenesis, fracturing, and plastic
flow.

7.1 Linear elasticity


7.1.1 Constitutive Equation
Let us derive the constitutive equation for isotropic elastic materials based on the representation
theorem (Eq. (5.5)). The equation is used in Chapters 7 and 8 to construct models for tectonics.
Elastic strains are very small so that the tensor is linearized (Eq. (2.6)) as V = I + E. Substituting
this into Eq. (5.5), we obtain the theorem for an infinitesimal strain (E ≈ O). Namely,
S = φ0 I + φ1 (I + E) + φ2 (I + E)2
≈ (φ0 + φ1 + φ2 )I + (φ1 + 2φ2 )E, (7.1)
where the approximation (I + E)2 ≈ I + 2E is used, and the coefficients φ0 , φ1 and φ2 are the
functions of basic invariants of E. Note that EI = trace E is similar to E in order of magnitude. Both
EII ∼ O(E2 ) and EIII ∼ O(E3 ) are negligible compared to EI . Therefore, the geometrical linearity
allows us to assume that the functions depend only on EI = trace E.
Experiments show that elastic stress and strain have a linear relationship for rocks. If we define
the unloaded shape of a rock for a reference configuration, Eq. (7.1) should be homogeneous—zero
stress must correspond to zero strain. Consequently, we have S = a(trace E)I + bE, where a and b
are scalar constants. Therefore, a linear elastic body should have a constitutive equation of the form
S = λ (trace E) I + 2GE, (7.2)

153
154 CHAPTER 7. ELASTICITY

where λ and G are material constants known as Lamé’s constants. As EI represents volume change,
λ indicates the resistance to the changes. Deformations without volume change are associated with
shear strains that induce stresses due to the second term of Eq. (7.2). G indicates the resistance to
shear strains, and is called the shear modulus. Equation (7.2) is rewritten with r and e as1

r = λ (trace e) I + 2G e. (7.3)

The tensorial equations of the linear elasticity (Eqs. (7.2) and (7.3)) relate not only to the mag-
nitude of stress and strain but also to their principal axes. We have assumed isotropy to use the
representation theorem. The assumption results in the correspondence—principal stress axes are
parallel to principal strain axes. Therefore, the relation of the principal values is often enough for
tectonics, provided that the axial directions are known a priori. If we take the axes of Cartesian coor-
dinates as parallel to the principal axes, the diagonal components of S and E are identical with their
principal values. Using the values, the constitutive equations (Eqs. (7.2) and (7.3)) are rewritten as
⎧ ⎫ ⎧ ⎫⎧ ⎫
⎨S1 ⎬ ⎨λ + 2G λ λ ⎬ ⎨E1 ⎬
S2 = λ λ + 2G λ E2
⎩ ⎭ ⎩ ⎭⎩ ⎭
S3 λ λ λ + 2G E3
⎧ ⎫ ⎧ ⎫⎧ ⎫
⎨σ1 ⎬ ⎨λ + 2G λ λ ⎬ ⎨ε1 ⎬
σ2 = λ λ + 2G λ ε2 . (7.4)
⎩ ⎭ ⎩ ⎭⎩ ⎭
σ3 λ λ λ + 2G ε3

The column vectors and coefficient matrices are put in braces here instead of being parenthesized,
because they are not first- or second-order tensors that are transformed as Eq. (C.34), but the prin-
cipal values are invariant to coordinate rotations and the components of the matrices are material
constants that do not depend on the rotation, either.
Sometimes we need to determine strain from stress. They are linearly related, so that the linear
equations
1+ν ν 1+ν ν
E= S − SI I, e = r − σI I (7.5)
Y Y Y Y
meet the necessity, where Y and ν are material constants called Young’s modulus2 and Poisson’s
ratio. Table 7.1 shows those constants of typical rocks. Equation (7.5) is rewritten by the relation
between the principal values as
⎧ ⎫ ⎧ ⎫⎧ ⎫ ⎧ ⎫ ⎧ ⎫⎧ ⎫
⎨E1 ⎬ 1 ⎨ 1 −ν −ν ⎬ ⎨S1 ⎬ ⎨ε1 ⎬ 1 ⎨ 1 −ν −ν ⎬ ⎨σ1 ⎬
E2 = −ν 1 −ν S2 , ε2 = −ν 1 −ν σ2 . (7.6)
⎩ ⎭ Y ⎩ ⎭⎩ ⎭ ⎩ ⎭ Y ⎩ ⎭⎩ ⎭
E3 −ν −ν 1 S4 ε3 −ν −ν 1 σ3

1 The signs of stresses S and r and of strains E and e are defined not to affect the form of constitutive equations as the

case of Eqs. (7.2) and (7.3).


2 Most literature assign the symbol E to Young’s modulus. However, we have appropriated the symbol to infinitesimal

strain, so that Y is used instead.


7.1. LINEAR ELASTICITY 155

Table 7.1: Elastic property of typical rocks [245].

ρ Y G ν
×103 kg m−3 GPa GPa
Sedimentary Rock
shale 2.1–2.7 10–30 14
sandstone 2.2–2.7 10–60 4–30 0.2–0.3
limestone (non-crystalline) 2.2–2.8 60–80 20–30 0.25–0.3
limestone (crystalline) 2.2–2.8 30–90 20–35 0.1–0.4
metamorphic rock
gneiss 2.7 4–70 10–35 0.4–0.15
amphibolite 3.0 50–100 0.4
igneous rock
basalt 2.95 60–80 30 0.25
granite 2.65 40–70 20–30 0.1–0.25
gabbro 2.95 60–100 20–35 0.15–0.2
diorite 2.80 60–80 30–35
mantle material* 3.359 160 60 0.3
*Inferred values at a depth of 100 km by seismological observations.

It is obvious that Eqs. (7.4) and (7.5) are the inverse of each other. Inverse of the coefficient matrix
of Eq. (7.6) is ⎧ ⎫
⎨ ν − 1 −ν −ν ⎬
Y
−ν ν − 1 −ν .
(ν + 1)(2ν − 1) ⎩ ⎭
−ν −ν ν−1
This should be equal to the coefficient matrix of Eq. (7.4), so that we have the following relations
among material constants:
Yν λ
λ= (7.7) ν= (7.8)
(1 − 2ν)(1 + ν) 2(λ + G)
G(3λ + 2G) Y
Y = (7.9) G= . (7.10)
λ+G 2(1 + ν)

7.1.2 Physical Interpretation of Y, ν and G


Suppose a uniaxial compression where σ1 = 0, σ2 = σ3 = 0. Combining these conditions and Eq.
(7.6), we obtain
σ1 = Y ε1 . (7.11)

The subscripts in this equation indicate that stress is proportional to strain in the extending direction
and Y is the constant of proportionality. This is compared to the behavior of springs that obeys the
equation F = −kx where F and x are force and length change of the spring and k is the spring
156 CHAPTER 7. ELASTICITY

constant. Linear elasticity is the three-dimensional extension of Hooke’s law3 , and Y is the stiffness
comparable to the spring constant.
Let us estimate elastic strains with using Eq. (7.11). Tectonic stresses are in the orders of 10–
100 MPa in tectonically active regions, and Young’s modulus of rocks is about 60 GPa (Table 7.1).
Therefore, elastic strain is σ/Y ≈ 10−4 –10−3 . Strains may be less than this in inactive regions.
The lithosphere cannot support stresses larger than the magnitude, but brittle or plastic deforma-
tions occur to create permanent deformations that may be observable as geologic structures. The
deformation relieves tectonic stresses, giving rise to the upper limit of the stresses.
Suppose a cylindrical sample is subject to the uniaxial compression σ1 > 0 and σ2 = σ3 = 0. If
the sample is made of a linear elastic material, combining Eqs. (7.6) we have ε1 = σ1 /Y and

ε2 = ε3 = −νσ1 /Y = −νε1 . (7.12)

Although O-2 and -3 directions are stress free, the cylindrical sample becomes thicker in those
directions. This is called the Poisson effect. It should be pointed out that plain stress does not result
in plain strain due to the effect. Plane strain does not cause plain stress.
Poisson’s ration is constrained in the range from 0 to 1/2. Equation (7.12) shows that ν indicates
how much Poisson effect appears. If ν = 0, no such effect occurs. Common materials including rocks
have the positive Poisson effect that a sample thickens in directions perpendicular to the direction of
uniaxial compression. This is not satisfied if ν < 0. Therefore, it is permitted to assume that ν ≥ 0. In
addition, elastic strains are infinitesimal, so that volume change is δV /V = −εI = −(ε1 +ε2 +ε3 ) (Eq.
(2.11)). Therefore, combining Eq. (7.12) we obtain δV /V = −(1 − 2ν) ε1 . We have assumed that
the sample is compressed, therefore the ratio should be negative in sign. In this case (ε1 > 0), so that
Poisson’s ratio must satisfy the inequality ν ≤ 1/2. If ν = 1/2, the elastic strain is incompressible—
no volume change occurs. Consequently, the ratio has a value of between 0 and 1/2.
Poisson’s ratio is often assumed at 0.25 in tectonic models because most rock types have similar
ratios (Table 7.1). Seismology shows that the representative Poisson ratio for the crust is 0.25 ±
0.04 on average, and at about 0.27 and 0.29 under Cenozoic orogenic belts and shields, respectively
[274]. In addition, the ratio at 0.25 is convenient for modeling. The reason is that although the
constitutive equation of linear elasticity (Eq. (7.2)) has two material parameters, λ and G, they have
a common value if ν = 1/4. In this case, Eq. (7.8) becomes 1/2 = λ/(λ + G), indicating that λ = G.
In the following sections, we shall often use the value ν = 0.25 as the representative Poisson ratio
for rocks.
The shear modulus G indicates the resistance of a material to simple shear (Exercise 7.1).

3 Ingeneral, linear elasticity is expressed by the equation S = C : E, where C = (Cijkl ) is a fourth-order tensor called the
tensor of elastic constants and represents the anisotropy of a material. This equation indicates proportionality between stress
and strain, called the generalized Hooke’s law. For isotropic materials, C is the fourth-order isotropic tensor, and the above
equation reduces to Eq. (7.2). See [61] for detail.
7.2. EARTH PRESSURE AT REST 157

7.1.3 Two-dimensional elasticity


Orogenic belts and rift zones have linear belts in which vigorous deformations take place. Accord-
ingly, it is sometimes convenient to model the deformations on the vertical section perpendicular to
the belts by means of two-dimensional elasticity, where the plane strain or plane stress condition is
combined with the linear elasticity.
First, let us derive the constitutive equations for plane-strain problems. Consider plane strain on
the O-13 plane, where the O-1 and -3 axes are taken to be parallel to the principal strain axes. In this
case we have ε2 = 0, so that the constitutive equation of linear elasticity (Eq. (7.4)) becomes
⎧ ⎫ ⎧ ⎫
⎨σ1 ⎬ ⎨(λ + 2G)ε1 + λε3 ⎬
σ2 = λ(ε1 + ε3 ) . (7.13)
⎩ ⎭ ⎩ ⎭
σ3 λε1 + (λ + 2G)ε3

Eliminating ε1 and ε3 from Eq. (7.13) and exchanging Lamé’s constants by Poisson’s ratio using
Eq. (7.8), we obtain the constitutive equation for the plane strain problem,

σ2 = ν (σ1 + σ3 ) . (7.14)

The deformation is incompressible if ν = 1/2. In this case, we have σ2 = (σ1 + σ3 )/2, i.e., the
incompressible plane strain of a linear elastic body results in the stress ratio, Φ = 1/2.
The state of plane stress parallel to the O-12 plane is sometimes assumed to model the elastic
behavior of the lithosphere, later. The state is represented by the principal stresses, S1 = 0, S2 = 0
and S3 = 0. Substituting these into Eq. (7.6), we obtain the constitutive equation for this states
⎧ ⎫ ⎧ ⎫
⎨E1 ⎬ 1 ⎨ S1 − νS2 ⎬
E2 = −νS1 + S2 . (7.15)
⎩ ⎭ Y ⎩ ⎭
E3 −(S1 + S2 )

It should be noted that plane stress does not necessarily result in plane strain. This discordance is
due to the Poisson effect which is explained in the following section.

7.2 Earth Pressure at Rest


We have introduced the lithostatic state of stress in Chapter 3 as a reference state, where rocks behave
as fluid at depths. Now we study the state of stress with the assumption that rocks behave as linear
elastic materials.
For simplicity, we derive the state of stress for a linearly elastic rock mass on five assumptions:
(1) we assume isotropy for rocks at depth: (2) we assume a layered structure: density below the
horizontal surface, ρ, is horizontally homogeneous but has vertical variations. We take the origin of
the Cartesian coordinates O-xyz at the surface and the z-axis downward. (3) We assume that gravity
is the only origin of stress. If so, one of the principal stress axes must be vertical and the rest lie
horizontal. Since the system has no horizontal variation at all, we expect axial symmetry around the
158 CHAPTER 7. ELASTICITY

vertical for the state of stress. Therefore, σxx , σyy and σzz equals principal stresses with σxx = σyy . If
rocks are isotropic, the strain field must have the same symmetry: εxx , εyy and εzz are the principal
strains and εxx = εyy . (4) Rocks are confined at depth in the crust, horizontal expansion is restricted
by adjacent rock. Therefore, let us assume that the layers are horizontally constrained so as not to
allow horizontal displacement: ux = uy = 0. Therefore, we have

∂ux ∂uy
εxx = = 0, εyy = = 0. (7.16)
∂x ∂y
(5) Rocks behave as linear elastic materials. This allows us to use Eq. (7.4), in which Eq. (7.16) is
substituted. The result is the equation
⎧ ⎫ ⎧ ⎫⎧ ⎫ ⎧ ⎫
⎨σxx ⎬ ⎨λ + 2G λ λ ⎬⎨ 0 ⎬ ⎨ λ ⎬
σyy = λ λ + 2G λ 0 = λ εzz .
⎩ ⎭ ⎩ ⎭⎩ ⎭ ⎩ ⎭
σzz λ λ λ + 2G εzz λ + 2G

It is found from this equation that


 
λ
σxx = σyy = σzz . (7.17)
λ + 2G

Using Eqs. (7.7) and (7.10), we have


Yν ν
λ (1−2ν)(1+ν) 1−2ν ν ν
= = = = .
λ + 2G 1−2ν + ν + (1 − 2ν) 1−ν
ν
(1−2ν)(1+ν) + (1+ν)
Yν Y 1

Using Eq. (7.17), we obtain


 ν   ν 
σxx = σyy = σzz = pL , (7.18)
1−ν 1−ν
where pL stands for overburden (Eq. (3.26)). This is called earth pressure at rest which is the state
of stress that is induced by gravity in a horizontally layered rock mass. Hence, let us write the state
as ⎛ ⎞
ν/(1 − ν) 0 0
rgravity = pL ⎝ 0 ν/(1 − ν) 0⎠ . (7.19)
0 0 1
We are able to use earth pressure at rest as the reference for the state of stress at depths instead of
the lithostatic state of stress. It is necessary to specify the reference state to discuss crustal stresses.
Since Poisson’s ratio is bounded as 0 < ν ≤ 1/2, the coefficient ν/(1 − ν) in Eqs. (7.18) and
(7.19) ranges between 0 and 1. Therefore, the inequality 0 < (σxx , σyy ) ≤ σzz holds, indicating that
horizontal stress is less than the overburden. Rocks are vertically constricted by overburden but are
horizontally extended by the Poisson effect. In the extreme case of ν = 1/2, Eq. (7.18) becomes
σxx = σyy = σzz . This is equivalent to the lithostatic state of stress, though the media is not assumed
to be fluid but elastic.
7.3. STABILITY OF ELASTIC ROCK MASSES 159

Since Poisson’s ratio of representative rocks can be approximated as ν ≈ 1/4 (p. 156), the
coefficient K = σHmean /σv is approximately equal to 1/3. Horizontal stress is about one-third of
overburden. In-situ stress measurements conducted in various regions of the world show that a few
regions show K ≈ 1/3 [24]. The ratio is considerably scattered in the shallow levels of the crust,
but seems to converge with increasing depth to ∼ 1, the lithostatic state of stress. The above model
predicts the differential stress Δσ = (1 − K)pL , which piles up unlimitedly with increasing depth.
However, real rocks have their own yield stress. The yielding may relief the differential stress to
cause convergence in the ratio. This is perhaps at least one of the reasons for the convergence. The
stress state discussed above is an idealized state, convenient for theoretical considerations, but we
should be careful to apply the model to real tectonics. Rocks behave as elastic materials in a human
timescale. However, elastic stress may be released in a geological timescale.

7.3 Stability of elastic rock masses


Is a rock stable if its behavior is linear elastic? Let us examine whether the Earth pressure at rest is
under the frictional strength of rocks. If it is not so, rocks may collapse through normal faulting on
inclined pre-existing fractures without tectonic stress. The Earth pressure at rest has the principal
stresses  ν 
σv = pL > σH = σh = pL . (7.20)
1−ν
Figure 7.1a shows this stress state.
In one extreme case, ν = 0, this is a uniaxial stress σH = σh = 0. The Mohr circles corresponding
to various depths pass the origin of the Mohr diagram so that the line of frictional strength always
cuts the circles, i.e., the elastic rock body with ν = 0 is unstable. In another extreme case, ν = 1/2,
we have lithostatic states σv = σH = σh = pL . Then no deformation occurs at all and the elastic rock
body is always stable.
For the general case where 0 < ν < 1/2, the Mohr circles has a linear envelope passing the
origin of the Mohr diagram for the proportionality of all the principal stresses with pL (Fig. 7.1(a)).
The slope of the line is given by means of the Pythagorean theorem about the triangle OCD. Namely,
using Eq. (7.20), we have
1 ν  1 1 ν  2ν − 1
OC = pL + pL = pL , CD = pL − pL = pL ,
2 1−ν 2(1 − ν) 2 1−ν 2(1 − ν)
  ν 1/2
∴ OD = c2 − r 2 = pL .
1−ν
Consequently, we obtain the slope
CD 2ν − 1
=  .
OD 2 ν(1 − ν)
Figure 7.1(b) shows the variation of this slope. If we use the value ν = 0.25 for the representative
Poisson ratio for rocks, we have a slope of 0.58. This is slightly smaller than the coefficient of
160 CHAPTER 7. ELASTICITY

Figure 7.1: (a) Mohr diagram showing the Earth pressure at rest. (b) Relationship between ν and
the slope in (a).

friction. Consequently, the rocks are stable under the Earth pressure at rest. Rocks with small
Poisson ratios may be unstable.
We have neglected the effect of pore pressure in the above argument. However, pore pressure
plays a critical role in the stability because increased pore pressure shifts the Mohr circle to the left
and destabilizes rocks with ordinary Poisson ratios. Therefore, earthquakes can occur on pre-existing
fracture surfaces only through a building up of pore pressure.

7.4 Thermal stress


Changes of temperature leads to thermal stress which is as large as a stress increment resulting from
other factors. Since temperature increases generally with depth, depth changes accompanied by
exhumation or basin subsidence affect the state of stress.

Basic equation

Rocks expand when heated. Consider a strain E of a rod accompanied by the temperature change
of ΔT . Experiments show that the proportionality E = α ΔT holds for a rod of rock if the length
of the rod is unconstrained. The constant of proportionality α is known as the coefficient of linear
expansion. Rocks have a coefficient of about 10−5 K−1 . The temperature change results in an
increased volume of a cube given by
 3  2  3
ΔV (L + ΔL)3 − L3 L + ΔL ΔL ΔL ΔL ΔL
= = −1=3 +3 + ≈3 ,
V L3 L L L L L

where L is the initial length of the sides of the cube and higher-order terms are neglected. Using
ΔL = LE, we have
ΔV
≈ 3E = 3α ΔT.
V
7.4. THERMAL STRESS 161

Accordingly, we define the coefficient of volume expansion or simply thermal expansion coefficient
α so as to satisfy the relationship ΔV /V = αΔT . Obviously, we have

3αl ≈ α. (7.21)

Rocks have the thermal expansion coefficients at around 3 × 10−5 K−1 .


The coefficient of linear expansion of a single crystal has anisotropy resulting from its crystal-
lographic structure. However, a rock mass consists of a polycrystalline material made of minerals
usually with random crystallographic orientations. Therefore, we assume that α is isotropic when
we consider the stress or strain of rocks. This is formulated as

ε1 = ε2 = ε3 = −αl ΔT. (7.22)

Incremental deformation against viscous forces dissipates kinetic energy (Eq. (3.41) to affect
the temperature field, i.e., mechanical and thermal processes are generally coupled. However, in
many situations their interaction can be neglected. The resulting analysis is known as the uncoupled
thermoelastic theory of continua. In this case, strain is considered to be the sum of the strains caused
by the incremantal stress and temperature. The constitutive equation of the former is given by Eq.
(7.5) or (7.3) and the latter is given by Eq. (7.22), so that we have

1+ν  νσ 
I
e= r− + αl ΔT I, (7.23)
Y  Y 
Y αl ΔT
r = 2Ge + λεI + I. (7.24)
1 − 2ν

Using principal stresses and strains, Eq. (7.23) is rewritten as

1 ν ν
ε1 = σ1 − σ2 − σ3 − αl ΔT,
Y Y Y
ν 1 ν
ε2 = − σ1 + σ2 − σ3 − αl ΔT , (7.25)
Y Y Y
ν ν 1
ε3 = − σ1 − σ2 + σ3 − αl ΔT .
Y Y Y

Prior to using Eq. (7.25), the principal orientations should be specified.


When magma cools down to be solidified, a temperature change of several hundreds degrees
leads to a stress change of Y α1 ΔT ∼ 102 MPa, as the representative Young’s modulus of rocks
is about Y = 60 GPa. Since the tensile strength of rocks is in the order of 101 MPa, this thermal
stress easily forms fractures in the solidified igneous body, called cooling joints. Cooling of a tabular
intrusive body splits the body into columns to form columnar joints perpendicular to the intruding
plane (Fig. 7.2).
162 CHAPTER 7. ELASTICITY

Figure 7.2: Doleritic dike penetrating purple granite and overlying volcanic rocks. The dike has
columnar joints. Oga peninsula, Northeast Japan.

Comparison with overburden stress

The state of stress in a linear elastic material at depth is described by the Earth pressure at rest,
which indicates a normal fault regime of stress. If a lava flow that was consolidated at the surface
is gradually buried in a sedimentary basin, the layer experiences horizontal extension. On the other
hand, the temperature of the layer increases with depth, leading to horizontal compression if the
layer is horizontally constrained as the condition from which the earth pressure at rest was derived.
Let us consider these competing effects on the stress state.
For this purpose, we assume a level surface and no horizontal density or tempeature variation.
Under these conditions, stress and strain have a vertical principal orientation. In addition, we as-
sume uncouple mechanical and thermal processes so that incremental stresses accompanied by an
increases of temperature and overburden are considered separately.
The layer is horizontally constrained but can freely displace the overlying strata by its thermal
expansion, i.e., the layer is not vertically constrained. Therefore, the thermal expansion does not
result in an increase of the vertical stress component, i.e., σzz = 0 for a contribution from the
7.4. THERMAL STRESS 163

temperature increase. Substituting this vertical stress into Eq. (7.25), we have

1
εxx = (σxx − νσyy ) − αl ΔT
Y
1
εyy = (−νσxx + σyy ) − αl ΔT (7.26)
Y
ν
εzz = (−σxx − σyy ) − αl ΔT
Y
Combining the condition of horizontal constraints εxx = εyy = 0, we obtain

Y αl ΔT
σxx = σyy = (7.27)
1−ν
for the incremental stress components due to the temperature increase. Let Γ be a geothermal gradi-
ent, then the temperature at depth z is given by T = T0 +Γz, where T0 is the surface temperature. The
layer had no strain at all at the surface, therefore ΔT = Γz. Substituting this temperature difference
into Eq. (7.27), we obtain
Y αl Γz
σxx = σyy = . (7.28)
1−ν
The differential stress due to the temperature difference equals
 
Y αl Γ
Δσ = z. (7.29)
1−ν

Geothermal gradients are usually a few degrees per 100 m except for geothermal areas. Using the
values Γ = 0.01 K m−1 , Y = 60 GPa, ν = 0.25, and α = 10−5 K−1 , we obtain Δσ/z = 8 kPa m−1 ,
and the gradient of overburden stress ρg = 2.4 kPa m−1 . Consequently, the thermal stress is as great
as othe verburden stress for a layer that was consolidated at the surface4 .

Polygonal fracture pattern on Venus

Venusian plain regions have polygonal terrains that are characterized polygonal fracture network by
relatively uniform spacing (Fig. 7.3). The random orientations of the fractures suggest isotropic
small-strain deformations, probably reflecting the thermal stress included by decreased surface tem-
perature over a billion years [93].
Venus was resurfaced by global massive volcanism a few billion years ago. Greenhouse effect
gases such as H2 O and SO2 emitted by the volcanism raised the surface temperature by more than
100 K. Later atmospheric cooling contracted rocks in the crust. Just like the half-space cooling
model of the oceanic lithosphere (§3.12), rocks at shallow levels in the crust are cooled swifter than
those at deep levels. The time lag induces a variation in the magnitude of horizontal contraction to
4 This model assumes that a horizontally extensive elastic layer subsided neither with jointing nor faulting in a geological

time scale. This is not probable in reality. Fractures between blocks may allow horizontal displacements, violating the
condition of horizontal constraints.
164 CHAPTER 7. ELASTICITY

Figure 7.3: Polygonal terrain on Venus [5].

eventually form the polygonal fractures eventually [5]. Let us estimate the contraction and associated
thermal stress by the climate change.
Consider the temperature changes with time as T = A sin ωt at the surface z = 0. The tempera-
ture in the half-space z ≥ 0 is calculated by the heat conduction equation ∂T/∂t = κ∂2 T/∂z2 . Here,
T represents the temperature anomaly. Substituting the solution of the form T = u(z)eiωt , we have
d2 u/dz2 = (iω/κ)u. Using the condition that the solution does not diverge to infinity at z = ∞, we
obtain
T = Ae−Bz cos(ωt − Bz),

where B = ω/2κ [36]. Therefore, the temperature change at z is delayed by Bz. The vertical
temperature variation induces thermal stresses. We have the variation

∂T  
= ABe−Bz sin(ωt − Bz) − cos(ωt − Bz) .
∂z
Hence, the near-surface geothermal gradient is represented by Γ ≈ AB. Using the values A = 100
K, ω = 1 rad/Ga ≈ 3 × 10−16 s−1 , and κ = 1 × 10−5 K−1 , we have Γ ≈ AB ≈ 10 K/km. Assuming
constrained horizontal displacement, the thermal stress between different depths Δz is evaluated as
Eq. (7.29), i.e., we have  
Y α Γ
σ≈ Δz.
1−ν
Using Y = 60 GPa and ν = 0.25, we have Y α Γ/(1 − ν) ≈ 1000 Pa m−1 . On the other hand, the
gravitational acceleration at on the Venusian surface is g ≈ 9 ms−1 so that the gradient of overburden
stress is ρg ≈ 3000 Pa m−1 . Therefore, the thermal stress is comparable and if the surface tempera-
ture drop was a few hundred degrees, the thermal stress becomes large enough to produce map-scale
fractures5 .
5 See [5] for detail.
7.5. GLOBAL THERMAL CHANGES AND SURFACE STRESS FIELD 165

7.5 Global thermal changes and surface stress field


Terrestrial planets and satellites experienced drastic global cooling in their early history. Except for
the Earth and Venus which have active geological processes including tectonism and erosion, tec-
tonic features are clues to their early thermal history. For example, it is suggested that the cessation
of extensional tectonics some 3.6 billion years ago was the surface manifestation of global cooling
[127]. Thrust faults on Mercury are thought to be the result of global cooling, also. In this section,
we investigate the linkage between the cooling and tectonics.
Consider a self-gravitating solid spherical body with a radius of R, in which all physical prop-
erties have spherical symmetry. The symmetry leads to a vertically axisymmetric stress field. Let
SV and SH be the vertical and horizontal stresses. Here, we assume a solid planetary body, so that
the vertical and horizontal stresses can be different. The unit volume in the body is subject to the
gravitational force −ρg, where the density ρ and gravitational acceleration g are functions of the
radial coordinate r. The body force and the stress components shown in Fig. 7.4 are found to lead
to an equillibrium equation in the form6

dSV 2
+ (SV − SH ) − ρg = 0. (7.30)
dr r
Thermal expansion of the body is very small compared to the radius of the planetary body R, so
that the former does not depend on temperature. Therefore, let us utilize the uncoupled thermoelastic
theory to investigate thermal stresses in the body. Stresses due to gravity and to temperature change
can be treated separately. Hence, the gravity term −ρg is deleted from Eq. (7.30) to estimate the
stress field due to temperature changes. Consequently, the equation of thermal stress (Eq. (7.25)) is
rewritten as
1
EV − αΔT = (SV − 2νSH ) , (7.31)
Y
1 
EH − αΔT = SH − ν(SV + SH ) , (7.32)
Y
where SV and SH are vertical and horizontal infinitesimal strain components due to temperature
change ΔT . These strain components are related to the vertical displacement u by the equations

du u
EV = , EH = . (7.33)
dr r
Rearranging Eqs. (7.31) and (7.32), we have
   
Y (1 − ν)EV + 2νEH − (1 + ν)αΔT Y EH + νEV − (1 + ν)αΔT
SV = , SH = . (7.34)
(1 + ν)(1 − 2ν) (1 + ν)(1 − 2ν)
6 Equation (3.46) that describes the acceleration due to gravity as a function of r lacks the second term of Eq. (7.30).

When we derived the former equation, we implicitly assumed that the planetary body is composed of fluid at rest. Therefore,
the differential stress SV − SH vanished.
166 CHAPTER 7. ELASTICITY

Figure 7.4: Force balance of a small circular truncated cone with a thickness of dr and an apical
angle of dφ.

Subsitituting these equations into Eq. (7.30), we have the differential equation of u,

d2 u 2 du 2u  1 + ν  dΔT
+ − = α .
dr2 r dr r 1 − ν dr
Rearranging this equation, we obtain
   1 + ν  dΔT
d 1 d  2 
r u =α ,
dr r dr 1 − ν dr
which is integrated to give
r
α 1 + ν  C2
u= ΔT r 2 dr + C1 r + , (7.35)
r2 1 − ν r0 r
where C1 and C2 are constants of integration and r0 designates the base of a spherical shell. Com-
bining this differential equation and Eqs. (7.33) and (7.34), we obtain the equations [246],
r
2αY Y C1 2Y C2
SV = − ΔT r 2 dr + − , (7.36)
(1 − ν)r3 r0 1 − 2ν (1 + ν)r3
r
αY Y C1 Y C2 αY ΔT
SH = ΔT r 2 dr + + − . (7.37)
(1 − ν)r r0
3 1 − 2ν (1 + ν)r 3 1−ν
7.5. GLOBAL THERMAL CHANGES AND SURFACE STRESS FIELD 167

The constant C2 is determined through Eq. (7.35) by the vanishing displacement at the center r = 0
as C2 = 0 because of 
1 r
lim 2 ΔT r 2 dr = 0.
r→0 r 0
Let ΔT0 be the temperature change at the center, then we have

1 r ΔT0
lim 3 ΔT r 2 dr = .
r→0 r 0 3
Therefore, the other constant is constrained by assuming a free surface at r = R as

Y C1 2αY 1 R
= ΔT r 2 dr.
1 − 2ν 1 − ν R3 0
Consequently, Eqs. (7.36) and (7.37) become
  r 
2αY 1 R 1
SV = ΔT r 2 dr − ΔT r dr ,
2
1 − ν R3 0 r3 0
  r 
αY 2 R 1
SH = ΔT r 2 dr + ΔT r dr − ΔT
2
.
1 − ν R3 0 r3 0

Now consider a planetary body being composed of n spherical shells, where the top is the 0th
layer. The shells have different material constants but the constants are uniform within each layer.
Let Yi and νi be Young’s modulus and Poisson’s ratio of the ith layer. Then, we have

R3
u(r) = rI (r) + Ai r + Bi , (7.38)
r2
2Yi I (r) Yi Ai 2Yi Bi R3
SV (r) = − + − ,
1 + νi 1 − 2νi (1 + νi )r3
Yi I (r) Yi Ai Yi Bi R3 Yi α(r)ΔT (r)
SH (r) = − + + − , (7.39)
1 + νi 1 − 2νi (1 + νi )r 3 3(1 − νi )
where r
1+ν
I (r) = α(r)ΔT (r)r2 dr (7.40)
3(1 − ν)r3 0
and Ai and Bi are dimensionless numbers corresponding to the following three conditions on the
stress field [218]. Namely, (1) there is no singularity at the center, (2) no discontinuity between
layers, and (3) the surface of the planetary body is free, SV = 0. The integral in Eq. (7.40) equals
the thermal expansion within the radius of r. 
The goal of this argument is to compare the surface horizontal stress SH r=R and observed
tectonic features. For brevity, we assume that the surface temperature is constant through time,
ΔT (R) = 0. Combining the free surface condition, we obtain the equation for the surface layer,
2Y0 I (R) Y0 A0 2Y0 B0
− + − = 0.
1 + ν0 1 − 2ν0 1 − ν0
168 CHAPTER 7. ELASTICITY

Rearranging this equation, we get


1 1 + ν0
B0 = A0 − I (R). (7.41)
2 1 − 2ν0
Combining Eqs. (7.38), (7.39) and (7.41), we obtain the displacement and horizontal stress of the
surface layer (r = R),
3 R(1 − ν0 )A0
u(R) = , (7.42)
2 1 − 2ν0
3 Y0 A0
SH (R) = . (7.43)
2 1 − 2ν0
The increase of the radius ΔR equals the upward displacement of the top layer u(R). Therefore,
combining Eqs. (7.42) and (7.43), we obtain
Y0 ΔR
SH (R) = . (7.44)
1 − ν0 R
Consequently, the surface horizontal stress determined only by the stretch of the radius ΔR/R and
the elastic constants of the surface layer, independent from the physical properties of the inside. The
ratio ΔR/R reflects the internal changes.
The activity of thrust faults (Fig. 2.14) contracted Mercury 3–4 billion years ago. The orientation
and distribution of these thrust faults appear to be rondom, the contraction in the radius ΔR was
roughly estimated via Eq. (2.56) at minus 1–2 km, and is interpreted as the surface manifestation
of global cooling [232]7 . |ΔR| is much smaller than the present radius of Mercury at 2439 km.
Therefore, the initial radius can be replaced by the present one, and we have the ratio −ΔR/R =
(4–9)×10−4 . Substituting this ratio with Y = 60 GPa and ν = 0.25 as the representative values of
rocks into Eq. (7.44), we obtain the surface horizontal stress at −SH = 30–60 MPa.
According to the theory of thermal evolution, that amount of contraction is explained well by
the heat loss just as the amount of heat production by core formation alone, and the total freezing of
the core can lead to an 8 km contraction of the radius [217]. The contraction estimated from thrust
faults may be the lower bound, so that Mercury probably has a large inner core [208].
The Moon has tectonic features indicating both horizontal contraction and expansion, which are
evidenced by wrinkle ridges and linear rilles, respectively. In addition, those features appear to
reflect regional tectonics characteristic of mare basins (§8.8). Global cooling left a less conspicuous
pattern of tectonic features on the Moon than on Mercury. The lack of a pattern suggests the global
expansion as small as |ΔR| < 1 km for the Moon [129]. This corresponds to |SH | < 50 MPa.
However, the wrinkle ridges still suggest global cooling, as extensional tectonics terminated at ∼3.6
Ga on the Moon, and horizontal compression prevailed thereafter. Too hot initial conditions lead
to a large contraction, which is inconsistent with the tectonic features on the Moon. Using these
constraints, Solomon et al. [219] estimated the depth of the magma ocean at about 200 km at the
beginning of the history of the Moon.
7 Several researchers raise objections to the randomness, as the surface of the Mercury has not been globally photographed

with enough resolution [147, 241].


7.6. EXERCISES 169

7.6 Exercises
7.1 Demonstrate that the shear modulus G indicates the resistance of a linearly elastic material to
simple shear.

7.2 The Venusian surface has polygonal fractures, each of which has a diameter of several kilo-
meters (Fig. 7.3). The spacing between the fractures is thought to reflect the depth extent of crack
propagation. Derive the equation
ΔΓαY  te 
ΔSH = −z ,
1−ν 2
describing the change in the surface horizontal stress due to the change in the geothermal gradient
ΔΓ, where te is the thickness of the surface elastic layer. Assuming the Coulomb-Navier criterion,
estimate the maximum depth for fracture propagation due to thermal stress [93].
Chapter 8

Flexure
The lithosphere behaves as a thin elastic plate in geological time scales. The elasticity of
the lithosphere affects the uplift of mountain ranges and the subsidence of sedimentary
basins to leave geological records of the behavior. We are able to estimate the elasticity
of the ancient lithosphere.

8.1 Regional isostasy


The lithosphere is deflected by several forces, including the weight of mountain ranges and plate
boundary forces [15]. The weight is supported by two forces—buoyancy from the underlying man-
tle and the elasticity of the lithosphere itself. A simple engineering model that was developed to
calculate the flexure of a thin elastic plate is useful in understanding the flexure of the lithosphere.
When we considered isostasy in Section 3.7, we did not take the elasticity into account for
the balance of forces, whereby the vertical density distribution was the only control for isostasy.
Horizontal variations are not significant in the model, and topography or vertical movement is locally
determined (Fig. 8.1(a)). Hence, the balancing condition is called local isostasy. If this is always
true and if the load of mountains is supported by the buoyancy of their deep crustal roots with the
density ρc , the altitude h(x) from a reference level should be related to the downward deflection of
the Moho relative to a reference depth d(x) as ρc h(x) = (ρm −ρc )d(x), where x indicates a horizontal
position. In this case, therefore, the topography of the surface h(x) and the Moho d(x) are mirror
images of each other. Does this relationship hold for any topographic relief? It is natural to consider
that it may hold for large mountain ranges and plateaus, but may not hold for mounds. The fact
is that the elasticity of rocks supports short-wavelength topography to cause a deviation from local
isostasy.
If a mountain is on a horizontally lying elastic plate, the load of the mountain not only pushes
down the plate just beneath the mountain but also its surrounding regions (Fig. 8.1(b)). The de-
pression provides accommodation for atmosphere, water, or sediment. Whatever the filling is, it has
generally a smaller density than the mantle. Accordingly, the basin works as a negative load that

171
172 CHAPTER 8. FLEXURE

Figure 8.1: Schematic picture showing two types of isostasy, (a) local and (b) regional. Squares
represent crustal blocks floating on the mantle (gray shading). The weight of the mountains pushes
down the underlying crustal block in (a). The subsided deep crustal root that shoves mantle ma-
terials aside causes buoyancy to support the weight at the top. Elasticity is symbolized by springs
connecting the blocks in (b). The weight of the mountains is distributed to neighboring blocks by
the elastic plate under the mountains, resulting in peripheral basins.

compensates the positive load of the mountain in the vertical force balance. The force balance is
called regional isostasy.
Elasticity of the lithosphere changes with time and region. The cross-sectional shape of the
basin bears the information on the elasticity. Namely, a topographic load makes a shallow but wide
depression around the load for an elastically stiff lithosphere, but a narrow and deep one for a flexible
lithosphere (Fig. 8.1(b)). Local isostasy may be viewed as regional isostasy with zero stiffness. The
stiff lithosphere is difficult to bend for horizontally short-wavelength distribution of positive and
negative loads. Long-wavelength distributions can flex the lithosphere. Accordingly, the flexure of
the lithosphere works as a low-pass filter for the horizontal distribution of loads.
The peripheral depressions may provide accommodation for water or sediments, resulting in
sedimentary basins. The subsidence and uplift of the basins may leave geologic records in strata as
transgressions and regressions. The ancient topography indicated by those phenomena are the clue
to the elasticity of the lithosphere in the geological past.

Example 1: giant submarine fan A large river has been suppling sediments to a continental mar-
gin for tens of millions of years to build a giant submarine fan, which pushes down the lithosphere.
Figure 8.2 shows the isopach map of the Amazon submarine fan, which has gathered clastics derived
from the Andes since the Middle Miocene. Driscoll et al. [51] estimate the amplitude of the litho-
spheric flexure on the assumption that the lithosphere is a thin elastic plate. The result shows that
the lithosphere downwarps more than 2 km beneath the center of the fan. In addition, the sediment
load results not only in downwarping but in a peripheral bulge with an amplitude of several tens of
meters around the fan, i.e., the gray region in Fig. 8.2 has been gradually uplifted since the Miocene.
The river mouth is located at a saddle point of the bulge.
8.1. REGIONAL ISOSTASY 173

Figure 8.2: Flexure of the lithosphere by the loading of the Amazon submarine fan (After [51]).
Solid lines show the isopach lines of the post-Middle Miocene deposits. The contour interval is 250
m. The thickness of the fan reaches 5,300 m. The gray region around the fan is the peripheral bulge
in which the contour lines show the amount of uplift in meters.

Since passive continental margins have relatively inactive tectonism, the margins have strati-
graphic records that evidence the relative sea-level curve for millions of years. The uplift of the
bulge is small, but the growth of the bulge may be observable in stratigraphic records as regression.
The record constrains the elastic constants of the lithospheric.

Example 2: foreland basin The topographic load of a rising orogene makes a foreland basin.
There was a wide and shallow seaway from the Gulf of Mexico to Canada to the east of the
Cordilleran orogenic belt in the Cretaceous, called the Western Interior Basin. Adjacent to the east-
ern front of the orogenic belt, a zone with a width of several hundred kilometers subsided to catch
a sedimentary pile that thickens westward1 . The thickening evidences the syndepositional tilting
of the basement toward the orogene. Figure 8.3 shows the cross-sections of the sedimentary pile
corrected through the backstripped technique2 , with taking into account the lithospheric flexure with
a flexural rigidity (p. 178) of 1.0 × 1024 Nm. The syndepositional tilting and basin subsidence was
due to the growing topographic load of the orogenic belt to the west of the basin. Consequently, the
flexure of the lithosphere can be recorded in sediments so that we can read out the ancient flexural
1 The Western Interior Basin was about 1,000 km wide, and this zone occupied the western margin of the basin. The

formation of the wide basin was an epeirogenic movement probably by the negative dynamic topography (§9.5).
2 The backstripping technique introduced in Section 3.10 assumed local isostasy. However, the technique employed here

assumes regional isostasy, so that it is called the flexural backstripping.


174 CHAPTER 8. FLEXURE

Figure 8.3: E–W trending cross-sections of the Cretaceous deposits in central North America [169].
The sections show the inferred structures at 80 Ma.

parameters of the lithosphere.

8.2 Flexure of thin elastic plate


Let us assume the lithosphere as a thin elastic plate, and examine the relationship between the forces
acting on the plate and its flexure. The flexure of the plate is assumed to be small for simplicity in
that displacements due to the flexure are much smaller than the thickness te . The governing equation
for the lithospheric flexure is obtained in the following discussions by combining the constitutive
equation of linear elasticity, force balance equation, and the approximation based on the assumption
that the flexure is small.
We use the Cartesian coordinates O-xyz with the z axis pointing downward. The central plane
between the top and base of the horizontally lying elastic plate is called the middle surface of the
plate. The xy plane is placed on the middle surface at the unloaded state of the plate. We treat the
flexure as a plane-strain problem on the xz plane.
Linear elasticity is assumed in the following discussion so that we are able to deal with the
elastic deformations due to gravity and flexure separately, and the law of superposition applies to
our analysis. Therefore, at first we neglect the gravitational force to obtain the governing equation
of the thin elastic plate.
8.2. FLEXURE OF THIN ELASTIC PLATE 175

Figure 8.4: Three nearby points A, B, and C, and the radius of curvature, R, of the curve z = w(x)
on the point B. The segment Δs, shown by a thick gray line, is bounded by the midpoints between
the points. R is the radius of the circle tangent to the curve at point B.

8.2.1 Curvature and bending moment


Consider a smooth and continuous function w(x) to represent the shape of the middle surface to
define the curvature of the plate
K = 1/R, (8.1)

where R is the radius of curvature of the graph z = w(x) (Fig. 8.4). We shall see how to relate the
curvature to the function. For this purpose, consider three nearby points A, B, and C on the graph.
The lines AB and BC intersect the x axis with the angles θ and θ + Δθ. The former is equal to w  (x),
and the latter is
tan(θ + Δθ) = w  (x + Δx). (8.2)

The right-hand side of Eq. (8.2) is approximated as w  (x+Δx) ≈ w  (x)+w  (x)Δx, and the left-hand
side becomes
Δθ Δθ
tan(θ + Δθ) ≈ tan θ + 2
= w  (x) + ,
cos θ cos2 θ
because (d/dφ) tan φ = 1/ cos2 θ. Therefore, we have w  (x)Δx = Δθ/ cos2 θ and the equation

1 1 + tan2 θ 1 + [w  (x)]2
Δx = Δθ = 
Δθ = Δθ. (8.3)
w  (x) cos2 θ w (x) w  (x)
 
Length of the arc Δs in Fig. 8.4 is Δs = (Δx)2 + (Δz)2 = Δx 1 + [w  (x)]2 . Using Eq. (8.3),
we get
Δθ w  (x)
=7 2 83/2 .
Δs 1 + w  (x)
176 CHAPTER 8. FLEXURE

Figure 8.5: (a) Small rectangular element ABCD at the unloaded state on the middle surface (dot-
bar line). (b) Deformed state. The corners A and B are assumed to be the right angle, as the plate is
thin and the flexure is small. The right-handed rectangular coordinate system αβγ, where the γ axis
points into the page, is used to calculate the state of stress. The element of the elastic plate (gray)
is bent with the convex upward, so that the convex side of the middle surface is extended parallel to
the surface and the concave side is shortened.

The left-hand side of this equation is equal to RΔθ so that


1 Δθ w  (x)
K= = =7 2 83/2 . (8.4)
R Δs 1 + w  (x)

We consider small flexure so that |w  (x)|  1. Thus,

K = 1/R = w  (x). (8.5)

The z axis is pointing downward, therefore, convex upward flexure has a positive curvature and
convex downward flexure has a negative one. In the latter case, R is also negative. Note that the
geometrically meaningful radius is 1/|K| rather than 1/K.
We shall derive the relationship between the curvature of the plate and bending moment. Suppose
that the plate was horizontal before loading, and the middle surface was on the x axis (Fig. 8.5(a)).
The curvature of this state is K = 0. An element of the plate is bent due to the bending moment M
acting on the both ends of the element (Fig. 8.5(b)). The curvature is proportional to the moment
for small flexures, so that
M = −DK, (8.6)
where −D is the constant of proportionality3 .
Assuming that the lithosphere is thin and the flexure is small, we use Kirchhoff’s hypothesis that
the planes normal to the middle surface at an unloaded state are kept as normal to the surface after
bending. For the flexed state, the coordinate system αβγ are defined as Fig. 8.5(b). Since we are
dealing with a plane-strain problem where there is no movement along the γ axis, we have Eγγ = 0
3 The reason why the negative sign is attached here is to be consistent with Eq. (8.13).
8.2. FLEXURE OF THIN ELASTIC PLATE 177

and the γ axis is the principal strain axis. In addition, as the flexure is small compared to the plate
thickness te , the normal strain Eββ is negligible (Eββ = 0), and the stretching and shortening parallel
to the middle surface are the major strains in the plate. It follows that the α axis is approximately
parallel to another principal strain axis. The plate is assumed to be composed of isotropic linear
elastic material, therefore, the principal strain axes coincide with the principal stress axes.
Note that the β axis is more or less vertical. The upper and lower surfaces of the elastic plate
are assumed to be free4 , so that we have Sββ = 0. Therefore, a state of plane stress exists. The
constitutive equation for the isotropic linear elastic body (Eq. (7.15)) in this case is
⎧ ⎫ ⎧ ⎫
⎨Eαα ⎬ 1 ⎨ Sαα − νSγγ ⎬
Eγγ = −νSαα + Sγγ . (8.7)
⎩ ⎭ Y ⎩ ⎭
Eββ −Sαα − Sγγ

Combining Eqs. (8.7) and Eββ = Eγγ = 0, we have the simultaneous equations

Y Eαα = Sαα − νSγγ , 0 = −νSαα + Sγγ (8.8)

and we obtain5
Y
Sαα = Eαα (8.9)
1 − ν2
The moment due to the normal stress is
 te /2
M= Sαα β dβ. (8.10)
−te /2

Substituting Eq. (8.9) into (8.10), we obtain


 te /2
Y Eαα β dβ
M= . (8.11)
−te /2 1 − ν2

The relationship between the normal strain Eαα and the curvature K is derived as follows. Sup-
pose a rectangle ABCD on the middle surface at the unloaded state (Fig. 8.5(a)). The plate is
assumed to be homogeneous in strength, so that the middle surface coincides with the neutral sur-
face. The line segment CD becomes the arc C D , so that we have


C D − CD C D − AB
Eαα ≈ = .
CD AB

However, the length along the middle surface does not change so that AB = A B = RΔθ, where
R is the radius of the curvature of the middle surface and Δθ is the small angle between the lines
4 Thevertical force balance and the effect of gravity are considered later.
5 Theother horizontal principal stress is Sγγ = [Y ν/(1 − ν 2 )]Eαα , indicating that Sγγ and Eαα have the same sign.
Accordingly, the sign of Sγγ changes across the middle surface.
178 CHAPTER 8. FLEXURE

A C and B D . The length of the arc between C D is (R − A C )Δθ. Therefore, the strain from the
segment CD to the arc C D is
(R − β)Δθ − RΔθ β
Eαα = = − = −Kβ. (8.12)
RΔθ R
Substituting this into Eq. (8.11), we obtain
 te /2
Y Kβ 2 dβ Y K  3 te /2 Y t3e
M=− = − β = − K. (8.13)
−te /2 1 − ν2 3(1 − ν 2 ) −te /2 12(1 − ν 2 )

This is compared with Eq. (8.6) to obtain the equation

Y t3e
D= . (8.14)
12(1 − ν 2 )
This parameter depends on the elastic parameters Y and ν and the thickness te , indicating that D is
defined for each plate to characterize the resistance to bending. This is called the flexural rigidity of
the plate. The negative sign was placed in Eq. (8.6) to make this quantity positive.
The flexural rigidity of the lithosphere is controlled largely by te . Common rocks have similar Y
and ν, whereas the lithosphere of the Earth has various thicknesses from a few km to >100 km. In
addition, D ∝ t3e , so that D has variations over six orders of magnitude.
However, the actual lithosphere does not behave as an elastic plate as a whole. Deformations
are not accommodated by elastic strain in the shallow and deep levels in the lithosphere, where
brittle and ductile deformations are dominant, respectively. Therefore, the above model of the elastic
lithosphere with uniform thickness and strength is quite ideal. The quantity te is the thickness if we
assume the lithosphere as an elastic plate. In this sense, te is called the equivalent or effective elastic
thickness of the lithosphere. The factors determining te of the lithosphere are discussed in Section
12.3.

8.2.2 Balance equations


The task of this subsection is to study the relationship between the flexure w(x) and vertical loads
on the elastic plate.

Vertical force balance Consider the coordinates O-αβ for the element of an elastic plate in a
deformed state (Fig. 8.6(a)). We define the shear force acting on the side of the element as
 te /2  te /2
V ≡ Sαβ dβ ≈ Sxz dz. (8.15)
−te /2 −te /2

Flexure is assumed to be small, so that the x- and z-axes are nearly parallel to the α- and β-axes,
respectively. Thus, the right-hand side of Eq. (8.15) is obtained by the approximation dx ≈ dα
and dz ≈ dβ. Since we use the sign convention that tension is the positive stress for Sxz , the left
8.2. FLEXURE OF THIN ELASTIC PLATE 179

Figure 8.6: (a) Shear forces V and V + dV acting on the sides of the element of an elastic plate
around the point O that is placed on the middle surface (dot-bar line). This element has the unit
length in the direction perpendicular to the page. The α and β axes are defined to be parallel and
normal to the surface. (b) Normal stress Sαα and moment M due to the normal stress. The force
Sαα dβ is oriented left on the left side of the element, whereas the torque Sαα β has opposite direction
on the both sides of the middle plane, resulting in the clockwise moment M.

and right sides are the surfaces of negative and positive directions, respectively. Consequently, the
positive shear force V on the left side is pointing upward (Fig. 8.6(a)).
Consider that the plate is subject to the vertical load q(x) for a unit area on the middle surface.
The total vertical load for the element of width dx (Fig. 8.6(a)) is
 x+dx
dx
q(t) dt ≈ q + (q + dq) ≈ qdx,
x 2
where the trapezoidal rule is applied to the integral and the second-order term dxdq is neglected.
The shear forces V and V + dV and the vertical load must be balanced so that we have the equation
dV + qdx = 0. (8.16)

Moment balance Consider the right-handed coordinate system O-xyz. We assume that there is
no variation in the y direction in our system. The force balance equation (3.25) in the α-direction is
∂Sαα ∂Sβα
+ = 0.
∂α ∂β
Multiplying both sides of this equation and βdβ and taking the integral across the thickness, we have
 te /2  te /2
∂Sαα ∂Sβα
β dβ + β dβ = 0.
−te /2 ∂α −te /2 ∂β

Note that the position of the pivot is arbitrary in considering the moment balance (p. 67). Replacing
dα and dβ by dx and dz, we get the equation
 te /2  te /2
∂Sxx ∂Szx
z dz + z dz = 0. (8.17)
−te /2 ∂x −te /2 ∂z
180 CHAPTER 8. FLEXURE

Exchanging the integral and differentiation, the first term becomes ∂M/∂x, where
 te /2
M≡ Sxx z dz.
−te /2

This is the clockwise moment acting on the left side of the element in Fig. 8.6(b). The opposite side
of the element is subject to the counterclockwise moment M + dM. On the other hand, the second
term in Eq. (8.17) becomes
 te /2  te /2  te /2
∂Szx
z dz = Szx z − Szx dz = −V.
−te /2 ∂z −te /2 −te /2

t /2
In this calculation, we use the relationship Sxz = Szx . The term Szx z −te /2 vanishes, because the top
   e

surface of the lithosphere is a free boundary Szx  z=−t /2 = 0 and the basal drag by asthenospheric
   e


flow is neglected Szx z=t /2 = 0 . Consequently, Eq. (8.17) becomes6
e

dM
− V = 0. (8.18)
dx
Differentiating Eq. (8.18) and combining with Eq. (8.16), we get the equation

d2 M
+ q = 0. (8.19)
dx2
Using Eq. (8.6), we can rewrite this equation to the differential equation of the curvature D(d2 K/dx2 ) =
q, and the curvature is further replaced by w  , so that

d4 w
D = q. (8.20)
dx4
In case where the flexural rigidity is a function of position x, Eq. (8.6) is replaced by the equation
M = −D(x)K. This is substituted into Eq. (8.19), and we have
 2 
d2 d w
2
D 2 = q.
dx dx

If flexure is the function of not only x but also y, the downward displacement w(x, y) obeys the
equation D∇4 w = q, where ∇4 is the differential operator given in Eq. (C.55).

8.2.3 Effect of buoyancy


The lithosphere is subject to buoyancy when it is displaced vertically, because the downward shift
of the lithosphere pulls the low density (ρ1 ) superstratum downward and pushes the high density
6 The moment by the vertical load q(x) is negligible here. See Excercise 8.1
8.3. FLEXURE OF THE OCEANIC LITHOSPHERE 181

Figure 8.7: Downward deflection of elastic plate pushes the high density substratum away and makes
way for the low density superstratum. This density difference cause buoyancy to the plate.

(ρ2 ) substratum away (Fig. 8.7). The density difference causes buoyancy to the lithospheric plate.
Upward deflection causes negative buoyancy. The effect of buoyancy can be incorporated as the
vertical load q(x). If the plate is displaced downward by w(x), the buoyancy force per unit area of
the plate is
q = −Bw + qa , (8.21)
where
B = (ρ2 − ρ1 )g (8.22)
designates the effect of buoyancy and qa (x) stands for vertical loads other than the buoyancy. The
minus sign is attached to the buoyancy term Bw in Eq. (8.21) because the load q is defined to be
positive downward. If the mantle underlies the plate, ρ2 should be ρm . If the downward deflection
of the plate creates an ocean basin, ρ1 should be ρw , whereas if the basin is filled with sediments, ρ1
should be ρs . In case of an airless extra terrestrial body, ρ1 = 0.
The effect of buoyancy is incorporated by substituting Eq. (8.21) into (8.20), we obtain the
governing equation of the elastic plate subject to buoyancy:
d4 w
D + Bw = qa . (8.23)
dx4

8.3 Flexure of the oceanic lithosphere


Let us apply the thin elastic plate model to the oceanic lithosphere. It is important here that topog-
raphy can constrain the effective elastic thickness of the lithosphere. Accordingly, we can evaluate
the elastic strength of the lithosphere in the geological past from stratigraphic evidence or that of
terrestrial planets and satellites from their detailed topography7 .
There are multiple seamounts that work as topographic loads to the oceanic lithosphere. For
example, the Hawaiian Islands are on a great submarine ridge standing on the ocean floor at a depth
of 4 km, and extends thousands of kilometers across the Pacific (Fig. 8.8(a)). The ridge pushes down
the Pacific lithosphere to produce the Hawaiian Deep at both sides of the Hawaiian Archipelago by
its load at ∼100 MPa. The submarine topography around the ridge provides a constraint to the
effective elastic thickness of the lithosphere.
7 See [255] for further reading.
182 CHAPTER 8. FLEXURE

Figure 8.8: Submarine topography around the Hawaiian Ridge and northwestern Pacific. Contour
interval is 500 m. HD, Hawaiian Deep; MFZ: Molokai Fracture Zone; OTS, outer trench swell; PFZ:
Pioneer Fracture Zone.

Figure 8.9: Moment M exerted at a locked fracture zone by the differential subsidence of the oceanic
lithospheres on the other sides of the fracture zone.

Many oceanic trenches have topographic swells oceanward that are parallel to the trenches, called
outer trench swells. The swells shed light on the elasticity of the subducting slab. Figure 8.8(b)
shows the swell accompanied by the Kuril and Japan Trenches. The effective elastic thickness of
the oceanic lithosphere can be also evaluated at locked fracture zones [200]. The subsidence rate
of the lithosphere is inversely correlated with oceanic age. Therefore, their difference in depth is
large when they are young (Fig. 8.9(a)), but decreases with age (Fig. 8.9(b)). If the fracture zone is
mechanically locked, the decrease leads to a torque at the boundary that bends the lithosphere.
8.3. FLEXURE OF THE OCEANIC LITHOSPHERE 183

Figure 8.10: Graphs of Eqs. (8.28) and (8.31), respectively, showing the response of the continuous
(Case 1) and broken lithosphere (Case 2) to the vertical load V0 at x = 0. The horizontal axis
designates the dimensionless distance from the load.

8.3.1 Theory
Consider that the Hawaiian Ridge has a transverse profile similar to either of the curves shown in Fig.
8.10, so that the thin plate model developed in the previous section is applicable to the topography.
The topography on the northern side of the archipelago is typical where there is a moat and an outer
peripheral bulge along the islands (Fig. 8.8(a)).
Flexure of the lithosphere by a linear load is estimated by Eq. (8.23) combined with the buoyancy
B = (ρm − ρw )g. Here, the horizontal tectonic force is neglected for brevity, and the topographic
load is represented by the delta function qa (x) = −V0 δ(x). Consequently, the equation we have to
solve is
Dw  + Bw = −V0 δ(x). (8.24)
This is a linear ordinary differential equation with constant coefficients. Hence, we first solve the
homogeneous equation
w  + 4ka4 w = 0,
where 4ka4 = B/D. We have the general solution for this equation,

w(x) = C1 eka x sin ka x + C2 eka x cos ka x + C3 e−ka x sin ka x + C4 e−ka x cos ka x.

The constants C1 through C4 are determined from boundary conditions. Since w(x) should be an
even function, so that we investigate only the range 0 ≤ x. Firstly, the effect of the load may vanish
at x = ∞. This is satisfied by the condition C1 = C2 = 0. Therefore, the solution has the form

w(x) = C3 e−ka x sin ka x + C4 e−ka x cos ka x (x ≥ 0). (8.25)

Case 1 For the case 1 shown in Fig. 8.10, the lithosphere is horizontal just under the load so that
we have w  (0) = 0. Differentiating Eq. (8.25), we obtain
   
w  (x) = −ka e−ka x C3 cos ka x + C4 sin ka x + ka e−ka x −C3 sin ka x + C4 cos ka x (x ≥ 0).
184 CHAPTER 8. FLEXURE

Consequently, we have w  (0) = −ka C3 + ka C4 = 0, namely, C3 = C4 . The solution satisfying the


boundary conditions becomes
 
w(x) = C3 e−ka x cos ka x + sin ka x (x ≥ 0). (8.26)

The remaining constant C3 is determined from the load V0 .


The second-order derivative w  (x) is transformed into moment through Eqs. (8.5) and (8.6), i.e.,

M = −Dw  . (8.27)

Combining Eqs. (8.18) and (8.27), we have V = −Dw  . Substituting Eq. (8.26), we obtain

V (x) = 4C3 Dka3 e−ka x cos ka x.

V0 /2 is the load acting at x = 0 and is supported by the one side of the load x > 0, so that we have
C3 = V0 /4Dka3 /2. Therefore, the flexure near the linear load has the form
 
V0 λ3f e−x/λf x x
w(x) = cos + sin (x ≥ 0), (8.28)
8D λf λf

where λf = 1/ka has dimensions of length, called the flexural parameter, defined by the equation
9
4D
λf = 4 . (8.29)
(ρm − ρw )g

This is the distance required for the effect of a vertical load to become 1/e. One other parameter
important for lithospheric bending is the wavelength

λw = 2πλf . (8.30)

Namely, the peaks of the peripheral bulges on both sides of the linear load is separated by this
distance. One of the peaks has the position xs = πλf , which is easily recognized from the topography.
Regarding the Hawaiian Ridge, xs ≈ 250 km. Therefore, the flexural rigidity and effective elastic
thickness of the Pacific plate is estimated at D = 2.4 × 1023 Nm and te = 34 km using the parameters
Y = 70 GPa, ν = 0.25, ρm = 3.3 × 103 kg m−3 , ρw = 1.0 × 103 kg m−3 , and g = 10 ms−1 .
3/4
It is seen from Eqs. (8.14) and (8.29) that the λf is proportional to te . A water-loaded litho-
3/4
sphere has the approximate equation λf ≈ 31te . For a lithosphere with te ∼ 100 km, te is a few
times larger than te . A sediment-loaded lithosphere has a constant of proportionality at around 45,
so that λf is ∼5 times greater than te if te ∼ 100 km.

Case 2 Hotspot magmatism may heat up and weaken the moving lithosphere, just like a gas burner
weakens an iron sheet. In order to take this effect into account, the Case 2 in Fig. 8.10 assumes the
thin elastic plate is broken beneath the load at x = 0. The flexure of this lithosphere in the range
8.4. BUCKLING 185

0 ≤ x is obtained from Eq. (8.25). The plate is broken at x = 0, so that the plates in the ranges
x < 0 and 0 < x are thought to be connected by a hinge joint at x = 0. That is, we assumed zero
moment at the point M (0) = 0. From Eq. (8.27) we have

M = −Dw  (0) = 2DC3 k 2 = 0 ∴ C3 = 0.

In addition, the shear force at the same point is V0 /2, which constrains w  (0) to give the equation
V0 λf e−x/λf x
w(x) = cos (x ≥ 0). (8.31)
4D λf
Differentiating Eq. (8.31) we obtain that the distance of the peak of the peripheral bulge is
V0 λf e−x/λf  x x
w  (x) = − cos + sin (x ≥ 0).
4Dλf λf λf
The equation w  (xx ) = 0 has the solution xs = (3/4)πλf , indicating the position of the peripheral
bulge. This is slightly smaller than that for Case 1.

8.3.2 Effective elastic thickness versus oceanic age


Figure 8.11 shows the effective elastic thickness te versus the age of the lithosphere estimated in
various regions. It is obvious that they have a good correlation. The thickness estimated from the
bathymetry around seamounts is about the depth of the 400◦ C isotherm (Fig. 8.11(a)), suggesting
that the thickness is controlled by the thermal regime of the lithosphere. Figure 8.11(b) shows the
same relationship estimated from the outer trench swells, supporting the correlation. However, the
data points are plotted around the 600◦ isotherm instead of 400◦ , i.e., te seems larger at the outer
trench swells than around the seamounts.
This discrepancy possibly reflects three factors; reheating of the lithosphere at hotspots, thermal
stress, and yielding of the lithosphere under seamounts [256]. The thermal history of the oceanic
lithosphere depicted in Fig. 8.11 does not take these effects into account. The reheating leads
not only to the rejuvenation of the lithosphere but also to the apparent decrease of te , because the
downwarp of the lithosphere due to the vertical load is additionally enhanced by thermal stress. The
yielding gradually decreases the effective elastic thickness (§10.4). However, the yielding may occur
at the outer trench swell to result in the scattering of the data points in Fig. 8.11(b).

8.4 Buckling
Engineers have studied deformation of elastic materials because structures including buildings and
machines are made of solid materials that should support possible loads. Buckling is one of the
important issues in that field, and is the failure of an elastic plate or rod by compression parallel to
the plate or rod.
Folds are basic geological structures. The model of thin elastic plates is applicable to this prob-
lem if horizontal stress is taken into account.
186 CHAPTER 8. FLEXURE

Figure 8.11: Oceanic age versus the effective elastic thickness of the lithosphere [256]. Parabolas
designate descending isotherms with age estimated from the half-space cooling model. (a) The
thicknesses determined around seamounts in French Polynesia () and in other regions (◦). (b) The
thicknesses determined from the topography () and gravity anomaly () of the outer trench swells.
The gray region indicates the thickness around the Mendocino Fracture Zone.

8.4.1 Basic equations


Consider the force balance in the x direction of the element of the thin elastic plate shown in Fig.
8.12. The element is thought to have a unit length in the y direction that is perpendicular to the page.
Assuming no variation of the system in the y direction, derivatives with respect to y should vanish.
The force balance equation (Eq. (3.25)) becomes

∂Sxx ∂Szx
+ = 0. (8.32)
∂x ∂z
Integrating normal stress Sxx (1 × dz) over a cross-section is the force normal to the section on which
it acts, where (1 × dz) is the area of the section. Let us write the force as
 te /2
N≡ Sxx dz. (8.33)
−te /2

Integrating both sides of Eq. (8.32), we have


 te /2  te /2
∂Sxx ∂Szx ∂N  te /2
dz + dz = + Szx (∗)
−te /2 ∂x −te /2 ∂z ∂x −te /2

∂N
= = 0. (8.34)
∂x
The last term in Eq. (∗) vanishes due to the following boundary
 condition. Firstly, the surface is a
free boundary so that the shear traction at the surface Szx −t /2 is zero. Secondly, the lithosphere
e
is assumed to be underlain by an inviscid asthenosphere, so that the shear traction at the interface
8.4. BUCKLING 187

Figure 8.12: Moment at the sides of inclined plate element due to the forces N and N + dN. The
white dot-bar line indicates the middle surface of the plate that is parallel to the xy plane when the
plate is unloaded.


Szx t /2 is negligible. Equation (8.34) indicates that N is constant in the x direction if all forces are
e
balanced.
Suppose a horizontal force N acting at the ends of the elements. The plate goes down to the right
if (dw/dx > 0). Hence, the forces result in a counterclockwise moment Ndw. The moment balance
equation derived in Section 8.2 should include this factor. Namely, Eq. (8.18) is replaced by

dM − V dx + Ndw = 0. (8.35)

Buckling occurs due to compression so that the compressive force F = −N is more convenient for
the present discussions than the tensile force N. Equation (8.35) is rewritten as dM − F dw − V dx =
0. Rearranging this equation, we have

d2 M dV d2 w
= + F . (8.36)
dx2 dx dx2
Further, combining Eqs. (8.5) and (8.6), we obtain

d4 w d2
D = q − F .
dx4 dx2
When there is no vertical load (q = 0), this equation is rewritten by replacing w  by K (Eq. (8.5)):
 2 
d F
+ K = 0.
dx2 D
The general solution of this equation has the form K(x) = C1 sin kx + C2 cos kx, where C1
and C2 are integration constants, and k = F/D. C1 = C2 = 0 is a possible but trivial solution,
corresponding to layer-parallel shortening. Non-trivial solutions represent buckling modes. If the
plate is constrained vertically at its ends, x = ±L/2, K(x) should be an even function (Fig. 8.13),
i.e., we have K(x) = C2 cos kx. In addition, we assume that there is no torque at the ends. This
leads to the equation, K(±L/2) = C2 cos(Kk/2) = 0. After the plate buckles, C2 = 0 and we have
Lk Lk πn
cos =0 ∴ =± (n = 1, 3, 5, . . . ).
2 2 2
188 CHAPTER 8. FLEXURE

Figure 8.13: Plate at buckling mode under a compressive force acting at the both ends, x = ±L/2.


Combining this equation and k = F/D, we obtain

π2D
F = n (n = 1, 3, 5, . . . ). (8.37)
L2
Therefore, the force depends on n, which distinguishes the different buckling modes. The smallest
value corresponding to n = 1 is called Euler’s critical buckling load8 . A thin elastic plate buckles
when layer-parallel force is piled up to reach the critical load F = π 2 D/L2 .
Consider a folded sedimentary rock with a thickness of a few kilometers to estimate the critical
load. Figure 1.15 shows an example. If the rock has the parameters Y = 30 GPa and μ = 0.25 and
the the thickness te = 1 km, then we have D = 2.7 × 1018 Nm. If the half wavelength is L = 5
km, the minimum force needed to fold is F = 1.1 × 1012 N. The average normal stress acting on the
sides of the sedimentary rock mass is F/(1 km × 1 m) = 1.1 GPa. On the other hand, the average
overburden stress is (ρs g × 1 km)/2 = 12 MPa, if ρs = 2.5 × 103 kg m−3 , i.e., the tectonic force
needed to buckle sedimentary rock with a thickness of 1 km is twice greater than a representative
overburden stress, suggesting that the sedimentary pile cannot be buckled by horizontal compression
if the above model is valid for the folding. The fact is that the model is not valid. What is wrong is
that we have assumed that the thickness of the sedimentary pile was equal to that of an elastic plate.
A sedimentary pile has bedding planes, some of which work as weak planes for frictional sliding
to cause flexural-shear folding (Fig. 8.14). It is obvious that a sheaf of paper is much more easily
bent compared to a plate of wood of the same thickness, although both paper and wood are made
from the same substance, i.e., cellulose fibers. Sliding between sheets of paper weaken the flexural
rigidity of the sheaf.

8.4.2 Effective elastic thickness of a laminated plate


Consider a plate composed of two thin elastic layers which are lubricated between. What is the
effective elastic thickness of this composite plate if each of the layers has the thickness t? The layers
have the same elastic constants.
The clue to this problem is Eq. (8.6), which represents the proportionality between bending
moment M and the curvature K. If the radius of the curvature of the inner layer is R, that of the outer
8 The critical buckling load depends on the boundary conditions at the ends of the plate [50]. However, the difference in

the critical load is less than one order of magnitude. The purpose of the above discussion is the order of magnitude estimation,
so that the difference is negligible here.
8.4. BUCKLING 189

Figure 8.14: Flexural-shear folding.

layer is approximated by R + t. However, the layer is assumed to be thin. Therefore, it is allowed


to take both radii to be equal. They have, therefore, the same curvature K. Moments are additive
quantities, so that the composite plate supports the moment M = −2DK, where D = Y t3 /12(1−ν 2 )
is the flexural rigidity of each of the plates (Eq. (8.14)). Let D = Y t3e /12(1 − ν 2 ) be that of the
composite plate, then we have
  3   3   t 3  3
KY te te KY e KY te
D= + = ·2 = √3 .
12(1 − ν )2 2 2 12(1 − ν )
2 2 12(1 − ν )
2
4
√3
As 1/ 4 ≈ 0.63, a plate composed of two equally thin elastic layers has a effective elastic thickness
of about 63% of the thickness of the composite plate. It is obvious from the above argument that if
many layers make up a composite plate, the effective elastic thickness of the latter te satisfies

t3e = t31 + t32 + · · · + t3n ,

where ti is the thickness of the ith layer. The effective elastic thickness decreases as te = n−2/3 (nt), if
the plate is divided into n layers with the same thickness t and the same elastic constants. Therefore,
flexural-shear folding is easy for sedimentary piles [185]. The approximate equation te ≈ max(t1 , t2 )
is sometimes used when n = 2.

8.4.3 Wavenumber selection


It was assumed that the elastic plate had a width of L in Section 8.4.1, and that the wavelength
of the buckled layer was prescribed at 2L. By contrast, the layer has a fluid substratum and the
system consisting of the layer and fluid determines the wavelength by itself as follows. The inverse
of wavelength is called a wavenumber, i.e., the number of peaks contained in a unit length. Hence,
the determination is called a phenomenon of wavenumber selection.
If a layer with a fluid substratum buckles under the critical load Fc , the folded layer pushes the
denser substratum downwards under synclines and pulls it up under anticlines (Fig. 8.15). Con-
sequently, folding makes horizontal density variations, which further lead to positive and negative
buoyancy of the layer. The equation of the plate flexure with buoyancy effects is

Dw  + F w  + Bw = 0. (8.38)
190 CHAPTER 8. FLEXURE

Figure 8.15: Wavelength selection by buckling of elastic plates with a fluid substratum.

The last term in the left-hand side is the buoyancy term and we have B = Δρg. Let λc be the
wavelength of the fold that is formed by the critical load. Here, a sine curve

w = A sin ax (8.39)

is assumed to represent the fold, where a is related to the wavelength as a = 2π/λc . Substituting Eq.
(8.39) into (8.38), we have Da4 A sin ax − F a2 A sin ax + BA sin ax = 0, and rearranging gives the
quadratic equation of (a2 ): D(a2 )2 − F (a2 ) + B = 0. We obtain its solutions
 2 
2π F ± F 2 − 4DB
a =
2
= . (8.40)
λ 2D

The critical load is defined by the minimum force F to make the discriminant (F 2 − 4DB) positive,
i.e., we have Fc2 − 4DB = 0, where 4DB is a constant specific to this system. Namely, the critical

load is determined by Fc = 2 DB. Substituting this force into Eq. (8.40), we obtain the critical
wavelength 
λc = 2π 4 D/B. (8.41)
This system is always unstable if the substratum is denser than the elastic layer. In this case,
we have B < 0 and the discriminant in Eq. (8.40) is positive regardless of the horizontal force
F . Such an inverted density stratification occurs at the top of a layer of natural rock salt or of
an uncompacted strata. Figure 8.16 shows an example of load casts, a train of depressions in a
bed of sediment developed by the differential sinking of the sediment, while still soft, into a less
dense sediment blow [2]. The white fine tuff in Fig. 6.16 was probably less dense than the overlying
massive tuff breccia, so that they have an undulating boundary. The fracturing in the breccia triggered
the instability to produce the clastic dike at the peak of an anticline of the boundary.
Gigantic folds are found in the northeastern Indian Ocean where the oceanic lithosphere makes
E–W trending folds that accommodate a intra-plate deformation with a wavelength of ∼200 km.
Undulations are observed in gravity anomaly also, suggesting that not only a deep sea sedimentary
pile but also the oceanic crust as a whole is folded. The folding probably began in the Late Miocene
[43]. The effective elastic thickness of the oceanic lithosphere has a good correlation with age. The
age of the region is about 55 Ma, which corresponds to te = 30–40 km if we regard the base of the
elastic plate at a 400–600 C◦ isotherm (Fig. 8.11). Using the parameters, Y = 60 GPa, ν = 0.25,
g = 10 ms−2 and ρm − ρw = 2.3 × 103 kg m−3 and Eq. (8.41), we obtain the critical wavelength at
8.4. BUCKLING 191

Figure 8.16: Load cast in Miocene shale, Morozaki, central Japan. Note the horizontal variation of
the thickness of the light-colored thick mudstone upon which a hammer is placed. The variation in-
dicates that the mudstone had not been consolidated when this structure was formed. The mudstone
was probably less denser than the overlying dark-colored one. This structure is truncated by an ero-
sional surface that represents a negligible interval in depositional age, suggesting that the structure
was formed just below the sea floor while the sediments were still soft.

310–390 km, significantly greater than the observed wavelength. That is, te appears to be smaller in
the region than that of the ordinary oceanic lithosphere. It is suggested that the plate is yielded by a
large compressive stress 9 [136].
Let us apply the model to the folds in Fig. 1.15 to see how the critical stress is reduced from
that estimated on p. 188. Using the values Δρ = 0.1 × 103 kg m−3 and D = 2.7 × 1018 Nm, which
were evaluated on p. 188, we have the critical wavelength λc = 81 km, which is much longer than
the actual wavelength at around 5 km. Hence, substituting λc = 5 km into Eq. (8.41), we obtain
D = 4.0 × 1015 Nm. This is 1.5 × 10−5 times the previous estimate. The corresponding effective
elastic thickness is 24 m. If the sedimentary rock mass is divided into n elastic layers, the effective
elastic thickness of the mass with a thickness of 1 km becomes n−2/3 × 1 km = 24 m. Therefore, we
have n ≈ 260, and the mass has slip surfaces every 1 km ÷ 260 = 3.8 m. The critical load is, in this

case, Fc = 2 DB ≈ 1.3 × 109 N, so that the average horizontal stress is Fc ÷ (1 km×1 m) ≈ 1.3
9 The yielding of the lithosphere is investigated on p. 322 after plasticity is introduced.
192 CHAPTER 8. FLEXURE

MPa. This is a possible magnitude for tectonic stress.


Flexural-shear drastically diminishes the force needed for folding. This idea is demonstrated in
the field by Jackson et al. [90]. They found slip surfaces on bedding planes with intervals of several
meters in an anticline upon a laccolith. In addition, fault striations (§11.2) had concordant directions
with those expected from the flexural-slip folding.

8.5 Rift shoulder uplift


An extensive basin or a wide valley acts as a negative load to pull the lithosphere upward. It is
important that uplift is not always the result of crustal thickening. Erosion can raise the lithosphere.
Therefore, many rift valleys are bounded on both sides by mountains. This upheaval is known as rift
shoulder uplift, the surface expression of elasticity of the lithosphere.

8.5.1 Rift on the Satellite Miranda


Continental rifting is associated with vertical movements by multiple factors including crustal thin-
ning, thermal isostasy, lithospheric elasticity, and magma genesis. Therefore, let us consider passive
margins, inactive or very slowly opening rifts for simplicity. Firstly, we deal with a rift on the
satellite Miranda (Fig. 8.17).
Miranda is a small moon with a radius of 236 km. However, its icy crust has a variety of
geological features. Among them is a grooved oval terrain, called corona. The less denser impact
craters indicate the younger age of a corona. Pappalardo et al. [172] regard the grooves of the Arden
Corona as normal faults. Figure 8.17(b) shows the topographic section through the grooved region
composed from Voyager images. The large valley in the center of this section has uplifts on both
sides, so that Pappalardo et al. were able to estimate the effective elastic thickness there.
The lithosphere under a rift zone suffers a large strain, which can decrease the effective elastic
thickness (§12.3). A raised geothermal gradient by rifting can also reduce the thickness. Therefore,
we assume the lithosphere is broken under the rift, as illustrated in Fig. 8.10(b), and apply Eq.
(8.31) with the negative load by the large valley. The load is written as V0 = −ρi W d/2, where
ρi ≈ 1.0 × 103 kg m−3 is the density of the ice. Using the parameters ν = 0.33, Y = 10 GPa and the
gravitational acceleration on Miranda g = 0.08 ms−2 , the flexural rigidity and the effective elastic
thickness are evaluated from this valley as D = 6 × 1018 Nm and te = 1.7–1.8 km, respectively.

8.5.2 Simple shear rift


If extension occurs by displacement along a normal detachment fault that extends completely through
the crust, the resultant asymmetrical rift is called a simple shear rift (simple shear rift). The Gulf of
Suez rift [227] and the one between East and West Antarctica (Fig. 8.19) are typical examples.
Consider a normal fault dipping at an angle of θ through the continental lithosphere with the
8.5. RIFT SHOULDER UPLIFT 193

Figure 8.17: Uranian satellite Miranda. Courtesy of NASA. Topographic profile along the white line
is shown in Fig. 8.18.

Figure 8.18: Topographic profile along the white line in Fig. 8.17 across the margin of the Arden
Corona, Miranda [172].
194 CHAPTER 8. FLEXURE

Figure 8.19: (a) The Ross Sea was formed as a simple shear rift between East and West Antarctica
[230], and the rift shoulder uplift has formed the Trans Antarctic Mountains [34]. (b) Topography
along the transects designated by dotted lines in (a) across the mountain range and the Wilkes Basin.
The profiles are tuned up in their vertical positions at the base of the basin. The origin of the abscissa
marks the coast.

horizontal displacement dh (Fig. 8.20(a)) [257]. Then, we have the surface topography

⎨ 0 (x < 0),
z0 (x) = x tan θ (0 ≤ x < dh ), (8.42)

dh tan θ (dh ≤ x),
where x is the horizontal coordinate perpendicular to the normal fault and z points downwards.
These difference levels are unstable because of the buoyancy from the asthenosphere. The right side
of the fault may be uplifted and the other side may subside, resulting in a counterclockwise moment
at the rift zone. Consequently, the topography is written as z = z0 + w, where the last term represents
the flexure of the lithosphere.
The flexure is calculated as follows. Since the density of rocks depends on temperature, the ther-
mal regime of the lithosphere has to be determined. To this end, we assume constant temperatures
T0 and Ta at the surface and base of the lithosphere, respectively, and obtain the thermal evolution
T (x, z, t) by solving the heat conduction equation ∂T/∂t = κ∂2 T/∂z2 . Domains 1 through 5 in Fig.
8.20 have different boundary conditions. For simplicity, we assume that the duration of faulting is
negligible compared to heat conduction through the lithosphere. Density distribution is obtained by
the formulas ρc = ρ0c (1 − αT ) and ρm = ρ0m (1 − αT ), which are transformed to the buoyancy force
V (x, t). Substituting this into Dw  + Bw = V , we obtain the flexure w(x, t). In the case where the
450◦ C isotherm is thought to mark the base of the elastic lithosphere, the flexural rigidity D is a
function of position x as the temperature depends on x and z. The flexure equation to be solved for
this case is
∂2  ∂2 w(x) 
2
D(x) + Bw(x) = V (x).
∂x ∂x2
Figure 8.20(b) shows the results for the case where θ and dh are assumed to be 20◦ and 10
8.6. THE DEGREE OF COMPENSATION 195

Figure 8.20: Topography associated with a simple shear rift [257]. (a) Schematic picture for the
rift. (b) Topographic profiles predicted by an elastic plate model with the parameters a = 125 km,
tc = 32 km, θ = 20◦ C, e = 10 km, and t = 0 Ma. The lithosphere is modelled as an elastic plate
with te = 0, 5 and 20 km and as a thermal boundary layer whose base is marked by the 450◦ C
isotherm. Uplift in the domains 1 and 5 for the case of te = 0 is due to thermal isostasy.

km, respectively. Consequently, rift shoulders can be uplifted by a few kilometers. The uplift is
enhanced by increasing te . The topographic profile through the Antarctic Mountains is consistent
with the flexural uplift of an elastic plate with te = 115 km. The mountains are high because the
Antarctic shield has a thick lithosphere (Fig. 8.19).

8.6 The degree of compensation


Long-wavelength topographic loads are supported by buoyancy, but short ones are compensated by
the elasticity of the lithosphere (§8.1). Let us estimate how the two factors bear part of the loads. To
this end, we assume sinusoidal topography

h(x) = h0 sin(2πx/λ),
196 CHAPTER 8. FLEXURE

Figure 8.21: (a) The relationship between te and λ0 . (b) The relationship between λ and C. Both
graphs are produced with the parameters, Y =70 GPa, ν=0.25, ρm = 3.3 × 103 kg m−3 , ρc = 2.8 × 103
kg m−3 and g=10 ms−2 .

where the amplitude h0 is much smaller than the thickness of the lithosphere. The corresponding
topographic load is
qa (x) = ρc gh0 sin(2πx/λ).

For simplicity, we assume a zero horizontal load (F = 0), and we have the governing equation for
the plate flexure,
2πx
Dw  + Bw = ρc gh0 sin , (8.43)
λ
where B = (ρm − ρc )g. The cause of the flexure is the sinusoidal load so that the resultant flexure
may have the same form. That is, we assume a solution of the form w = w0 sin(2πx/λ). Substituting
this into Eq. (8.43), we obtain
⎡ ⎛9 ⎞4 ⎤−1
ρc gh0 λ4 ρ D 2π
= h0 ⎣ +⎝4 ⎠ − 1⎦ .
m
w0 = · (8.44)
16Dπ 4 + Bλ4 ρc ρc g λ

The amplitude of flexure w0 is a function of the length λ. The amplitude has positive correlation with
λ, and we have limλ→+0 w0 = 0 and limλ→∞ w0 = h0 /(ρm /ρc − 1). The coefficient of the angular
frequency (2π/λ) in Eq. (8.44) is
9
D
λ0 = 2π 4 , (8.45)
ρc g

which is similar to the flexural parameter λf defined by Eq. (8.29). Equation (8.44) is rewritten as

 4 !−1
w0 ρm λ0
= + −1 . (8.46)
h0 ρc λ
8.7. EFFECTIVE ELASTIC THICKNESS ESTIMATED FROM GRAVITY ANOMALIES 197

The parameter λ0 indicates a length scale characteristic to the elastic plate with the flexural
rigidity D. Figure 8.21(a) shows the graph of
⎡ 9 ⎤
Y
λ0 = ⎣2π 4
⎦ t3/4
e .
12(1 − ν)ρc g

If the interval between mountains is much smaller than λ0 , the periodic loads hardly bend the elastic
plate because λ  λ0 leads to w0  h0 . Instead, the plate subsides as a whole by the loads. On
the other hand, if λ is much larger than λ0 , we have (w0 /h0 ) → ρc /(ρm − ρc ), indicating that the
loads are supported by local isostasy. Buoyancy from the mantle and the lithospheric elasticity play
comparable roles for the loads with the intervals λ ≈ λ0 .
The degree of compensation is a convenient measure to estimate their contributions and is defined
by the equation
 
w0 ρm − ρc ρ m − ρc
C= =  4 . (8.47)
h0 ρc ρ + ρ λ0 − 1
m c λ

The cases C = 0 and 1 correspond to regional and local isostasy, respectively. Local isostasy is
largely satisfied if λ is at least a few times larger than λ0 (Fig 8.21b).

8.7 Effective elastic thickness estimated from gravity anomalies


Lithospheric flexure is recognized from gravity anomalies, where the scale-dependence of the degree
of compensation is important. A long-wavelength topographic high is compensated by the buoyancy
of the underlying low-density crustal root (Fig. 8.22(a)). By contrast, small-scale undulations are
supported by the elasticity of the lithosphere or more locally by that of the crustal rocks.
Free-air correction is a mathematical operation to observe the gravity anomaly that is the devi-
ation of the observed gravitational acceleration from that predicted for the rotating Earth ellipsoid.
The observed one includes the attractions from the rocks above and below sea level. The former
decreases with altitude following the law of inverse square. Free-air correction compensates this
decrease to calculate the free-air anomaly, and the Bouguer correction removes the effect of the at-
traction from the rocks above sea level from the free-air anomaly. The result is known as the Bouguer
anomaly. However, the Bouguer correction leaves the effect of the low density root under the exten-
sive topographic high (Fig. 8.22), which results in a long-wavelength negative Bouguer anomaly.
On the other hand, the free-air anomaly is not only affected by the root but also by the mass above
sea level, so that the long-wavelength free-air anomaly vanishes.
A horizontally large-scale topographic high has a low density root, but small-scale ones do not.
This contrast is exhibited by the wavelength-dependence of the Bouguer gravity anomaly. At the
same time, the contrast reflects the degree of compensation of the long- and short-wavelength loads
which is determined by te of the lithosphere. Therefore, the relationship between the gravity anomaly
198 CHAPTER 8. FLEXURE

Figure 8.22: Gravity anomaly and the flexure of the lithosphere. (a) Elastic layer downwarped by
the topographic load of a mountain range with a long wavelength. (b) Bouguer correction mathe-
matically removes the effect of the mass of the mountain range but leaves the effect of the negative
density anomaly under the range.

and topographic load for various wavelengths is a clue to te . Let h(x) and g(x) represent the topogra-
phy and gravity anomaly, respectively. Then we have the cross-correlation and its Fourier transform
∞ ∞
: 1
C(x) = g(x)h(x − x ) dx , C(k) =√ C(x)e−kx dx.
−∞ 2π −∞

The latter is the cross-spectrum between the two functions representing the correlation for each
wavenumber k. The Fourier transform of the functions
∞ ∞
1 −kx : 1
:
g (k) = √ g(x)e dx, h(k) = √ h(x)e−kx dx
2π −∞ 2π −∞
are used to normalize the cross-spectrum as
 
 2 C(k)
: 2
Ch (k) =    2 .
g:(k)2 :
h(k)
8.7. EFFECTIVE ELASTIC THICKNESS ESTIMATED FROM GRAVITY ANOMALIES 199

Figure 8.23: Effective elastic thickness estimated by the coherence method [18]. (a) Coherence
versus wavenumber for southeast Canada and the theoretical curve that best-fits the observation. (b)
Map showing the distribution of the thickness.

Ch (k) is known as coherence which is related to the degree of compensation to estimate the effective
elastic thickness [64]. Figure 8.23(a) shows coherence versus wavenumber for the southwestern
Canada, where there is an upheaval in the coherence between the wavelengths from 200 to 800 km.
The effective elastic thickness was determined at 39 km by fitting a theoretical curve to the data.
The coherence method has revealed that the effective elastic thickness varies according to the
locality. Figure 8.23(b) shows the variation of te in North America. The Canadian shield has a
maximum te > 120 km in the continent. The thickness is the minimum in the southwestern United
States where te is smaller than 4 km. The thickness largely increases northwestward from this region.
This trend affected the architecture of Mesozoic foreland basins accompanied by the Cordilleran
orogenic belt [261], indicating the same trend also in the Mesozoic.
The thickest region coincides with the stable core of the continent where there has been little
volcanism and tectonic deformations in the Phanerozoic. In contrast, the thinnest region, including
the Basin and Range Province and California, has active tectonics. These results suggest that te
has a positive correlation with tectonic age, the length of time since the last major tectonic phase.
However, the correlation is not so good as the case of the oceanic lithosphere [255].
Active island arcs have the tectonic age at zero. However, it was found that there are variations
in te in island arcs in Japan. Kudo et al. [105] estimated te by the coherence method for 11 areas of
Japan (Fig. 8.24(a)). Figure 8.24(b) shows the thickness in the 11 areas. The estimated te values
exhibit a good correlation with the heat flow averaged over each of the areas (Fig. 8.24(b)). Hot
areas have small effective elastic thicknesses.
However, there are deviations. Areas H, I and J in Southwest Japan had not enough dimension to
catch the upheaval in their coherence-wavelength diagrams so that the lower bound for their te at 20
200 CHAPTER 8. FLEXURE

Figure 8.24: Effective elastic thickness of the island arc lithosphere [105]. (a) Eleven windows, A
through K, in which the thickness was estimated by the coherence method. (b) The thickness versus
average heat flow in the windows. In the regions H, I and J, only the lower bound of the thickness
was determined.

km is depicted there. The main island of Japan is too narrow for full use of the method. The large te
in those regions was attributed to the elasticity of the shallow dipping Philippine Sea slab and to the
negative dynamic topography (§9.5) induced by the slab [105, 106]. On the other hand, the region
K has a small te , about half of the thickness expected from the average heat flow. This is explained
by the extensional tectonics in the area10 [105].
Heat flow is found to have a correlation with te better than the tectonic age in the African Conti-
nent [72]. However, detailed investigation of te in the Canadian shield shows not so good correlation
with heat flow [254]. The factors that can control te of the continental lithosphere is discussed in the
last chapter.

8.8 Lunar mare tectonics


The Moon has various tectonic features, many of which have surface topographic expressions in-
cluding wrinkle ridges and linear rilles (Fig. 1.14). Circular mare basins on the Moon appear to
have characteristic tectonism as they have concentric wrinkle ridges and linear rilles. The Mare Se-
lenitatis is a typical one (Fig. 8.25). Those tectonic features were generated by the surface stress
field resulting from the flexure of the lithosphere by the loading of the mare deposits.
The circular mare basins were formed by massive impact cratering that blew the ancient lu-
nar crust away. The crustal thinning led to the formation of circular basins in which flood basalts
and ejecta from craters were accumulated up to a thickness of a few kilometers. Since the basalts
are denser than the anorthositic crust, the thick mare deposits rich in basaltic lavas as loads to the
10 See §12.3 for the relationship between effective elastic thickness and stress regimes.
8.8. LUNAR MARE TECTONICS 201

Figure 8.25: Geologic structures in and around the Mare Selenitatis on the Moon [221]. Mare
deposits are divided into three stratigraphic units, I, II and III. They are 3.5–3.7, 3.5, and 3.0–3.4
Gyr old. Figure 1.14 shows the southeastern part of the mare.

lithosphere. Most circular mare basins have positive free-air anomaly over 100 mgal, suggesting
the existence of high density layers supported by the elasticity of the lithosphere. Those basins are
called the mascon basins.
The mascon basins is as large as 1000 km in diameter, about half the radius of the Moon (R ≈
1740 km), therefore, we have to deal with the flexure of elastic spherical shell instead of that of
elastic plate. Let R be the radius of the thin spherical shell with the thickness te  R. We use
the spherical coordinates O-θφ with the colatitude θ is zero at the center of the basin. The vertical
flexure w  te by axially symmetric loading q(θ) is described by the equation [23]

Y te
D∇4 w + w + Bw = q. (8.48)
R2
Unlike elastic plates, formation of a dimple on a spherical shell leads to horizontal compression. The
second term in Eq. (8.48) represents this effect. Using the distance x that is nondimensionalized by
the representative length of this system
9
D
≡ 4 ,
Y te
R2
+B

Eq. (8.48) is rewritten as ∇4 w(x) + w(x) = q(x) 4 /D. Consider a disk with a constant area density
q0 for the load. Then, we have q(x) = q0 H (d − x), where H ( ) is Heaviside’s step function and
d is the radius of the disk. Since the load does not depend on φ, the separable solution of the
202 CHAPTER 8. FLEXURE

Figure 8.26: Kelvin-Bessel functions of the 0th order.

homogeneous equation ∇4 w(x) + w(x) = 0 has the solution of the form

w(x) = A0 Ber0 (x) + B0 Bei0 (x) + C0 Ker0 (x) + D0 Kei0 (x),

where A0 , B0 , C0 and D0 are integration constants, and Ber0 (x), Bei0 (x), Ker0 (x) and Kei0 (x) are
Kelvin-Bessel functions of the 0th order [223]. Figure 8.26 shows the graphs of the functions. We
obtain the solution satisfying the boundary condition q(x) = q0 H (d − x)

q0 d   
w= Ker0 (d) Ber0 (x) − Kei0 (d) Bei0 (x) + 1 (x ≤ d)
A+B
q0 d   
w= Ber0 (d) Ker0 (x) − Bei0 (d) Kei0 (x) (x > d),
A+B
where the primes indicate differentiation by x [220]. The resultant horizontal stress is given by

q0 Y te
)
1 d  *
Sφφ = + Ker0 (d) Bei0 (x) + Kei0 (d) Ber0 (x) (x ≤ d)
R(A + B) 2 x
q0 Y te
) 2
1d d  *
  
Sφφ = + Ber0 (d) Kei0 (x) + Bei0 (d) Ker0 (x) (x > d).
R(A + B) 2 x2 x

For both cases, the radial stress Sθθ satisfies

Sθθ = −Sφφ + ARw, (8.49)

where A = Y te /R2 . The radial stress raised by various q(θ) is obtained as follows. Figure 8.27(a)
shows the distribution of the load of Unit I in the Mare Selenitatis. The distribution is approximated
by a pile of disks, each of which has a constant area density. Since the governing equation (Eq.
(8.48)) is linear, the radial stress due to the pile is the superposition of the solutions for the disks.
The radial stress is visualized by the concentric tectonic features. Figure 8.27(b) shows the solutions
for the three cases te = 25, 50, and 100 km.
8.9. EXERCISES 203

Figure 8.27: Surface horizontal stress due to the load of Unit I in the Mare Selenitatis [220]. (a)
Distribution of the load. (b) Variation of the stress. Positive Sθθ indicates horizontal extension.
Grabens are formed in Unit I in the zone around r = 400 km.

The deposits in the Mare Selenitatis are divided into three stratigraphic units. Wrinkle ridges
affects all the units, but linear or arcuate rilles are formed only on the oldest Unit I. Detailed strati-
graphic and structural studies revealed that the rilles exist only on geological units older than ∼3.6
Ga, suggesting that extensional tectonics ceased at that time [127]. Figure 8.27 suggests the effec-
tive elastic thickness of the lithosphere in and around the Selenitatis region when Unit I deposited
as follows. The concentric linear or arcuate rilles, which indicates radially extensional stress, exist
only in the gray zone in Fig. 8.27(b). This is interpreted to suggest that extensional radial stress
peaked in the zone. This is best satisfied for the case of te = 50 km. Therefore, it is possible that the
effective elastic thickness of the lithosphere in this region was about 50 km when Unit I deposited at
36–37 Ga [220].
Solomon and Head [221] estimated te for several mare basins on the Moon for two ages, the
early Imbrian Period (∼3.6 Ga) and the Eratosthenian Period (1.0–3.2 Ga). Their results show that
te increased between the ages, suggesting cooling of the lithosphere in those regions.

8.9 Exercises
8.1 Show that the moment by the vertical load q(x) is negligible when we consider the moment
balance of the element of an elastic plate (p. 180).
204 CHAPTER 8. FLEXURE

8.2 Show that the flexture w(x) caused by a distributed load q(x) is given by the equation
  ∞ 
1 ∞ 1 −ikx
w(x) = q(x)e dx eikx dk,
2π −∞ (ρm − ρw )g + Dk 2 −∞

if the lithosphere is water-loaded [110].


Chapter 9

Fluid
Rocks behave like fluids at depths in geologic timescales. In this chapter, basic equa-
tions for a simple fluid are introduced. Then equations are used to consider vertical
movements. Linear stability analysis of the motions of the fluid is applied to periodic
geological structures.

9.1 Fluids
9.1.1 Definition
Fluids are distinguished from solids by their unlimited deformations [276]. Solids ‘remember’ their
initial shape—they tend to return to their original shapes once they are relieved of the force to
deform. By contrast, fluids change their shapes anyway. Water conforms to the shape of its container.
Accordingly, the constitutive equation of fluids does not contain a strain tensor such as V. The faster
we swim, the greater the resistance we feel from the surrounding water. The equation should have
a relationship with velocity instead of strain. However, there is no resistance if we move at the
same velocity as the water velocity. Difference in velocity causes resistance, i.e., velocity gradient
is important.
We will consider only isotropic fluid for simplicity. This allows us to use the representation
theorem (Eq. (5.5)), but we should replace V by the velocity gradient tensor D in the constitutive
equation
S = φ0 I + φ1 D + φ2 D2 ,

where φ0 , φ1 , and φ2 are the scalar functions of the basic invariants of D. As S is negative for
compression, the hydrostatic pressure is −pI. Hydrostatic pressure is often separated from the first
term of the above equation,
S = (−p + φ0 )I + φ1 D + φ2 D2 (9.1)

Materials that have this constitutive equation are called Reiner-Rivlin fluids. In this case, φ0 vanishes
for stationary fluids.

205
206 CHAPTER 9. FLUID

If the constitutive equation of a material is written as

S = −pI + f(D), (9.2)

then the material is called a Stokesian fluid. The function f has a value of a second-order tensor,
possibly anisotropic. If f is linear, the fluid is called Newtonian fluid. The fluid should be relaxed if
it is at rest. Therefore, f(O) = O. If it is at rest, stress in the fluid is just the pressure. However, due
to the term f(D), the fluid parcels feel different forces in motion, even if p is constant. Accordingly,
f is called dynamic pressure. Due to this pressure, flow in the mantle can push up or pull down the
surface (§9.5).

9.1.2 Newtonian fluid


Newtonian fluids have a linear and homogeneous function f in Eq. (9.2). For simplicity, we assume
incompressibility. The constitutive equation of incompressive Newtonian fluid is

S = −pI + 2ηD, (9.3)

where η is a material constant representing resistance against movement and is called viscosity.
Viscosity is half the constant of proportion between the velocity gradient and stress1 . Table 9.1
shows the viscosity of familiar and geologic fluids.
Let us derive the divergence of stress, ∇ · S, from Eq. (9.3). The ith component is

 ∂  
∂vi ∂vj

∂p  ∂2 vi  ∂2 vj
−pδij + η + =− +η 2
+ η . (9.4)
j
∂xj ∂xj ∂xi ∂xi j ∂xj j
∂xi ∂xj

We have assumed incompressibility so that Eq. (2.29) holds and density is constant everywhere.
%
Differentiating the equation by xi , we have j ∂2 vj /∂xj ∂xi = 0. The last term in Eq. (9.4) vanishes
accordingly. Substituting this ∇ · S into the equation of motion (3.23), we obtain the Navier-Stokes
equation
ρa = −∇p + η∇2 v + ρX. (9.5)

Long-term tectonic movements are very slow and the inertia term ρa is neglected. So we have the
Stokes equation
0 = −∇p + η∇2 v + ρX. (9.6)

The Navier-Stokes equation is rewritten with vorticity instead of velocity. Vorticity is defined as
ω = ∇ × v so that we apply ∇× to both sides of Eq. (9.5), and we have
 
∇ × ρa = −∇ × ∇p + ∇ × η∇2 v + ∇ × (ρX) . (9.7)
1 Viscosity is measured with the unit Pa s = Nm−2 s. In the CGS system, the viscosity unit is called Poise (P), and 1 P =

10 Pa s.
9.1. FLUIDS 207

Table 9.1: Viscosity of several fluids. As viscosity depends on temperature, rough values are shown.
Glacier, crust, and mantle are of the Earth.
Viscosity (Pa s) Reference
−5
Air 10
Water 10−3
Olive oil 10−1
Alcohol 100
Asphalt (15◦ C) 107
Columbia River Basalt (1250◦ C) 101 [210]
Lunar mare basalt (1200◦ C) 100 [159]
Glacier 1013 − 1015 [123]
Crust 1018 − 1030 [231]
Upper mantle 1020 [112, 113, 114]
Lower mantle 1021 [112, 113, 114]

This differential equation holds in every small parcel of fluid within which the continuum hypothesis
verifies that viscosity can be regarded as being constant. Density is also constant because we have
assumed incompressibility. The left-hand side of Eq. (9.7) becomes

 
Dv D∇ × v Dω
ρa = ∇ × ρ =ρ =ρ .
Dt Dt Dt

The term ∇×∇p vanishes as Eq. (C.57), and second term of the right-hand side of Eq. (9.7) becomes
 
∇ × η∇2 v = η∇2 ω. Accordingly, Eq. (9.7) is simplified to the equation

Dω η
= ∇2 ω + ∇ × X.
Dt ρ

This is equivalent to the Navier-Stokes equation. In the case of ∇ × X = 0, we have


= ν∇2 ω. (9.8)
Dt

This is the diffusion equation of vorticity and the diffusion coefficient ν ≡ η/ρ is called kinematic
viscosity. To emphasize the difference from this, η is called dynamic viscosity. Heavy liquids tend
to keep their velocity due to inertia, so that vorticity is difficult to change, resulting in a low diffusion
coefficient. In the case that inertia is negligible, Eq. (9.8) become

0 = η∇2 ω + ρ(∇ × X). (9.9)


208 CHAPTER 9. FLUID

9.2 Stream function


For two-dimensional problems, the scalar function, ψ (x1 , x2 ), that is related to velocity components
by the equations
∂ψ ∂ψ
v1 = − , v2 = . (9.10)
∂x2 ∂x1
is useful, and is called the stream function. The velocity field that is described by the stream function
is always incompressible, because

∂ ∂ψ ∂ ∂ψ
∇·v =− + = 0.
∂x1 ∂x2 ∂x2 ∂x1

Two-dimensional, incompressible, Newtonian fluids have a stream function.


The function is useful because the contour lines of ψ drawn on the O-12 plane show the stream
lines. In addition, flux is indicated by the difference in the value of ψ. To show these facts, suppose
a smooth curve C between points P and R on the plane (Fig. 9.1). A point Q is on the curve. Let
(x1 , x2 ) be the coordinates of the point and s be its distance from P along the curve. The unit vector
tangent to the curve at Q is s = (dx1 /ds, dx2 /ds)T . The unit vector perpendicular to s and points to
the left of s is n = R · s, where
   
cos 90◦ − sin 90◦ 0 −1
R= = .
sin 90◦ cos 90◦ 1 0

This is the orthogonal tensor which rotates vectors counterclockwise by 90◦ on a plane (Eq. (C.9)).
The components of n are
      
n1 0 −1 dx1 /ds −dx2 /ds
= = .
n2 1 0 dx2 /ds dx1 /ds

Given the flow velocity v at the point Q, the flux crossing the curve C at the point is v · n, therefore
the quantity of fluid that crosses the entire length of the curve in a unit of time is
 
F = v · n ds = (v1 n1 + v2 n2 ) ds
OP OP
   
∂ψ dx2 ∂ψ dx1 ∂ψ
= + ds = ds = ψ (P) − ψ (O).
OP ∂x 2 ds ∂x 1 ds OP ∂s

Accordingly F = 0 if the points P and R are on the same contour line of ψ, indicating that the
contour line lies along a stream line. Note that the flow direction is to the left2 of the vector ∇ψ. As
the above equation suggests, only the relative ψ is meaningful. Therefore, we can assume ψ = 0 at
convenient points, such as at infinitely far points or at the origin of the coordinates.
2 There are textbooks in which the stream function is defined with opposite signs to Eq. (9.10). In that case, the flow

direction is to the right of ∇ψ.


9.2. STREAM FUNCTION 209

Figure 9.1: Schematic picture showing the contour lines of stream function ψ.

The stream function has an important relationship with vorticity. Substituting Eq. (9.10) into
the equation that defines vorticity in two-dimensional flow, we have
     
∂ ∂ψ ∂ ∂ψ ∂2 ψ ∂2 ψ
ω= − − =− 2
+ 2 = −∇2 ψ.
∂x2 ∂x2 ∂x1 ∂x1 ∂x ∂x
1 2

Hence, if ω = 0, ψ is a harmonic equation satisfying ∇2 ψ = 0.

Biharmonic equation

The stream function is a harmonic equation if vorticity vanishes. However, the function is always a
biharmonic equation that satisfies the equation

∂4 ψ ∂4 ψ ∂4 ψ
4
+2 2 2 + 4 =0 or ∇4 ψ = 0. (9.11)
∂x ∂x ∂z ∂z
Let us consider that gravity is the only body force and let us derive this equation. We use the Carte-
sian coordinates O-xz with the z-axis taken vertically downward. Gravitational force is represented
by X = (0, 0, g)T , where g stands for the acceleration due to gravity that does not change horizon-
tally.
First, we will write D using the stream function:

∂vx ∂2 ψ ∂vy ∂vz ∂2 ψ


Dxx = =− , Dyy = = 0, Dzz = = .
∂x ∂x∂z ∂y ∂z ∂z∂z
   
1 ∂vx ∂vy 1 ∂vy ∂vz
Dxy = + = 0, Dyz = + = 0,
2 ∂y ∂x 2 ∂z ∂y
   2 
1 ∂ ∂ψ ∂ ∂ψ 1 ∂ ψ ∂2 ψ
Dxz = − + = − 2 ,
2 ∂z ∂z ∂x ∂x 2 ∂x2 ∂z
Dyx = Dxy , Dzy = Dyz , Dzx = Dxz .
210 CHAPTER 9. FLUID

In deriving these equations, we used the fact that vx and vz do not depend on y and that vy is zero or
constant anywhere. Substituting these equation into Eq. (9.3), we obtain
 
∂2 ψ ∂2 ψ ∂2 ψ
Sxx = −p − 2η , Sxz = η − 2 ,
 2 ∂x∂z  ∂x2 ∂z (9.12)
∂ ψ ∂2 ψ ∂2 ψ
Szx = η − 2 , Szz = −p + 2η .
∂x2 ∂z ∂x∂z

Other stress components are zero: Syy = Sxy = Szy = 0. Here, we neglect the inertia term. Hence,
combining X = (0, 0, g)T and the force balance equation (Eq. (3.24)), we have

∂Sxx ∂Sxz ∂Szx ∂Szz


0= + , 0= + + ρg.
∂x ∂z ∂x ∂z

Substituting Eq. (9.12), we obtain


 
∂p ∂3 ψ ∂3 ψ ∂3 ψ
0=− − 2η 2 + η − , (9.13)
∂x ∂x ∂z ∂x2 ∂z ∂z3
 3 
∂p ∂3 ψ ∂ ψ ∂3 ψ
0 = − + 2η +η − + ρg. (9.14)
∂z ∂x∂z2 ∂x3 ∂x∂z2

Gravity is supposed not to depend on x, so that ∂(ρg)/∂x = 0. Differentiating Eqs. (9.13) and (9.14)
by z and x, respectively, and subtracting both sides, we obtain the biharmonic equation (Eq. (9.11))
that describes very slow flows of Newtonian fluids. In the above derivation, we take gravity into
account, but the gravity term has been dropped.
We will often use the Fourier transform of the biharmonic equation. The nth-order differentia-
tion ∂n f (x)/∂xn is transformed into (ik)n f:(k), so that the Fourier transform can simplify differential
equations—ordinary and partial differential equations are replaced by algebraic and ordinary differ-
ential equations, respectively.
We use the Cartesian coordinates O-xz, of which the z-axis is taken as being downward. Let
: (k, z) be the Fourier transform of a stream function φ(x, z). They are transformed from each
ψ
other by the equations
∞
1
: (k) = √
ψ ψ (x)e−ikx dx, (9.15)
2π −∞
∞
1
ψ (x) = √ : (k)eikx dk,
ψ (9.16)
2π −∞

where i and k stand for the imaginary unit and horizontal wave number. The biharmonic equation is
transformed to the equation
: ∂4 ψ
∂2 ψ :
: − 2k 2
k4 ψ + = 0. (9.17)
∂z2 ∂z
9.2. STREAM FUNCTION 211

Figure 9.2: Velocity profile in Newtonian fluid flowing between parallel walls.

The general solution of this equation has several forms such as

: = Aekz + Bzekz + Ce−kz + Dze−kz ,


ψ (9.18)
: = A sinh kz + Bx sinh kz + C cosh kz + Dx cosh kz,
ψ (9.19)
: = (A + Bkz) cosh kz + (C + Dkz) sinh kz,
ψ (9.20)

where the coefficients A, B, C and D are the functions of k and z (and sometimes t), and are
determined under given boundary conditions. The forms (Eqs. (9.19) and (9.20)) are convenient

:  z=0 =
when boundary conditions are given at z = 0, because A is determined by the condition as ψ
A.

Simple velocity field

With a simple example, let us see how Newtonian fluids behave. Suppose that a Newtonian fluid is
flowing between parallel walls that are separated by a distance 2h (Fig. 9.2). Due to the viscosity,
the fluid clings to the walls so that we have the boundary condition for the velocity v = 0 at z = ±h.
We assume that the walls are very long in the x direction so that v does not change in direction.
Namely, ∂v/∂x = 0 and vz = 0. The question is how vx acts. In this case, the biharmonic equation
(Eq. (9.11)) becomes the ordinary differential equation d4 ψ/dz4 = 0. Its general solution is

ψ = Az3 + Bz2 + Cz + D.

Hence, we have
∂ψ
vx = − = −3Az2 − 2Bz − C.
∂z
The system is symmetric with respective to z = 0, therefore vx (z) should be an even function. This
is satisfied when B = 0. Let us replace −3A by A , then the solution becomes vx = A z2 − C. Using
212 CHAPTER 9. FLUID

Figure 9.3: Post-glacial uplift of Fennoscandia illustrated by the altitudes of the ancient strandline
between central Norway and Estonia [156].

the boundary condition, we obtain the velocity profile

vx = A (h2 − z2 ),
&h
a parabola with the maximum velocity at the center. The integral −h vx dz is the volume of fluid that
passes across a plane perpendicular to the x-axis in a unit time. Given the volume, A is determined.

9.3 Viscous relaxation


Continental ice sheets depress the continent by their sheer weight. Once an ice sheet disappeared
after the last glacial period, the underlying continental block rose rapidly to recover isostatic com-
pensation without the sheet. This is the phenomenon known as post-glacial rebound. The Holocene
uplift of Fennoscandia is a well-known example (Fig. 9.3).
Isostatic compensation is an important process for the deformation of large impact craters.
Shortly after large impact cratering, a crater tends to have a smooth, bowl-shaped interior with a
raised rim. Figure 1.14 shows fresh craters with such rims in the Mare Selenitatis on the Moon.
Fresh craters show a diameter-depth ratio of about 2:5. If the ambient rocks are hard, they can sup-
port a bowl-shaped slope. However, rocks have finite strengths. A large impact crater is isostatically
unstable so that the depression is gravitationally relaxed. The relaxation induces flows in a deep-
seated ductile layer if the impact basin is large compared to the surface brittle layer3 . Ice is softer
than rocks, so that impact craters on icy satellites are more easily relaxed than those on rocky planets
and satellites. Figure 9.4 shows craters on the icy satellite Ganymede. Among them, large craters
have flat or even upbowed floors (Fig. 9.5).
3 Large impact cratering fluidizes the shallow part of the target body, and gravitational collapse of the impact basin begins

while the target is excavated by the impact [146]. The above explanation gives a simplified picture of impact cratering. See
[145] for further reading.
9.3. VISCOUS RELAXATION 213

Figure 9.4: Large impact craters on the icy satellite Ganymede. Voyager Image 0550J2-001. The
topographic profile along the white line is shown in Fig. 9.5.

The speed of isostatic rebound depends on the horizontal dimension (wavelength) of a depres-
sion. Let us see this by the order of magnitude estimation by assuming the Newtonian viscosity for
the resistance for the gravitational collapse. Consider a bowl-shaped basin with a diameter D and
central depth H (Fig. 9.6). The stress under the basin may be in the order of

τ ≈ ρgH. (9.21)

If the center subsides by ΔH, the corresponding engineering shear strain for the subsidence may be
represented by ΔH/(D/2), where D/2 is the radius of the basin. Infinitesimal shear strain is half
214 CHAPTER 9. FLUID

Figure 9.5: Topographic profile of a crater along the line shown in Fig. 9.4 showing its upbowed
floor (dotted line). After [225].

Figure 9.6: Diagram showing the depth and diameter of a basin to explain its viscous relaxation.

the engineering shear strain, so that we have ε ≈ ΔH/D. Therefore, the representative strain rate
may be in the order of
˙ ≈ −Ḣ/D, (9.22)

where −Ḣ is the uplifting rate of the basin center. Substituting Eqs. (9.21) and (9.22) into the
constitutive equation of Newtonian fluid τ = 2η , ˙ we have the differential equation on H, Ḣ/H =
−2η/ρgD. Consequently, we obtain the solution H = H0 exp(−t/τ), where H0 is the initial depth
of the basin and τ = 2η/ρgD is the time constant for the relaxation. Therefore, τ is inversely
proportional to D, indicating that the amplitude of a horizontally large depression is attenuated
more rapidly than that of a small one. This is contrary to diffusion phenomena including thermal
conduction where short-wavelength temperature anomalies disappear faster than those with long
wavelengths.
Large impact craters on Ganymede are examples (Fig. 9.4). A raised crater rim accompanied by
a fresh crater is a short-wavelength topography so that the circular mountain range holds out against
gravitational spreading for a much longer time than the bowl-shaped depression of the crater itself.
Morphology of still larger craters have been subdued on Ganymede; only circular bright low-relief
circular patches called palimpsests are left (Fig. 3.6).
A degree of viscous relaxation was used to estimate the chemical composition of the icy crust of
the Neptunian satellite Triton, where impact craters smaller than 1.5 km in diameter have not been
relaxed. The reflectance spectrum indicates abundant CH4 and N2 ice on the surface of the satellite,
but they have too low viscosity to maintain the crater morphology. Therefore, the icy crust probably
consists of H2 O ice [206].
On the other hand, there are impact craters on Mars that show different modes of relaxation. The
subdued impact craters have a raised rim, whereas some of the Martian impact craters have pro-
9.4. SELF-ORGANIZING GEOLOGICAL STRUCTURES 215

Figure 9.7: Gravitational relaxation of surface undulations of a thin layer of viscous fluid on a solid
basement. (a) Model configuration of Squyres’ finite element simulation for the relaxation of a
crater with a diameter of 20 km. (b) Results of the simulation for different thicknesses H of the
layer [226].

nounced rounding of sharp topographic features. Short wavelength topographic features have been
already erased around those craters. Original crater morphology seems to have creeped, the phe-
nomenon called terrain softening [226]. Mars has a near-surface permafrost layer with a thickness
of 102 –103 m. The layer is underlain by a less viscous regolith layer or solid basement. Namely, ter-
rain softening is thought to be a near-surface phenomenon. A large depression relaxes by gathering
fluid from distant places around it. If a flat basement is sheeted by a viscous fluid, the gravitational
relaxation of a large depression must carry the fluid through a thin conduit (Fig. 9.7(a)). There-
fore, it takes longer to erase long wavelength topographic features than those with short wavelength
features4 . Figure 9.7(b) shows the results of finite element simulation by Squyres for the softening
of a crater with a different thickness of its viscous layer. The simulation demonstrates that terrain
softening occurs for a thin viscous layer lying on a solid basement. In addition, the layer is thick and
the relaxation is shown to lead to subdued craters with an upbowed floor like those on Ganymede.

9.4 Self-organizing geological structures


Stability analysis

Tectonic deformations include mechanisms of self-organization. There are multiple geological struc-
tures with spatial periodicity. The joint sets in Fig. 6.15, load cast in Fig. 8.16, and boudins5 in Fig.
9.8 are examples of mesoscale periodic structures. It is often observed that macroscale faults with
4 See Exercise 9.4 and its answer.
5 Boudins are the linear segments of a layer that has been pulled apart along periodically spaced lines of separation called
boudin lines [248].
216 CHAPTER 9. FLUID

Figure 9.8: Periodically changing thickness like sausages, which are known as boudins, of a sand-
stone bed. Quartz veins fill fractures at boudin lines (arrows). Melange zone in the Shimanto Belt,
Southwest Japan.

the same trend have constant intervals. The grooves on the icy crust of Ganymede are macroscale
samples (Fig. 9.4). Let us consider such a pattern formation associated with tectonic deformation.
The stability of fluid motions is a clue to the pattern formation by ductile deformations.
The theory of stability separates a fluid motion into a mean flow and perturbation around it. The
former is also called a basic flow. For example, flow velocity is expressed as v = v + v$, where v
is the velocity of the basic flow and v$ is the velocity of the perturbation. Likewise, other variables
are separated into two parts associated with the basic one and perturbation. For example, we deal
with pressure p = p + p$, stream function ψ = ψ + ψ $ , stress S = S + $
S, and so on. If the basic flow
is unstable, one or more perturbations grow to overwhelm the basic flow. Therefore, we investigate
the rate of growth or extinction of each perturbation using basic equations such as the Navier-Stokes
equation and given boundary conditions. If the fluid is incompressible, each mean and perturbation
satisfies the incompressibility condition as follows. The fluid is incompressible, so we have
 
∇ · v = ∇ · v + v$ = 0.
9.4. SELF-ORGANIZING GEOLOGICAL STRUCTURES 217

Figure 9.9: Deformation of a high viscosity layer between low viscosity fluids. The system is
unstable—even if pure-shear deformation is exerted, the high viscosity layer is folded by layer-
parallel compression (a) or is broken into boudins (b). For both cases, deformation proceeds in the
numbered order.

The mean flow must rigorously satisfy the condition: ∇ · v = 0. Therefore, we obtain

∇ · v$ = 0, (9.23)

which indicates that the incompressibility condition also holds for the perturbation. Likewise,

$ =0
∇4 ψ (9.24)

is obtained from Eq. (9.11). As for the Navier-Stokes equation (Eq. (9.5)), we have
   
ρa = −∇ p + p$ + η∇2 v + v$ + ρX

and
ρa = −∇p + η∇2 v + ρX.
Therefore, the pressure and velocity perturbations must satisfy

p + η∇2 v$.
0 = −∇$

Formation of boudins and folds

The ductility of rocks depends on the lithology. When a sedimentary pile of alternating consolidated
sand- and mudstones is subject to ductile deformation, their ductility contrast leads to instability
to form folds when they are compressed and boudins when they are extended in the orientation
parallel to the bedding (Fig. 9.9). Let us investigate the development of a spacing of bodins and fold
wavelength assuming incompressible Newtonian fluids [215].
Consider a thin layer of fluid with a viscosity of η (2) and a thickness of H is embedded in very
thick fluid layers with a viscosity η (1) (Fig. 9.10(a)). The layer is initially flat with infinitesimal
undulations. In what follows, we discriminate the plate and ambient fluids by the superscripts (1)
218 CHAPTER 9. FLUID

and (2). The origin of the rectangular Cartesian coordinates O-xz is placed on the middle surface
of the thin layer, and the x- and z-axes are defined as being parallel and perpendicular to the layer.
In order to develop a unified theory for the formation of boudins and folds, we assume pure shear in
the range −L ≤ z ≤ L as the basic flow, where L  H. The perturbation in vz leads to folds or
boudins.
The existence of the layer causes disturbance in the pure shear. Hence, it is appropriate that
the velocity field far from the layer equals that of the basic flow. For this reason, we assume the
boundary condition for the velocity to be

vz  z=±L = ±V, (9.25)

where V is a constant. Correspondingly, D = V /L is the representative velocity gradient in this


system. Pure shear is the basic flow so that we have the condition for the shear stress

Szx  z=±L = 0. (9.26)

We neglect both inertia force and surface tension at the interfaces, so that stress is continuous across
them:
(1)  (2) 
Szz  z=H/2 = Szz  z=H/2 , (9.27)
(1)  (2) 
Szx  z=H/2 = Szx  z=H/2 . (9.28)

Velocity is also assumed to be continuous there. Namely, we have


 
v (1)  z=H/2 = v (2)  z=H/2 . (9.29)

The other side of the layer at z = −H/2 has the same conditions. Equation (9.29) represents the
lack of gaps or overlaps of the two fluids. Force balance between them results in Eqs. (9.27) and
(9.28). It should be noted that Sxx can be discontinuous across the interface6 . It is shown later that
the difference in Sxx raises the undulation of the interface that eventually forms boudins or folds.
Let hT (x) = H/2 + h(x) be the top surface of the thin layer. It is difficult to deal with the layer
if the interfaces have undulations with a large amplitude. Hence, we consider this system only while
the condition |h(x)|  1 holds. The interface moves at the velocity
∂hT DH
≈ . (9.30)
∂t 2
The right-hand side of this equation designates the mean flow of the interface. Under the condition
|h(x)|  1, perturbations are small. Namely, the velocity field is approximated by that of pure shear

vx ≈ −Dx, vz ≈ Dz. (9.31)


6 Imagine that water and barley sugar consist of the fluids (1) and (2). The latter is much more viscous than water, so that

a layer of barley sugar carries much larger tensile stress or compression than water.
9.4. SELF-ORGANIZING GEOLOGICAL STRUCTURES 219

Figure 9.10: (a) Initial configuration of the model. L is much larger than H, the thickness of a thin
layer. The interfaces at z = ±H/2 have infinitesimal undulations. (b) Close-up of the top interface.
Forces acting at the top interface and force balance for small rectangle.

The basic flow has the constant rate of strain


 
−D 0
Ė = D = . (9.32)
0 D
Substituting Eq. (9.32) into the constitutive equation of Newtonian fluid, we have

Sxx = −p + 2ηDxx = −p − 2ηD, (9.33)


Szz = −p + 2ηDzz = −p + 2ηD, (9.34)
Szx = 2ηDxz = 0. (9.35)

Equation (9.35) indicates that this system always satisfies the boundary condition in Eq. (9.28).
Combining Eqs. (9.27) and (9.34), we obtain

−p(1) + 2η (1) D = −p(2) + 2η (2) D.

Therefore, there is a pressure difference between the fluids on the other side of the interface:

Δp ≡ p(2) − p(1) = 2 η (2) − η (1) D. (9.36)

The fluids have different Sxx . Equation (9.33) holds of each of the fluids:
(1) (2)
Sxx = −p(1) − 2η (1) D, Sxx = −p(2) − 2η (2) D. (9.37)

Combining Eqs. (9.36) and (9.37), we have the difference in the normal stress in the fluids:
(2) (1)
ΔSxx ≡ Sxx − Sxx = −Δp − 2 η (2) − η (1) D

= −4 η (2) − η (1) D. (9.38)
220 CHAPTER 9. FLUID

In this section, we use overlines and tildes to distinguish the basic flow and perturbation around
it. Namely, velocity is divided as
v =v+$ v. (9.39)
The basic flow of this system has the velocity
 
v = v x , v z T = (−Dx, Dz) T . (9.40)

The continuity of the velocity field across the interface leads to the boundary condition for the
velocity perturbation  
v$(1)  z=h = v$(2)  z=h . (9.41)
T T

The interface is not only driven by vz but also by the lateral flow vx . If the interface with an
inclination to the left (∂h/∂x > 0) moves to the right, the interface goes down (Fig. 9.10). Therefore,
we have  
∂h ∂h  
=− · vx  + vz  . (9.42)
∂t ∂x z=hT z=hT
Hence, substituting Eqs. (9.39) and (9.40), we obtain
∂hT   ∂h     ∂h  
= − vx + $
vx + vz + $
vz = − −Dx + $ vx + Dz + $vz
∂t  ∂x
  ∂x 
 
∂h ∂h  
=D x + hT + − ·$vx  +$
vz  . (9.43)
∂x ∂x z=H/2+h z=H/2+h
     
basic flow perturbation
The initial undulations of the interface is extended vertically and constricted laterally by the first
term in the last line. Therefore, the first term is not responsible for pattern formation. The second
term gives rise to instability.
Let us consider the force balance in the x-direction of the small rectangle shown in Fig. 9.10(b),
using the stress tensor
S = S+$ S. (9.44)
(1) (1)
The fluid (1) includes the left side of the rectangle so that the force −Sxx dz = −Sxx (∂h/∂x)dx
  that the amplitude |h(x)| is very small, (1)
acts there. Since weare assuming the stress perturbation is
    (1)
also very small S x   S$xx . Hence, the force is approximated by −S xx dz = −S xx (∂h/∂x)dx.
(2)
The right side of the rectangle is in the fluid (2), so that the normal force at the side is S xx dz =
(1)
S xx (∂h/∂x)dx+ΔSxx (∂h/∂x)dx, where ΔSxx is given by Eq. (9.38). Because of (9.32) and (9.35),
we have S zx = S zx = 0. However, the perturbation in those shear stresses does not necessarily
vanish. The top and base of the rectangle feel the shear forces S$zx dx and −S$zx dx, respectively.
(1) (2)

Therefore, the total force balances if


left right
      
top base

(1) ∂h (1) ∂h ∂h
+ S$zx − S$zx = 0.
(1) (2)
−S xx + S xx + ΔSxx
∂x ∂x ∂x
9.4. SELF-ORGANIZING GEOLOGICAL STRUCTURES 221

We have a similar equation for the z-direction. Using the difference in the stress components between
the fluids
ΔS$zx ≡ S$zx − S$zx , ΔS$zz ≡ S$zz − S$zz ,
(2) (1) (2) (1)

we have the force balance in the x- and z-directions:

∂h
ΔS$zx = ΔSxx , (9.45)
∂x
∂h
ΔS$zz = ΔS$zx . (9.46)
∂x
The governing
 equations for the development of undulations have higher-order terms such as
(∂h/∂x) · v z=H/2+h in Eq. (9.43) and the right-hand sides of Eqs. (9.45) and (9.46). We employ
linear stability analysis to neglect non-linear terms with the assumption that the amplitude is very
small |h|  1. Under this condition, we have |∂h/∂x|  1, and all perturbation terms designated
by a tilde are infinitesimal. Therefore, the second-order term v$x (∂h/∂x) is neglected in Eq. (9.43)
to give
∂$ 
h 
=$ vz  . (9.47)
∂t z=H/2+h

This boundary condition is defined at the level H/2 + h, which includes the unknown variable h
itself. Therefore, we have to linearize this condition also. The right-hand side of Eq. (9.47) is
expanded as



 ∂  
$
vz  ≈$vz  +h· $ vz  ,
z=H/2+h z=H/2 ∂z z=H/2

whereas the last term is a second-order term and is therefore negligible. Consequently, we have

∂$ 
h 
=$
vz  . (9.48)
∂t z=H/2

Likewise, neglecting non-linear terms, Eq. (9.41) is rewritten as a condition at z = H/2:


 
 
v$(1)  = v$(2)  . (9.49)
z=H/2 z=H/2

The discontinuity of the normal stress is also divided into mean and the infinitesimal fluctuation,
ΔSxx = ΔS xx + ΔS$xx . The jump conditions for stress components (Eqs. (9.45) and (9.46)) are
linearized as
  
(2)  (1)   ∂h
S$zx  − S$zx  = ΔS xx  · , (9.50)
z=H/2 z=H/2 z=H/2 ∂x
 
(2)  (1) 
S$zz  − S$zz  = 0. (9.51)
z=H/2 z=H/2
222 CHAPTER 9. FLUID

We investigate the wavenumber-dependence of the growth rate of fluctuations. To this end, we


assume solutions of the form
$
h(x, t) = a(t) sin kx, (9.52)
$ (x, z, t) = ϕ(z, t) sin kx
ψ (9.53)

for the linearized governing equations (Eqs. (9.24) and (9.48)–(9.51)). If an even function ϕ(2) (z) =
(2)
ϕ(2) (−z) is the result, vz = ∂ϕ(2) /∂x is also an even function of z. In this case, the top and basal
interfaces of the thin layer approach or depart to each other, leading to the formation of boudins. On
the other hand, if an odd function ϕ(2) (z) = −ϕ(2) (−z) is the result, the upwarping portions of the
top interface are associated with the upwarping portions of the basal interface. Therefore, folding
occurs in this case. Substituting Eq. (9.53) into the biharmonic equation (Eq. (9.24)), we have
 4 
∂ ∂2 ∂4
+ 2 2 2 + 4 ϕ(z, t) sin kx
∂z4 ∂x ∂z ∂x
= ϕzzzz sin kx − 2k 2 ϕzz sin kx + k 4 ϕ sin kx = 0,

where the subscripts indicate partial derivatives such as ϕzz = ∂2 ϕ/∂z2 . Each of the fluids satisfy
this equation so that we have
(i) (i)
ϕzzzz − 2k 2 ϕzz + k 4 ϕ(i) = 0 (9.54)
for the ith layer. This equation has the same form as Eq. (9.17). Hence, the general solution of Eq.
(9.54) is
ϕ(i) = a(i) e−kz + b(i) ze−kz + c (i) ekz + d(i) zekz . (9.55)
The condition for the formation of boudins and folds is indicated by the conditions among the coef-
ficients of this equation:

Boudins a(2) = c (2) and b(2) = d(2) , (9.56)


Folds a (2)
=c (2)
and b (2)
=d .
(2)
(9.57)

From the stream function in Eq. (9.53), we have the velocity components

(i) ∂ψ (i) (i) (i) ∂ψ (i)


$
vx = − = −ϕz sin kx, $
vz = = kϕ(i) cos kx. (9.58)
∂z ∂x
Substituting Eq. (9.58) into (9.49), we have

ϕ(1) = ϕ(2) , (9.59)


(1) (2)
ϕz = ϕz . (9.60)

Equation (9.58) allows us to write the stretching tensor as


   (i) (i)

$= ∂$
v /∂x ∂$
v /∂z −kϕ cos kx −ϕ sin kx
D x x
= z
(i)
zz
(i) .
∂$vz /∂x ∂$ vz /∂z −ϕzz sin kx kϕz cos kx
9.4. SELF-ORGANIZING GEOLOGICAL STRUCTURES 223

(i)
Hence, the constitutive equation $
(i)
S= −$p(i) I + 2η (i) D$ becomes
   
(i) (i)
$ p$ 0 −kϕz cos kx −ϕzz sin kx
S=− + 2η .
0 p$ (i) (i)
−ϕzz sin kx kϕz cos kx

Equations (9.50) and (9.51) are, therefore, rewritten as


 
(2) (1)
−2 η (2) ϕzz − η (1) ϕzz sin kx = ΔS xx ka cos kx, (9.61)
 
(2) (1)
η (2) ϕz − η (1) ϕz 2k cos kx = Δ$p. (9.62)

Combining Eqs. (9.48) and (9.58), we obtain


da
= kϕ(i) tan kx. (9.63)
dt
Now we introduce the following parameters:

η (2) 1 da 1 S xx
E≡ , G≡ , R≡ . (9.64)
η (1) a dt 2 2η (1) G

E stands for the viscosity ratio, G is the growth rate, and 2η (1) G represents the viscous resistance to
drive the fluid (1) for the growth. Using these parameters, Eqs. (9.59)–(9.62) are rewritten as

ϕ(1) = ϕ(2) , (9.65)


(1) (2)
ϕz = ϕz , (9.66)
   
(2) (1)
E ϕzz + k 2 ϕ(2) − + k 2 ϕ(1) = 4kRϕ(1) ,
ϕzz (9.67)
   
(2) (2) (1) (1)
E ϕzzz + 3k 2 ϕz = ϕzzz + 3k 2 ϕz . (9.68)

Assuming the boundary condition that perturbations vanish at infinity (z → ±∞), we readily obtain
 
 
c (1)  = d(1)  =0 (9.69)
z→±∞ z→±∞

from the general solution (Eq. (9.55)).


In the case of the formation of boudins (Eq. (9.56)), we have

ϕ(1) = a(1) e−kz + b(1) ze−kz , ϕ(2) = 2b(2) ze−kz .

Substituting these equations into Eqs. (9.65)–(9.68), we obtain the simultaneous equations
⎛ ⎞ ⎛ (1) ⎞ ⎛ ⎞
k  −fk −g a 0
⎜ −k 1− gk −g + f ⎟ ⎜b(1) ⎟ ⎜0⎟
=
⎝−k − Rk −Ef + Eg ⎠ ⎝a(2) ⎠ ⎝0⎠
. (9.70)
1 −  − R Efk
k  −Egk −Ef b(2) 0
224 CHAPTER 9. FLUID

Let A be the square matrix in this left-hand side. The newly introduced parameters , f and g
depend on k and H as
Hk e− − e e− + e
= , f= , g = . (9.71)
2 e− e−
The simultaneous equations (Eq. (9.70)) have non-trivial solutions if

det(A) = 0.

Solving this equation, we obtain the growth rate G for the formation of boudins as a function of 
and E,
1  
Rb = sinh2 (E +  − E 2 ) + sinh  cosh (1 + E 2 ) + cosh2 (E −  + E 2 ) . (9.72)
2E
On the other hand, in the case of folding (Eq. (9.57)), we obtain simultaneous equations with the
matrix ⎛ ⎞
k  −gk −f
⎜ −k 1− fk −f + g ⎟
A=⎝
−k − Rk 1 −  − R Egk −Eg + Ef ⎠
. (9.73)
k  −Efk −Eg
Consequently, we have the rate of folding
1  
Rf = cosh2 (E +  − E 2 ) + cosh  sinh (1 + E 2 ) + sinh2 (E −  + E 2 ) . (9.74)
2E
Combining Eqs. (9.38), (9.38), (9.72) and (9.74), the growth rate G is obtained as

G = [1 − E] [1/R] [D],

where Smith’s notation is used [215], i.e., the brackets indicate the different signs for the following
conditions:
)
+(1 − E) (boudins),
[1 − E] =
−(1 − E) (folds) ,
)  (2) 
+1/R (1)
1/R = η (2) > η (1)  ,
−1/R η <η ,
)
+D (Basic flow is layer-parallel shortening,
[D] =
−D (Basic flow is layer-parallel extension).

The growth rate G is a function of k and H through the parameter  (Eq. (9.71)). Boudins or
folds with a wavenumber that maximizes the growth rate are expected to emerge. The dimensionless
wavelength λ/H = 1/kH chosen by this theory is plotted in Fig. 9.11(a), indicating that the
wavelength gradually increases with the viscosity contrast E = η (2) /η (1) . This model predicts that
there is no single layer that folds with a λ/H ratio smaller than 6. However, field observations
indicate the dominance of the single-layer folding with a ratio of around 6 (Fig. 9.11(b)). Smith
explains the high frequency folding with non-Newtonian viscosity (Fig. 10.19).
9.5. DYNAMIC TOPOGRAPHY 225

Figure 9.11: (a) Dominant non-dimensional wavelength versus viscosity ratio for single layer fold-
ing predicted from Smith’s model [216]. (b) Histogram of observed wavelengths [211].

9.5 Dynamic topography


Mantle convection has surface expressions such as subduction zones, oceanic ridges, and hotspots.
They are the loci of large vertical movements that reflect sublithospheric flow. The Hawaiian Swell
is an example where the oceanic lithosphere is uplifted by ∼1 km by the mantle upwelling under the
hotspot (Fig. 8.8(a)). A vertical velocity gradient in viscous fluid leads to a dynamic pressure, which
raises or pulls down the lid of the fluid. The depression or uplift supported by the dynamic pressure
in the sublithospheric mantle is known as dynamic topography. When the flow dies away, dynamic
topography vanishes to leave isostatic compensation. If the vertical movement can be grasped from
stratigraphic records, we have a clue to the patterns of mantle convection over billions of years.
In the vicinity of mantle up or downwelling, various parameters, including velocity and tem-
perature, tend to have large variations. Therefore, elaborated numerical calculations are needed to
understand actual dynamic topography. However, the purpose of the following discussion is to inves-
tigate whether dynamic topography has enough amplitudes to infer the convection from stratigraphic
data so that we assume a very simple thermal convection in the mantle to estimate the amplitude.

9.5.1 Linearized boundary condition

First of all, it is demonstrated that the overburden stress can be derived from the Navier-Stokes
equation if the inertia term is omitted. To this end, we take the rectangular Cartesian coordinates
O-xz with the z-axis pointing downward. The body force significant for this problem is the gravity,
i.e., X = (0, 0, g)T . Consider that all forces are balanced and the fluid is at rest, v = 0. Then, Eq.
(9.6) becomes 0 = −∇p + (0, 0, ρg)T . Therefore, we obtain

dp
= ρg.
dz
226 CHAPTER 9. FLUID

Figure 9.12: Linearization of boundary condition.

Indeed, integrating both sides of this equation results in the equation of the overburden stress (Eq.
(3.29)), i.e., we have the pressure 
z
p= ρ(ζ)g dζ.
surface
Next, we consider dynamic pressure. From the equation for incompressible Newtonian fluid (Eq.
(9.3)), we have the vertical stress
∂vz
Szz = −p + 2η . (9.75)
∂z
When the local isostasy was introduced, overburden was assumed to be supported solely by p. The
second term in the right-hand side of Eq. (9.75) represents the dynamic pressure, through which
the vertical velocity gradient has an additional effect for topography. The term has a positive sign
because the z-axis is defined downward (Fig. 9.12). Equation (9.75) designates that the amplitude
of dynamic topography has a positive correlation with the viscosity η. For simplicity we assume a
constant viscosity.
Consider the origin of the z-axis is placed at the surface without dynamic topography (Fig.
9.12), and let z = h(x) be the surface topography. If dynamic topography leads to a depression
and a bulge, we have h > 0 and < 0, respectively. In addition,
 we assume that the surface is a free

boundary. Namely, we have the boundary condition S surface = O. However, this equation is not
convenient for further calculation, because this condition is defined at the surface of which level h
is unknown. In addition, the surface generally has an inclination dh/dx, which also includes the
unknown. Therefore, we have to simplify the boundary condition. Fortunately, dynamic topography
produces long wavelength undulations, so that the inclination is neglected for brevity. Then, the
horizontal gradient of S becomes negligible, and we have the approximate boundary condition for
the remaining stress component 

Szz  = 0. (9.76)
z=h

The left-hand side of this equation is expanded as


  



 ∂Szz 
 + ··· ,
Szz  = Szz  + h
z=h z=0 ∂z z=0

where the gradient ∂Szz /∂z can be approximated by Δρg. Δρ represents the density difference
between the mantle and overlying substance, seawater or sediments. Therefore, from Eq. (9.76) we
9.5. DYNAMIC TOPOGRAPHY 227

obtain the linearized boundary condition




Szz  = −Δρgh. (9.77)
z=0

The right-hand side represents topographic load (Fig. 9.12). Once the vertical stress at z = 0 is
obtained, dynamic topography h is calculated with this equation.

9.5.2 Simple convection model


We assume that a Newtonian fluid with a constant viscosity η and thickness d is convected by basal
heating. Parsons and Daly [173] presented the dynamic topography for this simple thermal convec-
tion.
Let ρm be the mantle density at the equilibrium state T = 0. Then, the density at the temperature
anomaly T is given by ρm (1 − αT ). Therefore, omitting the inertia term, we have the equation of
motion
−∇p + η∇2 v + ρm (1 − αT ) g = 0. (9.78)
The acceleration due to gravity is assumed to be constant throughout the convecting cell, g =
(0, 0, g)T . Temperature perturbation in the mantle may be of the order of 100 K. Density varia-
tion in the cell is on the order of αΔT , i.e., ∼10−5 K−1 ×102 K = 10−3  1. For this reason,
we employ the Boussinesq approximation to simplify governing equation; variations in density is
neglected except insofar as they are coupled to the gravitational acceleration and buoyancy forces
[65, §16.1]. We assume incompressibility of the fluid which allows us to use the stream function ψ.
When the function was introduced in Section 9.2, the buoyancy term ∂ρg/∂x vanished. However,
this gradient drives the convection. Therefore, the buoyancy term is indispensable here. Instead of
Eq. (9.11), the stream function satisfies
 4 
∂ ∂4 ∂4 ρm gα ∂T
4
+2 2 2 + 4 ψ = . (9.79)
∂x ∂x ∂z ∂z η ∂x

Taking the Fourier transform of both sides of Eq. (9.79) we obtain


 2 2  
d ρm gα :
−k 2
: = ik
ψ T, (9.80)
dz2 η

where i = −1 is the imaginary unit and k is horizontal wavenumber.
Dynamic topography is calculated from the vertical stress at z = 0 (Eq. (9.77)). According to
the constitutive equation of Newtonian fluid, this stress component satisfies
∂vz
Szz = −p0 − p1 + 2η , (9.81)
∂z
where hydrostatic pressure is divided into p0 = ρgz that is hydrostatic pressure when the fluid at rest
and p1 that is the correction term for the hydrostatic pressure due to convection. The former has no
228 CHAPTER 9. FLUID

horizontal variation so that Eq. (9.78) becomes


 2 
dp1 ∂ vx ∂2 vx
− +η + = 0.
dx ∂x2 ∂z2
Taking the Fourier transform of both sides of this equation, we have
iη 2 d:
ψ
p:1 = ∇ . (9.82)
k dz

 is water-loaded, we have Δρ = ρm −ρw . Then, the boundary condition (Eq. (9.77))


If the lithosphere
becomes Szz z=0 = (ρw − ρm )gh. Substituting Eq. (9.81) into this boundary condition, we have
 
 ∂vz 

ρw gh = ρm gh − p1  − 2η  . (9.83)
z=0 ∂z z=0

Combining Eqs. (9.82) and (9.83), we obtain


  
: 1 iη d2 ψ
d:  .
h= − 3k 2 (9.84)
(ρm − ρw )g k dz2 dz z=0

Let T1 be a representative temperature fluctuation. Using the dimensionless variables (Fig. 9.13),
x z T η
x = , z = , k  = dk, T = , ψ = ψ,
d d T1 ρm gαT1 d3
The equation of motion (Eq. (9.80)) is rewritten as
 2 2
d
−k 2
ψ: = ik  T: . (9.85)
dz 2
&
T: is a function of z. The solution of Eq. (9.85) includes the integral T: dz, i.e., the flow ψ
:  is af-
fected by the temperature distribution over the entire convecting cell. Thus, the dynamic topography
determined through (Eq. (9.84)) is affected by the entire temperature distribution, unlike the isostatic
compensation which bears little relation to variations under the depth of compensation. However,
the effect of temperature depends on depth z; shallow temperature anomalies are more important
than deep ones. Accordingly, let us use Green’s function7
1
ψ: (k  , z ) = ik  Ψ(k  , z , ζ)T: (k  , ζ) dζ. (9.86)
0

Ψ(k  , z , ζ) describes how the temperature distribution at a depth of ζ affects the flow ψ
: at the depth
z , and is the solution of the equation
 2 2
d 2
−k ψ  = δ(z − ζ), (9.87)
dz 2
7 See [173] for detail.
9.5. DYNAMIC TOPOGRAPHY 229

Figure 9.13: Diagrams for the dynamic topography induced by two-dimensional thermal convection
[173].

which describes the flow by a point source at z = ζ. In this equation, δ( ) is the delta function. The
dimensionless dynamic topography h is also related to the entire temperature distribution through
Green’s function H (k  , z ) as

ρm αT1 1
 
h (k ) = H (k  , ζ)T: (k  , ζ) dζ. (9.88)
ρm − ρ w 0
H is obtained from the equation
  
 d2 dΨ 
H (k , ζ) = − − 3k  2 , (9.89)
dz 2 dz z =0
and depends on the top and basal boundary conditions. If the convecting cell has a fluid substratum
with a density of ρb and the interface is deformed by the convection, the interface has dynamic
topography hb and also by its Green’s function Hb . They are related to the temperature distribution
as 
ρm gαT1 1
 
hb (k ) = Hb (k  , ζ)T:(k  , ζ) dζ, (9.90)
ρb − ρ m 0
where  2  
 d 2 dΨ 
Hb (k , ζ) = − 3k . (9.91)
dz 2 dz z =1
The dynamic topography of the surface h (k  ) is obtained from Eq. (9.88), which needs the
Green’s function given in Eq. (9.89). To this end, Eq. (9.87) has to be solved with tectonically
230 CHAPTER 9. FLUID

appropriate boundary conditions. If we think of dynamic topography around an oceanic ridge, the
top of the convecting cell may be a free boundary because the oceanic plate is a part of the cell. In
contrast, the base of the lithosphere may be a rigid boundary for a mantle plume under a hotspot. If
the lid of the cell is a stable continent, a rigid boundary may be appropriate, also. The base of the
convecting cell can be either a free or rigid boundary.
In case of the free boundary, we have the velocity and shear stress the conditions vz = 0 and
Szx = 0 at z = 0. The latter condition indicates that the flow is parallel to the x-axis at z = 0, i.e.,
the top surface coincides with the contour line of ψ so that ψ|z=0 is constant there. The absolute
value of ψ is meaningless, so that we define ψ to have this constant at zero. The other condition, the
shear stress Szx satisfies
∂vz ∂2 ψ
Szx = 2η = 2η 2 .
∂x ∂x
Consequently, the free boundary has the conditions
 
 ∂2 ψ 
 =0
ψ = 0, (9.92)
z=0 ∂x2 z=0

for the stream function at z = 0.


In the case of the rigid boundary, the stream function has the boundary condition ψ = 0 at the
base of the lithosphere z = 0, also. In addition, if the lithosphere is stationary, the horizontal velocity
component of the asthenospheric flow vanishes at z = 0. Consequently, the rigid boundary condition
is described by the equations 
 ∂ψ 
  = 0.
ψ  = 0, (9.93)
z=0 ∂z z=0

Equation (9.87) has the solution

ψ  = A1 sinh k  z + B1 cosh k  z + C1 z sinh k  z + D1 z cosh k  z (z < ζ), (9.94)


    
ψ = A2 sinh k (1 − z ) + B2 cosh k (1 − z )
+ C2 (1 − z ) sinh k  (1 − z ) + D2 (1 − z ) cosh k  (1 − z ) (z > ζ). (9.95)

The coefficients A1 , B1 , etc., are the functions of k  and z , and are determined from the boundary
conditions. We readily obtain B1 = B2 = 0 from the boundary condition ψ  = 0 at the base and top
of the cell.
The pressure perturbation p1 has discontinuity (Fig. 9.14) at z = ζ by the point source that is
represented by the delta function in Eq. (9.87). The discontinuity corresponds to the discontinuous
d3 ψ  /dz 3 in Eq. (9.14) at z = ζ. The first- and second-order derivatives are still continuous there

so that we integrate both sides of Eq. (9.87) over the infinitesimal interval ζ − , ζ + , where
0 <  1. The result is  
d3 ψ  
 d3 ψ  

− = 1. (9.96)
dz 3   z =ζ+ dz 3   z =ζ−
9.5. DYNAMIC TOPOGRAPHY 231

Figure 9.14: Schematic picture showing pressure perturbation induced by a sinking sphere. The
perturbation has a discontinuity at the sphere.

The other boundary conditions constrain the coefficients of Eqs. (9.94) and (9.95). For example, if
the convecting cell has free boundaries at its top and base (Eq. (9.92)), we obtain
1  
          
A1 = k z sinh k cosh k (z − 1) − k sinh k z − sinh k sinh k (z − 1) , (9.97)
2k  3 sinh2 k 
1  
          
A2 = k (z − 1) sinh k cosh k z − k sinh k (z − 1) − sinh k sinh k z , (9.98)
2k  3 sinh2 k 
C1 = C2 = 0, (9.99)
sinh k  (z − 1) sinh k  z
D1 = , D2 = . (9.100)
2k  2 sinh k  2k  2 sinh k 
Figure 9.13(c) shows H for this case, where H is plotted for three different k  . The dynamic topog-
raphy for other boundary conditions has similar amplitudes so that we omit the other cases8 .
Dynamic topography is controlled by the temperature distribution in the convecting mantle,
which is affected by the distribution of heat sources. Neglecting radioactive heating in the upper
mantle and assuming the basal heating only, the dimensionless Rayleigh number
ρm gαΔT d3
Ra =
ηκ
represents the relative magnitude of advective and conductive heat, where ΔT is the temperature
difference of the top and base of the convecting layer. The basal heating of constant viscosity fluid
induces thermal convection whose temperature is proportional to sin πz , and Eq. (9.85) becomes
 2 2
d 2
− k ψ  = ik  sin πz .
dz 2
8 See [173] for further reading.
232 CHAPTER 9. FLUID

Figure 9.15: Dynamic topography resulting from the thermal convection in the upper mantle (d =
600 km) with a constant viscosity corresponding to an assumed Rayleigh number of Ra = 1.4 × 106
[173].

This equation has the general solution

ik   
ψ = sin πz + A(z − 1) sinh k  z + Bz sinh k  (z − 1) .
(π 2 + k  2 )2
Substituting this into Eq. (9.84), we obtain

ρm αT1 π 3 + 3πk  2 − 2k  3 A ρm αT1 π 3 + 3πk  2 − 2k  3 B


h = , hb = . (9.101)
ρm − ρ w (π 2 + k 2 )2 ρb − ρ m (π 2 + k 2 )2
Figure 9.15 shows the dynamic topography with an amplitude of the order of 100 m generated by
the upper mantle convection with an appropriate Rayleigh number9 . This amplitude is large enough
to leave stratigraphic evidence such as regression or transgression if a shallow marine basin is above
the mantle convection.

9.5.3 Epeirogeny
Subduction zones are the surface expressions of the downgoing mantle flow, so that island arcs have
negative dynamic topography. The amplitude of the topography depends on the density anomaly
and downgoing velocity of the slab and the viscosity of the asthenosphere under the arc, The density
anomaly depends on the oceanic age of the slab and the velocity. A subduction zone is so com-
plex that numerical simulations are needed to estimate the dynamic topography. Among those, the
simulation by Zhong et al. [280] demonstrate that the dynamic topography can be more than 1 km
over an area of 1000 km from the trench axis. The amplitude has a positive correlation with slab
age. Dynamic topography is an important factor for the secular vertical movement of island arcs. It
affects vertical movements in vast backarc regions if an old slab subducts with a low angle.
9 Mantle convection induces vertical stresses at the base of the lithosphere, resulting in the dynamic topography. The
stresses act as vertical loads and bend the lithosphere. Therefore, the lithosphere masks the dynamic topography with a
wavelength much shorter than the flexural parameter of the lithosphere. In this respect, the lithosphere acts as a low-pass
filter for the dynamic topography.
9.5. DYNAMIC TOPOGRAPHY 233

Figure 9.16: Cretaceous Western Interior Basin in North America. (a) Isopach map of the
Campanian–Maastrichitian (65–83 Ma) deposits in the basin [46]. The contour interval is 1000
feet. Dot-bar line indicates the western border of the present distribution of the deposits. Distri-
bution of igneous rocks produced by island-arc magmatism is shown by hatching. (b) Slab model
for the calculation of dynamic topography at 65 Ma. Stable continent is exhibited by hatching. (c)
Dynamic topography at 65 Ma. Panels (b) and (c) are after [26].

The Cretaceous Western Interior Basin was produced not only by the topographic load of the
Cordilleran orogen (Fig. 8.3) but also by the negative dynamic topography by a shallow dipping slab
subducting from the Pacific [46, 152]. Figure 9.16(a) is the isopach map of Cretaceous strata. The
western part of the sedimentary basin has been eroded away by the upheaval of the Rockies. The
basin subsided with little tectonic deformation except for its western margin which was involved
through Cordilleran orogeny. The sedimentary basin was wider than 1000 km, too extensive for
the flexural subsidence of the lithosphere, even if the effective elastic thickness was as great as
100 km. Island-arc magmatism occurred at that time far away from the western coast of North
America, suggesting that the dip of the slab was shallow. Therefore, the extensive negative dynamic
topography left the stratigraphic record [46, 152]. Figures 9.16(c) and (d) show the modelled slabs
and the calculated dynamic topography from [26]. The result shows that the shallow subduction of
the Farallon Plate produced a depression deeper than 1 km in the vast backarc region.
Horizontally extensive vertical movements occur with little tectonic deformation. Those phe-
nomena are known as epeirogenic movements. The Western Interior Basin is an example. The
Russian platform is another one, where the Precambrian basement continued from the Baltic Shield
is blanketed by the latest Precambrian to the Mesozoic sedimentary rocks more than 3 km thick. A
single mechanism cannot explain the lengthy subsidence [88, 124]. However, the Devonian–Permian
subsidence was rapid for this region but accompanied by little magmatism and deformation, so that
it is attributed to the negative dynamic topography induced by the the subduction at the eastern and
southern margins of the Russian platform [153]. It is suggested that Devonian subsidence in the
Russian platform and in Canada was attributed to the avalanche of cold subducted materials piled up
on the 660 km boundary into the lower mantle [49, 187].
234 CHAPTER 9. FLUID

Figure 9.17: Hypothetical tectonic evolution around Japan from 20 to 15 Ma [272]. Note that slab
age abruptly changed from to Early Miocene for the Southwest Japan Arc. The Pacific Plate has
Cretaceous lithosphere in this region. The gross form of the paleotopography in these periods is
illustrated in Fig. 3.14. IBA, Izu-Bonin Arc; KP, Kyushu-Palau Ridge.

The sudden uplift of the Southwest Japan at 15 Ma (Fig. 3.14) was presumably due to a jump in
the dynamic topography brough about by the abrupt change in the slab age from Cretaceous to Early
Miocene (Fig. 9.17). The young Philippine Sea Plate began subduction at that time. The arc was
raised above sea-level simultaneously with the onset of widespread magmatism and the beginning
of the arc-perpendicular compressional stress field [272].
9.5. DYNAMIC TOPOGRAPHY 235

Figure 9.18: Surface topography of a viscous fluid layer with an average thickness of H. The layer
is underlain by a rigid basement with a horizontal interface.

Exercises
9.1 Show that gravity force X satisfies ∇ × X = 0. Due to this identity, the diffusion equation of
vorticity (Eq. (9.8)) holds if gravity is the only body force for Newtonian fluid.

9.2 Estimate the viscosity of the mantle under Fennoscandia from the ancient strandlines shown in
Fig. 9.3 with the assumption that the uplift began at 13 kyr ago. [156].

9.3 Assuming the crust as a viscous fluid, consider the relationship between the Argand number
(p. 86) and the viscosity.

9.4 Consider the relaxation of surface depression of a layer of viscous fluid covering a rigid base-
ment with a flat interface. Let H be the average thickness of the layer, and z = h0 (x) be the initial
surface topography, where  is defined downward with the level z = 0at the average
 z-axis  altitude
(Fig. 9.18). We assume h0 (x)  H, indicating gentle undulations, and h0 (±∞) = 0. Then,
the Fourier transform 
: 1 ∞
h0 (k) = h0 (x)eikx dx (9.102)
2π ∞
exists, where k is an horizontal wavenumber. The dimensionless number kH is regarded as the
dimensionless thickness of the layer. Namely, the two cases kH  1 and kH  1 correspond
to very thick and very thin layers relative to a wavenumber k, respectively. The time constant for
viscous relaxation τ can be nondimensionalized as τ = (ρgH/2η) τ. Show τ ∝ kH and τ ∝ (kH)−2
for very thick (kH  1) and very thin (kH  1) viscous layers, respectively [123].

9.5 Show that Argand number for the situation in Fig. 3.19 is given by Ar = ρgΔh/ηV , where Δh
is the difference in surface heights and V is the velocity of the plate under the fluid [29].
Chapter 10

Plasticity

Fluids with a non-linear constitutive equation are introduced in this chapter.


They include plastic body, Bingham, and power-law fluids. The equations of
plasticity are applied to the deformation of impact craters. Rocks behave as
power-law fluids whose constitutive equation is used in Chapter 12 to investi-
gate the mechanics of the lithosphere.

10.1 Quasilinear Fluids


The constitutive equation of Newtonian fluid (Eq. (9.3)) has a viscosity η that does not depend on
fluid motions. Let us broaden this limitation, i.e., we assume isotropy, incompressibility (DI = 0),
and the dependency of viscosity on D (Fig. 10.1). This fluid is used to model tectonic deformations
in Chapter 10. The function η(D) must satisfy the representation theorem so that the viscosity is a
function of basic invariants DII and DIII . The second law of thermodynamics applies to this fluid
flow. It is known that the law demands that the function does not include DIII [276]. Consequently,
the variable viscosity fluids have a constitutive equation of the form

T = 2η(DII )D (10.1)

and is spoken of as quasilinear fluids. In this case, η is called shear-rate dependent viscosity. Quasi-
linear fluids include power-law fluids, Bingham fluids, and plastic bodies which are collectively
called viscoplastic bodies.
Attention must be paid to the sign of DII . We have assumed incompressibility (DI = trace D =
0), so that we can presume D as a deviatoric tensor (see Appendix C.6). The same is true for
the strain rate tensor Ė, because D is identical with Ė (Eq. (2.27)). Therefore, their second basic

237
238 CHAPTER 10. PLASTICITY

Figure 10.1: Deviatoric stress T versus velocity gradient D for quasilinear fluids. The inclination
of the segment OP equals twice the effective viscosity, which decreases with increasing velocity
gradient for pseudoplastic fluids. That of dilatant fluids increases. SY is the yield stress.

invariants satisfy the inequalities (Eq. (C.50)):


1 1 
DII = (D : D) ≥ 0, ĖII = Ė : Ė ≥ 0. (10.2)
2 2
Consider a simple shearing motion on the O-12 coordinate plane to see the basic behavior of
quasilinear fluids. The stretching tensor for this flow is
⎛ ⎞
0 D 0
D = ⎝D 0 0 ⎠ . (10.3)
0 0 0

Substituting Eq. (10.3) into (10.1), we find that the two components T12 = T21 = T are the only
non-zero components of T. Accordingly, the equation
2k D
T = 2aD + tan−1 (10.4)
π b
is a constitutive equation for this flow, and the graph for the equation is given by the convex upward
curve in Fig. 10.1. Note that Eq. (10.4) includes the constitutive equation of Newtonian fluid as
the case where k = 0. In the case of a = 0 and b → 0, Eq. (10.4) is a constitutive equation of a
rigid-perfectly-plastic body.
The shear-rate-dependent viscosity is half of the slope of the line segment OP in this figure, so
that we have
2k D
2η(D) = 2a + tan−1 . (10.5)
πD b
This viscosity increases with D. This is spoken of as shear-thinning viscosity. Fluids with this type
of viscosity are called pseudoplastic fluids. By contrast, shear-thickening viscosity increases with
D, and fluids with this type of viscosity are called dilatant fluids.
10.2. PRINCIPAL STRESS SPACE 239

Taking the limit b → ∞ and replacing k by SY , Eq. (10.4) becomes

T = 2aD + SY , (10.6)

the graph of which is shown in the two line segments that meet on the vertical axis at S = SY . This
is called Bingham fluid, and SY is of that fluid. The fluid has a viscosity

2η(D) = 2a + SY /D (10.7)

that goes to infinity with D → 0. Therefore, the decrease of D increases the viscosity. The fluid flow
is consequently rapidly decelerated. Bingham fluids start to flow when a driving force overcomes
the yield stress. Under the critical stress, the fluid behaves as a rigid or elastic body.
For the simple shear flow, DII = 12 D : D = 2D2 . Accordingly, the expression for the viscosity of
Bingham fluids (Eq. (10.7)) may be broadened for general flow cases as
SY
2η(DII ) = 2a +  . (10.8)
DII /2

The constitutive equation of Bingham fluids is therefore


 
SY  
T=2 a+  D D = O . (10.9)
2DII

Let us consider the simple shear flow (Eq. (10.3)), again, to derive the relationship between SY
and T. Substituting DII = 2D2 into Eq. (10.9), we have
⎛ ⎞ ⎛ 2 ⎞
  0 2aD + SY 0 T12 0 0
SY
T = 2a + D = ⎝2aD + SY 0 0⎠ , T2 = ⎝ 0 T12 2
0⎠ .
D
0 0 0 0 0 0

The second basic invariant of the deviatoric stress is TII = − 21 (trace T)2 − trace T2 , so that

lim TII = SY2 .


D→0

This indicates the fluid yields when TII reaches a critical level1 .

10.2 Principal stress space


If a rock is isotropic, its mechanical properties are the same in any direction, only the magnitude
of the principal stresses plays a role in describing the yield or failure behavior. Hence, we use a
three-dimensional stress space where three principal stresses are the coordinate axes O-σ1 σ2 σ3 or
O-S1 S2 S3 (Fig. 10.2(a)). The states of hydrostatic and lithostatic stresses S1 = S2 = S3 are
1 This is the yield criterion of von Mises, which is introduced in Section 10.3.
240 CHAPTER 10. PLASTICITY

Figure 10.2: (a) Principal stress space. A point on the line S1 = S2 = S3 represents a hydrostatic
state of stress with the pressure S0 . The distance from the line indicates deviatoric stress . Perpen-
dicular to the line is called a deviatoric plane. The equation S1 + S2 + S3 = const. represents the
plane. If S0 = 0, the plane is called a π-plane. (b) Θ is an angle on the π-plane and represents the
symmetry of Lamé’s stress ellipsoid.


represented by the points that make up a line parallel to the unit vector e⊥ = (1, 1, 1)/ 3. Planes
perpendicular to the line are called deviatoric planes. The point P in Fig. 10.2 is an example. Those
points designate isotropic stresses, whereas points out of the line designate anisotropic stresses. The
−−→
point Q represents those points, and has the position vector OQ = (S1 , S2 , S3 )T . If the line PQ is
perpendicular to the line OP, we have
−→ −−→ ⊥ SI 
OP = OQ · e = √ = 3S0 . (10.10)
3
−→ −→ √ √
Hence, we have OP = OP e⊥ = 3S0 · (1, 1, 1)T / 3 = (S0 , S0 , S0 )T , and further
⎛ ⎞ ⎛ ⎞
S − S0 T1
−→ −−→ −→ ⎝ 1
PQ = OQ − OP = S2 − S0 ⎠ = ⎝T2 ⎠ . (10.11)
S3 − S0 T3
−→
Therefore, the length PQ is related to the second basic invariant of T and the octahedral shear
stress by the equation
−→  2  
PQ = T + T 2 + T 2 = 2TII = 3 S oct , (10.12)
1 2 3 S

where Eqs. (4.14) and (4.17) are used. Namely, TII and SSoct indicate the magnitude of anisotropy.
The plane passing through the origin O, and perpendicular to the line OP, is known as the π-
plane in the theory of plasticity. Let O-Si be the orthogonal projection of the coordinate axes O-Si
10.2. PRINCIPAL STRESS SPACE 241

onto the plane (Fig. 10.2(b)). Obviously, the S1 -, S2 - and S3 -axes meet at 120◦ . Let the point Q
be the foot of the base of the perpendicular dropped from Q (S1 , S2 , S3 ) onto the π-plane, then
Q has the coordinates (T1 , T2 , T3 ) (Eq. (10.11)). Let Θ be the angle between OQ and the S1 -
axis. This is known as Lode angle. The orthogonal projection has the projector (I − e⊥ e⊥ ). The
S1 -axis has the base vector e = (1, 0, 0)T , so that the unit vector e parallel to the S1 -axis satisfies
 
e ∝ (I − e⊥ e⊥ ) · e = e − e⊥ · e e⊥ = (2, −1, −1)T /3. Therefore, we have

1
e = √ (2, −1, −1)T , (10.13)
6
−−→ −−→ −→ 
   
and OQ · e = OQ  cos Θ = PQ cos Θ = 2TII cos Θ. Substituting Eq. (10.11), we have

−→ √
−−→ ⊥ PQ · (2, −1, −1) (T1 , T2 , T3 ) · (2, −1, −1) 2T1 − T2 − T3 3
OQ · e = √ = √ = √ = √ T1 ,
6 6 6 2

3 S1
∴ cos Θ = . (10.14)
2 TII
Using the formula cos 3Θ = 4 cos3 Θ − 3 cos Θ, this is rearranged as

3 3  
cos 3Θ = (TII )−3/2 T13 + T1 TII . (10.15)
2
Eliminating T1 from Eq. (10.15) with the three equations,

T1 + T2 + T3 = 0, TII = −(T1 T2 + T2 T3 + T3 T1 ), TIII = T1 T2 T3 ,

we obtain √
3 3
cos 3Θ = TIII (TII )−3/2 (0 ≤ Θ ≤ 60◦ ). (10.16)
2
Rearranging Eq. (10.14), we obtain
  
2 TII 2 TII 2 TII
T1 = √ cos Θ, T2 = √ cos(120 − Θ), T3 = √ cos(120◦ + Θ)

(10.17)
3 3 3
and
  
SI 2 TII SI 2 TII SI 2 TII
S1 = + √ cos Θ, S2 = + √ cos(120◦ − Θ), S3 = + √ cos(120◦ + Θ),
3 3 3 3 3 3
(10.18)
where 0 ≤ Θ ≤ 60◦ . Principal axes exchange their orientations when Θ value exceeds the multiples
of 60◦ . Stresses with these Θ values are axial stresses (Fig. 10.3(a)). Φ and Θ have a one-to-one
correspondence (Fig. 10.3(b)),
2 sin Θ
Φ= √ , (10.19)
3 cos Θ + sin Θ
which is derived from Eq. (10.17), provided that Θ is in the range [0, 60◦ ].
242 CHAPTER 10. PLASTICITY

Figure 10.3: (a) Diagram showing the relationship between Mohr circles and Lode angle. (b) Graph
of Eq. (10.19).

10.3 Yield conditions


At the limits of a → 0 and b → 0 in Eq. (10.4), the pseudoplastic fluid becomes a perfectly plastic
fluid that yields when TII = 2k 2 is satisfied. In this section, a phenomenological theory of yielding
is introduced.
For simplicity, we assume an isotropic and incompressible plastic material. The material yields
when the state of stress satisfies a yield condition, which is represented by the function of stress
tensor F (S) = 0, called the yield function. Since an isotropic material is assumed, this function is
depends only on the principal stresses. That is, the function is rearranged as F (S1 , S2 , S3 ) = 0,
which is a function of the points in the stress space and defines a surface in the space, known as the
yield surface. Using Eqs. (10.16) and (10.18), the principal stresses can be replaced by TII and TIII
so that
F (S0 , TII , TIII ) = 0 (10.20)

also represents the yield condition.


Laboratory experiments show that the yield conditions of metals do not depend on mean stress.
However, mean stress is not negligible for porous materials including rocks, indicating that overbur-
den stress is significant for yielding. S0 in Eq. (10.20) indicates this dependency. For simplicity,
10.3. YIELD CONDITIONS 243

Figure 10.4: Yield surfaces for yield criteria in three-dimensional stress space. Dot-bar lines desig-
nate the line σ1 = σ2 = σ3 .

we first introduce the yield conditions that have no such dependence. In that case, the yield function
becomes F (TII , TIII ) = 0.

Tresca yield criterion

The Tresca yield criterion assumes that plastic yielding occurs when the maximum shear stress
reaches the critical value cY . Since the maximum shear stress equals half the differential stress (Eq.
(4.18)), we have the yield condition

1 1
ΔS = (S1 − S3 ) = cY . (10.21)
2 2
When ΔS/2 < cY , deformation of the material is accommodated by elasticity. The Tresca crite-
rion does not include the intermediate principal stress in contrast to the following von Mises yield
criterion. Neglecting the inequality S3 ≤ S2 ≤ S1 , we obtain the yield condition
     
max S1 − S2 , S2 − S3 , S3 − S1  − 2cY = 0.

This is rearranged as
     
S1 − S2  = 2cY or S2 − S3  = 2cY or S3 − S1  = 2cY . (10.22)
 
Each of these equations designates a plane in the stress space. The plane represented by S1 −S2  =

2cY is perpendicular to the unit vectors ±(1, −1, 0)T / 2, which with the vector e in Eq. (10.13)
make an angle of
 
(1, −1, 0)T (2, −1, −1)T
cos−1 ± √ · √ = 30◦ or 150◦ .
2 6
244 CHAPTER 10. PLASTICITY

Figure 10.5: Yield locus on the π-plane of three yield criteria.


  
The vector e isparallel to the S1 -axis. The same angles are obtained from the equations S2 −
S3  = 2cY and S3 − S1  = 2cY in Eq. (10.22), so that the solid figure indicated by this equation
is a hexagonal prism with the axis perpendicular to the π-plane (Fig. 10.4(a)). If a stress state
is designated by a point within the prism, the deformation of the material is accommodated by
elasticity. The figure is the yield surface of the Tresca criterion. The regular pentagon in Fig. 10.5
shows the yield locus of the criterion on the π-plane.

Von Mises yield criterion

The condition known as von Mises yield criterion assumes that plastic yielding occurs when elastic
strain energy reaches a critical value. Consider a linear spring under a tensile force F , which is
related to the displacement u and the constant k as F = ku. The energy stored in the spring is
u u
1 1
U = F du = ku du = ku2 = F u
0 0 2 2

which is generalized for a material with linear elasticity as

1
U= S : E. (10.23)
2
This is divided into two parts corresponding to the changes of volume and shape. Substituting

1+ν ν
E= S − (trace S) I
Y Y
into Eq. (10.23), we obtain
 
1 1+ν ν 1+ν ν
U= S: S − (trace S) I = S:S− (trace S)2 .
2 Y Y 2Y 2Y
10.3. YIELD CONDITIONS 245

Combining S = T + S0 I = T + (trace S) I/3 we have


1+ν ν  2
U= (T + S0 I) : (T + S0 I) − trace (T + S0 I) .
2Y 2Y
Since TI = 0 (Eq. (4.12)), we obtain
US UV
     
1+ν 3(1 − 2ν)
U= T:T + S0 , (10.24)
2Y 2Y
where US and UV are strain anergy corresponding to changing the shape and volume, respectively.
Using the second basic invariant TII = 21 T : T, we have
1+ν
US = TII . (10.25)
Y
Using Eq. (4.17), this is rewritten as
3(1 + ν)  oct  2
US = SS . (10.26)
2Y
The von Mises criterion assumes that plastic yielding occurs when this energy reaches a critical
value. Namely, the yield condition for this criterion is
3  oct  2
TII − cY2 or S = cY2 . (10.27)
2 S
−−→ √
Combining Eqs. (10.12) and (10.27), we obtain |P Q| = 2cY , which does not depend on Θ, i.e.,

the yield surface is a cylinder with a radius of 2cY (Fig. 10.4).
Unlike the Tresca criterion, the yield condition of the von Mises criterion includes the inter-
mediate principal stress via TII (Fig. 4.3). If the Tresca and von Mises criteria are assumed for a
material, the critical differential stresses predicted by the criteria should be equal to each other for
axial stresses. Figure 10.5 shows the yield loci for the criteria on the π-plane. The critical differ-
ential stress predicted from the Tresca criterion is smaller than that from the von Mises criterion.
The loci of the former is represented by a regular hexagon inscribing the circle that represents the
latter. The difference reaches the maximum at Θ = 30◦ or, equivalently,
 −→ Φ = 1/2. At that time,
− →
PQ for the criteria is different by 1 − cos 30◦ ≈ 15%. Since PQ is proportional to SSoct (Eqs.
(10.12)), which is further proportional to ΔS, the difference in ΔS is about 15 %, also. Laboratory
experiments demonstrate that the von Mises criterion is more appropriate for rocks than the Tresca
criterion. However, the maximum difference is small and, therefore, is sometimes neglected.

Coulomb-Navier yield condition

The Coulomb-Navier criterion is designated by an inclined line in the Mohr diagram (Fig. 10.6),
and its yield condition is represented by

σ1 − σ3 = 2τ0 cos φ + (σ1 + σ3 ) sin φ. (10.28)


246 CHAPTER 10. PLASTICITY

Figure 10.6: Schematic picture showing the Coulomb-Navier criterion.

Using Θ and basic invariants, this is rewritten as

1 √  
−σ I sin φ + τII 3(3 − sin φ) cos Θ + 3(1 + sin φ) sin Θ − 3τ0 cos φ = 0, (10.29)
2
where 0 ≤ Θ ≤ π/3. The yield loci and yield surface corresponding to this criterion are represented
by a hexagon (Fig. 10.5) and a hexagonal pyramid (Fig. 10.4(c)), respectively.

Extended Tresca and extended von Mises criteria

Unlike metals, porous materials including rocks have yield a stress depending on the confining pres-
sure. An extended Tresca yield criterion assumes that the critical values is a function of the mean
stress as
ΔS = f (S0 ). (10.30)

Likewise, the yield condition for the extended von Mises yield criterion is represented by

TII = g(S0 ). (10.31)

If the function g(S0 ) is linear, it is called a Drucker-Prager yield criterion. Laboratory experiments
demonstrate the S2 dependence of the yield conditions, and the extended von Mises criterion is better
than the extended Tresca criterion [154]. It is shown that the function g in Eq. (10.31) is not linear
[212].

10.4 Plastic yielding by folding


Hinge zone of a fold is the most highly curved zone near the hinge line of the fold. Under Kirchhoff’s
hypothesis, layer-parallel normal stress is proportional to curvature (Eqs. (8.9) and (8.12)) (Fig.
10.7(a)). Accordingly, plastic yielding can occur in the hinge zone when a thin elastic-plastic plate
is folded. In order to consider the yielding, we firstly calculate stress in the zone. Let us use the
same coordinates with those in Section 8.2. Namely, O-xy plane is defined by the middle surface
10.4. PLASTIC YIELDING BY FOLDING 247

of the plate when it was flat. Plane strains on O-xz plane is assumed. The displacement of the
middle surface is assumed to be small (w  1), so that strain and stress have the principal axes
approximately parallel to the coordinate axes. Consider the correspondence, ε1 = εxx , ε2 = εyy and
ε3 = εzz . Neglecting the effect of overlying and underlying layers, we have σzz = 0. The curvature
is approximated as K = w  , so that we have

εxx = −w  z, (10.32)

where z is the distance from the middle surface and w  = d2 w/dx2 . In this case, Eq. (8.9) is
rewritten as σxx = [Y/(1 − ν 2 )]εxx . Substituting Eq. (10.32) into this equation, we have
Y z 
σxx = − w . (10.33)
1 − ν2
Equation (7.13) is expressed in the present coordinates as

σxx = (λ + 2G)εxx + λεzz , σyy = λ(εxx + εzz ), σzz = λεxx + (λ + 2G)εzz .

Substituting σzz = 0 and Eq. (10.33) into the above three equations, and combining Eqs. (7.7) and
(7.10) to eliminate λ and G, we obtain the intermediate principal stress
Y νz 
σyy = w = −νσxx .
1 − ν2
Now we know all the principal stresses and have the octahedral shear stress
 oct 2 1   2 
σS = (σx − σy )2 + (σy − σz )2 + (σz − σx )2 = 1 + ν + ν 2 σx2 .
9 9
The yield condition for the von Mises criterion (Eq. (10.27)) is indicated by 2SY2 = (σxx − σyy )2 +
(σyy − σzz )2 + (σzz − σxx )2 , where SY is the critical stress. Then, yielding occurs when σxx satisfies

SY
σxx =  .
1 + ν + ν2
Let H be the thickness of the plate, then the normal stress at its base (z = H/2) is given by Eq.
(10.33), i.e., 
 Y Hw 
σxx  =− . (10.34)
z=H/2 2(1 − ν 2 )
The zones of plastic deformation appear at the top and base of the plate along the hinge line, and
gradually expand toward the middle surface (Fig. 10.7(b)). The elastic core is the central part where
deformation is accommodated by elasticity. The elastic core is eroded atboth sides and eventually
disappears. Let w0 be the curvature when the yielding begin, and cY be σxx z=H/2 . Then, Eq. (10.34)
is rearranged as
2(1 − ν 2 )cY
w0 = − , (10.35)
YH
248 CHAPTER 10. PLASTICITY

Figure 10.7: (a) Stress states in the hinge zone of a fold. (b) Plastic yielding of a thin elastic-plastic
plate.

indicating the critical curvature for the yielding to begin.


The elastic core is attenuated with increasing curvature w  . If a vertical load exerted on the
lithosphere increases or abruptly appears, the lithosphere is firstly bent as an elastic plate. However,
the elastic core is gradually thinned. Whole lithosphere failure is said to occur if the elastic core in
the lithosphere eventually disappears. However, it takes a time to thin the elastic core so that yielding
may be negligible for outer trench swells that are accompanied by rapid subduction.

10.5 Slip line theory


If a rock mass has numberless fracture surfaces with various orientations, faulting with very small
displacements on the surfaces gives rise to a deformation of the mass that can be regarded as a plastic
deformation. In this case, which surfaces are activated as faults? Slip line theory gives a clue to this
problem.

Strain increments

Unlike elastic strain, plastic strain can be very large. It is not easy to relate stress and finite strain.
To this end, we have to know deformation history quantitatively. Instead, strain incremental theory
10.5. SLIP LINE THEORY 249

assumes that an incremental plastic strain dEp is proportional to deviatoric stress

dEp = Tdλ, (10.36)

where dλ is the constant of proportionality. It is assumed that the total strain is the sum of the plastic
and elastic strains
dE = dEp + dEe , (10.37)
where the superscripts “p” and “e” stand for the plastic and elastic parts, respectively.
When mechanical properties are measured, stresses are often applied to a test piece. If the mate-
rial has isotropic mechanical properties, the resultant strain has the same axial symmetry. Therefore,
it is convenient to make some formulas to convert triaxial stress and strain to axial stress and strain.
For this purpose, equivalent stress is defined based on the von Mises yield criterion:
(
 3 3
Te = 3TII = T : T = √ SSoct . (10.38)
2 2
For a uniaxial stress (S1 = S2 = S3 = 0), we have the mean and principal deviatoric stresses
S0 = S1 /3, T1 = 2S1 /3 and T2 = T3 = −S1 /3. In this case, therefore, the equivalent stress equals
the non-zero principal stress (S1 = Te ).
In relation to the energy dissipation by plastic strain dW p , the equivalent plastic strain increment
dEe is defined as dW e = Te dEe , i.e., equivalent plastic strain increment is given by
(
3
dEe = (dE) : (dE). (10.39)
2
An equivalent strain rate is defined as
(
p 3
Ėe = Ė : Ė. (10.40)
2
Equivalent strain and stress are sometimes defined as
( (
1  1
TE = T : T = TII , ĖE = Ė : Ė (10.41)
2 2
instead of as in Eqs. (10.38) and (10.40). These quantities are convenient for comparing the results
of a simple shear test that measures mechanical properties under simple shear. Consider a simple
shear ⎛ ⎞
1 2q 0
F = ⎝0 1 0 ⎠ .
0 0 1
Then, the stress tensor has only non-zero components S12 = S21 if the test piece is isotropic, and
has the principal stresses −S12 , 0 and S12 . Hence, we have ΔS = 2|S12 |. In this case, we have the
stress ratio Φ = 0, so that Eq. (4.15) reduces to TII = |S12 |. In addition, from Eq. (10.41) we obtain
TE = |S12 |. Therefore, TE equals the shear stress for simple shear tests.
250 CHAPTER 10. PLASTICITY

Figure 10.8: Slip lines and Mohr circles.

Slip line field

extitslip line field is introduced by assuming incompressible plane strain on the O-13 plane. Because
of the plane strain, the intermediate principal strain vanishes. Therefore, we have
p p p
dE1 = −dE3 , dE2 = 0. (10.42)

Substituting Eq. (10.42) into (10.36), we have


 
p 1 1
dE2 = T2 dλ = S2 − (S1 + S2 + S3 ) dλ = (−S1 + 2S2 − S3 ) dλ = 0.
3 3

S1 + S3
∴ S0 = S2 = . (10.43)
2
In this case, we have Φ = 1/2 so that the Tresca and von Mises criteria predict the same yield
condition (Fig. 10.4), i.e., yielding occurs on the surfaces with the maximum shear stress. They
make angles with the principal axes at 45◦ (Fig. 10.8(a)). The intersection between the surfaces and
the O-12 plane is known as a slip line. The slip line that makes a clockwise angle of 45◦ with the
S1 -axis is called an α-line. A slip line perpendicular to this line is called a β-line (Fig. 10.8(a)). The
trajectories of the α- and β-lines are called slip line field.
In this case, we have S12 = S23 = 0. Therefore, the yield condition for the von Mises criterion
is given by
 2  S11 + S33 2
S11 − S33 + 4S13
2
= 4k 2 or S11 − + S13
2
= k2 .
2
The latter designates the Mohr circle for the slip lines. Let φ be the angle shown in Fig. 10.8(b),
then we have

S1 = S0 + k, S2 = S0 , S3 = S0 − k, S11 = S0 − k sin 2φ, S33 = S0 + k sin 2φ


10.6. COLLAPSE OF IMPACT CRATERS 251

Figure 10.9: (a) Mosaic photograph around the south pole of the Moon (Clementine BJ90A000).
The upper half of this picture shows the near side. The Schödinger basin is about 320 km in diameter.
(b) The multi-ringed basin Orientale on the Moon (Lunar Orbiter IV194M). The Cordillera ring is
about 900 km in diameter.

and
S22 − S11
tan 2φ = . (10.44)
2S13
The trends of the slip lines are obtained through this equation.
The slip-line theory can be applied not only to plane strains but also to axisymmetric strain fields.
The theory is applied in the next section to the stability of impact craters. It should be noted that the
theory is not valid for triaxial strains. Therefore, the theory is not useful in regions where crustal
thickening or thinning is significant.

10.6 Collapse of impact craters


Impact craters have various shapes. However, the shapes can be classified by diameter [145]. Small
craters are bowl-shaped with raised rims, whereas larger craters have complex shapes. Large craters
are classified as central peaks (Fig. 1.8), peak-ring and multi-ring craters. The Schöredinger basin
is a peak-ring crater, and the Orientale basin has multiple rings (Fig. 10.9).
The depth H versus diameter D for lunar craters is shown in Fig. 10.10. Obviously, lunar craters
have two trends that intersect at about D = 15 km. Craters smaller than this are bowl-shaped and
called simple craters. Larger ones are called complex craters. On the Moon, craters greater than
about 15 km in diameter usually have landslides on their rims to reduce the H/D ratio. Craters with
252 CHAPTER 10. PLASTICITY

Figure 10.10: Depth H versus diameter D of craters in highland (dots) and maria (open circles) on
the Moon [182].

the sizes of D ∼ 101 km initially had bowl-shapes, which are called transient craters, but they were
immediately deformed by landslides.
The intersection at D = 15 km marks the critical diameter for a bowl-shaped basin to keep
its topography against gravity. A large crater has a deep basin that generates a large difference in
overburden stresses around the crater. The imbalance causes gravitational collapse of the basin.
Therefore, the critical diameter enables an order-of-magnitude estimation of the strength of the ma-
terials under craters [145]. To this end, the topography of a transient crater with the depth H and
diameter D is approximated by a paraboloid of revolution
 2 !
r
z=H 1− , (10.45)
D/2

where r is the horizontal distance from the crater center and z is depth. The mass defect of the basin
is the source of force for deformation. Using Eq. (10.45), the force is roughly estimated to be of the
order of H
1
F = ρg πr 2 dz = πρgHD2 .
0 8
The surface of the hemisphere with a radius of D/2 is S = πD2 /2. Therefore,
F 1
σ= = ρgH
S 4
is the representative stress under the crater. The transient crater has a D/H ratio of around 2.7, so
that a critical diameter of about 15 km corresponds to a critical depth of H ≈ 5.6 km. Using the
parameters ρ ≈ 3 × 103 kg m−3 , g ≈ 1.5 ms−1 , we obtain the critical stress σY ≈ 6 MPa. This is one
order of magnitude smaller than the tensile strength of rocks, and two orders of magnitudes smaller
10.6. COLLAPSE OF IMPACT CRATERS 253

Figure 10.11: Central peak transition diamter D and surface gravity g on icy satellites and terrestrial
planets [205].

than their compressive strength. The small strength suggests that the crashed rocks were fluidized by
impact cratering so that they behave as a plastic fluid. If it was the case, the critical diameter should
have inverse correlation with surface gravity g. Impact craters on icy satellites and terrestrial planets
demonstrate the inverse correlation. Figure 10.11 suggests that the fluid made from the icy crust was
weaker than that from rocks. The peak-rings and multi-rings are considered to be the waves on the
plastic fluid that were quenched when the fluidization came to an end (§10.7).

Width of landslide blocks

The strength can be estimated from the width of landslide blocks at the rim of small complex craters
using the slip line theory [143]. Among complex craters, smaller ones were merely subject to land-
slides (Fig. 1.8). The slopes of those craters are stepped by the blocks that slid down along the
arcuate fractures surfaces. In order for the estimation, we assume that the sliding was very slow
relative to freefall, so that inertia forces are negligible. Secondly, the materials under the crater of
a are rigid-perfectly-plastic body following the Tresca yield criterion (σ1 − σ3 )/2 = cY . Thirdly,
Assuming axial symmetry about the vertical for the transient crater whose shape is approximated by
the paraboloid of revolution (Eq. (10.45)), we define the cylindrical coordinates O-rφz. Fourthly,
254 CHAPTER 10. PLASTICITY

Figure 10.12: Slip line fields under a transient crater whose shape is approximated by a paraboloid of
revolution (Eq. (10.45)) with a D/H ratio of 0.2 [143]. Dot-bar lines indicate the axis of revolution.
Note the convergence of β-lines to a thick line near the crater rim. Stress state is under the yield
condition in the gray region in (a). Slip lines are not drawn under the thick lines in (b) and (c)
because this model fails there.

the Haar-von Kármàn hypothesis σφφ = σ1 = σ2 is employed to constrain the hoop stress σφφ [45].
Figure 10.12 shows the slip line fields obtained through numerical calculations under these con-
ditions2 , where ρgH/cY is the dimensionless depth of the transient crater. It was found that the stress
state near the crater rim reaches the yield condition for a dimensionless depth of greater than 5. In
addition, if the dimensionless depth is greater than 7, β-lines show convergence under the rim to the
thick lines in Fig. 10.12. Those lines may be utilized for landslides. If ρgH/cY is ∼10, the area
where the maximum shear stress exceeds cY becomes extensive under the rim, suggesting an entire
collapse of the slope and resultant upheaval of the fractured crater center.
Landslides may occur along the surface that is designated by thick lines in Fig. 10.12. Accord-
ingly, the stability of the parallelogram in Fig. 10.13 constrains the strength of the materials under
the rim. The inclination of the slope at its top φ0 is obtained by differentiating Eq. (10.45). The
result is tan φ0 = |dz/dr| = 4H/D. Let h be the height of the block that is shown in Fig. 10.13. The
shear stress at the base of the gray parallelogram is σS = ρgh cos φ0 sin φ0 . When this shear stress
reaches the critical value cY , the block slides down. Namely, the failure condition is

ρgh cos φ0 sin φ0 = cY .

The height is replaced by the width w = h/ tan φ0 , so that we obtain the critical width of the landslide
block  
cY 1
w= 1+ . (10.46)
ρg 16(H/D)2
Equation (10.46) allows us to estimate the critical value cY . Lunar craters have those blocks with
widths of several kilometers [177]. Since simple craters have an H/D ratio of 0.2, we assume the
order of magnitude to be H/D ∼ 10−1 for the ratio of complex craters. Then we have cY ∼ 100
MPa. This is consistent with the critical stress estimated on p. 252.
2 See [143] for details.
10.7. BINGHAM MODEL FOR COMPLEX CRATERS 255

Figure 10.13: Parallelogram approximating the cross-sectional shape of the landslide block under a
crater rim. The thick line approximates the line to which the β-lines converge in Fig. 10.12.

Stability of caldera walls

The above model can be applied to calderas. Assuming a cylindrical depression for the caldera,
the slip line field is shown in Fig. 10.14. The vertical wall is stable for shallow depressions with
ρgH/cY < 5. When this dimensionless depth reaches 6.8, the failed region reaches the center of the
depression, indicating an uplift of the center to accommodate the subsidence of the rim.
Figure 10.14(c) shows the large caldera called Creidne on Io. There are few landslides in the
caldera, although its wall is as high as 1 km. The surface of Io is covered by various sulphur-
rich compounds. However, this cover is probably thin because the compounds do not have enough
strength to support the wall [40].

10.7 Bingham model for complex craters


The shape of a complex crater depends on the diameter of the crater D. Lunar craters with D greater
than 20 km have central peaks and those with D  140 km have peak rings. Central peaks, peak
rings, and multirings are thought to be formed by the quench of waves of fluidized materials excited
by impacts. A moving Bingham fluid can exhibit rapid deceleration (§10.1) so that the final state
of the fluid is quenched. Assuming the damped oscillation of Bingham fluids, let us estimate the
mechanical properties of the fluids from those geological structures [144]. Quenching at half the
oscillation cycle leads to a central peak. If the oscillation was quenched at the end of the first cycle,
a peak ring is the result. A multiring structure is the result of cessation after a few cycles. If this
hypothesis is valid, the crater diameters for the onset of central peaks and that of peak rings should
depend on surface gravity g, because gravity is the driving force of the waves. If gravitational force
does not overcome the critical stress of the Bingham fluid, oscillation does not occur. Figure 10.15
demonstrates the existence of this correlation.
We are going to determine the effective values of the Bingham parameters. Instead of using the
parameters in considering the quenching, we utilize the critical depth of the transient crater to be
unstable. The critical dimensionless depth was presented as ρgH/cY ≈ 5 on p. 254. Accordingly,
256 CHAPTER 10. PLASTICITY

Figure 10.14: (a) Slip line field for a cylindrical depression with H/D = 0.05 [143]. (b) Modes
of gravitational collapse of the depression for three-dimensionless depth ρgH/cY . (c) Surface of
Jovian satellite Io (Voyager image 0145J1+000). The caldera called Creidne has a depth of  1 km
relative to the surrounding plain and a major axis of 100 km. Montes Haemus is about 10 km high.

let us use the critical amplitude


5cY
A∗ = . (10.47)
ρg

The oscillation is frozen when the amplitude of oscillation falls to A∗ .


Consider a hemisphere with a radius of L in which fluidization and oscillation occur. The hemi-
sphere includes a transient crater. Let A, ω and t be the representative amplitude, angular frequency,
and the time since the beginning of the oscillation. Then the representative vertical displacement is
expressed as z = A cos ωt, and P = 2π/ω is the period of oscillation. Let h be the root-mean-square
10.7. BINGHAM MODEL FOR COMPLEX CRATERS 257

Figure 10.15: The minimum diameter of central peak craters [205]. That of Venus is after [77].
Gray lines have an inclination of −1. Open circles: rocky planets and satellites; closed circles: icy
satellites.

of z over a cycle:
 2π/ω  2π  2π
ω 1 1 A2 dτ A2
h =
2
(A cos ωt) dt = A
2 2
(cos 2τ + 1) dτ = = .
2π 0 2π 0 2 2π 0 2 2
Taking damping into account, the root-mean-square is somewhat smaller than above. However,

h ≈ A/ 2 can be used as a rough estimate. This mean displacement occurs over the surface area
πL2 , so that the potential energy is represented by
    π
V ≈ − ρgh2 × πL2 = − ρgA2 L2 . (10.48)
2
The minus sign is attached here for the convenience of later arrangement. On the other hand, if

Ȧ sin ωt is the vertical velocity, we similarly obtain the representative velocity v ≈ Ȧ/ 2. The mass
of the hemisphere is M = (4π/6)L3 ρ, where ρ is the density of the Bingham fluid. Therefore, the
kinetic is roughly estimated as
1 π
U ≈ Mv 2 = ρL3 Ȧ2 . (10.49)
2 6
In order to deal with the damped oscillation, we utilize the Laglangian [117]
π π
L = U − V = ρL3 Ȧ2 + ρgA2 L2 , (10.50)
6 2
where Eqs. (10.48) and (10.49) are used. Let η be the viscosity of the fluid. The velocity of the
center is represented by Ȧ, but the velocity vanishes at a distance L from the center. Hence, we have
258 CHAPTER 10. PLASTICITY

the representative velocity gradient Ȧ/L and viscous resistance 2η Ȧ/L. The energy dissipation
within a unit volume per unit time is 2ηȦ2 /L2 . Therefore, we have the energy dissipation within the
hemisphere
2πL3 2ηȦ2 4π 2
D= × = η Ȧ L. (10.51)
3 L2 3
Since Laglange’s equation of motion with dissipation is [117]
d  ∂L  ∂L 1 ∂D
− + =0
dt ∂Ȧ ∂A 2 ∂Ȧ
so that substituting Eqs. (10.50) and (10.51), we obtain the equation of damped oscillation

mÄ + bȦ + kA = 0, (10.52)

where
π 3 4π
m=
ρL , b= ηL, k = πρgL2 . (10.53)
3 3
Note that these parameters do not include the assumed angular frequency ω. We define the angular
frequency for the free oscillation (b = 0),
( (
k 3g
ω0 = = (10.54)
m L
and the magnitude of damping
b 2η
γ= = . (10.55)
2m ρL2
Using parameters in Eqs. (10.54) and (10.55), we have the solution of Eq. (10.52) in the form of

A(t) = A0 e−γt cos Ωt, (10.56)



where A0 is the initial amplitude and Ω = ω0 2 − γ 2 .
The initial phase angle is assumed to be zero. The number of cycles until freezing depends on
the initial potential energy which determines A0 . The initial potential energy is obtained from the
shape of the transient crater
 that is approximated by a paraboloid of revolution. From Eq. (10.45)
we have the shape r = a (H − z)/H, where a is the radius of the transient crater. Then, we have
the initial potential energy3
H H
  z(H − z) 1
V0 = − ρgz πr 2 dz = − πρga2 dz = − πρga2 H 2 . (10.57)
0 0 H 6
Combining Eqs. (10.48) and (10.57), we obtain the initial amplitude
aH
A0 = √ . (10.58)
3L
3 In reality, gravitational collapse begins while the transient crater is excavated [146]. However, the initial kinetic energy

is neglected here for brevity.


10.8. POWER-LAW FLUID 259

If this is lower than A∗ in Eq. (10.47), the oscillation is not excited.


The number of cycles while the amplitude of the damped oscillation A(t) is greater than A∗
determines the number of rings of the complex crater. Let t∗ be the duration of oscillation. This is
determined by solving the equation A(t∗ ) = A∗ or

A∗ = A0 e−γt . (10.59)

Substituting Eqs. (10.47), (10.54) and (10.57) into Eq. (10.59), we obtain the duration
 
ρL2 ρgaH
t∗ = log √ .
2η 5 3cY L
The number of zero-crossing for A(t) is N = Ωt∗ /π, therefore we obtain4 .
 1/2  
1 3ρg2 L3 1 ρgH a
N= −1 log √ . (10.60)
π 4η2 5 3 cY L
Among complex craters, smaller ones have only landslides at their rims that characterize the com-
plexity. In those cases, it is possible to estimate the radius of transient crater a and the extent L
to which landslides occur from present crater morphology. Detailed analysis indicates L/a ≈ 1.5
[144]. On the other hand, H ≈ 0.4a is the depth/diameter ratio of the transient cavity of a crater.
Unknowns are the Bingham parameters cY and η in Eq. (10.60). These values can be estimated from
the crater morphology (Table 10.1).
Values in the two columns † and ‡ in Table 10.1 indicate observed smallest radii for central peak
and peak-ring craters on Earth, Moon, and other terrestrial planets and icy satellites. The density
values in the table are assigned for rock 3.0 × 103 kg m−3 and ice 0.9 × 103 kg m−3 . Assuming that
the radii correspond to a in Eq. (10.60) for N = 1 and 2, the values cY and η are evaluated and
shown in the table. Values in the column T designate the period of oscillation when the minimum
central peak craters were formed.
The estimated Bingham parameters in Table 10.1 show that the critical values cY of rocky plan-
etary bodies are 1–3 MPa. This is consistent with those estimated on p. 252 and p. 254, supporting
the Bingham model for the formation of complex craters. Large impacts gave rise to the fluids by
acoustic fluidization [144]. Interestingly, the critical values for icy satellites are one or two orders of
magnitude smaller than those of the rocky planets.

10.8 Power-law fluid


10.8.1 Definition
Rocks get fluidity at temperatures higher than about half their melting temperatures on a geologi-
cal timescale. Laboratory experiments have determined the rheological properties of rocks at high
4 The numeral “5” of the denominator in the parentheses of Eq. (10.60) comes from the numeral “5” in Eq. (10.47), which

depends on the distribution of the topographic load.


260 CHAPTER 10. PLASTICITY

Table 10.1: Bingham properties of planets and satellites [205]. T is the period of oscillation for the
formation of a minimum central-peak crater. †, the diameter at the onset of central peaks (N = 1);
‡, diameter onset of central rings or pits (N = 2).
g ρ a (km) cY η T
−2 −3
(ms ) (kg m ) † ‡ (MPa) (Pa s) (minute)
Earth
Crystalline rocks 9.8 3.0 × 103 1.9 4.8 2.23 1.6 × 107 1.6
Sedimentary rocks 9.8 3.0 × 103 1.9? 2.3 ca. 2.23 ca. 1.4 × 108 1.1
Moon
Highlands 1.62 3.0 × 103 10.9 27 2.12 1.5 × 108 9.6
Maria 1.62 3.0 × 103 8.6 19 1.68 2.9 × 108 8.0
Mercury 3.78 3.0 × 103 4.7 13 2.13 1.0 × 108 4.3
Mars 3.72 3.0 × 103 3.1 6 1.38 1.4 × 108 3.0
Icy satellites
Ganymede 1.43 0.9 × 103 < 10 ca. 4 < 0.33 < 7.8 × 107 < 4.0
Callisto 1.25 0.9 × 103 < 10 < 10 < 0.29 −
Titania 0.372 0.9 × 103 >1 < 20 > 0.01 −
Rhea 0.285 0.9 × 103 8.6 15 0.06 6.3 × 107 17
Ariel 0.251 0.9 × 103 17.0 15 0.10 1.5 × 108 18
Dione 0.224 0.9 × 103 17.0 20 0.09 1.6 × 108 22
Tethys 0.185 0.9 × 103 26.0 27.5 0.11 2.7 × 108 28
Miranda 0.083 0.9 × 103 > 25 > 25 > 0.05 −
Mimas 0.079 0.9 × 103 14.1 27.5 0.03 5.9 × 108 46

temperatures and very low strain rates. It was found that rocks with these conditions behave as
power-law fluids, equivalent strain rate and equivalent stress of which obey the equation
n
Ė E = AT E , (10.61)

where A > 0 and n > 1. Namely, the equivalent strain rate is proportional to some power of equiv-
alent stress. The parameter n is known as the power-law exponent. The power-law fluid includes
Newtonian fluids when n = 1. Differentiating both sides of Eq. (10.61) by TE , we obtain
dE˙E n  n  nĖE
= AnTEn−1 = ATE = . (10.62)
dTE TE TE
Since the viscosity of Newtonian fluids was defined as half the constant of proportionality between
the strain rate and stress (Eq. (9.3)),
TE
η= (10.63)
2ĖE
gives the effective viscosity. Combining Eqs. (10.62) and (10.63), the power-law exponent is related
to the effective viscosity as
dĖE
n = 2η . (10.64)
dTE
10.8. POWER-LAW FLUID 261

Figure 10.16: Equivalent stress versus equivalent strain rate for a power-law fluid with n > 1.
Effective viscosity η decreases with an increasing strain rate. The slope μ is smaller than 2η for this
fluid.

Pseudoplastic fluids includes power-law fluids with n > 1, and the effective viscosity decreases with
increasing ĖE , i.e., the fluids show shear-thinning behavior. Let μ = dTE /dĖE be the slope of the
graph in Fig. 10.16 and we have
n = 2η/μ (10.65)

from Eq. (10.64).


In order to apply the power-law rheology to tectonics, the constitutive equation (Eq. (10.61))
should be rewritten with tensors to account for various stress states. For this purpose and for simplic-
ity, we assume anisotropy in the rheological properties of rocks. Then, Eq. (10.36) is transformed
to proportionality Ė = λT. The equivalent strain rate and equivalent stress should satisfy this pro-
portionality so that we have ĖE = λTE . The constant is given by comparing this and Eq. (10.61):
λ = ATEn−1 . A depends on lithology and temperature. Consequently, power-law fluids have the
tensorial constitutive equation
Ė = ATEn−1 T. (10.66)

According to the strain incremental theory, strain rate and deviatoric stress have proportionality
Tij /Ėij = TE /ĖE (Eq. (10.36)). Combining this and Eq. (10.61), we obtain

−1 1 1
−1
T = T E Ė E Ė = A n Ė En Ė. (10.67)

Non-linear fluids have the effective viscosity that is defined through the equation T = 2η(D)D =
2η(Ė)Ė (Eq. (9.3)). Comparing this with Eq. (10.67), we obtain the effective viscosity of power-law
fluid
1 1 1 −1
η = A n Ė En . (10.68)
2
Definition of equivalent stress is based on the von Mises yield criterion (§10.5). When rocks
behave as power-law fluids, they obey the criterion. This is evidenced by the observation that ex-
perimental deformations by triaxial shear tests and shear tests show consistent relationships between
equivalent stresses and equivalent strain rates.
262 CHAPTER 10. PLASTICITY

Figure 10.17: Velocity profiles for pressure-driven flow of incompressible power-law fluids between
stationary parallel planes. The horizontal axis is the ratio vx /v x , where the denominator is the mean
velocity.

10.8.2 Velocity profiles for power-law fluids


Consider the velocity profile for the pressure-driven flow of an incompressible power-law fluid with
n > 1 between stationary parallel plates (Fig. 10.17). We take the x-axis in the flow direction, so that
the velocity vector has the velocity components vx > 0 and vy = vz = 0. The flow has the boundary
condition v = 0 at the walls z = ±h. The velocity profile for the fluid with n ≈ 1 may resemble
that of Newtonian fluids (Fig. 9.2), i.e., vx is the maximum and the velocity gradient dvx /dz is the
minimum at z = 0. Hence, the effective viscosity of the power-law fluid decreases towards the walls
from the flow center z = 0. Therefore, the velocity gradient near the walls may be greater than that
of Newtonian fluids, but that near the center may be smaller.
For the verification of the above argument, we firstly consider the force balance equation (Eq.
(3.23))
∂Sxx ∂Sxz
+ = 0. (10.69)
∂x ∂z
The first term in this equation equals the pressure gradient Δp/L, where L is the length along the
flow and Δp is the pressure difference between the ends of the length. The second term in Eq.
(10.69) equals Szx , which is the shear stress SS working at planes parallel to the flow. Therefore,
Eq. (10.69) is rewritten as
dSS Δp
=− . (10.70)
dz L
According to Eq. (10.61), the power-law fluid has vx = ASSn . Hence, we have
1  1
SS = A n vx n , (10.71)

where vx = dvx /dz. Substituting Eq. (10.71) into (10.70), we obtain
− 1n    1n Δp
A vx =− .
L
10.8. POWER-LAW FLUID 263

Combining the boundary conditions vx z=±h = 0, we have
 n
A Δp  n+1 
vx = h − |z|n+1 .
n+1 L

Figure 10.17 shows the graphs of this equation for various n. Velocity gradient increases toward the
walls. There is a zone at the flow center in which the velocity gradient is small. In the case of n  1,
the zone more or less behaves like a rigid body. This is called a plug, and the velocity gradient is
concentrated in the vicinity of the walls. Consequently, faults are sometimes simulated by those
narrow zones of high velocity gradients.

10.8.3 Temperature and other dependence


Laboratory experiments have demonstrated that when rocks behave as power-law fluids, their rheo-
logical parameters change significantly with temperature, obeying Dorn’s equation
 
Q
ε̇ = AD (Δσ) exp −
n
, (10.72)
RT
 
1 Q
Δσ = BD ε̇ n exp , (10.73)
nRT

where R is the gas constant, T the absolute temperature, and Q the activation energy. Q depends on
lithology. AD and BD are constants and are related to each other through the equation

BDn = 1/AD . (10.74)

AD and BD indicate the predisposition and difficulty for flow, respectively.


Power-law parameters have some pressure p-dependence. If that is significant, the scalar consti-
tutive equation is expressed as
 
 H ∗ + pV ∗
ε̇ = A (Δσ) exp −
n
, (10.75)
RT

where H ∗ and V ∗ are known as activation enthalpy and activation volume, respectively.
Unlike dislocation creep by the gliding motion along crystalline dislocations, diffusion creep by
the diffusion of individual atoms across individual crystals is easy for fine grained rocks, i.e., flow
by the diffusion has grain-size d-dependence of the form
 
H ∗ + pV ∗
ε̇ = A (Δσ)n dm exp − , (10.76)
RT

where m is a material constant slightly larger than unity5 .


5 See [97] for further reading on the rheology of rocks.
264 CHAPTER 10. PLASTICITY

Table 10.2: Rheological parameter values of selected rock types and minerals. †, abbreviations in
Fig. 10.18. ‡, MPa−n s−1 .

† AD n Q or H ∗ V∗ Reference
(GPa−n s−1 ) (kJ mol−1 ) (cm mol−1 )
3

Granite
dry Gr 5.0 3.2 123 [99, 192]
wet Gr(w) 100 1.9 137 [99, 192]
Quartzite
dry Qz 100 2.4 156 [99, 192]
wet Qz(w) 2.0 × 103 2.3 154 [99, 192]
Albite rock Ab 1.3 × 106 3.9 234 [99, 192]
Anorthosite An 1.3 × 106 3.2 238 [99, 192]
Granodiolite Qd 2 × 104 2.4 219 [99, 192]
Diabase Db ‡190, ‡8 4.7 487 [130]
Olivine
dry Ol ‡6.3 × 104 3.5 533 [99]
wet Ol(w) ‡1.9 × 103 3.0 420 [98]
Halite (rock salt) RS 5.0 × 1016 5.3 102 [99, 192]
Ice
≤ 195 K I 1.6 × 10−7 6.0 39 –13 [53]
195–240 K I 126 4.0 61 –13 [53]
240–258 K I 6.3 × 108 4.0 91 –13 [53]

Power-law parameters of several rock types are listed in Table 10.2. The values of AD are differ-
ent in order of magnitude, indicating that creep strengths are different for rock types by several orders
of magnitude. This further suggests that prediction of tectonics requires detailed knowledge about
the subsurface distribution of lithology. Continental crust has an intricate structure shaped through
its long history so that continental tectonics is more difficult to understand than oceans where the
crust has a simpler constitution.
Ductile strength is obtained by eliminating BD from Eqs. (10.74) and (10.73). Consequently,
we have  
−1 1 Q
Δσ = AD n ε̇ n exp , (10.77)
nRT
designating the viscous resistance to exerted strain rate ε̇. The rate is more or less in the order of
10−15 s−1 for the tectonics of the present Earth. Substituting ε̇ = 10−15 s−1 as a representative strain
rate into Eq. (10.77), Fig. 10.18 shows the strength of the rocks listed in Table 10.2.
Olivine is the most abundant mineral in the upper mantle. The figure designates that the mantle
is much stronger than crustal rocks if they are at the same temperature. However, Olivine has larger
temperature dependence than the crustal rocks, so that the mantle lithosphere is more weakened by
heating than the crust. This leads to the mechanical instability of the lithosphere (§12). Recent
10.8. POWER-LAW FLUID 265

Figure 10.18: Ductile strength and effective viscosity of rocks and minerals when they behave as
power-law fluids at the strain rate of 10−15 s−1 under axial stresses. The strengths (Δσ) are calculated
with Eq. (10.77). Abbreviations designating rock types and minerals are listed in Table 10.2.

experiments showed that diabase, a constituent of the crust, has a similar rheology to olivine [130].
On the other hand, rock salt is commonly deposited in thick layers of evaporite. It is not a rare
constituent of the crust, but has peculiar properties. Namely, it is much weaker than rocks by about
10 orders of magnitude, and has a smaller density (2.2 × 103 kg m−3 ). Hence, rock salt often make
diapirs and domes. Due to the sharp difference in the mechanical properties of salt and silicate rocks,
a salt dome can disturb the stress field around it [22].
Ice is also a very weak material compared to rocks. However, ice is as strong as some kind rocks
including granite at the surface temperatures of icy satellites, Outer Solar System6 . The temperature
is slightly over 100 K on Jovian moons, and some 70 K on Saturnian moons [119]. At those very
low temperatures, ice scarcely flows thus maintaining geological structures over billions of years.

6 The ice listed in Table 10.2 had grain sizes of ∼100 μm. Ice made of grains smaller than several microns has a smaller

strength than that illustrated in Fig. 10.18.


266 CHAPTER 10. PLASTICITY

10.8.4 Stability
Let us derive equations for describing the growth of perturbations in incompressible power-law
fluids. The fluids can be regarded as Newtonian fluids with the effective viscosity η(DII ) and with
the constitutive equation S = −pI + 2η(DII )D. Dividing variables in this equation into means and
perturbations, we have
    
S+$ $ II D + D
S = − p + p$ I + 2η DII + D $ . (10.78)
     
Substituting Taylor expansion η DII = η DII + η DII D $ II + · · · into Eq. (10.78), we have

        
S+$ S ≈ − p + p$ I + 2 η DII + η DII D $ II D + D $
       
≈ − p + p$ I + 2η DII D + 2η DII D $ + 2η DII D $ II D, (10.79)

where higher-order terms are neglected and η = dη/dDII . Combining Eq. (10.79) and the constitu-
 
tive equation satisfied by the mean flow S = −pI + 2η DII D, we obtain

$    
$ + 2η DII D
pI + 2η DII D
S = −$ $ II D. (10.80)

Now we assume incompressible plane strains on the O-12 plane, i.e., the stretching tensor has
the form ⎛ ⎞
D11 D12 0
D = ⎝D21 D22 0⎠ , D12 = D21 .
0 0 0
The flow is incompressible so that D is a deviatoric tensor. Hence, the second basic invariant of D
satisfies the equation with the same form as Eq. (4.13). Accordingly, we have
1 2 
DII = D11 + D22
2
+ D12
2
. (10.81)
2
Incompressibility is designated in this case by the equation, trace D = D11 + D22 = 0, so that the
above equation is transformed to
DII = D11
2
+ D12
2
. (10.82)
From (10.81), we have
   2  2  2
$ II = D11 + D
2 DII + D $ 11 + 2 D12 + D
$ 12 + D22 + D
$ 22 .

Rearranging this equation,


   
$ II = 1 D211 + 2D212 + D222 + D11 D
DII + D $ 11 + 2D12 D
$ 12 + D22 D
$ 22
2
1  $2 
+ D11 + 2D $2 + D $2 . (10.83)
12 22
2
10.8. POWER-LAW FLUID 267

The content of the last pair of parentheses is composed of second-order terms of perturbations, so
that they are negligible. On the other hand, from Eq. (10.82) we have
2 2 2
2DII = D11 + D12 + D22 . (10.84)

Comparing Eqs. (10.83) and (10.84), we have

D $ 11 + 2D12 D
$ II = D11 D $ 12 + D22 D
$ 22 . (10.85)

If pure-shear is the mean flow, we take the coordinate axes parallel to the principal axes of the
pure shear. Then, we have D12 = 0. Therefore, Eq. (10.85) is simplified to
$ II = D11 D
D $ 11 + D22 D
$ 22 . (10.86)

Substituting Eq. (10.86) into (10.80), we have


      
$
S = −$ pI + 2η DII D$ + 2 η DII · D11 D
$ 11 + D22 D
$ 22 D, (10.87)
 
$ 11 , T$22 = 2μD
or, equivalently, T$11 = 2μD $ 22 and S$12 = 2η DII D
$ 12 , where
    2
μ = η DII + 2η DII D11 . (10.88)

Using these equations, the biharmonic equation in Eq. (9.11) is rewritten as


  4
:
∂4 ψ 2μ ∂ ψ: :
∂4 ψ
4
+ 2 − 1 2 2
+ =0 (10.89)
∂x1 η ∂x1 x2 ∂x42
for power-law fluid7 [216].
Using these equations, the solution of single layer folding in Section 9.4 was extended to power-
law fluids [57, 216]. Consequently, the dominant dimensionless wavelength λ/H was found to have
strong dependence on the ratio of power-law exponents of the layer and its matrix, and that short
wavelength folds can be formed (Fig. 10.19). The flow stability of power-law fluids was employed
to explain an imbricate thrust system [68].

Exercises
10.1 Locate a point for the octahedral plane on the Mohr diagram for a triaxial stress (0 < σ3 <
σ2 < σ3 ).

10.2 To what extent does a dense rigid body sink within a low density matrix for each of the cases
where the matrix behaves as an elastic, Newtonian and Bingham fluids?

10.3 Estimate the effective elastic thickness of the folded thin elastic-plastic plate shown in Fig.
10.7.
7 The parameters η and μ are oppositely defined in [216].
268 CHAPTER 10. PLASTICITY

Figure 10.19: Dominant dimensionless wavelength λ/H of single-layer folding for power-law fluids
[216]. The horizontal axis is the ratio of the effective viscosity of the layers (1) and (2), which are
illustrated in Fig. 9.10. H is the thickness of the layer (2).
Chapter 11

Determination of Stress from Faults


Structural geologists observe deformations of rocks. However, they infer or constrain
(paleo)stresses from faults. It is important to know not only the deformation but also the
state of stress when the deformation occurs for the understanding of the mechanics of
tectonics. In this chapter we study the methods for (paleo)stress analysis. The methods
are useful to understand the present state of stress when they are applied to seismologi-
cal data. In addition, the (paleo)stress field is important to understand the migration of
fluids, including hydrocarbons and thermal water. Theories based on the Wallace-Bott
hypothesis are introduced in the first half of this chapter. The sections after Section 11.6
introduce other theories on the relationship between faulting, stress, and deformation.

11.1 Mesoscale faults


We observe fault on various scales at outcrops. Mesoscale faults are minor faults with displacements
that can be grasped at one outcrop (Fig. 11.1), i.e., their displacements range from a few millimeters
to several meters. The slip orientation of a fault is recognized by the scratches or grooves produced
by fault movement on a fault surface. They are called fault striations or simply striae (Fig. 11.2).
Offset of pre-faulting features such as strata tells the sense of shear (Fig. 11.1). Mesoscale faults
can be reactivated. Those faults have two or a few sets of striae with different orientations. For the
overlapping striae, the sense of shear can be identified for each set by asymmetrical minor structures
along the striae [181]. It is possible to determine the relative chronology of the sets by their overlap-
ping relationship. Consequently, one can recognize not only the orientation of a fault surface but the
sense of shear and slip direction for each mesoscale fault. These are referred to as fault-slip data.
For the following reasons, those faults are more often used than large-scale faults to investigate
the stresses responsible for the movement of the faults. Firstly, the much greater number density
of those faults than that of large-scale faults allows us to infer a stress field with a higher spatial
resolution. Secondly, most large faults have a complicated history as they are sometimes reactivated
and a large amount of work is needed to understand that history. Minor faults can be reactivated, but
their history is believed to be simple. Repeated reactivation of a fault increases its total displacement,

271
272 CHAPTER 11. DETERMINATION OF STRESS FROM FAULTS

Figure 11.1: Mesoscale faults (arrows) in Pliocene forearc sediments, the Miyazaki Group, northern
Ryukyu arc, Japan [266]. The sediments and the faults are truncated by the angular unconformity
(dotted line) on which Quaternary gravel layers lie, indicating that the faulting is older than the
formation of the unconformity.

but mesoscale faults have small displacements. Thirdly, the deformation of a rock mass caused by
a mesoscale fault within the mass can be treated as an infinitesimal deformation if the rock mass is
very large (§2.7).
The techniques introduced in this chapter utilize fault-slip data obtained from outcrops and ori-
ented borehole cores to constrain (paleo)stresses. If those faults were generated by the present stress
field, the stress that they indicate is the present stress. The same techniques can be applied to seis-
mological data to infer the present stress.
Several methods have been proposed to estimate the state of stress from faults. Among them,
the most popular one is Anderson’s theory of faulting [4]. The theory is still used, as it is very
simple. However, one can trace modern methods back to the papers of the 1950s by Wallace and
Bott [20, 252]. Bott [20] explained the abundance of oblique slip faults, which were incompatible
with Anderson’s theory (§6.3). McKenzie utilized Bott’s principle to infer the state of stress at
an earthquake source [137]. The principle was used to formulate a mathematical inverse method
by Carey and Brunier [35] and by Angelier [6] in the 1970s to infer paleostresses from geological
11.2. WALLACE-BOTT HYPOTHESIS 273

Figure 11.2: Fault striations on a fault surface in Early Miocene strata, the Atsumi Formation, North-
east Japan. There are comet-like structures heading right. They were produced by the drag of hard
particles between fault surfaces. These asymmetric structures indicate that this fault is dextral in
sense, i.e., the block on this side of the fault moved left.

faults. The inverse method was applied to many areas in the world. Since the mid 1990s, theoretical
investigations have been carried out on the numerical techniques to separate stresses from fault data
that record plural stresses [60, 161, 209, 264].

11.2 Wallace-Bott hypothesis


11.2.1 Basic equations
Modern methods for determining (paleo)stresses are based on the Wallace-Bott hypothesis [20, 252],
stating that the shear traction applied on a given fault plane causes a slip in the direction and orien-
tation of that shear traction, irrespective of the faults created in an intact rock or along a pre-existing
fracture.
The traction vector at the fault plane whose unit normal is n is given by the equation

t(n) = r · n. (11.1)

From Eqs. (3.16) and (3.17), the normal and tangential components of this vector are the normal
and shear traction vectors

σN = N · r · n = n n · (r · n) (11.2)

σS = (I − N) · r · n = r · n − n n · (r · n) . (11.3)
274 CHAPTER 11. DETERMINATION OF STRESS FROM FAULTS

Faulting occurs to release the shear stress so the slip direction is indicated by the unit vector −σS /|σS |
instead of σS .
The slip direction of a given fault has a linear relationship with the stress tensor r. To see
this, let us rewrite Eq. (11.3) as σS = f (r, n). Then the linearity is represented by the equation
(1) (2)
f (a1 r(1) + a2 r(2) , n) = a1 f (r(1) , n) + a2 f (r(2) , n), where σS and σS are stress tensors and a1
and a2 are arbitrary scalar variables. Equation (11.3) is rewritten for these tensors as
(1)
σS = (I − N) · r(1) · n,
(2)
σS = (I − N) · r(2) · n,

so that we have
(1) (2)
a1 σS + a2 σS = a1 (I − N) · r(1) · n + a2 (I − N) · r(2) · n

= σS = (I − N) · a1 r(1) + a2 r(2) · n.

Consequently, the slip direction predicted by the Wallace-Bott hypothesis has a linear relationship
with stress tensors.

11.2.2 Graphical expressions of slip directions


The slip direction predicted by the Wallace-Bott hypothesis for a given fault and a stress is shown by
Means’ graphical method1 [142]. A tangent-lineation diagram is a graphical expression of fault-slip
data and is convenient for theoretical considerations.

Means’ graphical method Suppose that a state of stress and a fault orientation are given, and that
the vector s points to the slip direction on the fault by the stress. The vector is obtained through the
following steps. Firstly, the stress axes are plotted on a stereonet whose center indicates the pole to
the fault (Fig. 11.3). The point P on the stereonet denotes the pole. Secondly, let δ1 be the angular
distance between the σ1 -axis and the pole, and δ3 be that between the σ3 -axis and the pole. We
calculate the parameters
1
a(1) = (σ1 − σ2 ) sin 2δ1 , (11.4)
2
1
a(3) = (σ2 − σ3 ) sin 2δ3 . (11.5)
2
Thirdly, the arrows for the unit vectors u(1) and u(2) are drawn from the point P. The former is directed
on the opposite side of the σ1 -orientation on the stereonet. The latter is directed from the point P to
the σ3 -orientation. Using the parameters defined by Eqs. (11.4) and (11.5), we produce the vectors
s(1) = a(1) u(1) and s(3) = a(3) u(3) and draw them on the stereonet (Fig. 11.3). Finally, the point S is
plotted on the base circle of the stereonet, where the ray PS is parallel to the vector s = s(1) + s(3) .
The point S indicates the slip direction on the fault plane by the given stress.
1 An exercise at the end of this chapter takes up the theoretical basis for this method.
11.2. WALLACE-BOTT HYPOTHESIS 275

Figure 11.3: Graphical method to determine the slip direction of a fault whose pole is plotted at the
center of a stereonet.

Equations (11.4) and (11.5) demonstrate that the slip direction depends not only on the orienta-
tion of stress axes but also on the ratio
σ1 − σ 2
R= . (11.6)
σ2 − σ3
This is related to the stress ratio as R = Φ/(1 − Φ), i.e., the slip direction depends on Φ. If the ratio
is zero, the ray PS0 indicates the slip direction. The ray PS1 indicates the slip direction for the case
of Φ = 1. It should be stressed that Φ influences the slip direction as much as the principal stress
orientations do.
The graphical method demonstrates that slip directions are directed away from the σ1 -axis and
toward the σ3 -axis. If the fault in Fig. 11.3 is nearly perpendicular to the σ2 -axis, the angles δ1 and
δ3 are approximately equal to 90◦ . Therefore, a(1) ≈ a(3) ≈ 0, though their signs depend on the tiny
difference of the angles from 90◦ . This further indicates that the slip direction, s, is swerved around
the σ2 -axis.

Tangent-lineation diagram A tangent-lineation diagram is a convenient graphical method to il-


lustrate the slip directions.
The plane containing the slip vector and the pole to the fault plane is called the M-plane of the
fault. The plane perpendicular to the M- and fault planes is know as an auxiliary plane (Fig. 11.4(a)).
A fault-slip datum is expressed in the diagram by an arrow plotted on a lower-hemisphere stereonet
(Fig. 11.4(b)). The position of the arrow in the stereonet indicates the pole to the fault plane. The
arrow is drawn parallel to the M-plane, and is pointing at the slip direction of the footwall.
Figure 11.5 shows tangent-lineation diagrams for five different stress ratios but with common
principal stress axes. The pattern made up of arrows on a diagram shows a source and sink at the
σ1 - and σ3 -axes, respectively2 . As the cases Φ = 0 and 1 indicate axial stresses, the pattern for those
2 The source and sink are called fabric attractors in metamorphic geology [175].
276 CHAPTER 11. DETERMINATION OF STRESS FROM FAULTS

Figure 11.4: (a) Schematic picture showing the orientation of a reverse fault plane and the direction
of fault movement (bold arrows). The line OQ indicates the orientation of striae, and the point P
is the pole to the fault plane. The arrow attached to the point P indicates the slip direction of the
footwall block. The plane containing the triangle OPQ is called the M-plane for the fault. The plane
perpendicular to the fault plane and M-plane is called the auxiliary plane. (b) Lower-hemisphere
stereonet showing the three planes and the slip direction of the footwall that is depicted by the arrow
attached to the point P. The line OB is the intersection of the fault and auxiliary plane.

cases shows axial symmetry about the σ1 - and σ3 -axis, respectively. The pattern exhibits orthorhom-
bic symmetry for triaxial stresses, and a saddle point appears at the σ2 -orientation. Tangent-lineation
diagram for a set of natural fault-slip data is shown in Fig. 11.6.

11.2.3 Indeterminacy of stress from faults


Fault-slip data constrain the state of stress that was responsible for the faults. However, not all the
stress components are determined. The Wallace-Bott hypothesis indicates that slip directions are
not affected by pore fluid pressure, although it controls the strength of faults. The effective stress
is given by the equation r = r − pI. Thanks to the linearity, the slip direction due to r is the
linear combination of those by r and −pI. However, the latter is an isotropic stress and causes no
shear traction. Therefore, the pore fluid pressure does not affect the slip direction of a fault. This is
convenient for paleostress analysis, because it is difficult to know the pore fluid pressure on a fault
surface when the fault moved.
It is also difficult to estimate the depth of burial of a fault when it was activated. However, the
depth has little affect on the slip direction. Crustal stress is limited by the brittle strength (§6.7),
and faulting occurs when stress is brought to the limit. Equation (6.21) indicates that the limiting
stress is proportional to the depth z. Let r0 be an arbitrary stress tensor. Since the slip directions
due to stresses r0 and r = qr0 are the same, the slip directions are independent from the depth
of burial. However, this is valid only as an approximation, because stress states inferred by in-situ
measurements more or less deviate from the proportionality.
11.2. WALLACE-BOTT HYPOTHESIS 277

Figure 11.5: Tangent-lineation diagrams showing the slip directions of faults by stresses with the
same principal orientations but different stress ratios. The principal orientations are shown in the
lower right box. Open circles in the stereonets indicate the poles to the fault plane.

The slip directions predicted by the Wallace-Bott hypothesis for the stresses r and

r = qr0 − pI (11.7)

are the same for a given fault, whether or not p and q are interpreted as pore pressure and depth.
When the state of stress is illustrated by Mohr circles, p and q indicate the position of the circles on
the abscissa and the size of the circles, respectively. The stresses that have common principal orien-
tations and similar stress ellipsoids have similar Mohr circles, and result in identical slip directions.
The tensor that represents those stresses is termed reduced stress tensor.
Due to the insensibility of the slip direction to these parameters, fault-slip it is not enough to
constrain the mean and differential stresses. However, this is convenient for paleostress analysis,
because we do not need to worry about the depth and pore fluid pressure when each fault was
activated3 . These quantities are difficult to estimate.
It is convenient to assume a special form for r0 to consider slip directions. For example, we can
3 Researchers have attempted to determine all the stress components from fault-slip data with assumption such as the brittle

strength of rocks [8].


278 CHAPTER 11. DETERMINATION OF STRESS FROM FAULTS

Figure 11.6: Natural fault-slip data obtained from the Otadai Formation, mid Quaternary forearc
basin deposit in central Japan [151]. The left panel is the tangent-lineation diagram showing the
data. The right panel is the lower-hemisphere stereonet in which the orientations of fault planes
are indicated by great circles. Open circles on the great circles designate the orientation of fault
striations, and arrows attached to the open circles show the sense of shear. If an arrow is attached,
it indicates the slip direction of the hanging wall. If two arrows are attached to an open circle, the
arrows depict the sinistral or dextral sense of shear.

express any stress tensor by combining Eq. (11.7) and


⎛ ⎞
1 0 0
r0 = RT · ⎝0 Φ 0⎠ · R, (11.8)
0 0 0

where R indicates the principal stress orientations and Φ = (σ2 − σ3 )/(σ1 − σ3 ) is the stress ratio.
The tensor between RT and R in Eq. (11.8) is a reduced stress tensor, but the definition of a reduced
stress tensor is not unique [7]. Different shapes of reduced tensor is introduced in Section 11.5 and
in Appendix B.
The tensor r0 in the shape of Eq. (11.8) is enough to calculate the slip directions of faults. The
orthogonal tensor R has three degrees of freedom corresponding to the three Euler angles (Fig. C.5).
The reduced stress tensor has one degree of freedom. Therefore, r0 has four degrees of freedom.
However, the stress tensor has six degrees of freedom. The difference corresponds to p and q in Eq.
(11.7). The tensor r0 represents the principal orientations and stress ratio. In this chapter, the term
“a state of stress” is used to denote the stresses with a common r0 .
A fault-slip datum does not tightly constrain the principal stress orientations. It should be noted
that different stresses can result in a fault-slip datum. The admissible orientations for the datum are
illustrated in Fig. 11.7. The possible principal σ1 orientation reduces by a decrease of stress ratio,
which in turn expands the possible σ3 orientation. It is known that both the σ1 - and σ3 -axes exist on
one side of the plane AOC [122]. Namely, if the domain OBC in Fig. 11.7 has the σ1 -axis, then the
11.2. WALLACE-BOTT HYPOTHESIS 279

Figure 11.7: Stereogram (upper hemisphere) showing the principal stress orientations compatible
with a fault-slip datum. The upper block is assumed to move rightwards on the plane ABCD. P-
and T-axes are contained in the plane AOC and make angles of 45◦ with O. They approximate
the σ1 - and σ3 -orientations if the Coulomb-Navier criterion holds for this fault. The Wallace-Bott
hypothesis predicts the σ1 -axis in the region OBCD for this fault. The σ3 -axis is oriented in the
region OBAD. The σ1 -orientation is constrained more tightly for a prescribed stress ratio. Dotted
lines with decimal fractions show the rightward limit of the permissible region for the σ1 -axis, where
the fractions indicate stress ratios. The lines are obtained from Eq. D.45.

σ3 -axis must exist in the domain OAB.

For example, if there are conjugated N–S trending normal faults formed by the E–W extensional
stress with Φ = 0.5 (Fig. 11.8(a)), the stress compatible with the fault-slip data is not unique.
Stresses with a wide variation of principal orientations and stress ratios are compatible with the data.
However, if there are varieties of fault attitudes and of slip directions, the admissible stresses are
constrained as the overlapping region of the dotted and hatched regions in Fig. 11.8(a).

Figure 11.8(b) shows the case of two conjugated normal faults with different ages and different
extensional orientations. If we do not know the difference and we regard them as the results of the
same state of stress, the vertical axial compression is a possible stress compatible with the data, i.e.,
it is not possible to distinguish from the data which scenario is correct, the two-phase or the single-
phase stress history. The set of stress states compatible with given data set are said to be associated
stresses [264]. Other lines of evidence are needed to determine the correct stresses.
280 CHAPTER 11. DETERMINATION OF STRESS FROM FAULTS

Figure 11.8: Schematic illustrations showing the ambiguity of stresses inferred from fault-slip data.
(a) Conjugated dip-slip normal faults and the principal orientations (squares) of the stress that
formed the faults. Dotted and hatched regions designate the σ1 orientations compatible with the
west- and east-dipping normal faults, respectively. These regions correspond to the solid region in
Fig. 11.7. (b) Two sets of conjugated dip-slip normal faults. The sets were formed successively
by NE–SW and NW–SE extensional stresses. Both stereograms are lower-hemisphere equal-angle
projections.

11.3 Stress inversion


It is possible to determine the state of stress from fault-slip data. The tangent-lineation diagram in
Fig. 11.6 shows a set of natural data collected from mesoscale faults in mid-Quaternary sedimentary
rocks in central Japan. Obviously, many arrows in the diagram are directed along a WNW–ESE
orientation, suggesting that the faults corresponding to the arrows were activated by WNW–ESE
trending extensional stress.
This inference is mathematically formulated as the following inverse method, by which the opti-
mal state of stress is determined from given fault-slip data [6, 7, 190]. Assuming the principal stress
orientations and a stress ratio, we compose a stress tensor r0 using Eq. (11.8). Secondly, the theo-
retical slip direction is calculated by Eq. (11.3) for each fault. Thirdly, let Δ(i) be the angular misfit
between the theoretical and observed slip directions of the ith fault. As we assumed the principal
orientations and stress ratio to be arbitral, it is natural that the directions are not parallel. Then, the
sum

N
S= f (Δ(i) ) (11.9)
i=1

can be used as the measure of misfit of the assumed stress state to the data set, where N is the
number of faults and f ( ) is a monotonously decreasing function such as that shown in Fig. 11.9. S
is a function of not only the data but also of the principal orientations and stress ratio. Accordingly,
the optimal state of stress is determined by maximizing S(r0 ). In inverse theory, this is called an
11.3. STRESS INVERSION 281

Figure 11.9: Graph of a monotonously decreasing function f (Δ) used in stress inversion.

object function.
This method is referred to as stress inversion4 . Contrasting with the recent numerical techniques
that are introduced in the next section, this is sometimes called the classical inverse method. The
stress inversion is a non-linear inverse method, because the slip direction is denoted by the unit
vector −σS /|σS |. The non-linearity comes from this division [150].
The optimal stress determined by the stress inversion has uncertainty resulting from the indepen-
dence of the slip direction from p and q in Eq. (11.7). The absolute values of stress components
are not known. Instead, the tensor defined by Eq. (11.8) is determined, i.e., the optimal principal
orientations and the optimal stress ratio are obtained by the inverse method.
Figure 11.10 shows the optimal stress determined through the stress inversion for the natural data
in Fig. 11.6. The result is a WNW–ESE trending extensional stress, which coincides with the stress
roughly inferred by eye from the tangent-lineation diagram of the data.
However, the optimal solution is not satisfactory as the histogram of angular misfits is definitely
bimodal. About two-fifths of the faults have misfits greater than 60◦ . The optimal stress cannot
explain the observed slip directions of those faults, indicating that the fault-slip data are heteroge-
neous. Fault-slip data is called heterogeneous if plural stress states were responsible for the faults.
Consequently, the forearc region in central Japan experienced a polyphase stress history since the
Otadai Formation, from which the data were obtained, deposited in the mid Quaternary [151].
Given heterogeneous fault-slip data the object function S(r0 ) has maxima, and the optimal so-
lution might be a meaningless solution [265]. Thus, we have to input homogeneous fault-slip data
for the calculation. For this purpose, researchers try to sort faults at outcrops by their apparent ages.
There are several criteria [10, 39, 250]. The principal guides for distinguishing the different fault
sets are (1) consideration of stratigraphy or of the age of the rocks affected by a certain deformation,
(2) characterization of the synsedimentary faults such as fault drag, (3) cross-cutting relationships,
(4) superimposed striae on the same fault plane and (5) association of mineral veins. Unfortunately,
these criteria are not always available in the field, especially in young geologic units such as the mid-
Quaternary Otadai Formation. In addition, the resemblance of the orientation of fault planes and slip
directions is sometimes employed for fault sorting. However, many faults with the same orientations
4 See [9] for further reading.
282 CHAPTER 11. DETERMINATION OF STRESS FROM FAULTS

Figure 11.10: (a) Lower-hemisphere stereonet showing the optimal stress determined from the data
shown in Fig. 11.6 via the classical inverse method. The number of data is N = 53. The σ1 - and
σ3 -axes are nearly vertical and horizontal so that this is a normal faulting regime of stress. The
optimal stress ratio is Φ = 0.09 so that the optimal stress is approximately of axial compression. (b)
Histogram showing the angular misfits for the optimal stress illustrated in the left stereonet.

and the same slip directions can be activated by different stresses (Fig. 11.7). Therefore, such fault
classification implicitly restricts solutions prior to using the stress inversion.
Researchers have tackled the problem how to separate stresses from heterogeneous fault slip
data, and several numerical techniques have been proposed. Before explaining one of the techniques
in Section 11.5, we will introduce an other formulation of the classical stress inversion in the next
section.

11.4 Fry’s formulation of stress inversion


The classical stress inversion is a non-linear inverse method [150]. Fry (1999) [60] divided the
inversion into linear and non-linear parts as follows.
Given a fault-slip datum, a unit vector b on the fault plane is defined perpendicular to the ob-
served slip direction, i.e., b is the null vector [80]. The vector is parallel to the line OB in Fig. 11.4.
If the theoretical slip direction is parallel to the observed one, the traction vector upon the fault plane
should lie on the M-plane. Therefore, b should be perpendicular to the traction vector t(n), i.e.,
b · (r · n) = 0. This is rewritten using components of these vectors and tensors as

3
bi σij nj = 0. (11.10)
i,j=1

The symmetry of r allows us to rewrite the left-hand side of this equation as



3 
3 
3
σij bi nj = σji bi nj = σij bj ni . (11.11)
i,j=1 i,j=1 i,j=1
11.4. FRY’S FORMULATION OF STRESS INVERSION 283

In the last transformation, the dummy indices i and j are exchanged. Consequently, Eq. (11.10) is
rewritten as
3
bj σij ni = 0. (11.12)
i,j=1

Adding both sides of Eqs. (11.10) and (11.12) gives5


3
σij fij = 0, (11.13)
i,j=1

where
fij = bi nj + bj ni .

Note that n and b are observable quantities but r is unknown. Accordingly, we can compose the
tensor fij from the fault-slip datum. In addition, fij has six degrees of freedom, because of the
symmetry fij = fji . Therefore, Eq. (11.13) is further rewritten as the scalar product

s · f = 0, (11.14)

where s = (σ11 , σ12 , σ13 , σ22 , σ23 , σ33 )T and f = (f11 , f12 , f13 , f22 , f23 , f33 )T . The components
of these vectors are independent components of the tensors σij and fij .
Equation (11.14) indicates that the vectors s and f meet at right-angles in six-dimensional space
if the assumed stress r fits the fault-slip datum. If |s · f | does not vanish, r is not appropriate for the
datum. Accordingly, we define the unit vector f = f /|f | so that the endpoint of the position vector
f is represented by a point on the six-dimensional unit sphere.
Let f (k) be the unit vector for the kth datum. If the stress is compatible with all the fault-slip
data, then the stress tensor must satisfy the equation


N
s · f (k) = 0. (11.15)
k=1

(1)
This equation has the geometrical interpretation that the end points of the unit vectors, f ,...,
(N)
f , are on the great circle whose pole is parallel to the vector s. The scalar product s · f becomes
negative in sign if the angle between the two vectors is greater than 90◦ . Therefore, the stress
inversion can be reformulated as the least-squares method to minimize the quantity


N
2
s · f (k) .
k=1

5 Comparing the following equation with Eqs. (2.49) and (2.54) in Section 2.7, it is seen that Eq. (11.13) indicates that

the work done by the shear traction in the orientation perpendicular to the observed slip direction is nil.
284 CHAPTER 11. DETERMINATION OF STRESS FROM FAULTS

This is equivalent withto


N     !
 6
(k)
6
(k)
N 
6
(k) (k)

6 
N
(k) (k)
si fi sj fj = si fi fj sj = si fi fj sj = s · g · s,
k=1 i=1 j=1 k=1 i,j=1 i,j=1 k=1

where  

N
(k) (k)
g= fi fj
k=1
is a six-dimensional, second-order, symmetric tensor, and is composed from given fault-slip data.
The optimal stress determined by the stress inversion has uncertainty resulting from the indepen-
dence of the slip direction from p and q in Eq. (11.7). The absolute values of the stress components
are not known. Accordingly, the uncertainty allows us to assume s to be a unit vector.
The least-squares method that we should solve is to search the optimal vector s that minimizes
the quadratic form s · g · s with the constraint s · s = 1. The constraint designate that the end point
of the vector is on the six-dimensional unit sphere. This extreme value problem is solved through
Lagrange’s method of the undetermined multiplier, i.e., we seek the extremal of the function,

L = s · g · s − λ(s · s − 1) = 2s · (g − λI) · s − λ,

where λ is the Lagrange multiplier. Since the tensor (g − λI) is symmetric, the formula in Eq. (C.61)
applies to this tensor, and we obtain
∂L
= 2(g · s − λs) = 0.
∂s
Consequently, we obtain the characteristic equation,

g · s = λs. (11.16)

A symmetric tensor has real eigenvalues. Therefore, all the roots of this equation are real. We are
seeking the eigenvector s that minimizes L. Noticing s · s = 1, we obtain s · g · s = s · λs = λ from
Eq. (11.16). The left-hand side of this equation equals L so that the minimum absolute eigenvalue
corresponds to the minimum L, and the corresponding eigenvector gives the optimal stress [209].
It was shown that the optimal vector s is parallel to the pole to the great circle along which the
(1) (N)
endpoints of the data vectors f , . . . , f are aligned. Fitting of a great circle to the data vectors
is a eigenproblem to solve Eq. (11.16). This is a linear inverse problem.
Once a great circle is fitted, we have two poles for that circle. If one of the poles is represented by
s, the other is −s. These vectors correspond to the stresses r and −r. Therefore, we have to choose
one of them. The slip directions corresponding to the stresses are given by −σS = ±(N − I) · r · n
(Eq. (11.3)). That is, the stresses causes the opposite sense of movement for entire faults so that
we can choose one of the stresses corresponding to the correct sense of movement. The choice of
the correct pole is a non-linear operation. The stress inversion is, therefore, divided into linear and
non-linear parts6 .
6 See Appendix B and [203, 269].
11.5. MULTIPLE INVERSE METHOD 285

Figure 11.11: Schematic illustrations showing the principle of the multiple inverse method. (a) Data
points are aligned on two great circles (dotted lines) on the six-dimensional unit sphere. (b) White
arcs showing parts of the great circles that are fitted to k-element subsets of the data points, where k
generally equals 5. (c) The poles of the great circles are plotted to exhibit the optimal stress for the
subsets. The poles make clusters at the poles for the great circles indicated by the dotted lines in (a).

11.5 Multiple inverse method


The purpose of this section is to explain the multiple inverse method, a numerical technique used for
separating stresses from heterogeneous fault-slip data [264].

11.5.1 Principle
It was shown in the previous section that the stress inversion is compared to fitting a great circle to
data points on the six-dimensional unit sphere, and that one of the poles of the great circle represents
the optimal stress. Therefore, heterogeneous fault-slip data are represented by data points that are
aligned on plural great circles on the unit sphere. Suppose that a set of observed faults is a mixture
of two assemblages that were activated by two different stresses. Then, the fault-slip data are repre-
sented by two great circles on the unit sphere (Fig. 11.11(a)). Consequently, it is the problem how
to identify the great circles. The basic idea of the method is that the generalized Hough transform
is applicable to this problem. The transform is a technique of artificial intelligence to detect objects
with arbitrary shapes and orientations [31]. The objects in our problem are stress ellipsoids with
different shapes and different orientations.
The method firstly constructs k-element subsets from the entire fault-slip data. Given N faults,
we have N Ck subsets, where
N!
N Ck =
k!(N − k)!
is the binomial coefficient. Secondly, great circles are fitted to the subsets7 . Thirdly, the poles of the
great circles that represent the optimal stresses for the subsets are plotted on the unit sphere. There
7 It was shown that the resolution of this method is greatly improved by choosing a particular kind of subsets [168].
286 CHAPTER 11. DETERMINATION OF STRESS FROM FAULTS

Figure 11.12: Left and right stereograms (lower-hemisphere equal-area projection) showing the σ1 -
and σ3 -orientations, respectively, of the optimal stresses determined by the multiple inverse method
with k = 5 applied to the fault-slip data shown in Fig. 11.6. Stress ratios are indicated by different
colors. Clusters (shaded) indicates that WNW–ESE and NNE–SSW trending extensional stresses
affected the faulting in the Otadai Formation.

are eventually two clusters of the poles to the two great circles. The clusters are expected to appear at
the correct positions if the data points are plotted within narrow belts along the two great circles, as
most of the great circles fitted to the k-element subsets are nearly parallel to either of the two great
circles (Fig. 11.11(b)). However, if the belts are wide or the number of data points is small, the
density of the clusters decreases. The density designates the statistical significance of the stresses
represented by the clusters.
Tthe final task is to detect the clusters. For this purpose, the optimal stresses are mapped onto
a stereogram. The stress ratios of the optimal stresses are depicted by colors. Figure 11.12 shows
an example, where WNW–ESE and NNE–SSW trending extensional stresses are detected from the
fault-slip data exhibited in Fig. 11.6. The former is represented by the cluster of blue squares in
Fig. 11.12, and the color indicates Φ ≈ 0.1. Therefore, the classical inverse method detected the
WNW–ESE extension (Fig. 11.10).

11.5.2 Average and spread of clusters


The right stereogram in Fig. 11.6 demonstrates that the density and spread of the clusters indicate
the WNW–ESE extension is statistically more significant than the NNE–SSW extension. The spread
of a cluster is quantified as follows.
Firstly, we introduce the stress difference, a parameter to quantify the difference between two
stress states [164]. The uncertainty related to p and q in Eq. (11.7) demonstrates that the mean
and differential stresses are indeterminable for the stress inversion. In addition, the definition of
11.5. MULTIPLE INVERSE METHOD 287

a reduced stress tensor is not unique. Accordingly, we define the reduced stress tensor r so as to
satisfy the two conditions
trace r = σ 11 + σ 22 + σ 33 = 0 (11.17)
and the octahedral shear stress of unity. Equation (11.17) indicates that r is a deviatoric tensor so
that the second basic invariant of this tensor is given by Eq. (4.13). Following Eq. (4.17), we have
the second condition as
1 2 
σ 11 + σ 222 + σ 233 + 2σ 212 + 2σ 223 + 2σ 231 = 1. (11.18)
3
Using the reduced stress, we have the stress tensor R · r · RT for the stress tensor inversion.
(1) (2)
Secondly, given two reduced stress tensors, r and r , the difference between the stress states
(1) (2)
represented by the tensors is defined by the octahedral shear stress of the tensor D = r − r .
Namely, the difference is given by

1
D= (Δ11 − Δ22 )2 + (Δ22 − Δ33 )2 + (Δ33 − Δ11 )2 + 6 (Δ12 )2 + 6 (Δ23 )2 + 6 (Δ31 )2 . (11.19)
3
Orife and Lisle [164] term this parameter “stress difference”, and demonstrate that D is in the range
between 0 and 2 for any combination of reduced stress tensors provided that they satisfy the con-
ditions indicated by Eqs. (11.17) and (11.18). The similarlity between two reduced stress tensors
is indicated by (2 − D). If D = 0, the two stress states result in the same slip directions whatever
the orientations of faults are. If D = 2, the two stress states cause faulting with the opposite slip
directions. The couple of stresses with D = 2 are called inverse stresses.
(1)
Finally, the spread of stress states is defined as follows. Given m reduced stress tensors, r , . . . ,
(m)
r , their average is defined by
1  (1) (m)

r = r + ··· + r . (11.20)
m
The average principal orientations are obtained as the eigenvectors of this tensor, and the average
stress ratio is given by the equation
σ2 − σ3
Φ = ,
σ1 − σ3
where σi is the ith largest eigenvalue of the tensor r. The spread of the reduced stress tensors
around the average is defined by the average stress difference
1 (1)
D = D + · · · + D (m) , (11.21)
m
(i)
where D (i) is the stress difference between r and the reduced stress tensor derived r. The greater
the scatter of the stress states the larger the average stress difference. Therefore, D is a measure of
the spread of stress states8 .
8 he average of the values of 2 sin−1 (D/2) is more useful than that of D values. See Appendix B.
288 CHAPTER 11. DETERMINATION OF STRESS FROM FAULTS

Figure 11.13: Mohr circles for a triaxial stress and the slope tan χ.

The multiple inverse method yields clusters of stress states, and the center and spread of a cluster
are evaluated by Eqs. (11.20) and (11.21). The identification of those clusters can be computerized
using the k-means algorithm [167], a technique of artificial intelligence [52], with the average and
spread of the clusters.
It is difficult to apply stress inversion to fold belts as the timing of faulting relative to folding is
difficult to know. Therefore, it is sometimes assumed a-priori that a reverse faulting regime of stress
accompanied folding.
The bedding tilt test used in paleomagnetism can be applied to this problem using the multiple
inverse method and the average and spread of clusters [270]. Paleomagnetists uses the statistic α95 to
evaluate the spread of paleomagnetic vectors determined at various parts of a fold [32]. The larger the
statistic, the more the vectors are scattered. If rocks in the fold acquired the paleomagnetism before
folding, untilting of the vectors makes α95 smaller. If, on the other hand, the rocks were magnetized
after folding, the statistic becomes larger by the untilting. Therefore, the statistic works as an object
function to be minimized in the inverse problem to determine the optimal timing magnetization
relative to folding.
Similarly, we are able to use D instead of α95 to be minimized in the inverse problem for
determining the optimal timing of faulting relative to folding. This test procedure was applied to
mesoscale faults in a Quaternary fold in central Japan and revealed that the majority of the faults
were activated by a strike-slip faulting regime of stress at the middle of folding [270].

11.6 Slip tendency


The Wallace-Bott hypothesis concerns the fact that any faults can be activated, unless the fault plane
is one of the principal planes of stress. Shear stress on the plane vanishes so that those faults cannot
move. However, if fault planes have the same coefficient of friction, faults nearly parallel to the
principal plane may be difficult to move compared to those oblique to the principal planes. The
magnitude of shear stress is depicted by the Mohr diagram (Figs. 4.6 and 11.13).
Slip tendency [157] is a yardstick for the mobility of fault surfaces under a given stress r. The
11.6. SLIP TENDENCY 289

Figure 11.14: Stereonets showing slip tendency Ts for three stresses with the same principal orien-
tations (open circles) but with different stress ratios, Φ = 0, 0.5 and 1.

frictional resistance of the fault with the unit normal n is μf σN = μf (n · r · n). A fault is thought
to be preferable for r if the shear stress on the fault, σS = |(I − N) · r · n|, is much large compared
with Fs , i.e., the slip tendency is defined as
   
 σS  1  (I − N) · r · n 

Ts ≡   =  .
μ f σN  μ f  n · r · n 

Ts is a function of n, r and μf . Given r and μf , Ts has the maximum Tsmax for the fault orientations.
For simplicity, we assume that all faults have the same coefficient of friction to investigate the
relationship of the mobility with fault orientations. Then, the angle χ shown in Fig. 11.13 is related
to the slip tendency as tan χ = Ts . The points A, B and C have a common normal stress but different
shear stresses. The points A and B have maximum and minimum χs, respectively. It is obvious that
the fault plane designated by point D has the maximum χ, because the line OD is tangent to the outer
Mohr circle. However, the Mohr diagram is not convenient to see how the slip tendency changes
with different orientations.
Therefore, the ratio Ts /Tsmax is a convenient measure to consider the orientation-dependence of
the slip tendency. The variation of this ratio is shown by the gray scales in Fig. 11.14. The central
panel in this figure illustrates the relationship between the slip tendency and fault orientations for
triaxial stresses. There are two dark spots in the stereogram, demonstrating that the fault surfaces
parallel to the σ2 -axis and making an acute angle with σ1 -axis have a maximum slip tendency under
triaxial stresses. The two spots represent conjugate faults. The slip tendency has a orthorhombic
symmetry with respect to the principal planes of stress.
In contrast, the slip tendency for axial compression and axial tension has axial symmetry with
respect to the σ1 - and σ3 -axes, respectively (Fig. 11.14). The planes tangent to a cone whose axis is
parallel to the symmetry axis of stress are referentially activated as faults [95].
In reality, every fault may have its own coefficient of friction, and pore pressure may be different
for each faulting event. Accordingly, the slip tendency is not the only factor to indicate the mobility.
However, Fig. 11.14 helps in considering fault activity under a specified state of stress.
290 CHAPTER 11. DETERMINATION OF STRESS FROM FAULTS

Figure 11.15: Strain of a rock mass by the activity of fault sets in the mass, where faults belonging
to a set have similar fault orientations and slip directions. Rectangle ABCD indicates the initial top
surface of the mass. The z-axis is defined as being vertically downward. Although displacement is
shown along two or several faults for each set, the parallel faults carry similar displacements. (a)
The activity of a set of normal faults parallel to the y-axis and slip direction normal to the y-axis, the
mass suffers a plane strain on the xz-plane. The points B and C are transferred to B and C by the
faulting, resulting in an inclined top surface AB C D. (b) Two sets of faults making up conjugated
normal faults. If the total displacements of each set are equal to each other, the top surface remains
horizontal only to subside to the rectangle A B C D . The deformation of the mass is approximated
by a pure shear. (c) Four sets of faults leading to a three-dimensional coaxial deformation, i.e., the
top surface is kept horizontal.

11.7 Faulting controlled by strain


A rock mass encompassing faults deforms by the fault activity (Fig. 1.17). If there are a lot of faults
with much smaller displacements than the size of the mass, the deformation can be regarded as a
continuous plastic deformation.
Faulting obeying Anderson’s theory leads only to a plane strain (Exercise 11.1). However, strains
may be generally three-dimensional. The faulting predicted by the Wallace-Bott hypothesis allows
three-dimensional strains. When we used the hypothesis, stress was assumed to control faulting.
However, stress is not observable. We observe the strain of a rock mass through faults, and the
hypothesis transforms strain into stress. In contrast, what would happen if strain controls faulting?
The present and next sections deal with this problem.
Natural faults often form groups with similar orientations and slip directions. Here, we sort faults
by the orientations and directions into sets of faults (Fig. 11.15). One region may have multiple sets
of faults, which may or may not have been activated at the same time.
For simplicity, we assume that all faults belonging to a set have the same displacements and that
they are uniformly distributed in the rock mass. Then, the deformation of the mass due to internal
fault activity can be regarded as a uniform deformation. The deformation by a single set of faults
11.8. RECHES’ MODEL 291

equals an infinitesimal simple shear in the M-plane of the faults. The top surface ABCE in Fig.
11.15(a) is transformed into the inclined rectangle AB C D by this simple shear unless the fault
planes are horizontal. Two sets of faults enable pure shears that keep the top surface horizontal (Fig.
11.15(b)). If there are four sets of faults, coaxial deformations are possible (Fig. 11.15(c)).
Not only coaxial but also any three-dimensional deformations are possible if there are at least
four fault sets [193]. Here, faulting is assumed not to accompany any volume change of the rock
mass, so that we have |F| = 1, where F is the deformation gradient of the rock mass. F has nine in-
dependent components in three dimensions. However, the volume conservation reduces the degrees
of freedom of F to eight. One set has two degrees of freedom, the rake of slip direction on the fault
planes and the parameter q of the simple shear (Eq. (1.13)), if the orientations of fault planes are
given. Therefore, four fault sets are sufficient to accommodate a general incompressive deformation
in three dimensions.
If a three-dimensional deformation controls faulting, at least a few sets of faults should be acti-
vated to accommodate the deformation. Fault systems comprised of four fault sets are often observed
in the field and in clay cake experiments [131, 163, 193, 195].

11.8 Reches’ model


We now consider faulting controlled by a prescribed strain of a rock mass that includes the faults.
Reches [193, 194] applied a classical theory of the plastic deformation of single crystals. The plastic
deformation is due to the movement of dislocations through the crystal lattice. The lattice structure
allows a limited number of slip systems whose glide planes and slip directions are constrained.
For example, a face-centered cubic lattice has 12 slip systems. Minimizing energy dissipation,
Taylor (1938) linked the slippage on the planes with the macroscopic deformation of the crystal
containing the slip systems [239]. Unlike single crystals, a rock mass has numberless planes of
fractures with various orientations that are capable of faulting. However, the following argument
reduces the possible number of fault sets to four.
It is assumed that there are numberless pre-existing fractures capable of accommodating an ap-
plied three-dimensional coaxial strain, but that most preferable surfaces are activated. The rock
mass is also assumed to be homogeneous and isotropic in the broad view so that the stress and
strain have the same principal orientations. Since the controlling factor is the coaxial strain that has
an orthorhombic symmetry about the principal planes of strain, the resultant faulting may have the
same symmetry, i.e., the four sets of faults the orientations of which are arranged symmetrically
with respect to the planes (Fig. 11.16(a)). Four sets are necessary and sufficient to accommodate the
prescribed strain of the rock mass. To keep the symmetry, they carry the same magnitude of strain.
Here, they are called orthorhombic fault sets.
Each fault set has many faults but the displacements are very small compared to the dimension
of the rock mass. In addition, faults belonging to a set are located with constant intervals. Therefore,
the deformation of the mass is regarded as continuous and homogeneous. Let the unit vectors n(i) =
292 CHAPTER 11. DETERMINATION OF STRESS FROM FAULTS

Figure 11.16: (a) Orthorhombic fault sets arranged symmetrical with respect to the principal planes
of stress. Equal-angle projection. The rectangular Cartesian coordinates O-123 are defined as being
parallel to the principal axes of strain. The O-1 axis is directed out of the page. Ordinal numbers of
the fault sets are encircled. Gray arrows points the slip directions. (b) The rock mass encompassing
the fault sets is assumed to suffer a coaxial deformation by the slippage on the faults, each of which
carries a simple shear. The unit vectors n(i) and s(i) represent the pole to the fault planes and slip
direction of the ith set, respectively.
   
(i) (i) (i) (i) (i) (i)
n1 , n2 , n3 T and s(i) = s1 , s2 , s3 T be the unit normal and slip direction of the ith fault
set (Fig. 11.16(b)), where the coordinates O-123 are defined parallel to the principal axes of stress
of the rock mass (Fig. 11.16(a)). From the orthorhombic symmetry, we have
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
n1 n1 −n1 −n1
n(1) = ⎝ n2 ⎠ , n(2) = ⎝−n2 ⎠ , n(3) = ⎝−n2 ⎠ , n(4) = ⎝ n2 ⎠ (11.22)
n3 n3 n3 n3
and ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
s1 s1 −s1 −s1
s(1) = ⎝ s2 ⎠ , s(2) = ⎝−s2 ⎠ , s(3) = ⎝−s2 ⎠ , s(4) = ⎝ s2 ⎠ , (11.23)
s3 s3 s3 s3
where 0 ≤ n1 , n2 , n3 . The hypothesis of orthorhombic fault sets results in a tight constraint between
n and s.
Suppose the rock mass has M faults. Let f(i) = I + δf(i) be the deformation gradient due to the
ith fault, where i represents the sequence of faulting. Then the total strain of the rock mass is given

by the multiplication F = f(M) · · · f(1) = I + δf(M) · · · I + δf(1) ≈ I + f(M) + · · · + f(1) , where
second-order terms are neglected. Since the addition of matrices is commutative, the total strain is
independent from the order of fault activity. Therefore, we define F(i) = I + δF(i) as the deformation
gradient due to the faults belonging to the ith fault set, when we have the deviation of the total strain
from I,
δF = δF(1) + δF(2) + δF(3) + δF(3) . (11.24)
11.8. RECHES’ MODEL 293

Comparing Eqs. (2.53) and (11.24), we have

δF(i) = γ (i) s(i) n(i) , (11.25)

where γ (i) represents the contribution of the ith set to the total strain. F(i) represents an infinitesimal
simple shear. Therefore, taking a proper coordinate system, we have also the following expression
(Eq. (1.13)): ⎛ ⎞ ⎛ ⎞
1 2q (i) 0 0 2q (i) 0
F(i) = ⎝0 1 0⎠ ∴ δF(i) = ⎝0 0 0⎠ .
0 0 1 0 0 0
Since q in these matrices represents the infinitesimal shear strain that is half of the engineering
shear strain (p. 34), γ (i) in Eq. (11.25) represents the engineering shear strain for the ith set. The
orthorhombic symmetry requires

γ ≡ γ (1) = γ (2) = γ (3) = γ (4) . (11.26)

Combining Eqs. (11.22)–(11.26), we obtain the total deformation


⎛ ⎞ ⎛ ⎞
n1 s1 n2 s1 n3 s1 n1 s1 −n2 s1 n3 s1
δF = γ ⎝n1 s2 n2 s2 n3 s2 ⎠ + γ ⎝−n1 s2 n2 s2 −n3 s2 ⎠
n1 s3 n2 s3 n3 s3 n1 s3 −n2 s3 n3 s3
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
n1 s1 n2 s1 −n3 s1 n1 s1 −n2 s1 −n3 s1 n1 s1 0 0
+ γ ⎝ n1 s2 n2 s2 −n3 s2 ⎠ + γ ⎝−n1 s2 n2 s2 n3 s2 ⎠ = 4γ ⎝ 0 n2 s2 0 ⎠.
−n1 s3 −n2 s3 n3 s3 −n1 s3 n2 s3 n3 s3 0 0 n3 s3

Hence, the symmetry of this tensor indicates that the deformation of the rock mass due to the or-
thorhombic fault sets is irrotational X = O, as we have so assumed. Therefore, from Eq. (2.10), we
obtain the infinitesimal strain tensor
⎛ ⎞
n1 s1 0 0
e = 4γ ⎝ 0 n2 s2 0 ⎠. (11.27)
0 0 n3 s3

We have defined the coordinate system parallel to the principal axes of strain. Hence, from Eq.
(11.27) we have

ε1 = 4γn1 s1 , (11.28)
ε2 = 4γn2 s2 , (11.29)
ε3 = 4γn3 s3 . (11.30)

The volume conservation is expressed by

ε1 + ε2 + ε3 = 0. (11.31)
294 CHAPTER 11. DETERMINATION OF STRESS FROM FAULTS

Table 11.1: Strain states and the parameters Φε and k.


k Φε Principal strains Strain state
1 1 ε1 = ε2 Constrictional strain
0 1/2 ε2 = 0, ε1 = −ε3 Plane strain
–1/2 0 ε2 = ε3 Flattening strain

The principal strains are defined to satisfy ε3 ≤ ε2 ≤ ε1 . Therefore, ε1 ≥ 0, ε3 ≤ 0. The


intermediate principal strain ε2 has positive and negative signs depending on the strain of the rock
mass. Now, we introduce the parameter
ε2 − ε3
Φε = (0 ≤ Φε ≤ 1)
ε1 − ε3
to indicate the shape of the strain ellipsoid. Φε = 0 and 1 designate flattening (ε3 = ε2 < ε1 ) and
constrictional (ε3 < ε2 = ε1 ) strains, respectively. Φε in-between indicates a triaxial strain ellipsoid,
whereas Φε = 1/2 represents plane strain. Since Φε is defined as the same form as the stress ratio
(Eq. (4.5)), Φε is convenient to envisage a strain ellipsoid. However, the parameter

k = ε2 /ε1 (−1/2 ≤ k ≤ 1) (11.32)

is convenient for the following arguments. This is related with Φε as


1 + 2k 2Φε − 1
Φε = , k=− .
2+k Φε − 2
The correspondence between k and the strain states are shown in Table 11.1. The parameter k is
controlled from outside the system.
From Eqs. (11.28) and (11.29), we have ε1 n2 s2 = ε2 n1 s1 and, using the parameter k, we obtain

kn1 s1 − n2 s2 = 0. (11.33)

On the other hand, n and s satisfies the relations

n1 s1 + n2 s2 + n3 s3 = 0, (11.34)
n21 + n22 + n23 = 1, (11.35)
s21 + s22 + s23 =1 (11.36)

by definition. The total number of components of n and s is six, but Eqs. (11.33)–(11.36) designate
that only two of them are free. Accordingly, here we choose n1 and n2 as the independent variables
for the following argument. Eliminating s2 , s3 and n3 from Eqs. (11.33)–(11.36), we obtain
;
<
< 1 − n21 − n22
s1 = ±<<   !. (11.37)
< 1 − n 2
= 1 − n2 − n 2 1
k 2 + 2k
2 1
n22
11.8. RECHES’ MODEL 295

The eliminated values s2 , s3 and n3 are calculated using the four equations, i.e., s is linked with n
through the orthorhombic symmetry. Once the two variables n1 and n2 are specified, the fault activity
to accommodate the prescribed strain is constrained.
The next problem is how the variables n1 and n2 are chosen. To solve this problem, we assume
that the frictional resistance
τR = τ0 + (tan φ) σN (11.38)

works on a fault plane with the parameters τ0 and φ common to all faults. Namely, while a fault is
moving, the shear stress on the fault plane is equal to

σS = τR (11.39)

Note that the rock mass is assumed to be isotropic. Therefore, the principal axes of stress coincide
with the coordinate system, and the normal and shear stresses satisfy the following relationship with
principal stresses:

σN = t(n) · n = σ1 n21 + σ2 n22 + σ3 n23 , (11.40)


σS = t(n) · s = σ1 n1 s1 + σ2 n2 s2 + σ3 n3 s3 . (11.41)

Rearranging Eqs. (11.33)–(11.41), we obtain


     
σ1 − σ3 n1 s1 − n21 tan φ + σ2 − σ3 kn1 s1 − n22 tan φ = τ0 + (tan φ)σ3 . (11.42)

This relationship is satisfied when a fault is slipping.


We have assumed a uniform spacing between faults. Consider a set to have n faults within a cube
that has a unit length of sides. If one of the faces of the cube is parallel to the faults, the spacing
equals 1/n. Since the engineering shear strain of the cube is γ, the displacement of each fault is γ/n.
The cube deforms in a similar way to a deck of cards. The n faults move by the distance γ/n against
the resisting force τR × (unit area) = τR . Therefore, the energy dissipation by the n faults within the
unit volume of the cube is
w = σS γ. (11.43)

Combining Eqs. (11.28), (11.33), (11.34), (11.41) and (11.43), we obtain the total dissipation due
to the four sets in a unit volume The four sets of faults dissipate in a unit volume of the rock mass
  ε1
4w = 4σS γ = 4 σ1 n1 s1 + σ2 n2 s2 + σ3 n3 s3 ·
4n1 s1
ε1
= σ1 n1 s1 + σ2 n2 s2 + σ3 n3 s3 − (σ3 n1 s1 + σ3 n2 s2 + σ3 n3 s3 )
n1 s1
    ε1
= σ1 − σ3 n1 s1 + σ2 − σ3 n2 s2
n1 s1

= (σ1 − σ3 ) + (σ2 − σ3 )k ε1 .
296 CHAPTER 11. DETERMINATION OF STRESS FROM FAULTS

Accordingly, the total dissipation is proportional to the strain ε1 . Let us define the symbol W to
denote the constant of proportionality

W = (σ1 − σ3 ) + (σ2 − σ3 )k. (11.44)

This is the dissipation for every incremental strain ∂(4w)/∂ε1 = W . Using Φ = (σ2 − σ3 )/(σ1 − σ3 )
and Δσ = σ1 − σ3 , Eq. (11.44) is rewritten as

W = (1 + Φk)Δσ (11.45)

and Eq. (11.42) is rewritten as


 
(1 + Φk) n1 s1 − n21 + Φn22 tan φ Δσ = τ0 + (tan φ)σ3 . (11.46)

The friction is thought to be dependent on confining pressure (Eq. (11.38)), so we regard σ3 as a


variable controlled from outside this system instead of the confining pressure.
Since there are pre-existing fractures with various orientations in the rock mass, some of which
are chosen as faults, then the faults that can be activated by the minimum differential stress may be
chosen. The right-hand side of Eq. (11.46) and Δσ are, by definition, positive in sign. In addition,
the right-hand side is given. Hence, the minimization of Δσ is achieved by determining the optimal
combination of n1 , n2 and Φ that maximizes the content of the brackets in the left-hand side of Eq.
(11.46). We have the analytical solutions for the cases k = −1/2, 0 and 1.
Constrictional strain is represented by k = 1. Then, the stress tensor and the fault sets are
expected to have axial symmetry. Therefore,

σ1 = σ2 , n1 = n2 . (11.47)

Substituting these into Eq. (11.37), we have


9
1 − 2n21
s1 = . (11.48)
2
Substituting Eqs. (11.47) and (11.48) into Eqs. (11.45) and (11.46), we obtain

2Δσ = W, (11.49)
⎛ 9 ⎞
1 − 2n21
2Δσ ⎝n1 − n21 tan φ⎠ = τ0 + (tan φ)σ3 . (11.50)
2

Therefore, the dissipation W is minimized when Δσ is minimized (Eq. (11.49)). The optimal n1 is
obtained by maximizing the content of the parentheses in the left-hand side of Eq. (11.50). Solving
the equation
⎛ 9 ⎞ 9
d ⎝ 1 − 2n1
2
1 − 3n1
n1 − n21 tan φ⎠ = − 2n1 tan φ = 0,
dn 2 2(1 − 2n1 )
11.8. RECHES’ MODEL 297

and using Eqs. (11.47) and (11.48), we obtain the fault orientation and slip direction of the case of
constrictional strain:
1 1
n1 = 1 − sin φ, n2 = n1 , s1 1 + sin φ. (11.51)
2 2
Plane strain is indicated by k = 0 and ε2 = 0. Due to the symmetry,

n2 = 0. (11.52)

Combining Eqs. (11.37) and (11.52), we have



s1 = 1 − n21 (11.53)

and combining Eqs. (11.45) and (11.46), we obtain

Δσ = W, (11.54)
  
Δσ n1 1 − n1 − n1 tan φ = τ0 + (tan φ)σ3 .
2 2
(11.55)

Δσ fluctuates with W , again. Similar to the above discussion, it is found that the minimum Δσ
requires √ √
2 2
n1 = 1 − sin φ, n2 = 0, s1 = 1 − sin φ. (11.56)
2 2
Flattening is indicated by k = −1/2 and ε2 = ε3 . Due to symmetry, we have

n2 = n3 , σ2 = σ3 . (11.57)

Combining Eqs. (11.35), (11.37) and (11.57),


 
n1 = 1 − 2n22 , s1 = 3(1 − n21 ). (11.58)

In this case, we have

Δσ = W, (11.59)
  
Δσ n1 1 − n1 − n1 tan φ = τ0 + (tan φ)σ3 .
2 2
(11.60)

Δσ is minimized when
√ √
2 1 2
n1 = 1 − sin φ, n2 = 1 + sin φ, s1 = 1 + sin φ. (11.61)
2 2 2
Interestingly, the three cases have a common minimum differential stress, which is obtained from
Eqs. (11.49)–(11.51), (11.54)–(11.56) and (11.59)–(11.61):
2 cos φ
Δσmin = [τ0 + (tan φ)σ3 ] . (11.62)
1 − sin φ
298 CHAPTER 11. DETERMINATION OF STRESS FROM FAULTS

Figure 11.17: Equal-area projection showing the pole of the fault surfaces activated to accommodate
a prescribed coaxial strain of the rock mass that includes the faults [194]. The pole of one of the
orthorhombic fault sets is shown on a quadrant.

The resistance is assumed to be proportional to σ3 (Eq. (11.38)), which is the substitution of con-
fining pressure. The differential stress inherits the proportionality. The constant of proportionality
equals 2 cos φ tan φ/(1 − sin φ). Substituting a representative value of rocks, e.g. φ = 35◦ , we have
(σ1 − σ3 ) ∝ 2.7σ3 , i.e., σ1 is about 1.7 times greater than σ3 .
The stress conditions and optimal fault sets were determined numerically for general triaxial
strain cases −1/2 < k < 0 and 0 < k < 1 by Reches [194]. It was found that the optimal parameters
are continuous over the above intervals of k. Consequently, the optimal fault orientations and slip
directions derived above are assembled into the equations
√ ( ( T
2 (1 − sin φ) k(1 − sin φ) 
n = , , 1 + sin φ
2 1+k 1+k
√ ( ( T (11.63)
2 (1 + sin φ) k(1 + sin φ) 
s = , , 1 − sin φ
2 1+k 1+k
for 0 ≤ k ≤ 1 and

2    T
n = 1 − sin φ, |k|(1 + sin φ), (1 − |k|)(1 + sin φ)
2
√ (11.64)
2    T
s = 1 + sin φ, |k|(1 − sin φ), (1 − |k|)(1 − sin φ)
2
for −1/2 ≤ k < 0.
The orientations of n and s predicted by Eqs. (11.63)–(11.64) are illustrated in Fig. 11.17. In
the case of plane strain (k = 0), we have
√ √
2   T 2   T
n= 1 − sin φ, 0, 1 + sin φ , s = 1 + sin φ, 0, 1 − sin φ , (11.65)
2 2
11.8. RECHES’ MODEL 299

indicating that all faults are parallel to the σ2 -axis and slip directions are perpendicular to this axis.
Therefore, the four fault sets degenerate into a conjugate pair (Fig. 11.18(a)). Let θ be the angle
between n and the σ3 -axis. This component is given in Eq. (11.65), and is rearranged as

1
n23 = (1 + sin φ) (11.66)
2
1 sin φ
 √ 2        √ 2  2
2 2 φ 2 φ φ φ 2 φ φ
= cos + sin + 2 sin cos = cos + sin
2 2 2 2 2 2 2 2
 2    
π φ π φ π φ π/2 − φ
= cos cos + sin sin = cos2 − = cos2 .
4 2 4 2 4 2 2

Since n3 is the direction cosine of n, cos θ = n3 , we obtain

π
2θ = − φ. (11.67)
2

This is identical to the shear plane angle predicted by the Coulomb-Mohr criterion (Eq. (6.4)).
Consequently, plane strain leads to conjugate faulting.
As k departs from zero, faults make greater angles with σ2 -axis (Eqs. (11.63)–(11.64)), while
the angle between n and ε3 -axis is constant for any k ≥ 0. In this case, n23 has the same form as the
right-hand side of Eq. (11.66). Therefore, the constant angle equals θ in Eq. (11.67).
The angle between n and ε1 -axis is constant for any k < 0. Let θ  be the constant angle (Fig.
11.17). From Eq. (11.64), we have

 √ 2  √ 2  
2 2 2 φ 2 φ φ φ
n21 = (1 − sin φ) = cos + sin − 2 sin cos
2 2 2 2 2 2
 √ 2  2
2 φ φ
= cos − sin
2 2 2
 2    
π φ π φ π φ π/2 + φ
= sin cos − cos sin = sin2 − = cos2 .
2 2 2 2 4 2 2

Combining this and the equation cos θ = n1 , we obtain 2θ  = π/2 + φ. Therefore, we have Eq.
(11.67) again, where 2θ + 2θ  = π, i.e., 2θ  is the supplementary angle of 2θ which is equal to ∠CBF
in Fig. 6.5(a).
In the field, orthorhombic fault sets may be recognized by the fault-slip data shown in Fig. 11.17
and by the cross-cutting relationship between faults of different sets, which warrants the simultane-
ous activity of the fault sets. Krantz [104] identified orthorhombic fault sets in Jurassic sedimentary
rocks. Reches [195] estimated the strain of a rock mass that encompasses a well-exposed system of
orthorhombic fault sets, in good agreement with the theory.
300 CHAPTER 11. DETERMINATION OF STRESS FROM FAULTS

Figure 11.18: Fault sets accommodating a prescribed strain.

11.9 Block rotation and faulting


Highly deformed terrains have intricate fault systems, some of which were accompanied by the
rotation of fault blocks, which is sometime indicated by curved fault striations9 . Orogenic and
accretionary belts may have rotated macroscale blocks. If fault activity was abruptly quenched in
a rock mass, fault striations found in the mass may exhibit slip directions just before the end of
the deformation. This section deals with a method to estimate a macro-deformation from internal
mesoscale faults.
Instantaneous deformation is described by LΔt, where L is the velocity gradient tensor and Δt is
a short time interval. The tensor can be divided into two parts (Eq. (2.26)), indicating coaxial and
rigid body rotation, L = D + W. In the previous sections in this chapter, we have assumed that W
is negligible compared to D for determining paleo stress or strain. If this is valid, it can be expected
that an isotropic rock mass has a linear constitutive equation that links D and stress. However, it is
9 Curved fault striations do not always indicate block rotations [224].
11.9. BLOCK ROTATION AND FAULTING 301

the present problem to incorporate the rotation in the fault striation analysis.
To this end, the Twiss-Protzman-Hurst model utilizes the mechanics of micropolar continua
[249]. Micropolar continua are physically idealized macroscopic continua that have microscopic
rigid particles and can model such substances as liquid crystals, rigid suspensions, concrete with
sand, and muddy fluids. Within the present context, the rigid particles simulate fault blocks that can
rotate. The mechanics of micropolar continua is beyond the scope of this book. Interested readers
can consult textbooks [58, 128]. Here, the results of the Twiss-Protzman-Hurst model is briefly
introduced [249].
Micropolar continua is a limited version of micromorphic continua that has deformable mi-
crostructures. The mechanics of micromorphic continua has macro and micro fields of kinematic
variables including position, velocity, and rotation. The macro velocity gradient tensor is written as
usual: L = D + W. The micro velocity gradient tensor is also defined as L = D + W, i.e., we distin-
guish macro and micro quantities by overlines. Fault blocks are assumed to be rigid so that we use
the mechanics of micropolar continua. Then we have D = O, and the micro velocity gradient tensor
equals the micro spin tensor (L = W). Here, the principal values of D are numbered in descending
order, D1 ≥ D2 ≥ D3 . Obviously,
1 1 
ΔD ≡ D1 − D3
2 2
is the maximum shearing rate, just like the maximum shear stress equals half the differential stress.
The dimensionless quantity
D2 − D3
ΦD ≡ (0 ≤ ΦD ≤ 1)
D1 − D3
designates the shape of the instantaneous strain ellipsoid. Here, elongation is positive so that ΦD = 0
and 1 indicate constriction and flattening, respectively. ΦD = 1/2 corresponds to plane strain.
Taking the principal axes of D to define the rectangular Cartesian coordinates O-123, the max-
imum shear rate is on the O-13 plane. The non-diagonal components of a spin tensor make up a
vorticity vector (Eq. (2.43)), and the vorticity is related to an angular velocity (Eq. (2.44)). The
macro and micro angular velocities
 within the O-13 plane are equal to W 13 and W13 , respectively.
The quantity W 13 − W13 indicates the difference in the velocities. Accordingly, we define the
dimensionless number
W 13 − W13
Ω≡
ΔD/2
to indicate the difference.
Let N be the outward unit normal to a small part on the surface of a fault block. The surface is
rubbed by the relative velocity of the surface of the neighboring block. Twiss et al. [249] derived
the direction of the resultant striae, being parallel to the vector
⎛ ⎞T ⎛ ⎞ ⎛ ⎞
1 − N12 − ΦD N22 N1 N3
Ω
v = ⎝ ΦD − N12 − ΦD N22 ⎠ ⎝N2 ⎠ + ⎝ 0 ⎠ . (11.68)
2
−N1 − ΦD N2
2 2 N 3 −N 1
302 CHAPTER 11. DETERMINATION OF STRESS FROM FAULTS

Figure 11.19: Tangent-lineation diagrams showing fault-slip data predicted by the Twiss-Protzman-
Hurst model with different Ω and the same ΦD = 0.5 and the same principal axes of D [249]. The
D-axis indicates the orientation of the maximum instantaneous elongation.

Given the dimensionless numbers ΦD and Ω, the striating direction is calculated with this equation
for any surface with N. If the micro rotation is quenched by the cessation of macro deformation, the
striation on the surface of fault blocks may give constraints to the kinematic parameters ΦD and Ω
and the principal orientations of D.
The fault-slip data predicted by this model is shown in Fig. 11.19. The tangent-lineation dia-
grams in Fig. 11.5 show the fault-slip data specifically for the case of Ω = 0. When |Ω| is small, the
gross pattern in the tangent-lineation diagram is similar to that of the case of Ω = 0. However, it no
longer has the orthorhombic symmetry but makes up a triclinic system. For a large Ω, arrows in the
tangent-lineation diagram show a rotational pattern about D2 -axis.

Exercises
11.1 Suppose a rock mass that is cut by a great number of conjugate sets of small-scale faults.
Show that the strain of the mass caused by the internal fault activities is a plain strain. Assume that
the orientation of stress axes does not change in the body, and that faulting are generated following
the Coulomb-Navier criterion.

11.2 Show that the point S in Fig. 11.3 moves from S0 to S1 with an increase of ΦB from 0 to
1, where the points S0 and S1 represent the orientation of the σ1 and σ3 axes, respectively, on the
stereonet.

11.3 Verify that the σ1 -axis with a specific stress ratio Φ is constrained by a fault-slip datum to be
in the region to the left of the dotted line in the stereonet in Fig. 11.7.
Chapter 12

Dynamics of the Lithosphere

In this chapter we study how the lithosphere deforms in response to external forces, and
how factors such as the thermal regime and lithology affect the deformation. First, we
study several factors controlling the strength of the lithosphere. Then, the strength is
related to the instability of the lithosphere that determines the fate of continental rifts
and the creation of spatially periodic deformation zones.

12.1 Strength profile


We have studied several deformation mechanisms, including elasticity, brittle faulting, and ductile
flow, that are relevant to describe the deformation of rocks. They exhibit different stresses for a
given tectonic condition with a specific depth, temperature, strain, strain rate, etc. When a rock mass
deforms, it has options for the mechanisms. Then, the deformation mechanism that minimizes the
resistance to the deformation is chosen. If the stress accompanied by brittle deformation is less than
that of ductile deformation, the latter deformation mechanism is chosen. As the product of force and
distance is equal to energy, tectonic deformation proceeds with the minimum energy dissipation.
Consider large-scale deformations such as a continental breakup resulting from extensional de-
formation of the lithosphere. In those cases we do not take elasticity into account because those
deformations surpass the elastic limit. Brittle and ductile deformations are important. If the defor-
mation of a rift can be treated as plane strain on the vertical section across the rift zone, the resisting
stress is described by Eq. (6.21) as follows. No brittle faulting occurs if the differential stress,
σh − σv , is less than the critical one1 :
!
2μf (1 − λf )
Δσbrittle = − σv , (12.1)
(1 + μ2f )1/2 + μf

where σv is the vertical stress that is usually equal to the overburden stress (Fig. 12.1). In the case
1 Note that rifting occurs when σh < σv , so that σh − σv < 0.

303
304 CHAPTER 12. DYNAMICS OF THE LITHOSPHERE

of compressional tectonics, the critical stress is given by the equation


!
2μf (1 − λf )
Δσbrittle = + σv , (12.2)
(1 + μ2f )1/2 − μf

and faulting occurs when the differential stress exceeds this. A rock mass breaks when it is subject
to stresses over this limit. Therefore, Eqs. (12.1) and (12.2) represent the brittle strength of the
rock mass under extensional and compressional tectonic regimes, respectively. The brittle strength
is insensitive to lithology.
The vertical stress, σv , is determined by overburden so that the critical differential stress de-
scribes the magnitude of horizontal stress. If tectonic stress is gradually built up and the horizontal
stress reaches the critical stress, faulting ‘immediately’ occurs in the geological sense. The brittle
deformation relieves the state of stress under the critical state. Consequently, the state of stress is
at or just below the strength under tectonically active regions: Eqs. (12.1) and (12.2) describe the
approximate magnitude of horizontal stress.
On the other hand, the stress due to ductile deformation depends on the lithology, strain rate, and
temperature. The deviatoric stress for the ductile deformation is given by the equation (§10.8.3),
 1 1  Q 
− −1
s = A n ε̇En exp ė, (12.3)
nRT
where ε̇E is the equivalent strain rate (
1
ε̇E = ė : ė. (12.4)
2
Δσductile = τ1 − τ3 decreases rapidly with increasing temperature. If the lithology and strain rate do
not change within a rock mass, temperature rises with depth and, consequently, Δσductile falls rapidly.
Because of the temperature increase with depth, the ductile strength decreases with it. Hence, the
abscissa in Fig. 10.18 that displays temperatures can be translated into depths. Taking the horizontal
x- and vertically downward z-axes, and assuming a pure shear
⎛ ⎞
ε̇xx 0 0
ė = ⎝ 0 0 0 ⎠ (ε̇xx = −ε̇zz ),
0 0 ε̇zz

we have ε̇E = 2 |ε̇xx | from Eq. (12.4). If rocks are isotropic, we have Δσ = |τxx − τzz | and the
ductile strength
 − 1   1 −1  Q  
Δσ = 2A n ε̇xx  n exp ε̇xx − ε̇zz . (12.5)
nRT
Temperature T increases generally with depth, therefore Δσ decreases exponentially with it. The
ductile strength strongly depends on the lithology, as Eq. (12.5) includes material constants A and
Q. When differential stress applied to the lithosphere is smaller than the ductile strength or brittle
strength of rocks over the entire profile of the lithosphere, the lithosphere supports the stress by
elasticity.
12.1. STRENGTH PROFILE 305

Figure 12.1: Schematic strength profile of the lithosphere with a uniform lithology. Brittle and duc-
tile strengths are shown by solid lines. Rocks deform by brittle failure above the brittle-ductile
transition (BDT) because brittle deformation is easier than ductile deformation in that the former
needs less differential stresses than the latter, whereas ductile deformation prevails under BDT. The
strength of quasi-plastic deformation is shown by dotted lines. Since the brittle strength in exten-
sional tectonic regimes is smaller than that in compressional tectonic regimes (§6.7), the lithosphere
is more vulnerable to extensional than compressional tectonics.

The brittle strength increases and the ductile one decreases with increasing depth. Deformations
are accommodated either by a brittle or ductile mechanism that needs lower differential stress for the
deformation. Hence, there must be a transition region at which brittle deformation gives way to duc-
tile deformation. This is known as the brittle-ductile transition2 . It is often found that earthquakes
occur in the crust no deeper than 15–20 km. This is probably because the transition occurs at that
depth.
The solid lines in Fig. 12.1 shows the yield strength of constituent rocks at each depth, so that the
graph exhibiting such strengths is called the strength profile of the lithosphere. There is a difference
in the brittle strengths in normal and reverse faulting stress regimes (Fig. 6.14). The former strength
is smaller than the latter one so that the brittle-ductile transition in the normal faulting regime is
deeper than the other.
The oceanic lithosphere has abundant olivine minerals which control the mechanical strength of
both the crust and the lithospheric mantle. Therefore, the oceanic lithosphere has strength profiles
like those in Fig. 12.1. By contrast, the continental crust has a much more complex structure that
has been shaped in the long history of the crust. It has significant heterogeneity in lithology. Ductile
strength strongly depends on the lithology. In addition, impermeable rocks make overpressured rock
2 The fact is that there is a regime where the deformation of rocks is accommodated by ductile and brittle mechanisms at

the same time, known as quasi-plasticity , around the brittle-ductile transition (Fig. 12.1) [213]. However, the quasi-plastic
regime is not well understood, so that we do not refer to that regime any more.
306 CHAPTER 12. DYNAMICS OF THE LITHOSPHERE

Figure 12.2: Schematic picture showing the strength heterogeneity of the crust associated with the
heterogeneity in lithology and other factors including pore fluid pressure and temperature. Strength
is designated by the gray scale.

Figure 12.3: Diagram showing the strength profile of continental lithosphere with homogeneous
crust. Dashed lines indicate brittle strengths for two stress regimes. In this figure a brittle regime
exists in the upper mantle for the case of extensional tectonics. However, the existence depends on
the temperature of the depths.

masses. Rocks with abundant radioactive elements affect the thermal regime in the crust. Therefore,
the continental crust must have an intricate strength distribution, which is schematically shown in
Fig. 12.2. However, we assume simple structures for the following considerations. Namely, we
assume that the entire crust is composed of one type of rock, or that the crust has two layers with
different lithology. If the crust has a single layer, the continental lithosphere has a strength profile
like those in Fig. 12.3.
12.2. STRETCHING INSTABILITY OF THE LITHOSPHERE 307

12.2 Stretching instability of the lithosphere


Infinitesimal strain until whole lithosphere failure

Consider that Δσ(z) denotes the strength profile, then the integral
 tL
Fc = Δσ dz (12.6)
0

is the maximum force for the lithosphere to support without finite pure shear deformation, where
the depth tL denotes the base of the lithosphere. The force is designated by the area bounded by the
strength profile.
Unfortunately, the lithosphere-asthenosphere boundary is fuzzy, i.e., the strength decreases con-

tinuously as exp Q/nRT (z) in the boundary zone. Accordingly, the base of the lithosphere is
sometimes defined for convenience as the depth where Δσ  σzz is satisfied. More simply, the
depth is sometimes defined as the depth where the differential stress reaches a small specific value
in the range 10–20 MPa [191].
What if a tectonic force greater than Fc in Eq. (12.6) is applied to a continental lithosphere
that was initially at rest? Consider the initial distribution of Δσ being equal to that of the Earth
pressure at rest down to a certain depth that defines the base of the lithosphere (Fig. 12.4). Under
that depth, the initial state of stress is assumed to be lithostatic. Once the force is applied, rocks near
the surface yield because of the near-surface low brittle strengths. The differential stress in the rocks
near the base of the lithosphere exceeds the ductile strength so that the rocks yield immediately
after the force is applied. Namely, the elastic core begins to be eroded at the top and base. The
reduced thickness of the elastic core leads to an increase of force that the core have to support. The
increase further reduces the elastic core to form a positive feedback cycle [109] and, the entire depth
of the lithosphere eventually yields. This is called whole lithosphere failure. Until this phenomenon
occurs, the existence of the elastic regime prohibits finite deformation of the lithosphere. After the
failure of the whole lithosphere, finite deformation of the lithosphere begins.

Finite strain

The history of finite deformation of a rift can be revealed from syn- and post-rift sedimentary records.
Continental rifting can lead to continental breakup. However, some rifting is sometimes aborted. The
North Sea rift is an example of failed rifts. Intra-arc rifts have the two ends, also. Intra-arc rifting
leads to back-arc spreading as the rifting in the Izu-Bonin-Mariana arc resulted in the spreading
in the Shikoku and the Mariana basins of the Philippine Sea plate. On the other hand, there are
the aborted rift basins of the Paleogene under the continental shelf behind the Ryukyu arc to the
northwest of the plate.
The duration of intra-arc rifting is one order of magnitude smaller than that of intra-continental
rifting [263]. If rifting has several phases, each phase lasts tens of million years in continents. The
activity of the Triassic rift in Fig. 3.24 is an example. By contrast, intra-arc rifting lasts a few million
308 CHAPTER 12. DYNAMICS OF THE LITHOSPHERE

Figure 12.4: Schematic picture showing the evolution of differential stress in the continental litho-
sphere leading to whole lithosphere failure.

years. The extensional tectonics that led to the opening of the Japan Sea backarc basin lasted 5 m.y.
[263].
The difference is explained by the initial thermal structure of the lithosphere. Intra-arc rifting
occurs in volcanic arcs because magmatism has heated up and reduced the ductile strength of the
lithosphere before rifting. For example, temperature is estimated by a petrological method in the
Northeast Japan volcanic arc at 850 and 1400◦ at 25 and 80 km, respectively [107, 238]. The crust
has a steep geothermal gradient there. On the other hand, temperatures at depths of 25 and 40 km are
estimated to be 525 and 850◦ in the Basin and Range Province [179], indicating a lower geothermal
gradient.
Based on a plausible structure of the lithosphere, the duration of rifting can be estimated by
numerical simulations [55, 236]. For this purpose, we take the present thermal structure under
Northeast Japan and the Basin and Range Province as the references for volcanic arcs and continents.
In addition, we assume pure-shear rifting by a constant force. The evolution of a pure-shear rift is
calculated by the following one-dimensional model. The upper and lower crust is considered to
be composed of granite and gabbro, respectively, so that their ductile strengths are determined by
those of quartz and plagioclase. That of the mantle lithosphere is determined by olivine. Thermal
evolution is calculated by the one-dimensional equation
∂T ∂T ∂2 T
+v =κ 2, (12.7)
∂t ∂z ∂z
where T is the temperature, z is depth, t is the time since the beginning of rifting, v is the downward
velocity and κ is thermal diffusivity. Because of the pure-shear, v = zĖzz = −zĖxx(Eq. (3.81)).
Based on the petrological estimates, the boundary conditions T z=0 = 0◦ C and T z=a = 1300◦
12.2. STRETCHING INSTABILITY OF THE LITHOSPHERE 309

Figure 12.5: Strength profiles for model rifts [236].

C, where a is a depth of 50 km for the intra-arc rift and 74 km for the intra-continental rift. The
temperature distributions designated by the dotted lines in Fig. 12.5 are used for the initial conditions
for the rifts. The strength profiles for the strain rates of 1.432 × 10−15 s−1 and 7.954 × 10−17 s−1 ,
respectively, are shown in the same figure. The rate of the intra-continental rift is two orders of
magnitude smaller than that of the intra-arc rift, merely because the former has a smaller geothermal
gradient. Although the strength profile for the former has the maximum in the upper mantle, the
strength of the mantle part for the intra-arc rift is significantly smaller than the strength of the crust.
This is the most conspicuous difference in the strength profiles, and is due to the sensitivity of the
ductile strength of olivine to temperature compared to those of quartz and plagioclase. It is assumed
that the asthenosphere is passively uplifted under the rift by the same amount as the thinning of the
lithosphere. That is, mantle rocks are supplied into the region between the surface (z = 0) and the
depth z = a from below to compensate the thinning of the crust. We neglect the vertical movement
of the surface by rifting because the amount of movement is smaller than those of the upper-lower
crustal interface and of the Moho interface.
The thinning of the lithosphere raises isotherms to steepen the geothermal gradient. The tem-
perature changes are calculated with Eq. (12.7) for every time step in the numerical simulation.
Consequently, the transient temperature structure is designated by a convex upward curve in Fig.
12.6. If rifting proceeds very slowly, the change in the temperature structure is small. In this case, a
largely isothermal thinning of the crust increases the occupation of the olivine-rich layer in the range
between z = 0 and a. Because the layer is stronger than quartz and plagioclase if the temperature
is the same, the increased occupation augments the strength of the lithosphere that is defined by
&a
0
Δσ(z) dz. The applied force is assumed to be constant, so that the raised strength decelerates the
rifting. The slowdown of the rifting decreases the rate of temperature change to complete a negative
310 CHAPTER 12. DYNAMICS OF THE LITHOSPHERE

Figure 12.6: Schematic picture showing the temporal variation of temperature structure.

feedback cycle to stop rifting.


On the other hand, if the lithosphere thins much more rapidly than thermal conduction through
the lithosphere, the thinning is substantially adiabatic. The result is a increased geothermal gradient
which further leads to a weakening of the lithosphere. If the tectonic force is the constant through
time, the reduced strength accelerates the rifting. Accordingly, we have a positive feedback cycle in
this case. Then, the rate of lithospheric attenuation goes to infinity. We interpret this as the breakup
of the lithosphere.
Consequently, the fate of rifting, either leading to breakup or cessation, depends on the rate
of thinning relative to thermal conduction through the lithosphere [55]. The lithosphere exhibits
stretching instability that determines the fate of rifting. The rate depends on several factors including
the tectonic force, the composition of the lithosphere, geothermal gradient, and pore fluid pressure.
Figure 12.7(a) shows the strain rate versus time from the beginning of the intra-arc rift. The rate
shows divergence if the force applied to the lithosphere is greater than a critical value of about 0.396
TN−1 [236]. In this case, the duration of rifting is a few m.y. If the force is just under this value, the
rifting is aborted after several m.y. from the initiation of rifting. The duration is consistent with the
ones observed for intra-arc rifts in the world.
The critical force for intra-continental rifts is one order of magnitude greater than that for intra-
arc rifts3 . Namely, the island arc lithosphere is much weaker than the continental lithosphere, so that
inter-plate deformations are concentrated in the arcs (Fig. 12.7(b)). The critical force depends not
only on the strength of the lithosphere but also on the thermal diffusivity κ, because the transition
is determined by the speed of lithospheric thinning relative to thermal conduction through the litho-
sphere. Rising magmas carry heat toward the surface to increase the effective thermal diffusivity of
3 The critical force shown in Fig. 12.7(b) is valid specifically for the conditions assumed in the numerical simulation.

It is suggested that diffusion creep of olivine instead of dislocation creep greatly reduces the critical force by one order of
magnitude for intra-continental rifts [82].
12.2. STRETCHING INSTABILITY OF THE LITHOSPHERE 311


   

  
   



      



Figure 12.7: Unstable behavior of the model rifts [236]. Thermal diffusivity κ is assumed to be
3.5 × 10−6 m2 s−1 . (a) Temporal variation of the strain rate of the intra-arc rift for different tectonic
forces indicated in the dimensions of TN−1 . (b) The relationship between thermal diffusivity and the
critical force to separate the fate of intra-continental and intra-arc rifts.

the lithosphere4 . The numerical simulation showed that intra-arc rifts tend to be formed in volcanic
arcs. However, the critical stress is insensitive to κ for intra-arc rifts compared to intra-continental
rifts (Fig. 12.7(b)).
When the Japan Sea back-arc basin opened, the lithosphere under NE Japan was extended to form
an intra-arc rift. The rapid subsidence of the rift is designated by the increase of paleobathymetry
from ∼16 Ma to ∼15 Ma shown in Fig. 3.14. Figure 12.8 shows the temporal change in the rate
of subsidence estimated from the stratigraphic records of the rift. The rifting began at around 20
Ma. Since then, the subsidence accelerated to reach ∼3 km/m.y. The rapid subsidence was abruptly
stopped at 15 Ma when the rifting was aborted. The duration of rifting was 3–5 m.y. Assuming
the local isostasy and initial thickness of the crust to be 35 km, which is estimated from the present
crustal thickness of the Shikhote Alin (Fig. 2.12), the amount of subsidence can be transformed into
the reduced thickness of the crust. The total subsidence at around 2 km corresponds to the crustal
thickness at 25–30 km at the end of rifting5 . The total subsidence corresponds to an extensional
strain of 20%. Therefore, the average strain rate was 2 × 10−15 s−1 . This value agrees roughly with
the strain rate for the failed intra-arc rift that is estimated by the numerical simulation (Fig. 12.7).
4 Numerical simulation by Honda to estimate the thermal structure under the volcanic arc uses the effective thermal dif-

fusivity at 10−5 ms−2 , which is one order of magnitude larger than that of rocks [81]. The intra-arc rifting in Northeast
Japan was associated with massive volcanism. The representative thickness of the volcanics is about 2 km. Assuming
a heat capacity of 1.0 × 103 J kg−1 and a density of 2.5 × 103 kg m−3 , we have the heat capacity per unit volume at
(1.0 × 103 ) × (2.5 × 103 ) = 2.5 × 106 Jm−3 . In addition, assuming the temperature of magma to be 1000◦ C, the heat
transported by the volcanism was 5.0 × 1012 Jm−2 . Dividing this by the duration of volcanism (ca. 3 m.y.), we obtain the heat
flux of 50 mWm−2 . This is as great as the average heat flow from the solid Earth. A significant amount of intrusion may have
accompanied the volcanism so that the increased effective thermal diffusivity by one order of magnitude seems reasonable
for the Early Miocene rifting in Northeast Japan.
5 The present crustal thickness of Northeast Japan is greater than this thickness by 5–10 km, probably due to igneous

underplating since 15 Ma.


312 CHAPTER 12. DYNAMICS OF THE LITHOSPHERE

Figure 12.8: Temporal change in the subsidence rate of the Northeast Japan volcanic arc when the
Japan Sea back-arc basin opened [263].

In addition, the duration is consistent with the calculated one.

12.3 Factors controlling te of the continental lithosphere


It was seen in Section 8.3 that the effective elastic thickness of the oceanic lithosphere is approxi-
mated by the depth of the 400–600◦ isotherm, and that of the continental lithosphere has a positive
correlation with tectonic age and heat flow (§8.7). Figure 12.9 indicates that the latter has a much
better correlation. Those observations suggest that the effective elastic thickness of the lithosphere
is controlled by its thermal regime. Stable continents are stable because they generally have a low
average heat flow. However, the relationship between heat flow and te is complicated by the strength
profile of the continental lithosphere which is more intricate than the oceanic lithosphere. In this sec-
tion, we consider the effective elastic thickness of the continental lithosphere in terms of its strength
profile.
We assume that the continental crust is composed of a single rock type for brevity, and that the
lithosphere behaves as a elastic-perfectly-plastic body, i.e., we deal with the lithosphere that has
strength profiles as shown in Fig. 12.10. The figure shows the vertical distribution of differential
stress in the flexed lithosphere for three cases. While the curvature is small, yielding occurs only
near the top and base of the lithosphere to leave a thick elastic core in the lithosphere. However, as
the curvature increases, the differential stress generated by the bending exceeds the strength of the
lower crust so that the elastic core is divided into two layers (Fig. 12.10(b)). If the interface between
the elastic layers with thicknesses of t1 and t2 is lubricated, the whole system has the effective elastic
thickness te ≈ max(t1 , t2 ) (§8.4.2). The effective elastic thickness before the separation was t1 + t2 ,
which is larger than max(t1 , t2 ). Therefore, te may decrease abruptly at a critical curvature [28].
12.3. FACTORS CONTROLLING T E OF THE CONTINENTAL LITHOSPHERE 313

Figure 12.9: Effective elastic thickness of the continental lithosphere te . (a) te versus tectonic age
[28]. (b) te versus heat flow in Africa [72]. The error bars in this plot that are designated in the figure
in [72] are omitted here for simplicity.

Figure 12.10: Strength profiles of a model continental lithosphere and the vertical distribution of
the differential stress (gray) generated by the concave upward bending of the lithosphere, where
Δσ = σH − σv . (a) Bending with a small curvature K of the lithosphere gives rise to yielding near
the top and base of the lithosphere. There is a thick elastic core indicated by the linear part of
the stress distribution in the lithosphere. (b) As the curvature increases, the linear part reaches the
curve that designates the ductile strength of the lower crust (dashed line). Yielding in the lower crust
divides the elastic core into two parts, one in the crust and the other in the lithospheric mantle (gray).
Namely, the crust and mantle is mechanically decoupled. (c) Vertical distribution of the differential
stress caused by the bending and simultaneous horizontal compressive force F .

This phenomenon is observed in collisional orogens [140]. Figure 12.10(c) shows a case where a
horizontal tectonic force F is applied to the bent lithosphere.
Given the flexure w(x) of the lithosphere, it is straightforward to calculate the curvature K = w  .
Then, the horizontal strain εxx is obtained via Eq. (8.12). Accordingly, if ẇ(x) is also given, ε̇ xx
314 CHAPTER 12. DYNAMICS OF THE LITHOSPHERE

is obtained. The rate of strain is transformed to ductile strength. Therefore, we can calculate stress
from w(x) and ẇ(x). The bending moment is obtained from the stress as
 tL
M= σxx (z − zn ) dz. (12.8)
0

Then, using the formula M = −DK, the flexural rigidity D and the effective elastic thickness te of
the flexed lithosphere can be estimated.
There may be depth ranges of yielding for given w(x) and ẇ(x) (Fig. 12.10(b)). In that case,
Eq. (12.8) is replaced by the piecewise integrals
N 
  
(i)
M= σxx z − zn dz, (12.9)
i=1 ith layer

(i)
where zn is the depth of the ith layer in which deformation is accommodated by a brittle ductile or
elastic mechanism. Using the bending moment in Eq. (12.9), the effective elastic thickness of this
rheologically multi-layered lithosphere is estimated. It should be noted that te is different from the
thickness of the elastic core.
If the horizontal force  tL
F = σxx dz (12.10)
0
does not vanish, the horizontal stress σxx in Eq. (12.9) should be replaced by (σxx − Fmean ), where
Fmean = F/tL is the mean horizontal stress in the lithosphere. Figure 12.11 shows (σxx − Fmean )
for regions of compressive and extensional tectonics. Since the brittle strength for the former region
is larger than that for the latter if pore fluid pressure is the same, te of the horizontally stretched
lithosphere is smaller than that of the horizontally constricted lithosphere [141]. Whole lithosphere
failure occurs more easily in extensional than in compressive tectonics.
Figure 12.12 shows the effect of the curvature, geothermal gradient and tectonic force Fmean on
te for a model continental lithosphere. In addition, te is also affected by curvature, even if other
factors are the same. Increased curvature increases differential stress to widen the yielded zones,
which reduces te . Figure 12.12(a) shows that sensitivity on the geothermal gradient decreases with
increasing curvature. The radius of curvature at the right side of this graph is R = 1/K = 100 km,
comparable with the thickness of the continental lithosphere. Thus, the model of a thin elastic plate
has a large error at this end. Figure 12.12(b) shows te versus Fmean . The effective elastic thickness
declines more rapidly in a normal fault regime than in a reverse fault regime with increasing tectonic
force |Fmean |.
The continental lithosphere has complex structures, leading to an intricate strength profile. Thus,
the effective elastic thickness of the lithosphere depends on various factors including lithology, cur-
vature, thermal structure, horizontal tectonic force, radioactive heat generation in the crust, etc.
[140]. The Canadian shield has a considerable spatial variation in te , although those factors have not
so large differences. The variation suggests that geological structures such as old faults and suture
zones play important roles in determining te [27, 254].
12.4. PERIODIC DEFORMATION OF THE LITHOSPHERE 315

Figure 12.11: Schematic strength profiles of the oceanic lithosphere and vertical distribution of dif-
ferential stresses in regions of (a) compressive and (b) extensional tectonics. Due to the asymmetric
brittle strength for those cases, the effective elastic thickness for the former case is larger than for
the latter. The width of the gray area indicates Δσ − Fmean .

Figure 12.12: Effective elastic thickness te of a model continental lithosphere that has a dry granitic
crust [126]. The strain rate is assumed to be 10−15 s−1 . (a) The effect of curvature K and geothermal
gradient. (b) Effect of horizontal stress Fmean with an assumed curvature K = 10−5 km−1 .

12.4 Periodic deformation of the lithosphere


Faults and folds are sometimes aligned parallel and uniform intervals in wide deformation zones.
Horst and grabens in the Basin and Range Province are examples. The wide rift zone called sulcus
on Jovian satellite Ganymede have grooves that are surface manifestations of normal faults (Fig.
3.6). Many icy satellites have such zones. Figure 12.13 shows a rift zone separating the heavily
cratered, therefore older, terrains on Enceradus.
The Fletcher-Hallet model explains those spatially periodic structures using the linear stability
analysis of a multi-layered system with a free boundary [63]. The model is also applied to the
periodic surface undulations of ropy lavas [62].
316 CHAPTER 12. DYNAMICS OF THE LITHOSPHERE

Figure 12.13: Saturnian satellite Enceradus (Voyager image 1715S2-001). The diameter of this
moon is about 500 km. Note the difference in the number density of craters. Smooth planes with a
smaller crater density have grooves. The heavily cratered terrain is divided by a rift zone (arrow).

Consider the crust of an icy satellite instead of the Earth’s crust, because the latter has complex
structures. For example, Ganymede has a very thick icy layer (§9.3), so that the shallow part of the
layer may be brittle as grooves and other geological structures indicate, but deep-seated ice may be
ductile. Accordingly, we assume a surface brittle layer floating on a viscous fluid (Fig. 12.14(a)).
The brittle behavior of the layer (1) is simulated by a power-law fluid with a very large power-law
exponent n(1)  1. The layer (2) is also a power-law fluid but has the exponent n(2) ≈ 3. Plane
strains on the xz-plane across the wide deformation zone are assumed, where the z-axis is defined
upward with the origin O at the base of the initial brittle layer (Fig. 12.14). H is the initial thickness
of the layer (1). All the layers are assumed to be incompressible.
Following Section 9.4, horizontally extending pure shear is regarded as the mean flow for this
12.4. PERIODIC DEFORMATION OF THE LITHOSPHERE 317

Figure 12.14: Schematic illustration showing the Fletcher-Hallet model. The layer (1) is a brittle
layer with a free boundary at its top and is underlain by the viscous layer (2). The layer (0) is
composed of inviscid fluid or space. The interface between the layers (1) and (2) is the brittle-
ductile transition. Dashed lines indicate the initial levels of the surface and brittle-ductile transition.

multi-layered system, and we have assumed incompressibility. Therefore, the results in Section 10.8
are applicable to this problem. Namely, the differential equation in Eq. (10.89) holds for each of
layers (1) and (2), so that the growth rates of the coefficients of their separable solutions gives the
stability of this system.
In this model, infinitesimal strains are assumed so that we can ignore temperature perturbations,
i.e., temperature is a function of z in layer (2):

T = T0 − Γz (z < 0), (12.11)

where Γ is the geothermal gradient. Since the perturbation is very small, the effective viscosity in
layer (2) may be approximated by
η (2) = η0 exp(−γz). (12.12)
The effective viscosity of layer (1) is assumed to be constant.
In the layers (1) and (2), we obtain the stress ratio Φ = 1/2 using the same argument as we did
to derive the slip line theory in Section 10.5. Substituting this stress ratio into Eq. (4.15), we have
TII = (ΔS/2)2 and the Mohr circles in Fig. 12.15. Therefore, we have
1
TII = (Sxx − Szz )2 + Sxz
2
. (12.13)
4
The perturbations of the pure shear parallel to the coordinate axes are very small, therefore, (Sxx −
Szz )2  Sxz
2
(Fig. 12.15). The effective viscosity of the fluid η satisfies
 
(n−1)/2 −1
2η = BTII (12.14)
B = B ∗ exp(−Q/RT ) . (12.15)

Layer (2) has the constitutive equation

Sxx = −p + 2ηDxx , Szz = −p + 2ηDzz , Sxz = 2ηDxz . (12.16)


318 CHAPTER 12. DYNAMICS OF THE LITHOSPHERE

Combining the incompressibility Dxx + Dzz = 0 and Eq. (12.16), we obtain

S xx − S zz = 4ηDxx . (12.17)

The principal axes of the mean flow are parallel to the coordinate axes. Hence, we have S xz = 0.
Combining Eq. (12.15), we obtain
1 2 2 1 
T II = S xx − S zz = S , S≡ S xx − S zz > 0. (12.18)
4 2
Thus, the effective viscosity satisfies
−1 (1−n) − 1n ( 1 −1)
2η = B S =B Dxxn . (12.19)

Substituting Eq. (12.15) into (12.19), and using the temperature in Eq. (12.11), we have
   
 ∗  1n (1− 1n ) −1 Q
η= 2 B Dxx exp .
nR(T0 − Γz)
The operand of the exponential function is expanded about z = 0 as
Q Q ΓQz
≈ + .
nR(T0 − Γz) nRT0 nRT 2
0

Therefore, using Eq. (12.12), we obtain


   Q 
 ∗  1n (1− 1n ) −1
η0 = 2 B Dxx exp , (12.20)
nRT0
ΓQ
γ= . (12.21)
nRT02

These equations give the effective viscosity of layer (2). That of layer (1) is assumed to be constant at
η0 , and the state of stress in this layer is determined by Eq. (12.17) and on the Mises yield criterion.
For a linear analysis of this system, we have to solve Eq. (10.89) under appropriate boundary
conditions. Unlike single layer folding, the boundary between layers (0) and (1) is free. The local
rectangular coordinates are defined at the boundary to which the s- and n-axes are taken to be parallel
and perpendicular (Fig. 12.16). The function h(x) denotes the height of the boundary. Then, the
components of S satisfy the conditions at the boundary
 
Ssn z=h = 0, Snn z=H = ρgh, (12.22)

where the normal stress is linearized as Eq. (D.31). These equations are transformed into the
expressions with the coordinates x and z by considering the force balance at the boundary. Let N
be the unit normal to the boundary, then the force balance is expressed as

S(0) · N = S(1) · (−N ). (12.23)


12.4. PERIODIC DEFORMATION OF THE LITHOSPHERE 319

Figure 12.15: Mohr circles for the stress state with Φ = 1/2.

Let θ be the inclination of the s-axis with respect to the x-axis (Fig. 12.16). The stress components
are transformed as
     (i) 
(i) (i) (i)  
Sss Ssn cos θ sin θ Sxx Sxz cos θ − sin θ
= .
(i)
Sns Snn
(i)
− sin θ cos θ (i)
Szx Szz
(i)
sin θ cos θ

Therefore,
(i) (i) (i)
Snn = Sxx sin2 θ − 2Sxz sin θ cos θ + Szz cos2 θ, (12.24)
 
(i) (i) (i)  
Sns = − Sxx − Szz sin θ cos θ + Sxz cos2 θ − sin2 θ . (12.25)

These trigonometric functions are rewritten as sin θ = ∂$


h/∂s and cos θ = ∂x/∂s. Ignoring higher-
order terms, we have
∂$
h
sin θ ≈ , cos θ ≈ 1.
∂x
Substituting these equations into Eqs. (12.24) and (12.25), and again ignoring the higher-order
terms, we obtain
$   $
(i) ∂h (i) (i) (i) ∂h (i)
Snn = −2Sxz + Szz , Sns = − Sxx − Szz + Sxz . (12.26)
∂x ∂x
$ $
The stress tensor is divided
  the mean and perturbation, S = S + S. Because of S ≈ O, the
into
higher-order terms of ∂$ h/∂x S can be neglected so that Eq. (12.26) is transformed into

(i) ∂$
h (i)
+ S zz + S$zz ,
(i)
Snn = −2S xz (12.27)
∂x
 (i)  h
(i) ∂$ (i)
+ S xz + S$xz .
(i)
Sns = − S xx − S zz (12.28)
∂x
320 CHAPTER 12. DYNAMICS OF THE LITHOSPHERE

Figure 12.16: Schematic picture showing the boundary condition between layers (1) and (2).

On the other hand, a non-slip condition and force balance are appropriate bondary conditions at
the boundary between layers (0) and (1):
   
(1)  (2)  (1)  (2) 
$
vx  = $ vx  , $
vz  = $ vz 
z=0 z=0 z=0 z=0
   
S$zz  = S$zz  , S$xz  = S$xz 
(1) (2) (1) (2)
.
z=0 z=0 z=0 z=0

Note that the layers are of the same density. If they are not, buoyancy force should be taken into
account.
Under these boundary conditions, the differential equation in Eq. (10.89) is solved. We use a
solution of the form
1 dW
$
vx = − sin kx, $
vz = W cos kx, h − H = A(t) cos kx,
λ dz
where k is the horizontal wavenumber. Consequently, we obtain

dA 
= −Dxx A + $
vz  = (q − 1)Dxx A, (12.29)
dt z=H

where q(k) denotes the growth rate6 . The waves with q(k) > 1 are amplified, but those with q(k) < 1
are erased. The waves with the maximum growth rate may emerge as macroscale boudins. Accord-
ing to Fletcher and Hallet [63], q(k) is determined by the four dimensionless parameters, n(1) , n(2) ,
γH and ρgH/2cY . The third one is the characteristic depth over which the effective viscosity of
the layer (2) decreases by 1/e. The fourth one designates the strength of layer (1). Figure 12.17(a)
shows a graph of q(k), where kd indicates the wavenumber that maximizes the growth rate. It
is demonstrated that kd is insensitive to n(1) if this power-law exponent is greater than 102 (Fig.
12.17(b)). This is convenient to simulate the brittle behavior of layer (1). Rocks and ice have a
value of n(2) about 3. Therefore, the remaining parameters γH and ρgH/2cY determine the kd . The
6 See [62] for the concrete expression of q(k).
12.4. PERIODIC DEFORMATION OF THE LITHOSPHERE 321

Figure 12.17: Growth rate of crustal boudins predicted by the Fletcher-Hallet model [63]. (a) Growth
rate versus dimensionless wavenumber. The values n(1) =104 , n(2) =3, γH=10 and ρgH/2cY =3 are
assumed. (b) Relationship between the dominant wavenumber kd and the power-law exponent n(1) .
The parameter values n(2) =5, γH=10 and ρgH/2cY =3 are assumed here.

brittle-ductile boundary goes down with an increasing strain rate. Therefore, H depends on Dxx
(Eq. (12.20)). Likewise, the critical stress cY depends on the strain rate. On the other hand, γ is
determined by Γ via Eq. (12.21). Consequently, Dxx and Γ control the dominant wavenumber kd .
Many icy satellites have grooves with roughly uniform intervals. Thus, the dominant wavelength
is observable there. However, we need one more line of evidence to constrain Dxx and Γ in the
Fletcher-Hallet model. Let qd be the growth rate at the dominant wavelength kd . Then, from Eq.
(12.29), we obtain
ln(A/A0 ) = (qd − 1)Dxx t, (12.30)
where A0 and A are the initial amplitude and that of the time t. The left-hand side of Eq. (12.30)
represents the logarithmic strain of the crust.
This equation enables an order-of-magnitude estimation of the controlling factors. The initial
amplitude is unknown, but A0 ∼ 100 m may be acceptable. A is known from the present topography.
For example, grooves on icy satellites have amplitudes of the order of A ∼ 102 m. The global strain
of the moons is in the order of 1%. Thus, we estimate Dxx t ∼ 10−2 . Using these estimates, we
obtain qd ∼ 500 from Eq. (12.30). The Fletcher-Hallet model gives appropriate values of Dxx and
Γ from this growth rate at the observed kd . If the values are unrealistic, the model is found to be
inapplicable to the object. It was found that the model is applicable to Ganymede, Enceradus, and
Miranda [44, 76]. As for Ganymede, the values Γ ≈ 20 K km−1 and Dxx ≈ 10−14 s−1 were obtained.
Similar values were determined for Enceradus and Miranda. Thus, the extensional tectonics lasted
1%/10−14 s−1 ∼ 104 –105 years, surprisingly short for the global tectonics compared to the present
tectonics of the Earth.
The lithosphere of the Earth has a much more complex structure than that assumed in the
322 CHAPTER 12. DYNAMICS OF THE LITHOSPHERE

Fletcher-Hallet model. The density difference of the crust and mantle gives rise to buoyancy. The
strength of the brittle layer depends on depth, unlike the strength in the model. In some active
regions, the hot and fluent lower crust mechanically separates the upper crust and the mantle litho-
sphere, so that there are two strong layers there. If the upper and lower crusts have significantly
different rheology, such as granite and gabbro, the lithosphere may exhibit complex behavior. Mod-
els have been developed to take account of those factors [16, 17, 196, 282]. Such a model was
successfully applied to the folded oceanic lithosphere in the northeastern Indian Ocean [281], where
the simple elastic model failed to explain the dominant wavelength (§8.4.3). Linear stability analy-
sis assumes very small perturbations. Numerical simulations are used to link the analysis and finite
amplitude deformations as the rifts discussed in Section 12.2 [25].

Exercises
12.1 In the model introduced in §12.2, the fate of rifts is determined by the competition between
the rate of lithospheric thinning and vertical heat transfer through it. Consider how to define the
Péclet number, which is a dimensionless number indicating the representative time of motion of a
mass relative to the time constant of heat conduction.

12.2 Consider that pure shear rifting begins in the lithosphere with a negligible strength of the
crust compared to that of the mantle lithosphere, and that the geothermal gradient Γ is constant in
the lithosphere. Let  
− 1n 1
−1 Q
T = A Ė n exp Ė (12.31)
nRT
be the constitutive equation of the mantle, where
⎛ ⎞
Ė 0 0
Ė = ⎝ 0 0 0 ⎠ (12.32)
0 0 −Ė

is the rate of strain tensor. Show that the approximate equation


a  
A− n Ė 1/n TM2 nR
1
Q
Txx dz ≈ exp
0 ΓQ nRTM

indicates the tectonic force needed for the rifting [55], where Txx is the horizontal component of the
deviatoric stress, TM is the Moho temperature.
Appendix A

List of Symbols

0 zero vector
A Almansi’s finite strain tensor
B left Cauchy-Green tensor
B buoyancy coefficient (= Δρg)
C right Cauchy-Green tensor
C degree of compensation
D stretching tensor
D flexural rigidity
dc compensation depth
D/Dt material derivative
E infinitesimal strain tensor
Ė strain rate tensor
ĖE , Ėe equivalent strain rate
e unit vector
e(i) ith base vector
F deformation gradiant tensor
F (S) yield function
G Green’s finite strain tensor
G universal constant of gravitation
g accelaration due to gravity
I unit matrix, unit tensor
J Jacobian (= |F|)
K curvature
L velocity gradient tensor
Mg geometric seismic moment
M moment, torque

323
324 APPENDIX A. LIST OF SYMBOLS

Ma asymmetric moment tensor


N orthogonal projector onto the line parallel to the unit vector n
N unit normal vector
n unit normal vector
n power-law index
O origin of coordinates
O zero matrix, zero tensor
Pe Péclet number
pL lithospheric pressure
pf pore pressure
P (e) projector onto a line parallel to a unit vector e
P⊥ (e) projector onto a plane normal to a unit vector e
Q orthogonal tensor
Q activation energy
q heat flux
R orthogonal tensor
R radius, radius of curvature
Ra Rayleigh number
S stress tensor
S0 means stress
S1 , S2 , S3 maximum, intermediate and minimum principal stresses
Si ith principal stress
SNoct octahedral normal stress
SSoct octahedral shear stress
T deviatoric stress tensor
T temperature
AT transpose of tensor A
t time, thickness
Ta athenospheric temperature
tc crustal thickness
TE , Te equivalent deviatoric stress
te equivalent elastic thickness
trace A trace of tensor A
tL lithospheric thickness
t(n), t(N ) stress vector
U right stretch tensor
u displacement vector
V left stretch tensor
V volume
v velocity
325

W spin tensor
w angular velocity
w displacement due to flexure
X body force
x position vector
x1 , x2 , x3 spatial coordinates
X maximum principal strain
x horizontal coordinate
Y Young’s modulus, intermediate principal strain
y horizontal coordinate
Z minimum principal strain
z vertical coordinate
α volumetric thermal expansion coefficient
αl linear thermal expansion coefficient
β stretching factor of the lithosphere
γ engineering shear strain
Γ geothermal gradient
δF deviation of F from I
δij Kronecker delta
ΔS, Δσ differential stress
δ(x) delta function
e infinitesimal strain tensor
ė strain rate tensor
ijk permutation symbol
η viscosity, effective viscosity
Θ Lode angle
θ the angle of shear
κ thermal diffusivity
λ eigenvalue, longitudinal modulus, wavelength
λf pore fluid pressure ratio, flexural parameter
λw flexural wavelength
μ coefficient of internal friction
μf coefficient of friction
ν Poisson’s ratio
ξ material expression of position vector
ξ1 , ξ2 , ξ3 material coordinates
ρ density
ρc density of the crust or of the core
ρm density of the mantle
326 APPENDIX A. LIST OF SYMBOLS

ρs density of sediment
ρw density of water
r stress tensor
σ0 mean stress
σ1 , σ2 , σ3 maximum, intermediate and minimum principal stresses
σH , σHmax maximum horizontal stress
σh , σHmin minimum horizontal stress
σN normal stress
σS shear stress
σNoct octahedral normal stress
σSoct octahedral shear stress
σV vertical stress
σY yield stress, yield strength
s deviatoric stress tensor
Φ stress ratio
φ internal friction angle
ψ stream function
X rotation tensor
ω vorticity
 a, b, c  triple vector procuct
∇ nabla operator
( )· material derivative
Appendix B

Stress Inversion

This appendix supplements Chapter 11, and introduces a new formulation of the stress inversion, a
method for determining the stress from fault-slip data or from seismic focal mechanisms [203, 269].
Reduced stress tensor S is defined in the formulation to be deviatoric, and its first and second
basic invariants (Eq. (C.51)) have the prescribed values,
1 2 
SI = S11 + S22 + S33 = 0, SII = S + S22
2
+ S33
2
+ S23
2
+ S31
2
+ S12
2
= 1. (B.1)
2 11
Given an orthogonal
 tensor R for principal orientations and stress ratio Φ, the tensor has the form,
S = R ·S0 / 3Φ − 3Φ + 3·R, where S0 is the diagonal matrix with the three diagonal elements 2−
T 2
 √ √ √
Φ, 2Φ−1 and −Φ−1. Obviously, the 6-dimensional vector y = S11 / 2, S22 / 2, S33 / 2, S23 , S31 ,
T
S12 has a one-to-one correspondence with S, allowing us to identify y and S. Then, Eq. (B.1) is
rewritten as
(1, 1, 1, 0, 0, 0) · y = 0, |y|2 = 1. (B.2)
These equations indicate two figures in a 6-dimensional Euclidean space: a hyperplane through the
origin and the unit sphere, respectively. Endpoint of y exists in the intersection of the figures, i.e., the
surface of the unit sphere in a 5-dimensional space, which is known as Ilyushin’s deviatoric stress
space. Consequently, a 5-dimensional unit vector x can be identified with y and S.
Faulting gives rise to a strain of the surrounding rock mass E = γ(bn + nb)/2 (§2.7), where
the symbols in the right-hand side of this equation are illustrated in Fig. 2.13. A fault-slip datum
includes b and n. The faulting is assumed to be volume constant, EI = 0. The stress inversion does
not take absolute value of the shear strain γ into consideration. Accordingly, we assume the second
basic invariant of E to satisfy EII = 1. The two conditions of the invariants allow us to substitute a
√ √ √ 
6-dimensional unit vector e = − 2b1 n1 , 2b2 n2 , 2b3 n3 , b2 n3 + b3 n2 , b3 n1 + b1 n3 , b1 n2 + b2 n1 T
for E, and this vector satisfies the conditions similar to those of y in Eq. (B.2). It follows that a
5-dimensional unit vector  can be identified with e and E. Energy dissipation associated with the
faulting E : S is positive in sign. That is, e · y > 0. Note that x and y represent an identical unit
vector in the 5- and 6-dimensional spaces, respectively, and the former is included in the latter space.

327
328 APPENDIX B. STRESS INVERSION

Similary, e and  are different expressions of an identical unit vector. Accordingly, the inequality
 · x > 0 is equivalent with e · y > 0.
Finally, we define the strain tensor E = γ  (an + na)/2 with the conditions EI = 0 and EII = 1,
√ √ √
where a = n × b. Then, the 6-dimensional unit vector e = − 2a1 n1 , 2a2 n2 , 2a3 n3 , a2 n3 +
T
a3 n2 , a3 n1 + a1 n3 , a1 n2 + a2 n1 satisfies the conditions similar to those of y in Eq. (B.2), allowing
us to identify a 5-dimensional unit vector   with e and E . The vector a stands for the intersection
of the fault and auxiliary planes, and E indicates the shear strain in this orientation. The Wallace-
Bott hypothesis says that the fault does not move in the orientation at all. Hence, energy dissipation
vanishes, E : S = 0. It follows that e · y =   · x = 0. Eq. (11.14) is equivalent with e · y = 0.
The unit vectors a, b and n represent a fault-slip datum, the information of which is carried by
the vectors  and   . Elements of the vectors x,  and   are given in terms of the elements of S, a, b
and n in [203]. The unknown vector x that is compatible with the datum satisfies the conditions
|x| = 1,   · x = 0,  · x > 0, (B.3)
where || = |  | = 1. It is easy to show that  ·   = 0 from the orthonormality of a, b and n. That
is,  and   meet at a right angle. The three conditions in Eq. (B.3) represent the unit sphere, the
hyperplane normal to   and the hemisphere centered by , respectively. The endpoint of x exists,
therefore, on a great-circle arc with a length of 180◦ in the hemisphere. Pole and center of the arc are
denoted by   and   , respectively. We call it a Fry arc. Given a set of fault-slip data, intersections
of their corresponding Fry arcs make a cluster on the sphere, and the cluster center represents the
optimal stress for the data [268]. This geometrical interpretation of the stress inversion is not affected
by coordinate rotations in the physical space, because our stress and strain tensors are normalized by
their invariants, i.e., SI = EI = EI = 0 and SII = EII = EII = 1.

If the vectors x(1) and x(2) that denote reduced stress tensors are given, Ψ = cos−1 x(1) · x(2)
equals the great-circle distance between their endpoints on the unit sphere. This is a useful measure
of dissimilarity between reduced stress tensors, and is termed angular stress distance [269]. Ψ is
invariant under coordinate rotations in the physical space, and has the range of 0 ≤ Ψ ≤ 180◦ . The
value of Ψ equals the Lode angle difference of the stress tensors ΔΘ, provided that the tensors have
common principal orientations. In this case, rearranging Eq. (10.19), we have the expression of ΔΘ
in terms of the stress ratios Φ(1) and Φ(2) of the tensors corresponding to x(1) and x(2) :
 " >  (1) # " >  (2) #
 
ΔΘ = cot−1 2 − Φ(1) 3Φ − cot−1 2 − Φ(2) 3Φ .
Any combination of stresses with common principal orientations does not have a Ψ value over 60◦ ,
and vice versa. Ψ is related with stress difference (Eq. (11.19)) as D = 2 sin(Ψ/2).
If we have obtained n solutions, denoted by x(1) , . . . , x(n) , from a homogeneous data set by
the multiple inverse method, bootstrap or other resampling techniqe, the vectors make a cluster,
the center of which denotes the optimal solution. The uncertainty of the optimal solution can be
%
evaluated by Ψ = 1n ni=1 Ψ(i) , where Ψ(i) is the angular stress distance between the cluster center
and the endpoint of x(i) . It is known that Ψ is approximately equal to the noise level of the data,
where the level is represented by the standard error of slip directions on the fault planes [269].
Appendix C

Basic Equations

This appendix introduces mathematical tools that are needed to understand the topics in
this book.

C.1 Vectors
A vector has both magnitude and direction. Vectors are used to represent directional data such as
force and velocity. Such a physical entity does not change by the choice of coordinates if we choose,
for example, the x-axis in the north or east direction. However, vector components depend on the
choice. Consider that e(1) , e(2) , and e(3) are the unit vectors parallel to the coordinate axes. If a
vector a can be written as
a = a1 e(1) + a2 e(2) + a3 e(3) , (C.1)

the coefficients a1 , a2 , and a3 are the components of the vector. The direction of the unit vectors
depend on the choice of coordinates so that the components are also affected by the choice, although
the entity a is constant.

Scalar product

There are in general two types of products between two vectors. They are scalar and vector products.
Let a = (a1 , a2 , a3 )T and b = (b1 , b2 , b3 )T be artibrary vectors, then a · b ≡ |a| |b| cos θ is their
scalar product, where θ is the angle between them.
If we write the angle between e(i) an e(j) as
 
θij = e (i) , e(j) , (C.2)

the axes of Cartesian coordinates cross at right angles, so that θij = 90◦ and 0◦ for i = j and i = j,
respectively. Therefore,
e(i) · e(j) = δij , (C.3)

329
330 APPENDIX C. BASIC EQUATIONS

where δij is the Kronecker delta. Using this relationship and the expansion in Eq. (C.1), we obtain

a · b = a1 b1 + a2 b2 + a3 b3 .

This equation contains components in the right-hand side, and the scalar product does not depend on
the orientation of the coordinates because the product is also written only with the physical quantities
|a|, |b|, and θ.
Orthogonal projection of a vector onto a line is often used. Let a and e be an arbitrary vector and
the unit vector parallel to the line (Fig. C.1). The component a parallel to e is equal to their scalar
product.

Vector product

The other type of product between the vectors a and b is defined by the equation
⎛ ⎞
a2 b3 − a3 b2
a × b ≡ ⎝a3 b1 − a1 b3 ⎠ . (C.4)
a1 b2 − a2 b1

The result is a vector perpendicular to the plane defined by the vectors a and b, and |a × b| =
|a| |b| sin θ, where θ is the angle between the vectors. The absolute value is equal to the area of the
paralellogram that is spanned by the vectors. The direction of (a × b) may be remembered by the
right-hand rule: a and b are represented by the first and second fingers of the right hand. The thumb
then points in the direction of (a × b). The vector product can also be written as
 
  
a×b= ijk aj bk e(i) = ijk aj bk e(i) . (C.5)
i j,k i,j,k

The coefficient ijk is the permutation symbol with the value ±1 or zero as

123 = 231 = 312 = 1, 213 = 321 = 132 = −1,


ijk = 0 (othewise).

The symbol satisfies the identity [134]:


 
ijk kmn = δim δjn − δin δjm or ijk mnk = δim δjn − δin δjm . (C.6)
k k

Triple products

Consider a parallelepiped with the sides defined by three vectors a, b, and c. The magnitude of b × c
is identical to the area of one face parallel to the two vectors, and the distance between the face and
the other side of the parallelopiped is equal to a · e where e is the unit vector parallel to the vector
C.2. DUMY AND FREE INDICES 331

Figure C.1: The scalar product between an arbitrary vector a and a unit vector e is equal to the length
of the shadow of a projected onto the line parallel to e.

product: e = (b × c)/|b × c|. Therefore, the triple scalar product a · (b × c) is equal to the volume
of the solid figure. On the other hand, we call a × (b × c) a triple vector product.
Let us write the triple scalar product as a, b, c ≡ a · (b × c), then
 
  a1 a2 a3 
 
a, b, c = ijk ai bj ck =  
 b1 b2 b3  .
i,j,k  c1 c2 c3 

Accordingly, the triple scalar product is equivalent to the determinant of the matrix that is composed
by the vector components. The symbol ijk indicate that the sign of product depends on the order of
the three vectors:

a, b, c = b, c, a = c, a, b = − b, a, c = − c, b, a = − a, c, b .

C.2 Dumy and free indices


There are two types in indices. Let us use a × b + c as an example. We have the ith component of
the terms as

ijk aj bk + ci .
j,k

Since the symbols k and j in the first term are placed only for the summation, they can be replaced
by any other symbol:
 
ijk aj bk + ci = ipq ap bq + ci .
j,k p,q

The replacement does not affect terms out of the summation. In this case, ci is not affected. Such a
index is called a dummy index. By contrast, the replacement of i in the first term affects the second
one. Such an index is called a free index. Replacement of dummy indices sometimes helps to
simplify equations1 .
1 An example is shown in the process to derive Eq. (11.11).
332 APPENDIX C. BASIC EQUATIONS

Figure C.2: (a) Rotation of coordinates with fixed vector a. (b) Rotation of the vector with fixed
coordinates.

C.3 Rotation of coordinates


Vector components depend on the orientation of coordinate axes. Let us consider the dependence.
Let O-12 and O-1 2 be two Cartesian coordinates on a plane with the identical origin, and the
components of the vector a are ai and ai for the coordinates, respectively. If the vector is fixed and
the O-1 axis is rotated counterclockwise by the angle θ, the transformation between the components
is written by the famous formula
    
a1 cos θ sin θ a1
= . (C.7)
a2 − sin θ cos θ a2

We write the coefficent matrix as


 
cos θ sin θ
Q= . (C.8)
− sin θ cos θ

Now consider the same amount of rotation. However, the coordinates are fixed but the vector a is
rotated counterclockwise. This is the opposite situation to the above discussion, but the components
of the vector are also changed by this procedure. Their change is described by Eq. (C.7) with the
opposite sense of rotation. Namely, the matrix
 
cos θ − sin θ
R= (C.9)
sin θ cos θ

describes the rotation of the vector.


It is important to distinguish whether coordinates or vectors are rotated. We distinguish them by
the symbols Q and R. They are related by the equations

Q = R−1 , R = Q−1 . (C.10)

The equation
a = Q · a (C.11)
C.4. MATRICES 333

represents, accordingly, the rotation of coordinates. The physical entities of a and a are identical in
this case. The equation a = Q · a On the other hand, the equation
a = R · a (C.12)
indicates the rotation of the entity.
The angles between the axes are written as in Eq. (C.2). For the two-dimensional case (Fig.
C.2(a)), we have
θ11 = θ, θ12 = π/2 − θ,
θ21 = π/2 + θ, θ22 = θ.
Using the formula cos(π/2 ± θ) = ∓ sin θ, we obtain
   
cos θ11 cos θ12 cos θ sin θ
= = Q, (C.13)
cos θ21 cos θ22 − sin θ cos θ
   
∴ Q = cos θij = e (i) · e(j) . (C.14)

This relationship also holds for rotations in three-dimensional space. Q and R are called orthogonal
matrices which satisfiy the following formulas:
Q−1 = QT , Q · QT = I,
R−1 = RT , R · RT = I. (C.15)
Combining Eqs. (C.10), (C.13) and (C.15), we see that R has the form
⎛ ⎞
cos θ − sin θ 0
R = ⎝ sin θ cos θ 0⎠ . (C.16)
0 0 1

C.4 Matrices
Many physical quantities are represented by matrices in the models of tectonics. Let A and B be
arbitrary square matrices, then their products are defined by the equations:
   
 
A·B= Aik Bkj , A : B = Aij Bij .
k i,j

The following formulas of matrices are used in this book:


A · I = A, (C.17)
|A · B| = |A||B|, (C.18)
 
A · B T = BT · AT , (C.19)
|A | = |A|,
T
(C.20)
  −1  −1 
AT = A T
, (C.21)
 −1
A·B = B−1 · A−1 . (C.22)
334 APPENDIX C. BASIC EQUATIONS

Symmetry and antisymmetry

Consider a square matrix with the shape


⎛ ⎞
a b c
A = ⎝d e f⎠ .
g h i

This can be divided into two parts: A = B + C, where


⎛ ⎞
2a b+d c+g
A + AT 1⎝
B= = d+b 2e f + h⎠ ,
2 2
g+c h+f 2i
⎛ ⎞
0 b−d c−g
A − AT 1
C= = ⎝d − b 0 f − h⎠ .
2 2
g−c h−f 0
As B = BT , B is a symmetric matrix. On the other hand, C = −CT . This is called an antisymmetric
matrix. Every square matrix can be decomposed into symmetric and antisymmetric matrices.

C.5 Eigenvalues and eigenvectors


If A is a square matrix and u is a vector, the dot product A · u is a vector which is parallel to u for
special u. Those are the eigenvectors of the matrix. If they are parallel, λ refers to the ratio between
u and A · u, and the ratio is called an eigenvalue. Linear algebra says that an n × n matrix has n
pairs of eigenvectors and eigenvalues, and the latter is obtained by solving the following algebraic
equation, called a characteristic equation,

|A − λI| = 0. (C.23)

The matrix has real eigenvalues if the matrix components are real and A is symmetric. Then all the
eigenvectors are perpendicular to each other. The above equation does not explicitly contain compo-
nents that depend on the choice of coordinates, so that neither the eigenvalues nor the eigenvectors
depend on the choice. Specifically, the characteristic equation of the 2 × 2 matrix
 
a c
A= (C.24)
c b

is λ2 − (a + b)λ + ab − c2 = 0, so that

(a + b) +
(a − b)2 + 4c2
λ1 = ,
2
(a + b) − (a − b)2 + 4c2
λ2 =
2
C.5. EIGENVALUES AND EIGENVECTORS 335

are the eigenvalues, and the eigenvectors are


  
−a + b + (a − b)2 + 4c2
− , 1 T
,
2c
  
−a + b − (a − b)2 + 4c2
− , 1 T
.
2c

How about the eigenvalues and eigenvectors of A = A − pI, where p is an arbitrary scalar
parameter? Let λ be the eigenvalues of A , then we have
  
A − λ I = |A − (p + λ)I|.

Therefore, A and A have the same eigenvectors, but their eigenvalues differ by the constant p.
Let λ1 , λ2 , and λ3 be the real eigenvectors of the 3 × 3 symmetric matrix A, and u(1) , u(2) , and u(3)
are the unit vectors parallel to the eigenvectors, then the unit vectors make right angles to each other.
The unit vectors can be used as base vectors of the new orthogonal coordinates. Their components
make up an orthogonal matrix
⎛ (1) (1) (1) ⎞ ⎛ (1) T ⎞
u1 u2 u3 u
Q = ⎝u1 u3 ⎠ = ⎝ u(2) T ⎠ ,
(2) (2) (2)
u2
u
(3)
u
(3)
u
(3) u(3) T
1 2 3

(j)
where ui is the ith component of the jth vector that is evaluated in the old coordinates. Then the
coordinate transformation from the old to new is represented by the matrix Q (see Exercise A1).
Using this, we can reshape the matrix A to a diagonal matrix
⎛ ⎞
λ1 0 0
Q · A · Q T = ⎝ 0 λ2 0 ⎠ , (C.25)
0 0 λ3

in which the diagonal components are identical to the eigenvalues of A. If we use the new coordi-
nates, the matrix becomes the diagonal form. In this case, the power of A is
⎛ n ⎞
λ1 0 0
An = ⎝ 0 λn2 0 ⎠ . (C.26)
0 0 λn3

Accordingly, the diagonal form holds for any n, indicating that eigenvectors do not change.
If f (A) is a 3×3 matrix and is a function of A, and if the two matrices have the same eigenvectors,
the nth term of Taylor expansion of this function encompasses An with the above shape. Therefore,
⎛ ⎞
f (λ1 ) 0 0
f (A) = ⎝ 0 f (λ2 ) 0 ⎠. (C.27)
0 0 f (λ3 )
336 APPENDIX C. BASIC EQUATIONS

Given positive eigenvalues, we have


⎛ ⎞
log λ1 0 0
log A = ⎝ 0 log λ2 0 ⎠, (C.28)
0 0 log λ3
⎛n

 λ1 0 0
A=⎝ 0 λ2 0 ⎠
n
n
(C.29)
n
0 0 λ3
Given a non-diagonal matrix A, we get the matrix-valued function of A from the formula
⎛ ⎞
f (λ1 ) 0 0
f (A) = QT · ⎝ 0 f (λ2 ) 0 ⎠ · Q,
0 0 f (λ3 )
if A is real and symmetric, or simply diagonalizable [148].

Ellipse and Ellipsoid

The equation of an ellipsoid is


x21 x22 x23
+ + = 1.
a2 b2 c2
This is rewritten to the simple form
x · A · x = 1, (C.30)
where x = (x1 , x2 , x3 )T and ⎛ ⎞
1/a2 0 0
A= ⎝ 0 1/b2 0 ⎠.
0 0 1/c2
Equation (C.30) does not explicitly contain components so the equation does not depend on the
choice of coordinates. The matrix A prescribes the shape and principal axis of the ellipsoid. If A = I,
the equation represents a unit sphere. If a 3 × 3 matrix B with positive eigenvalues,
⎛ ⎞
λ1 0 0
B = ⎝ 0 λ2 0 ⎠
0 0 λ3
represents an ellipsoid, its principal radii a, b, and c are determined by the equations
  
a = 1/ λ1 , b = 1/ λ2 , c = 1/ λ3 . (C.31)

The principal axes are identical with the eigenvectors of the matrix B−1/2 .
Consider two vectors x and y that are related through the equation y = Q · x, where Q is an
orthogonal matrix. Then x = QT y. Substituting this one into the equation of ellipse x · A · x = 1,
we obtain
 
y · Q · A · QT · y = 1.
C.6. TENSORS 337

This equation represents the same ellipsoid with the equation x · A · x = 1. Therefore, the matrix
components are transformed by the rotation of coordinates as

A  = Q · A · QT . (C.32)

If the coordinates are fixed and the entity that the matrix represents is rotated, the result is the entity
represented by the matrix
A = RT · A · R.

C.6 Tensors
Mathematically, a tensor is defined as a linear transformation from between vectors. Within this
book we are able to regard tensors as physical entities that are represented by 3 × 3 or 2 × 2 matrices,
and the equations shown in the previous sections in this appendix apply to tensors. Symmetric
and antisymmetric tensors are represented by symmetric and antisymmetric matrices. Zero and
identical matrices, O and I, represent zero and unit tensors. Orthogonal matrices Q and R stands for
orthogonal tensors.

Higher-order tensors

Consider vectors a and b and the product of their components Tij = ai bi . Let us write this product
as T = ab. According to Eq. (C.11), the vectors are transformed by the rotation of coordinates as


3 
3
ai = Qip ap , bj = Qjq bq .
p=1 q=1

Therefore, the product T is transformed as


   
Tij = ai bj = Qip Qjq ap bq = Qip Tpq Qjq T . (C.33)
p,q p,q

This is T = Q · T · QT , and is in the same form as Eq. (C.32)—the quantity is a tensor. One can
see by comparing Eqs. (C.33) and (C.11) that a vector with a single index is transformed by one Q
and a tensor with two indices is transformed by two Qs. Generalizing this observation, we define a
nth-order tensor Tijk... by the behavior when the coordinates are rotated as


Tijk... = Qip Qjq Qkr · · · Tpqr... , (C.34)
p,q,r,...

where Tijk... has n indices and the right-hand side of this equation has n Qs. Scalar quantities like
temperature are not changed by the rotation, so that the quantities are 0th-order tensors. Vectors
are first-order tensors, and tensors that are represented by square matrices are second-order tensors.
338 APPENDIX C. BASIC EQUATIONS

This is the definition of tensors with their behavior for rotation. The product of a vector and tensor
makes a third-order tensor such as Tij ak and ai Tjk .
Tensors of the second-order are transformed as
T  = Q · T · QT . (C.35)
when coordinates are rotated. The equation
T = R · T · RT
stands for the rotation of the second order tensor T itself.

Projector

In general, the product of two vectors becomes a second-order tensor


⎛ ⎞
a1 b1 a1 b2 a1 b3
ab = ⎝a2 b1 a2 b2 a2 b3 ⎠ .
a3 b1 a3 b2 a3 b3
Using the base vectors of coordinates e(i) , we have a = a1 e(1) + a2 e(2) + a3 e(3) and b = b1 e(1) +
b2 e(2) + b3 e(3) , so that  
ab = ai bj e(i) e(j) .
ij
The products of the base vectors, for example
⎛ ⎞
0 1 0
e(1) e(2) = ⎝0 0 0⎠ ,
0 0 0
act as the bases for second-order tensors:

T= Tij e(i) e(j) . (C.36)
i,j

For a unit vector e = (1, 0, 0)T ,


⎛ ⎞ ⎛ ⎞
1 0 0 0 0 0
ee = ⎝0 0 0⎠ , I − ee = ⎝0 1 0⎠ .
0 0 0 0 0 1
Therefore, the tensor-valued functions
P (e) = ee, (C.37)

P (e) = I − ee (C.38)
are called projectors: if a is an arbitrary vector, P (e)·a and P⊥ (e)·a are the vector quantities that are
the orthogonal projection of a onto the line parallel to e and onto the plane normal to e, respectively
(Fig.C.1). P⊥ is also called an elementary orthogonal projector [148]. Note that P (e) = I − P⊥ (e)
and P (e) · a = e · a.
C.6. TENSORS 339

Isotropic and anisotropic tensors

There are tensors that are invariant for rotation. They are the scalar multiplication of the unit tensor
T = pI. For any rotation Q, T = Q · T · QT = pQ · QT = pI. This indicates that the tensors look
identical regardless of viewpoint. Such tensors are called isotropic. Most tensors are not isotropic
and are called anisotropic. Quartz single crystals are examples that have crystallographic anisotropy.

Dot and colon products

Tensors with an order greater than two have free indices. The operation of identifying two indices
of a tensor and so summing on them is known as contraction. Linear transformation of vectors
%
T · a = j Tij aj is a contraction of the third-order tensor Tij aj . The contraction of this tensor is not
%
unique: j Tji aJ is the other one. The scalar product of vectors is a contraction and is obviously
commutative: a · b = b · a. If a and b are matrices with 3 rows and 1 column, this is aT b. The matrix
components depend on the frame of reference. The vectors are physical quantities independent of
the frames so that the superscript “T” is not included in a · b. It is also an important formula that
x · A · y = y · A · x if A is symmetric.
The matrix multiplication A · B can be regarded as a contraction of the fourth-order tensor Aij Bkl
by identifying the inner and adjacent suffixes j and k, and a derivation of a lower-order tensor. In this
respect, the dot is placed similarly to the scalar product of vectors. The contraction of tensors results
in lower tensors. If two indices in the fourth-order tensor are identified, we write the contraction as
A : B, resulting in a zeroth-order tensor (scalar),

A:B= Aij Bji .
i,j

Second-order tensors only have the contraction



trace T ≡ Tii = T : I = I : T.
i

Let us look over the relationship between the trace of tensors and triple scalar products. We write
the triple scalar product of vectors a, b, and c as a, b, c ≡ a · (b × c), while the following equation
holds for an arbitrary second-order tensor T:
 
   

(T · a) , b, c = T · ai e (i)
, bj e ,
(j)
ck e =
(k)
ai bj ck T · e(i) , e(j) , e(k)
i j

k i,j,k
  (i) (j) (k)
 (i) (j) (k)
= ai bj ck Tip pqr ep eq er = ai bj ck Tii ijk ei ej ek
i,j,k pqr i,j,k

= ijk Tii ai bj ck
i,j,k
340 APPENDIX C. BASIC EQUATIONS


∴ (T · a) , b, c + a, (T · b) , c + a, b, (T · c) = (Tii + Tjj + Tkk ) ijk ai bj ck
i,j,k

= (T11 + T22 + T33 )a1 b2 c3 + (T22 + T33 + T11 )a2 b3 c1


+ (T33 + T11 + T22 )a3 b1 c2 − (T22 + T11 + T33 )a2 b1 c3
− (T33 + T22 + T11 )a3 b2 c1 − (T11 + T33 + T22 )a1 b3 c2

= trace T ijk ai bj ck
ijk

Accordingly, we arrive at an expression of the trace,


1  
trace T = (T · a) , b, c + (a, T · b) , c + (a, b, T · c) . (C.39)
a, b, c

Invariants

Let Ai be the eigenvalues of tensor A. It should be noted that Ai s are not the components of a vector,
because coordinate rotations do not affect eigenvalues. Namely, Ai s are invariant for coordinate
rotations.
Now suppose that the cubic equation

λ3 − AI λ2 + AII λ − AIII = 0 (C.40)

is the characteristic equation2 of the tensor A. The eigenvalues Ai s are the solutions of this equation.
According to the relationship between the coefficients and solutions of cubic equation, we have

AI = A1 + A2 + A3 , (C.41)
AII = A1 A2 + A2 A3 + A3 A1 , (C.42)
AIII = A1 A2 A3 . (C.43)

Given a symmetric tensor, all the eigenvalues are real and the coefficients AI , AII , and AIII become
real. The above three equations indicate that the coefficients are independent of the coordinate
system. Such scalar quantities are called the invariants of the tensor. Since Ai s are invariants, any
scalar-valued function of Ai s is invariant. Especially, those defined by Eqs. (C.41)–(C.43) are called
the first, second, and third basic invariants. They can be also written as

AI = A11 + A22 + A33 = trace A,


     
A22 A23  A33 A31  A11 A12 
AII =   +  + ,
A32 A33  A13 A11  A21 A22 
1  2 
= trace A − trace A2
2
AIII = |A|. (C.44)
2 Compare the sign of the third term in the left-hand side of Eq. (C.40) with that of Eq. (C.48).
C.6. TENSORS 341

The cubic equation


−λ3 + AI λ2 − AII λ + AIII = 0 (C.45)
is identical to the characteristic equation of A [148].

Cayley-Hamilton theorem

The Cayley-Hamilton theorem provides the basis to consider the relation between stress and strain
tensors. The cubic equation
A3 − AI A2 + AII A − AIII I = O (C.46)
stands for the theorem, where the coefficients in the left-hand side are the basic invariants of the
tensor A. Note that by replacing A, I and O in the above equation with λ, 1 and 0, respectively, we
obtain the characteristic equation (Eq. (C.40)).
The theorem is demonstrated as follows. According to Eq. (C.27), the tensor T = A3 − AI A2 +
AII A − AIII I has the eigenvalues Ti = λ3i − AI λ2i + AII λi − AIII , where i = 1, 2 and 3. Equation (C.45)
indicates that all the eigenvalues Ti vanish. Thus T = O.

Deviatoric tensors

Symmetric tensors with zero trace are called deviatoric tensors. Let A be an arbitrary 3×3 symmetric
tensor, then  
 trace A
A =A− I (C.47)
3
is a deviatoric tensor. This equation indicates that symmetric tensors can be decomposed into
isotropic and anisotropic tensors, A = (trace A/3)I + A . Therefore, deviatric tensors derived from
isotropic tensors vanish.
It is a convention of continuum mechanics that the characteristic equation of deviatoric tensors
is written as
λ3 − AI λ2 − AII λ − AIII = 0. (C.48)
Contrary to the characteristic equation of ordinary tensors (Eq. (C.40)), the third term in the left-
hand side of this equation has negative sign. Accordingly, the second basic invariant of deviatoric
tensors has the opposite sign to that of ordinary tensors. The first invariant of A always vanishes,
AI = trace A = 0. The second invariant satisfies
 
AII = − A1 A2 + A2 A3 + A3 A1 (C.49)
        
A
 A23 

A
 A31 

A
 A12 

= −  22 − 33 − 11
A32 A33  A13 A11  A21 A22 
1  2  2  1    
=− trace A − trace A = A:A (C.50)
2 2

1   2   2   2   2  2  2
= A11 + A22 + A33 + A12 + A23 + A31 ≥ 0, (C.51)
2
342 APPENDIX C. BASIC EQUATIONS

Figure C.3: Schematic illustration to explain the sign of the second basic invariant of deviatoric
tensors. Closed cicles lie in the gray zone that runs from upper left to lower right.

where A1 , A2 , and A3 are the eigenvalues of A and, according to Eq. (C.47),

trace A
Ai = Ai − . (C.52)
3
One can see this inequality by considering the correlation coefficient of two variables x and
y. Suppose that six pairs (A1 , A2 ), (A2 , A1 ), (A2 , A3 ), (A3 , A2 ), (A3 , A1 ), (A1 , A3 ) represent six
measurements of two phenomena. If we regard the pairs as points on a coordinate plane, we have
a picture like Fig. C.3. The figure is drawn taking into account the constraint AI = 0 which means
that the mean of A1 , A2 , and A3 is zero. The points are plotted on both sides of the abscissa and
ordinate, and in the gray zone in the figure. The zone indicates a negative correlation between
the observed phenomena. If the averages of x and y are both zero, the correlation coefficient is
%
(constant) × i xi yj , where xi and yi are the ith measurements of x and y. This summation has the
same form as in Eq. (C.49). The correlation is negative so we conclude that AII ≥ 0.
The above comparison of AII with the correlation coefficient indicates that the second invariant
of a deviatoric tensor is a measure of how the eigenvalues of the tensor disperse. Due to Eq. (C.52),
the invariant is also the dispersion of eigenvalues of A. If the eigenvalues are identical, then AII = 0
and A = O, meaning the tensor A is isotropic.

C.7 Polar decomposition


If a real second-order tensor F satisfies |F| > 0, then there are the orthogonal tensor3 R and symmetric
tensors U and V such that
F = R · U = V · R.
3 This should not be represented by Q which indicates the rotation of coordinates.
C.7. POLAR DECOMPOSITION 343

It is important that R, U and V are uniquely determined from F. These statements are collectively
called the polar decomposition theorem. R · U and V · R are called right and left polar decomposition,
respectively.
To understand the theorem, let v = (v1 , v2 , v3 )T be an arbitrary vector. Using a given F, we define
another vector w = F · v. As |F| > 0, w vanishes only if v = 0. We assume that v = 0. Considering
Eq. (C.19),
 
w · w = (F · v) · (F · v) = v · FT · F · v > 0. (C.53)
 
Accordingly, we define the tensor C ≡ FT ·F. This is symmetric, because CT = FT ·F T = FT ·F = C.
In this derivation, we have used the distributive property of transposition (Eq. (C.19)). The real
symmetric tensor C has real eigenvalues and three eigenvectors, so that we choose the Cartesian
coordinates parallel to the eigenvectors, and the tensor has the diagonal form
⎛ ⎞
C1 0 0
C=⎝0 C2 0 ⎠.
0 0 C3

Equation (C.53) becomes C1 v12 + C2 v22 + C3 v32 > 0. This inequality holds only if all the diagonal
components are positive. So, we introduce the tensors

U= C and R = F · U−1 .

The former is a real symmetric tensor and satisfies C = U2 , and the latter is a second-order tensor. If
this is shown to be an orthogonal tensor, the right polar decomposition may be prooved. Using the
definition of R, we have
 −1  −1 −1 −1
R · RT = U T
· FT · F · U = U · C · U
−1 −1
=U ·U·U·U = I.

Therefore, R is indeed an orthogonal tensor. The left polar decomposition is proved in the same way.
There are two remaining problems: (1) the uniqueness of the decomposition and (2) the relation
between U and V. Suppose that it is not unique and that we have F = R · U = R · U. Using the first
decomposition, we obtain
FT · F = UT · RT · R · U = U · U.

Similarly, FT · F = U · U, so that U · U = U · U. This indicates U = U. U is unique. Hence,


R = F · U−1 is also unique. The uniqueness of the left polar decomposition is proved in the same
way. Their uniquenss leads to the relationship between U and V:

RT · V · R = U, V = R · U · RT . (C.54)
344 APPENDIX C. BASIC EQUATIONS

C.8 Calculus
Vectors and tensors can depend on position and form vector and tensor fields. We use the nabla
operator
 
∂ ∂ ∂ ∂
∇ ≡ e(i) = , , T
∂xi ∂x1 ∂x2 ∂x3
to get the spatial derivatives of the field. ∇· and ∇× are divergence and curl operators, respectively,
and are defined as
 ∂  ∂
∇· ≡ e(i) , ∇× ≡ ijk e(j) .
i
∂x i j
∂x j

Applying the ∇ and ∇2 operators to a scalar function F (x), we obtain

  T
∂ ∂F ∂F ∂F
∇F = e(i)
F = , , ,
i
∂xi ∂x1 ∂x2 ∂x3
 ∂2 F ∂2 F ∂2 F ∂2 F
∇2 F = = + + ,
i ∂x2i ∂x21 ∂x22 ∂x23

where
 ∂  (i) ∂   ∂2
∇2 = ∇ · ∇ = e(i) e = 2
i
∂xi ∂xi i ∂xi

is the Laplacian operator. The vector quantity ∇F is the gradient of F , namely, F becomes larger
in the direction indicated by the vector. The operator ∇4 appears when the flexure of the lithosphere
and viscous flow are considered. In the two dimensional case, it reads

∂4 ∂4 ∂4
∇4 = +2 + . (C.55)
∂x41 ∂x21 x22 ∂x42

We use the relationship e(i) · e(j) = δij and the independence of the base vectors on the position
 (i)  %
∂e /∂xj = 0 to differentiate vector field a(x) = i ai (x)e(i) :
 ∂   (j)   ∂ai  ∂ai
∇·a= e(i) · e aj = δij = ,
i
∂xi j i,j
∂xj i
∂xi
 
 ∂   (k)   ∂ak
∇×a= e(j) × e ak = ijk e(j) e(k)
j
∂x j ∂x j
k j,k
 
∂a3 ∂a2 ∂a1 ∂a3 ∂a2 ∂a1 T
= − , − , − ,
∂x2 ∂x3 ∂x3 ∂x1 ∂x1 ∂x2
 
 ∂  (j)   ∂aj
∇a = e(i) e aj = e(i) e(j) .
i
∂x i j i,j
∂x i
C.8. CALCULUS 345

The results are a scalar, vector, and second-order tensor field, respectively. Similarly, the nabla
operator is contracted with a second-order tensor r = (σij ), to yield a vector
 ∂     ∂σji
∇·r= e(i) · σjk e(j) e(k) = e(i) . (C.56)
i
∂xi i,j
∂xj
j,k

The curl of the ∇F always vanishes:


∇ × ∇F = 0, (C.57)

because its first component is


∂ ∂F ∂ ∂F
− =0
∂x2 ∂x3 ∂x3 ∂x2
and, similarly, the other components vanish. Applying the Laplacian operator to a vector a yields a
vector
 ∂ 2 ai
∇2 a = e(i) 2 .
i ∂xi
Since x · A · x is a scalar, ∇x · A · x is a vector. If x is an independent variable, its components
are also independent of each other. Therefore, we have ∂xi /∂xj = δij or

∂x
∇x = = I.
∂x
The ith component of ∇x · A · x is

∂     
xj Ajk xk = δij Ajk xk + xj Ajk δki = Aik xk + xj Aji (C.58)
∂xi j
j,k j,k k
   
= Aij xj + AT ij xj . (C.59)
j

Therefore, we have
   
∇ x · A · x = A + AT · x. (C.60)

If A is symmetric, then
 
∇ x · A · x = 2A · x. (C.61)

We use formulae that exchange line, surface, and volume integrals4 . Gauss’s divergence theorem
provides a formula to interchange volume and surface integrals. Suppose V is a volume with a
smooth closed surface S (Fig. C.4). On the surface, we define the surface element dS where the
unit vector N is the outward normal. If A is a differentiable quantity, the theorem says that
 
∇A dV = NA dS. (C.62)
V S

4 See [134] for derivation of integral formulae.


346 APPENDIX C. BASIC EQUATIONS

Figure C.4: Schematic picture for an explanation of Gauss’s divergence theorem and Stokes’ theo-
rem. (a) A rock mass with a volume V is surrounded by a closed and smooth surface S. The unit
vector N is the outward normal on the surface element dS. (b) Closed curve C bounds a surface S.
Unit vector t is tangent to C.

Any scalar, vector, or tensor field that A represents satisfies this equation. In the case of a vector or
tensor field, we have  
∇ · A dV = N · A dS. (C.63)
V S

We also obtain the following formula from Gauss’s divergence theorem:


 
 2  ∂ψ
ϕ∇ ψ + ∇ϕ · ∇ψ dV = ϕ dS
∂N
  
V S

 2  ∂ψ ∂ϕ
= ϕ∇ ψ + ϕ∇ ψ dV =
2
ϕ −ψ dS,
V S ∂N ∂N

where φ and ψ are scalar fields and ∂/∂N ≡ (NN ) · ∇ is the gradient operator in the direction of
N.
Stokes’ theorem exchanges a surface integral along a closed curve C for a surface integral over
the open surface S that is bordered by the closed curve (Fig. C.4). For the vector field a, it reads
? 
(a · t) ds = (∇ × a) · N dS,
C S

where the unit vector t is tangent to the curve and ds is an element of the curve. The unit N in the
above equation is locally normal to the surface.

C.9 Rotation
Tectonic rotations are observable phenomana via geologic methods such as paleomagnetism and
geologic mapping of twisted rock masses. Here we consider how to treat rotations.
Suppose the Cartesian coordinates O-123 and O-1 2 3 that have the origin in common. Their
relationship is described by three angles, θ, φ, ψ (Fig. C.5). These angles are called Euler angles5 .
5 There are several definitions on Euler angles. See reference [69].
C.9. ROTATION 347

Figure C.5: Cartesian coordinates with the origin in common and the Euler angles describing their
relationship.

The orthogonal tensor that stands for the transformation from vector and tensor components of O-
123 to those of O-1 2 3 is
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
cos ψ sin ψ 0 1 0 0 cos φ sin φ 0
Q = ⎝− sin ψ cos ψ 0⎠ · ⎝0 cos θ sin θ ⎠ · ⎝− sin φ cos φ 0⎠ (C.64)
0 0 1 0 − sin θ cos θ 0 0 1
⎛ ⎞
cos ψ cos φ − cos θ sin φ sin ψ cos ψ sin φ + cos θ cos φ sin ψ sin ψ sin θ
= ⎝sin ψ cos φ − cos θ sin φ cos ψ − sin ψ sin φ + cos θ cos φ cos ψ cos ψ sin θ ⎠ . (C.65)
sin θ sin φ − sin θ cos φ cos θ

The three matrices from left to right in Eq. (C.64) represent the rotation about O3 , O1 , and O3 axes,
respectively. The equation indicates that the three rotations are equal to the single rotation shown by
the matrix in Eq. (C.65). If an axis of rotation is specified with θ and φ, ψ indicates the angle of
rotation.
If a rotation matrix is given, how can we calculate the direction of the pivot and rotation angle α
from the matrix? If we choose the third coordinate axis as being parallel to the pivot, the orthogonal
matrix has the form ⎛ ⎞
cos α sin α 0
Q = ⎝− sin α cos α 0⎠ .
0 0 1
Therefore, we have the trace of this matrix:

trace Q = 1 + 2 cos α. (C.66)

The trace of a matrix is an invariant, so that the value of this trace is not affected by the choice
of coordinates. The rotation angle α is, therefore, obtained by substituting the trance of the given
orthogonal tensor into Eq. (C.66). If a vector u is parallel to the pivot of rotation, we have u = I · u =
Q · u, indicating that Eq. (C.23) holds. Accordingly, the pivot is determined as an eigenvector of Q.
The vector corresponds to the only real eigenvalue of the non-symmetric tensor Q.
348 APPENDIX C. BASIC EQUATIONS

Rotation is also represented by the antisymmetric tensor


⎛ ⎞
0 −W3 W2
W = ⎝ W3 0 −W1 ⎠ . (C.67)
−W2 W1 0
%
Let us define a vector by these matrix components as W = i Wi e(i) , then we have

Wij = − ijk Wk . (C.68)
k

Contrary, we obtain the components of the vector from the matrix by the equation
1
Wi = − ijk Wjk . (C.69)
2
j,k

Such vectors that correspond to antisymmetric tensors are called axial vectors. In the case that the
axial vector W represents angular velocity, W × a indicates the velocity of the end point of a whose
initial point is fixed on the axis of rotation (Fig. 2.3). Accordingly, W should be related to the
%
rotation. For an arbitrary vector a = i ai e(i) , we have Q · a = − · W · a from Eq. (C.68):

Wij = − ijk Wk .
k

However, j and k are dummy indices. Replacing them by k and j, respectively, we have
  
Wij aj = − ikj Wj ak = ijk Wj ak .
j k,j j,k

This indicates that W · a = W × a. The angle of rotation during a time interval t is

ψ = |W |t. (C.70)

Then a question arises: how is the antisymmetric tensor W related to the orthogonal tensor R?
Suppose that the terminal point of a vector a is rotated to that of b during the time interval, and the
rotation is represented by R, namely
b = R · a. (C.71)
A broken line approximates the circular orbit of the point (Fig. 2.3). The arc length of one segment
is |W |dt. Let t be equal to n · dt, where n is a natural number. Taking the limit n → ∞, the broken
line approaches the circular orbit. Rotation during the interval dt is written as

a = a + (W × a)dt = a + (W · a)dt = (I + Wdt) · a,


a = (I + Wdt) · a = (I + Wdt) · (I + Wdt) · a.

The above two equations indicate that


 t n
b = lim I + W · a.
n→∞ n
C.10. EXERCISE 349

Comparing this and Eq. (C.71), we have


 t n
R = lim I + W .
n→∞ n
Taylor expansion of the nth power is
 t n n(n − 1)  t 2
I + W = I + tW + W
n 2! n
n(n − 1)(n − 2)  t 3
+ W + ···
3! n
Taking the limit n → ∞, the right-hand side is simplified, and we have
1 1
R = I + tW + (tW)2 + (tW)3 + · · · (C.72)
2! 3!
= exp (tW)

Recalling the Cayley-Hamilton theorem (Eq. (C.46)), the higher-order terms in Eq. (C.72) are
replaced by lower terms, and we obtain
  ψ !
sin ψ 1 sin 2
R=I+ tW + ψ  (tW)2 , (C.73)
ψ 2
2

where ψ is defined by Eq. (C.70). This is called Rodrigues’s equation. This is also written as
⎛ ⎞ ⎛ ⎞2
0 −3 2 0 −3 2
R = I + sin α ⎝ 3 0 −1 ⎠ + (1 − cos α) ⎝ 3 0 −1 ⎠ , (C.74)
−2 1 0 −2 1 0

when an angle of rotation α about the fixed axis denoted by the unit vector  = (1 , 2 , 3 )T is
given6 .

C.10 Exercise
B.1 Fault-slip data are represented by two unit vectors, one for the perpendicular to fault plane,
and the other for slip direction. Let the unit vector u = (u1 , u2 , u3 )T point to the slip direction
of the hangingwall, and let v = (v1 , v2 , v3 )T be the unit vector normal to the fault plane. If the
plane is horizontal, the vector points upward. The vector components are evaluated in the Cartesian
coordinates O-123. The third unit vector w is defined to be normal to the two vectors and to form
a right-hand system in the order of u, v, and w. Show the orthogonal tensor that stands for the
coordinate transformation from O-123 to O-uvw.

6 See Exercise 2.5 in p. 56 and its answer in p. 355.


Appendix D

Answers to Selected Problems


 
1.2 Substituting x = F · ξ into x · x = 1, we have (F · ξ) · (F · ξ) = ξ · FT · F · ξ = 1. Therefore,
the reciprocal ellipsoid is represented by FT · F = C = U2 (Eq. (1.19)). It is explained inSection
C.5 that the principal radii of the ellipsoid is given by the reciprocal of the eigenvalues of C. The
radii are equal to the eigenvalues of U−1 .

     
1.3 Substituting ξ = F−1 ·x into ξ·A·ξ = 1, we have F−1 · x ·A· F−1 · x = x· F−1 T ·A·F−1 ·x = 1.
 −1 T
Consequently, the resultant ellipsoid is represented by the symmetric tensor F · A · F−1 .

1.4 A homogeneous deformation transforms an ellipse into another. Let R and φ be the aspect ratio
and long-axis orientation of an ellipse, respectively. Let us distinguish those of pre- and post-strain
ellipses by the subscripts “i” and “f”, then the post-strain orientation is related to the aspect ratio of
strain ellipse, Rs , and the pre-strain parameters by Ramsay’s formula [188]
 
2Rs Ri2 − 1 sin 2φi
tan 2φf =  2      . (D.1)
Ri + 1 R22 − 1 + Ri2 − 1 Rs2 + 1 cos 2φi

The major axis of the strain ellipse is used as the reference orientation for these formulas. Therefore,
if another orientation is chosen for the reference, φi and φf should be corrected by the angle between
the major axis and the reference orientation. In order to determine the optimal strain ellipse for
deformed elliptical grains, we assume a uniform  frequency distribution
 of φi . If this is valid and if
(1) (N)
we have many grains, the mean orientation φi + · · · + φi /N is parallel to the major axis of
the strain ellipse, where N is the number of grains and the parenthesized superscript (k) stands for
the kth grain. Lisle proposes tests to check the validity [121].
Equation (D.1) describes the change of orientation by strain. The same equation can be used
for destraining, i.e., the initial orientation φi is calculated through the equation by exchanging Rs
(1) (N)
for 1/Rs and (Ri , φi ) for (Rf , φf ). Accordingly, a set of the unknown variables, φi , . . . , φi , is
calculated with an assumed Rs .

351
352 APPENDIX D. ANSWERS TO SELECTED PROBLEMS

If the frequency distribution is represented by a histogram with n bins for φi , the uniformity
requires that each bin has N/n grains. The discrepancy from this uniform distribution is quantified
by the χ 2 -statistic
 2
n
Fj − N/n
χ =
2
,
N/n
j=1

where Fj is the"
number of grains
 contained
 in the#
jth bin. Note that this is a function of not only the
(1) (1) (N) (N)
given data set Rf , φf ,..., Rf , φf but also the unknown variable Rs . Therefore,
parameters of the optimal strain ellipse are determined by minimizing χ 2 . This is known as the
θ-curve method1 [121].

2.1 The velocity field for the simple shearing motion shown in Fig. 1.11 is v = (2q̇x2 , 0, 0)T ,
where q̇ is a constant. We have the velocity gradient
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 2 0 0 1 0 0 1 0
L = q̇ ⎝0 0 0⎠ ∴ D = q̇ ⎝1 0 0⎠ , W = q̇ ⎝−1 0 0⎠ .
0 0 0 0 0 0 0 0 0

%
2.2 The material derivative of the right-hand side of |F| = i,j,k ijk Fi1 Fj2 Fk3 is
  
ijk Ḟi1 Fj2 Fk3 + Fi1 Ḟj2 Fk3 + Fi1 Fj2 Ḟk3
i,j,k
 
  ∂vi ∂xl  ∂vj ∂xl  ∂vk ∂xl
= ijk Fj2 Fk3 + Fi1 Fk3 + Fi1 Fj2
∂xl ∂ξ1 ∂xl ∂ξ2 ∂xl ∂ξ3
i,j,k l l l
  
= ijk Lil Fl1 Fj2 Fk3 + ijk Fi1 Ljl Fl2 Fk3 + ijk Fi1 Fj2 Lkl Fl3 .
i,j,k,l

Note that i, j, k and l are dummy indices; any symbol can replace them without affecting the sum-
mation. Accordingly, we exchange i and l in the first term, j and l in the second one, and k and l in
the third one in the last parentheses. The result is

D|F |  
= ljk Lli Fi1 Fj2 Fk3 + ilk Fi1 Llj Fj2 Fk3 + ijl Fi1 Fj2 Llk Fk3
Dt
i,j,k,l
 
= ljk Lli + ilk Llj + ijl Llk Fi1 Fj2 Fk3 .
i,j,k.l

1 Solutions of the θ-curve method are numerically unstable. A recent method [267] determines a stable solution and its
uncertainty from elliptical strain markers with initially anisotropic grain fabric. From the 18 ellipses in Fig. 1.19, aspect ratio
and long-axis orientation of the optimal strain ellipse were evaluated by the method as Rs = 1.70 ± 0.08 and φs = 15.0 ± 2.7◦ ,
respectively, where errors are indicated by 95% confidence intervals.
353

If l = i, then l = j or l = k. In both cases, we have ljk = 0. Therefore, the identities l = i, l = j


and l = k are needed for ljk , ilk and ijl to vanish, respectively.

D|F |    
∴ = L11 + L22 + L33 ijk Fi1 Fj2 Fk3 = trace L |F |.
Dt
i,j,k

%    
Using J = |F| and trace L = i (∂vi /∂xi ) = ∇ · v, we obtain J˙ = trace L J = ∇ · v J .

2.3 Suppose that the rate of strain is represented by the function

G(t) = Ėxx = −Ėzz . (D.2)

Substituting Eq. (D.2) into (2.39), we have β̇/β = G(t). Integrating this equation, we get

ln β = G(t) dt. (D.3)

Using the initial condition that β = 1 when t = 0, we obtain


 t 
β = exp G(s) ds + 1.
0

If G(t) = const., from the above equation we get that

β = exp(G0 t). (D.4)

Therefore, β is not proportional to time t even for a constant strain rate. Instead, Eq. (D.3) shows
that logarithmic strain piles up proportionally with time2 . Accordingly, Ė and D indicates the time
rate of logarithmic strain.
Suppose that w is the width of a pure shear rift that was initially w0 . Because of w = βw0 , a
constant strain rate results in the evolution of the width as w = w0 exp(G0 t). The width increases
exponentially. On the other hand, if the rate decreases as G(t) = G0 e−t/τ , the β factor evolves as
 
ln β = G0 τ 1 − e−t/τ . The content of the parentheses approaches the unity, so that the rift stops its
activity when the width reaches w0 exp(G0 τ). After the strain rate becomes  10−17 s−1 , we may
think of the rifting as being abandoned.

2.5 Using Eq. (2.41) and referring to Fig. 2.3, we see that a has became a+(W·a)dt = (I+Wdt)·a
at the time t = dt. Similarly, at the next step (t = 2dt), the result is (I + Wdt)2 a. If the time needed
to carry a to b is t and t = ndt, we have
 n
$ · a = lim I + t W · a.
b=R
n→∞ n
2 This is true only for coaxial deformations.
354 APPENDIX D. ANSWERS TO SELECTED PROBLEMS

The power in the right-hand side is expanded to


    2  3
t n tW n(n − 1) tW n(n − 1)(n − 2) tW
I+ W =I+ + + + ···
n n 2! n 3! n

which converges to
$ = I + tW + 1 (tW)2 + 1 (tW)3 + · · ·
R (D.5)
2! 3!
if we let n approach infinity. The right-hand side can be written as exp(tW). The Cayley-Hamilton
theorem allows us to replace the infinite series in this right-hand side with the sum of a few terms.
W has the components
⎛ ⎞ ⎧
0 −w3 w2 ⎨ WI = trace W=0
W = ⎝ w3 0 −w1 ⎠ ∴ WII = 21 (trace W)2 − trace W2 ,

−w2 w1 0 WIII = det W = 0

where WI , WII and WIII are the basic invariants of W. The theorem warrants the identity

W3 = −|w|2 W, (D.6)

where w = (w1 , w2 , w3 )T is the axial vector of W. From Eq. (D.6), we have


'
(−1)n−1 |w|2(n−1) W (n is odd),
W =
n
(−1)n−1 |w|2(n−1) W2 (n is even).

Therefore, Eq. (D.5) becomes


∞ 
 
$ =I+ t2n−1 t2n
R W 2n−1
+ W2n
(2n − 1)! (2n)!
n=1
! !
∞
|w| 2(n−1) 2n−1
t ∞
|w| 2(n−1) 2n
t
=I+ (−1)n−1 W+ (−1)n−1 W2
(2n − 1)! (2n)!
n=1 n=1
! !
∞ 2n−2 ∞ 2n−2
n−1 θ n−1 θ
=I+ (−1) tW + (−1) t2 W2 ,
(2n − 1)! (2n)!
n=1 n=1

where θ = |w|t is the angle of rotation in the time interval t. Using these equations and the equations
 
sin θ θ2 θ4 1 sin(θ/2) 1 θ2 θ4
=1− + − · · · and = − + − ··· ,
θ 3! 5! 2 θ/2 2! 4! 6!

we obtain Rodrigues’ equation


    !2
sin θ 1 sin θ2
$ =I+
R tW + θ (tW)2 . (D.7)
θ 2
2
355

Substituting θ = wt into Eq. (D.7), we get


    
$ = I + (sin wt)W + 2 sin2 wt W2 = I + (sin wt)W + 2 1 − cos2 wt W2
R
w2 2 w2 2
1 − cos wt 2
= I + (sin wt)W + W .
w2
Now, let us take the third axis of our Cartesian coordinates parallel to the rotation axis, then
w1 = w2 = 0 and w = w3 > 0, and we have
⎛ ⎞ ⎛ ⎞
0 −w 0 −w 0 0
W = ⎝w 0 0⎠ , W2 = ⎝ 0 −w 0⎠ . (D.8)
0 0 0 0 0 0

Combining these equation and Eq. (D.8) we obtain


⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 −1 0 −1 0 0 cos wt − sin wt 0
$ = I + (sin wt) ⎝1 0 0⎠ + (1 − cos wt) ⎝ 0
R −1 0⎠ = ⎝ sin wt cos wt 0⎠ .
0 0 0 0 0 0 0 0 1

The last matrix has the form common to that of Eq. (C.16), rotation around the third coordinate
axis by the angle wt.

2.6 Consider the ith dip-slip normal fault with the parameters shown in Fig. D.1. The horizontal
displacement (heave) of this fault is hi / tan di , so that the increase of the surface area due to this
%
fault is ΔSi = Li hi / tan di . If there are N normal faults, the total area increase is ΔS = N i ΔSi .
Let S be the initial surface area of the planet, then we have S = 4πR . The fault activity increases
2

the surface area and the radius by ΔS and ΔR, respectively. Because of |ΔR/R|  1, the left-hand
side of 4π(R + ΔR)2 = S + ΔS is approximated as
 
ΔR ΔR ΔR
4π(R + ΔR) ≈ 4πR 1 + 2
2 2
= 4πR2 + 8πR2 · = S + 8πR2 · .
R R R

Rearranging this equation, we obtain Eq. (2.56).

3.1 First of all, we define the x- and y-axes in the due east and north, respectively, and the z-axis as
vertically upward (Fig. D.2). We also use spherical coordinates with colatitude from the z direction
and longitude from the x direction, respectively. Axes of the Cartesian coordinates are parallel to
the principal axes of the given state of stress. Therefore, the components of the stress tensor vanish
except for the diagonal components σ11 = 4, σ22 = 5 and σ33 = 8 MPa.
The pole to the plane of fracture has the azimuth N210◦ E and plunge 30◦ . The rock mass for
which we are going to calculate the stress vector t(n) lies under the specified plane of fracture, so
that the inward unit vector n should have a negative z component. The colatitude of the direction
356 APPENDIX D. ANSWERS TO SELECTED PROBLEMS

Figure D.1: Increase of surface area ΔS by a normal fault.

Figure D.2: Cartesian coordinates O-xyz and spherical coordinates O-θϕ.

that the vector is pointing is 90◦ + 30◦ . Therefore, we have


⎛ ⎞ ⎛ ⎞ ⎛ ⎞
sin(90◦ + 30◦ ) cos 210◦ √3 √3
1
n = ⎝ sin(90◦ + 30◦ ) sin 210◦ ⎠ = − ⎝ 3⎠ , t(n) = − ⎝5 3/4⎠ .
4
cos(90◦ + 30◦ ) 2 4

The orthogonal projector onto the line parallel to n is


⎛ √ ⎞
9
√ 3 3 6

1 ⎝3 3
P (n) = nn = 3
√ 2 3⎠ ,
16
6 2 3 4

and that onto the plane perpendicular to the vector is P⊥ (n) = I − P (n). Therefore, the vectorial
form of the normal and shear stresses are
1   1  
tN (n) = P (n) · t(n) = − 249, 83 3, 166 T , tS (n) = P⊥ (n) · t(n) = 57, 3 3, −90 T ,
64 64
respectively. Thus, we obtain the normal stress |tN (n)| = 83/16 ≈ 5.2 MPa. Since the shear stress
357


is |tS (n)| = 3 79/16 ≈ 1.7 MPa, the unit vector parallel to the latter vector is
√   
tS (n) 79
a= = 19, 3, 30 T ,
|tS (n)| 316

which indicates the direction of shear. The x and y components of the latter vector (shear stress
vector) are positive, and the former is larger than the latter, so that the trend of the vector is between
northeast and due east. The rake of the vector is the angle between the vector and N030◦ , the
trendof the fracture. Thelongitude
 of latter from the x direction is 60◦ so that the unit vector
the 

b = cos 60◦ , sin 60◦ , 0 T = 21 1, 3, 0 T indicates the trend. The rake θ is obtained from the

equation cos θ = |a| |b| cos θ = 11 79/316 ≈ 0.31 and finally we obtain θ ≈ 72◦ .

4.1 Let us use the principal stress space so that the traction has the components t(n) = (σ1 n1 , σ2 n2 ,
σ3 n3 )T . The tangential component of this vector on the surface element on which the traction is
considered is the shear stress
σS2 = |t(n)|2 − σN2 , (D.9)

where the normal stress is


σN = t(n) · n = σ1 n21 + σ2 n22 + σ3 n23 . (D.10)

From Eq. (D.10) into (D.9), we have

|t(n)|2 = σN2 + σS2 = σ12 n21 + σ22 n22 + σ32 n23 , (D.11)
σS2 = σ12 n21 + σ22 n22 + σ32 n23 − (σ1 n21 + σ2 n22 + σ3 n23 )2 . (D.12)

Combining Eqs. (D.10), (D.11) and n21 + n22 + n23 = 1, we obtain

(σN − σ2 )(σN − σ3 ) + σS2


n21 = , (D.13)
(σ1 − σ2 )(σ1 − σ3 )
(σN − σ3 )(σN − σ1 ) + σS2
−n22 = , (D.14)
(σ1 − σ2 )(σ2 − σ3 )
(σN − σ2 )(σN − σ1 ) + σS2
n23 = . (D.15)
(σ2 − σ3 )(σ1 − σ3 )

Note that the denominator and numerator of Eq. (D.13) can be rearranged as
 σ2 + σ3  2  σ2 − σ3  2
(σ1 − σ2 )(σ1 − σ3 ) = σ12 − σ2 σ1 − σ3 σ1 + σ3 σ2 = σ1 − − ,
2 2
 σ2 + σ3  2  σ2 − σ3  2
(σN − σ2 )(σN − σ3 ) = −σN σ2 + σ2 σ3 + σN2 − σ3 σN = σN − − .
2 2
358 APPENDIX D. ANSWERS TO SELECTED PROBLEMS

Similar rearrangements can be done for those of Eqs. (D.14) and (D.15) and, consequently, we
obtain
 
 σ2 + σ3  2  σ − σ 2
2 3 σ2 + σ3  2  σ2 − σ3  2
σN − + σS2 = + n21 σ1 − − , (D.16)
2 2 2 2
  
σ1 + σ3  2  σ − σ 2
1 3 σ1 + σ3  2  σ1 − σ3  2
σN − + σS =
2
− n2 σ 2 −
2
− , (D.17)
2 2 2 2
  
σ1 + σ2  2  σ − σ 2
1 2
 σ1 + σ2  2  σ1 − σ2  2
σN − + σS2 = + n23 σ3 − − . (D.18)
2 2 2 2

These are obviously the equations of three circles on the σN σS plane. They are centered at ( σ2 +σ3
2 , 0),
σ3 +σ1 σ2 +σ1
( 2 , 0) and ( 2 , 0), respectively. The normal and shear stresses σN and σS have to simultane-
ously satisfy Eqs. (D.16)–(D.18). The square of the right-hand side of these equations is the radius
of the circles. The radius indicated by that of Eq. (D.16) has the minimum (σ2 + σ3 )/2 for n1 = 0.
The circle with this minimum radius is illustrated as the circle C1 in Fig. 4.5. The circle represented
by Eq. (D.17) has the maximum radius (σ1 − σ3 )/2 for n2 = 0. This largest circle is shown as the
circle C2 in the figure. Equation (D.18) indicates a circle whose minimum radius is (σ1 − σ2 )/2, and
this smallest circle is shown as the circle C3 , also. For the direction cosines n1 , n2 and n3 greater
than 0, the solution (σN , σS ) for the simultaneous equations (Eqs. (D.16)–(D.18)) is plotted in the
shaded area in Fig. 4.5.

4.3 The time needed for a significant temperature change in the rising mass is τ ∼ L2 /κ ≈ 1012
kg m−3 . Let v be the rising velocity, then the time to traverse the radius Δz = 10 km is Δz/v.
Therefore, the mass can go up while it is not significantly cooled if Δz/v  L2 /κ. Accordingly,
Péclet number
κΔz
Pe = 2
L v
is a convenient dimensionless number to judge whether the mass can carry significant heat. In this
case, Pe  1 is satisfied if v  10−8 ms−1 = 0.32 m/year.

5.1 According to Cauchy’s stress formula, we have t = ST · N and t = (S )T · N  . The vectors on
the both sides of this equation are transformed as t = Q · t and N  = Q · N. Combining these four
equations, we have t = Q · t = Q · (ST · N ) = (S )T · Q · N. Rearrainging this equation, we obtain
the equation {ST − QT · [(S )T · Q]} · N = 0 for any N. Therefore, using Eq. (C.19), we obtain
S = [(S )T · Q]T · Q. Therefore, multiplying the both sides of this equation with QT from the right,
we have S · QT = QT · S , and further S = Q · S · QT .

5.2 See Fig. D.3.


359

Figure D.3: Principal axes of stress and strain are not always parallel for anisotropic materials. This
schematic picture illustrates the obliqueness for a material composed of laminated veneers.

6.1 From Eqs. (4.21) and (4.22), we have

1 1
σN = (σ1 + σ3 ) − (σ1 − σ3 ) cos 2θ (D.19)
2 2
1
σS = − (σ1 − σ3 ) sin 2θ. (D.20)
2
Note that the range 0 ≤ θ ≤ 90◦ is significant for this problem. Hence, combining Eqs. (6.2), (D.19)
and (D.20), we obtain

μ 1
τ0 + (σ1 + σ3 ) + (σ1 − σ3 )(sin 2θ − μ cos 2θ) = 0.
2 2
Eliminating θ by combining this equation and Eq. (6.5), we obtain
   
σ1 μ2 + 1 − μ − σ3 μ2 + 1 + μ = 2τ0 . (D.21)

This is the expression of the Coulomb-Navier criterion using principal stresses. Substituting the
principal stresses for uniaxial compression (σ3 = 0 and σ1 = σ C and uniaxial tension (σ3 = σ T and
σ1 = 0) into Eq. (D.21) gives
 −1
σC = 2τ0 μ2 + 1 − μ ,
 −1 (D.22)
σT = 2τ0 μ2 + 1 + μ .

Combining Eqs. (D.21) and (D.22), we obtain another expression of the criterion
σ1 σ3
− = 1.
σT σC

6.2 Extension fracturing occurs when the outer Mohr circle connecting the maximum and mini-
mum principal values of the effective stress tensor is inscribed in the failure envelope at σN = σT
360 APPENDIX D. ANSWERS TO SELECTED PROBLEMS

Figure D.4: Parabola representing the Griffith theory of failure and Mohr circle inscribed in the
parabola.

(Fig. D.4). The effective stress represented by a Mohr circle with a radius smaller than the radius of
curvature of the envelope at this point can form extension fractures.
Non-dimensionalizing normal and shear stresses by the tensile strength x = σS /|σT | and y =
σN /|σT |, Eq. (6.1) is converted into the equation, y = 14 x2 − 1. The curvature of this parabola is (Eq.
(8.4))
  2 −3/2 
K = 1 + y y .
 
Substituting y  = x/2 and y  = 1/2 into this equation, we have K = 1
16 4 + x2 3/2 . Therefore, we
have the radius of curvature
1 16
R= =  .
K 4 + x2 3/2
The radius of curvature at the turning point of the parabola (x = 0) equals 2.
Therefore, extension fracturing is possible if Δσ < 4|σT |. The depth extent z to which the
fracturing is allowed is given by solving the equation ρgz = 4|σT |. That is, z = 4|σT |/ρg. Using
ρ = 2.3 × 103 kg m−3 and σT = −0.1 MPa, we have a maximum depth of z ≈ 17 m. If we use the
Earth pressure at rest (Eq. (7.20)), the maximum depth becomes a few times deeper.

7.1 We assume an infinitesimal simple shear,


⎛ ⎞
1 2q 0
F = ⎝0 1 0⎠ , (D.23)
0 0 1

where q ≈ 0 and the rectangular Cartesian coordinates O-12 in Fig. D.5 is used. This is an infinites-
imal deformation, so that Eq. (2.10) applies to this case, i.e., F = I + E + X. Combining this and Eq.
(D.23), we obtain ⎛ ⎞
0 q 0
E = ⎝ q 0 0⎠ .
0 0 0
361

Figure D.5: Infinitesimal simple shear and its principal axes. The gray square shows the shape before
deformation, and the parallelogram indicates that afterwards.

The characteristic equation of E is λ3 − q 2 λ = 0, so that the principal strains are E1 = q, E2 =


−q and E3 = 0. Substituting these into Eq. (7.4), we obtain {S1 , S2 , S3 } = 2q{G, −G, 0} =
γ{G, −G, 0}, where γ = tan 2q ≈ 2q is the engineering shear strain. Taking the coordinate system
O-1 2 parallel to the principal axes (Fig. D.5), the stress tensor is written in the primed coordinates
as ⎛ ⎞
γ 0 0
S = G ⎝0 −γ 0⎠ .
0 0 0
The stress tensor has the expression in the first coordinate system as
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
cos 45◦ − sin 45◦ 0 cos 45◦ sin 45◦ 0 0 Gγ 0

S = sin 45◦ cos 45◦ ⎠ ⎝
0 S − sin 45◦ cos 45◦ 0⎠ = ⎝Gγ 0 0⎠ .
0 0 1 0 0 1 0 0 0

Figure 3.2(a) indicates that the traction on a plane perpendicular to the O-2 axis is (S12 , S22 , S32 )T ,
and S12 = Gγ is the only non-zero component of the traction. The component is parallel to the
motion by the simple shear, and is proportional to the engineering shear strain, γ. The shear modulus
G is the constant of proportionality.

9.4 Let h(x, t) be the surface topography at time t. Because of |h0 (x)|  H, the amplitude of
h(x, t) is kept very small. Therefore, the surface materials may move only vertically, so that we
assume the boundary condition for the velocity:
 

 ∂ψ 
 = 0.
vx  = − (D.24)
z=0 ∂z  z=0
362 APPENDIX D. ANSWERS TO SELECTED PROBLEMS

The fluid adheres the rigid basement at the interface, so we have the basal boundary conditions
   
 ∂ψ   ∂ψ 
vx  =− 
 = 0, vz 
 =  = 0. (D.25)
z=H ∂z z=H z=H ∂x  z=H
There is no flux across the interface so we also have


ψ = 0. (D.26)
z=H
 
The initial topography is assumed to vanish at infinity, h0 (±∞) = 0, therefore, we have the
boundary conditions
   

 ∂ψ 

 ∂ψ 

vx  =− = 0, vz  = = 0.
x=±∞ ∂z  x=±∞ x=±∞∂x  x=±∞

These conditions warrant the existence of the Fourier transform of the stream function, ψ: . The
: succeeds
boundary conditions in Eqs. (D.24), (D.25) and (D.26) are independent from x, so that ψ
the conditions   
∂ψ: : 
∂ψ 

 =  =ψ:  = 0. (D.27)
∂z z=0 ∂z z=H z=H
Under these conditions, the solution of Eq. (9.17) has the form

: = (A + Bkz) cosh kz + (C + Dkz) sinh kz


ψ (D.28)

and we obtain
:
∂φ  
= Ak sinh kz + B k cosh kz + k 2 z sinh kz
∂z
 
+ Ck cosh kz + D k sinh kz + k 2 z cosh kz . (D.29)

The coefficients A, B, C and D are the functions of k and t. Combining Eqs. (D.27)) and (D.29),
we obtain the simultaneous equations
⎛ ⎞ ⎛A⎞
0 1 1 0
⎝k sinh kH k cosh kH + Hk 2 sinh kH k cosh kH k sinh kH + Hk 2 cosh kH ⎠ ⎜ B⎟
⎝ C ⎠ = 0.
cosh kH kH cosh kH sinh kH kH sinh kH
D

Noting the identity cosh2 kH − sinh2 kH = 1, we have


sinh2 kH − (kH)2 sinh2 kH
A= B, C = −B, D=− B. (D.30)
sinh kH cosh kH + kH sinh kH cosh kH + kH
Accordingly, once B is known, other coefficients are readily calculated through these equations.
B is determined from the surface boundary condition. The surface is a free boundary so that the
linearlized boundary condition in Eq. (9.77) can be applied to this problem, i.e., we have


Szz  = −ρgh. (D.31)
z=0
363

Equation (9.6) has the z component


∂p ∂ψ
= −η∇2 + ρg. (D.32)
∂z ∂x
Combining Eqs. (D.31) and (D.32) and the Szz in Eq. (9.12), p is eliminated to give
∞
1 Bηk 2 eikx
h(x, t) = − √ dk. (D.33)
2π −∞ ρg
B is determined by comparing the Fourier transform of the initial topography in Eq. (9.102) with
Eq. (D.33). Namely, we have
: ηk 2
h0 (k) = − B(k, 0).
ρg
The initial surface has very small amplitudes h0 (x) ≈ 0 so that the subsidence rate of the surface
∂h/∂t is approximated as  
∂h  ∂ψ 

≈ vz  =  . (D.34)
∂t z=0 ∂x z=0
Substituting Eq. (D.28) into (D.34) and rearranging, we obtain
  ∞ 
∂h ∂ψ  = √ 1 
=  : e dk 
ik ψ ikx

∂t ∂z z=0 2π ∞
" 1 
z=0
 
∞ ikx # A ∞
= √ ik (A + Bkz) cosh kz +(C + Dkz) sinh kz e dk  = √ ikeikx dk
2π ∞ z=0 2π ∞
Combining this with Eqs. (D.30), we have
∞
∂h 1 sinh2 kH − (kH)2
=√ Bikeikx dk.
∂t 2π −∞ sinh kH cosh kH + kH
Further, combining this with Eq. (D.33), we obtain

B(t, k) = B(0, k)e−t/τ , (D.35)


2η k(sinh kH cosh kH + kH)
τ= . (D.36)
ρg sinh2 kH − (kH)2
Equation (D.30) designates that A, B, C and D decay with the same time constant τ. Equation
(D.28) warrants that surface undulations decay at the same rate. The time constant (Eq. (D.36)) is
non-dimensionalized as
ρgH kH (sinh kH cosh kH + kH)
τ= τ= . (D.37)
2η sinh2 kH − (kH)2
Figure D.6 designates that the dimensionless time constant is proportional to the wavenumber if the
viscous layer is very thick. Namely, long and short wavelength topographic undulations decay fast
and slow, respectively. The figure also shows that the time constant is proportional to (kH)−2 , i.e.,
long wavelength undulations relax very slowly, whereas short ones decay rapidly.
364 APPENDIX D. ANSWERS TO SELECTED PROBLEMS

Figure D.6: Dimensionless time constant τ for gravitational relaxation of topographic undulations
as a function of the dimensionless thickness of a viscous layer kH.

11.1 Taking the O-1 and -3 axes of rectangular Cartesian coordinates as parallel to the σ1 and σ3
axes, respectively, and assuming 2θ to be the acute angle between the conjugate faults, we have the
unit vectors n(1) = (sin θ, 0, cos θ)T and n(2) = (sin θ, 0, − cos θ)T as the normals to the plane of
conjugate faults (Fig. D.7). The slip directions of these faults are u :(1) = (− cos θ, 0, sin θ)T and
:(2) = (− cos θ, 0, − sin θ)T , respectively. Therefore, the asymmetric moment tensor Ma = u
u :n for
each of the faults is
⎛ ⎞ ⎛ ⎞
− sin θ cos θ 0 − cos2 θ − cos θ sin θ 0 cos2 θ
Ma = ⎝ ⎠ , Ma = ⎝ ⎠.
(1) (2)
0 0 0 0 0 0
2
sin θ 0 cos θ sin θ − sin θ
2
0 cos θ sin θ
Combining these and Eq. (2.54), we have the strains
⎛ ⎞ ⎛ ⎞
M
(1) − sin 2θ 0 − cos 2θ M
(2) − sin 2θ 0 cos 2θ
E(1) =
g
⎝ 0 0 0 ⎠ , E(2) =
g
⎝ 0 0 0 ⎠.
2V 2V
− cos 2θ 0 sin 2θ cos 2θ 0 sin 2θ

The total strain E due to the fault activities is the linear combination of these tensors so that E12 =
E21 = E22 = 0. Therefore, the total strain is a plane strain on the O-13 plane.
From, Eq. (2.55), we have the rotation tensors
⎛ ⎞ ⎛ ⎞
Mg
(1) 0 0 −1 M
(2) 0 0 1
X(1) = ⎝0 0 0 ⎠ , X(2) = g ⎝ 0 0 0⎠ .
2V 2V
1 0 0 −1 0 0

O-123 is assumed to be a right-hand system and Mg is defined to be positive (Eq. (2.51)). There-
fore, comparison of the above equations with Eq. (C.67) indicates that X(1) and X(2) represent the
counterclockwise and clockwise rotations, respectively, in Fig. D.7.
365

Figure D.7: Unit normals for conjugate faults.

11.2 Using principal stress space for simplicity, the stress tensor can be divided into three parts:
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
σ1 0 0 σ2 0 0 σ1 − σ2 0 0 0 0 0
⎝ 0 σ2 0 ⎠ = ⎝ 0 σ2 0 ⎠ + ⎝ 0 0 0 ⎠ + ⎝0 0 0 ⎠, (D.38)
0 0 σ3 0 0 σ2 0 0 0 0 0 σ 3 − σ2
where the three terms in the right-hand side of Eq. (D.38) represent hydrostatic stress, uniaxial
compression, and uniaxial tension (∵ σ3 < σ2 ), respectively. The linearity of Eq. (11.3) with respect
to r indicates that the vectorial sum of the shear stresses resulting from these tensors is equivalent
to that from the stress tensor in the left-hand side of Eq. (D.38). However, the hydrostatic stress
causes no shear traction. If n = (n1 , n2 , n3 )T is the unit normal to the fault plane, the tractions by the
uniaxial compression and tension are
⎛ ⎞⎛ ⎞ ⎛ ⎞
σ1 − σ 2 0 0 n1 n1
⎝ 0 0 0⎠ ⎝n2 ⎠ = (σ1 − σ2 ) ⎝ 0 ⎠
0 0 0 n3 0

and
⎛ ⎞⎛ ⎞ ⎛ ⎞
0 0 0 n1 0
⎝0 0 0 ⎠ ⎝n2 ⎠ = (σ2 − σ3 ) ⎝ 0 ⎠ ,
0 0 σ3 − σ2 n3 −n3
respectively. The implications of these equations are as follows. The remaining uniaxial stresses
have axial symmetry. Traction by the uniaxial compression must lie on the plane containing the
σ1 -axis and the normal to the fault plane that is indicated by the point P in Fig. 11.3. The uniaxial
compression pushes the blocks on the sides of the fault plane, and the blocks go away from the
σ1 -axis. On the other hand, the uniaxial tension attracts the blocks and, consequently, the blocks
approach to the σ3 -axis.
The shear stress due to the uniaxial compression is given by Eq. (4.22):
 (1)  1
s  = (σ1 − σ2 ) sin 2δ1 ,
2
where δ1 is the angle shown in Fig. 11.3. A similar equation is obtained for s(3) . The shear traction
vector caused by the left-hand side of Eq. (D.38) equals s = s(1) + s(3) .
366 APPENDIX D. ANSWERS TO SELECTED PROBLEMS

11.3 The rectangular Cartesian coordinates O-123 is taken as in Fig. D.8. Here, the stress tensor
r is defined in this coordinate system. According to the Wallace-Bott hypothesis, we have
 
σ31 = σ13 = 0. (D.39)

Irrespective of the datum, we consider a stress tensor of the form


⎛ ⎞
σ1 0 0
r = Q · ⎝ 0 σ2 0 ⎠ · Q T
0 0 0
⎛ ⎞
Q11 σ1 + Q212 σ2
2
Q11 σ1 Q21 + Q12 σ2 Q22 Q11 σ1 Q31 + Q12 σ2 Q32
= ⎝Q11 σ1 Q21 + Q12 σ2 Q22 Q221 σ1 + Q222 σ2 Q21 σ1 Q31 + Q22 σ2 Q32 ⎠ , (D.40)
Q11 σ1 Q31 + Q12 σ2 Q32 Q21 σ1 Q31 + Q22 σ2 Q32 Q231 σ1 + Q232 σ2
where 0 ≤ σ2 ≤ σ1 . In addition, differential stress must not vanish for faulting, so that σ1 > 0. The
orthogonal tensor Q may be expressed in terms of successive rotations Q(1) , Q(2) and Q(3) through
angles θ, −φ and ψ about the O-1, -2 and -3 axes, respectively (Fig. D.8). Namely, we have

Q = Q(3) · Q(2) · Q(1)


⎛ ⎞⎛ ⎞⎛ ⎞
cos ψ − sin ψ 0 cos φ 0 sin φ 1 0 0
= ⎝ sin ψ cos ψ 0⎠ ⎝ 0 1 0 ⎠ ⎝0 cos θ sin θ ⎠
0 0 1 − sin φ 0 cos φ 0 − sin θ cos θ
⎛ ⎞
cos ψ cos φ − cos ψ sin φ sin θ − sin ψ cos θ − cos ψ sin φ cos θ + sin ψ sin θ
= ⎝ sin ψ cos φ − sin ψ sin φ sin θ + cos ψ cos θ − cos ψ sin φ cos θ − cos ψ sin θ ⎠ .
sin φ cos φ sin θ cos φ cos θ
(D.41)
Note that the σ1 -orientation is only designated by φ and ψ.
Combining Eqs. (D.39), (D.40) and (D.41), we obtain

σ13 = Q11 σ1 Q31 + Q12 σ2 Q32

= σ1 cos ψ sin φ − σ2 (cos ψ sin φ sin θ + sin ψ cos θ) sin θ cos φ = 0. (D.42)

This equation has three sets of solutions. (1) The first one is φ = 90◦ . However, this is a meaningless
solution because shear stress upon the fault surface vanishes. (2) The second one is σ2 = 0 and
ψ = 90◦ . In this case, the σ1 -axis lies on the O-23 plane and Φ = 0. This is an axial stress so that
θ is meaningless. The shear stress in the O-2 direction should be negative in sign for the sense of
shear matching the assumed one. This requires the inequality
 1
σ32 = σ1 cos φ sin φ = σ1 sin 2φ < 0.
2
Therefore, we obtain φ < 0, indicating that the σ1 -axis of axial compression is oriented on the line
OC in Fig. 11.7. (3) The third solution of Eq. (D.42) is
σ1 tan ψ
= sin2 θ + sin θ cos θ (D.43)
σ2 sin φ
367

Figure D.8: Rectangular Cartesian coordinates O-123 taken to be parallel to the fault plane coordi-
nate system. The O-12 plane is the fault surface, and the slip direction of the upper block (x3 > 0)
is the O-2 direction.

for triaxial stresses. Since the least principal value of y was defined to be zero, Φ = σ2 /σ1 . Intro-
ducing the parameter
2
α = − 1 ≥ 1,
Φ
Eq. (D.43) is rewritten as
 1
tan2 ψ 2
α = 1+ sin(2θ − χ), (D.44)
sin2 φ
where
sin φ
tan χ = .
tan ψ
Since the σ1 -orientation is independent of θ, Eq. (D.44) gives the assessible values of φ and ψ as
tan ψ  2
≥ α − 1. (D.45)
sin φ
Figure 11.7 illustrates this inequality3 .

12.1 The strength profile and strain rate ε̇ 0 at t = 0 is readily obtained from the given lithological
layering of the lithosphere and the applied tectonic force F from the constraint
a
Δσ(ε̇ 0 , z) dz = F.
0

Figure 12.5 shows the profiles just after the commencement of rifting. Accordingly, the Péclet
number is defined in this case as
a2 /κ ε̇ a2
Pe = = 0 .
a/(ε̇ 0 a) κ
Numerical experiments by [236] demonstrate that the fate of rifts is determined with the threshold
Pe ≈ 1.
3 See [137] for details.
368 APPENDIX D. ANSWERS TO SELECTED PROBLEMS

12.2 Substituting Eq. (12.32) into (12.31), we have the horizontal deviatoric stress
 
− 1n 1 Q
Txx = A Ė exp
n .
nRT

Let zM be the depth of the Moho interface, then


a a  
− 1n 1 Q
Txx dz ≈ A Ė n exp dz. (D.46)
0 zM nRT

Since the geothermal gradient is constant, we have

z = (T − T0 )/Γ, (D.47)

where T0 is the temperature of the surface. Substituting Eq. (D.47) into (D.46), we obtain
a  Ta  
A− n Ė 1/n
1
Q
Txx dz ≈ exp dT, (D.48)
0 Γ TM nRT

where TM and Ta are the temperatures at the Moho and the base of the lithosphere at z = a. Using
the exponential integral  ∞ xt
e
Ei(x) = − dt (x > 0),
−x t

Eq. (D.48) is transformed to


a  Q/nRTa
A− n Ė 1/n Q
1
ex
Txx dz ≈ Ei(x) − .
0 ΓnR x Q/nRTM

Using the parameters Q ∼ 105 J mol−1 , n ≈ 3, TM ≈ 1000 K and Ta ≈ 1600 K, we have Q/nRTM ∼
101 and Q/nRTa > Q/nRTM . Hence, the range of integration is beyond 10. The exponential integral
for this range is approximated as
 
ex 1
Ei(x) ≈ 1+ (x > 10).
x x

Therefore, we obtain Eq. (12.4). England and Mckenzie [56] derived the Argand number for the
case where the lithospheric mantle determines the strength of the lithosphere.

B.1 ⎛ ⎞ ⎛ ⎞
uT u1 u2 u3
Q = ⎝ v T ⎠ = ⎝ v1 v2 v3 ⎠ .
wT w1 w2 w3
Index

Activation Body torque, 68


— energy, 263 Bookshelf faulting, 48
— enthalphy, 263 Borehole breakout, 115
— volume, 263 Boudin, 215, 217
Almansi’s finite strain tensor, 17, 32 — line, 215
α-line, 250 Boundary condition, 113, 114
Amazon submarine fan, 172 Boussinesq approximation, 227
Anderson’s theory of faulting, 137 Breakup unconformity, 98
Angle of internal friction, 135 Brittle
Angle of shear, 135 — deformation, 126, 156, 178, 303
Angular stress distance, 328 — ductile transition, 305
Anisotropic, 64, 127, 339, 341, 359 — failure, 126, 305
— tensor, 341 — faulting, 303
Antisymmetric — material, 126
— matrix, 334 — strength, 132, 135, 276, 304, 305
— part, 33, 38, 45 Brittle-ductile transition, 305
— tensor, 35, 337, 348 Buckling, 185
Archimedes’ principle, 75 — load, 188
Argand number, 86, 235 — mode, 188
Asymmetric moment tensor, 53 Euler’s critical — load, 188
Auxiliary plane, 275 Buoyancy, 171, 172, 180, 181
Average stress difference, 287
Axial Caldera, 255
— compression, 109 Cayley-Hamilton theorem, 129, 341, 349
— stress, 125 Central peak crater, 251
— tension, 109 Characteristic equation, 45, 334, 340, 341
— vector, 35, 43–46, 56, 348, 354 Clastic dike, 148–150, 190
Coaxial deformation, 11, 13, 14, 17, 41, 42, 44,
Backstripping technique, 80 353
Balanced cross-section, 25, 27 Coefficient of
Basic flow, 216 — friction, 104, 138
Basin and Range Province, 120, 308 — internal friction, 135
Basin inversion, 141 — linear expansion, 160
β factor, 42, 56, 353 — volume expansion, 161
β-line, 250 Coherence, 199
Biaxial stress, 109 Cohesion, 135
Biharmonic equation, 209, 210 Columnar joint, 161
Bingham fluid, 237, 239 Compaction, 80
Body force, 64, 67, 209, 235 Complementary error function, 90

379
380 INDEX

Complex crater, 251 — deformation, 178, 303, 304


Compressive strength, 134 — strength, 87
Confining pressure, 125, 126 Dummy index, 331, 348, 352
Conjugate Dynamic pressure, 206
— faults, 49, 135 Dynamic topography, 225
— sets, 302
— shear planes, 135 Earth pressure at rest, 158
Constitutive equation, 128, 129, 153, 154, 156, 174, Effective elastic thickness, 178
177, 205, 206, 237–239 — of oceanic lithosphere, 181
Contraction, 339 Effective stress, 139
Cooling joint, 161 Effective viscosity, 260, 261
Coulomb-Navier criterion, 131, 135, 279, 302 Eigenvalue, 334
Couple stress, 68 Eigenvector, 334
Crater Ejecta, 70
central peak —, 251 Elastic, 125
Euler —, 13 — limit, 125, 303
impact —, 12, 212 Elastic core, 247
multi-ring —, 251 Elasticity
peak-ring —, 251 linear —, 125, 153, 156, 174, 177
Cross-section balancing, 25 Elementary orthogonal projector, 338
Curvature, 175–177, 180, 360 Elongation, 3
radius of —, 175, 177, 360 quadratic —, 3, 15
Enceradus, 315, 321
Deformation, 3 Energy equation, 70
— gradient tensor, 7, 9, 10, 12, 14, 31, 32, 40 Engineering shear strain, 4, 5, 18, 20, 28, 34
coaxial —, 11, 13, 14, 17, 41, 42, 44, 353 Epeirogenic movement, 233
ductile —, 304 Epeirogeny, 83, 232
homogeneous —, 10 Equation
infinitesimal —, 10, 31–34, 37, 39, 40, 44, 45 — of motion, 65
zero —, 10 biharmonic—, 209, 210
Degree of compensation, 197 characteristic —, 45, 334, 340, 341
Depth of compensation, 76 constitutive —, 128, 129, 153, 154, 156, 174,
Deviatoric 177, 205, 206, 237–239
— plane, 240 energy —, 70
— stress, 103, 239, 240, 304 Navier-Stokes —, 206, 207
— stress space, 327 Rodrigues’ —, 349, 354
— tensor, 237, 341, 342 Stokes —, 206
Differential stress, 101 Equivalent
Diffusion lenfth, 89 — plastic strain increment, 249
Dike, 116–118 — strain rate, 249
— swarm, 12, 48, 118, 119 — stress, 249
clastic —, 148–150 Equivalent elastic thickness, 178
ring —, 119 Error function, 89
Dilatant fluids, 238 Euler
Dilatation, 15, 34 — angle, 347
Dipole moment of vertical density profile, 85 — angles, 346
Dorn’s equation, 263 — coordinates, 6
Drucker-Prager yield criterion, 246 — description, 6
Ductile Extended Tresca yield criterion, 246
INDEX 381

Extended von Mises yield criterion, 246 — Ridge, 183


— Swell, 225
Fabric attractor, 275 Heat conduction equation, 88
Failure Heat flux, 69
— envelope, 131 Heterogeneous fault-slip data, 281
brittle —, 126, 305 Hinge
Fault — line, 151
— striation, 271 — zone, 246
conjugate —, 49 Homogeneous fault-slip data, 281
mesoscale —, 271 Hooke’s law, 156
orthorhombic — sets, 291 generalized —, 156
Fletcher-Hallet model, 315 Hubbert-Rubey model, 104
Flexural Hydraulic fracture, 115
— backstripping, 83, 173 Hydrofracturing, 115
— parameter, 184 Hydrostatic
— rigidity, 178, 180 — pressure, 64, 108
Flexural-shear folding, 188 — state of stress, 64
Fold, 217 Hypsometry, 77
Foreland basin, 173
Fourier transform, 210 Icy satellite, 136
Fracture Ilyushin’s deviatoric stress space, 327
— plane, 131 Impact crater, 212, 251
— plane angle, 131 In-situ stress measurement, 115
— strength, 132 Incompressible, 9, 208
Free Inelastic deformation, 126
— boundary, 109 Infinitesimal
— index, 331 — deformation, 10, 31–34, 37, 39, 40, 44, 45
— surface, 109 — strain, 31, 33, 34, 153, 154
Frictional — strain tensor, 4, 32–34, 38
— sliding, 138 Initial subsidence, 94
— strength, 138 Invariant, 154, 340
Fry arc, 328 basic —, 34, 129, 153, 205, 237–239, 327,
341, 354
Ganymede, 71, 213 Inverse stress, 287
Gauss’s divergence theorem, 345 Io, 255
Geohistory analysis, 80 Isostasy, 75
Geometric seismic moment, 53 local —, 171
Geometrical linearity, 32, 153 regional —, 172
Graben, 22 thermal —, 87
Grain fabric, 28, 352 Isostatic compensation, 212
Gravitational Isotropic, 127, 129, 339
— acceleration, 70
— collupse, 85 Jacobian, 8, 39, 65
Green’s finite strain tensor, 16, 17, 32 Japan Sea, 24, 25, 52, 81, 116
Griffith theory of fracture, 134 Joint, 147
— set, 148
Haar-von Kármàn hypothesis, 254 — system, 148
Hawaiian columnar —, 161
— Deep, 181, 182 cooling —, 161
382 INDEX

Jump condition, 113 Mechanics of micropolar continua, 301


Mercury, 56, 168
Kinematic viscosity, 207 Mesoscale fault, 271
Kirchhoff’s hypothesis, 176 Micro cracks, 134
Kyushu, 149 Middle surface, 174
Mineral vein, 23
Laglangian Miranda, 321
— description, 6 Mohr
— Coordinates, 6 — diagram, 106
Lamé’s — envelope, 131
— constants, 154 Mohr’s circle, 107
— stress ellipsoid, 102 Moon, 77, 168, 200, 251
Laplacian, 344 M-plane, 275
Layer-parallel shortening, 187 Multi-ring crater, 251
Left Cauchy-Green tensor, 14
Left stretch tensor, 13 Nabla, 344
Linear Natural strain, 3
— elastic body, 126, 127, 153 Navier-Stokes equation, 206, 207
— elastic material, 157, 158 Newtonian fluid, 206, 237
— elasticity, 153, 154, 156, 174, 177 Normal fault regime, 138
— momentum, 59, 65 Normal stress, 60
— rille, 22, 168, 200 Northeast Japan, 22, 39, 52, 116, 273, 308
— stability analysis, 221, 315 Null vector, 282
Lithosphere Number
Effective elastic thickness of —, 181 Argand —, 86, 235
Lithostatic Péclet —, 322, 358, 367
— pressure, 66, 108 Rayleigh —, 231
— stress, 66
Load cast, 190 Object function, 281
Lobate scarp, 56 Oceanic lithosphere
Local isostasy, 76, 171 elastic thickness of —, 181
Lode angle, 241, 328 Octahedral
Logarithmic strain, 3 — normal stress, 106
Longitudinal splitting, 133 — plane, 106
Lunar — shear stress, 106
— highland, 77 Ooid, 28
— mare, 77 Orientale basin, 251
— mare basin, 22 Orthogonal
— maria, 22 — matrix, 333
— tensor, 337
Magma ocean, 168 Orthorhombic
Mare Selenitatis, 22, 201 — fault set, 291
Mascon basin, 201 — symmetry, 8, 103, 107, 109, 276, 289
Material Outer trench swell, 182
— coordinates, 6 Overburden
— derivative, 36 — stress, 66, 114, 138, 140, 163
— description, 6 Overpressure, 140
Mean flow, 216
Mean stress, 103 Péclet number, 322, 358, 367
INDEX 383

Paleobathymetry, 78, 80 — stress, 101


Paleomagnetic rotation, 48 Principle of
Palimpsest, 71, 214 — material frame-indefference, 128
Passive continental margin, 78, 173 — material objectivity, 128
P-axis, 279 Projector, 338
Peak-ring crater, 251 elementary orthogonal —, 338
Peripheral bulge, 172 Proportional limit, 125
Permanent strain, 126 Pseudoplastic fluid, 238
Perturbation, 216 Pure shear, 11
π-plane, 240 — rift, 41
Plane — stress, 109
— strain, 10, 42, 47, 68, 157, 174, 297, 303
— stress, 68, 102, 109, 157 Quadratic elongation, 3, 15
Planetary differentiation, 70 Quantitative stratigraphy, 80
Plastic, 126 Quasi-plasticity, 305
— body, 237 Quasilinear fluid, 237
— deformation, 126, 156
pseudo— fluid, 238 Radius of curvature, 175, 177, 360
Rayleigh number, 231
rigid-perfectly — body, 238
Reciprocal strain ellipsoid, 8, 16
rigid-perfectly— body, 127
Reduced stress tensor, 277, 278, 287, 327
visco— body, 237
Reference configuration, 38, 40
Platform, 233
Regional isostasy, 172
Poisson effect, 156, 157
Reiner-Rivlin fluid, 205
Poisson’s ratio, 154
Representation theorem, 129
Polar decomposition, 19, 343
Retrodeformable, 27
— theorem, 12, 343
Reverse fault regime, 138
left —, 343
Ridge push, 91
right —, 343
Rift
Polygonal terrain, 163 — shoulder, 192
Pore — stage subsidence, 92
— fluid, 139 simple shear —, 41, 192
— fluid pressure, 139 Right Cauchy-Green tensor, 14, 16
— fluid pressure ratio, 140 Right stretch tensor, 13
— pressure, 139 Rigid body rotation, 10
— water, 139 Rigid-perfectly-plastic body, 127, 238
Post-glacial rebound, 212 Rodrigues’ equation, 349, 354
Post-rift subsidence, 92, 95 Rotation
Potential energy of the lithosphere, 92 — tensor, 33, 364
Power-law — vector, 35
— exponent, 260
— fluid, 237, 260 Second law of thermodynamics, 237
Primary structure, 12 Shear
Principal — force, 178, 179
— axes of stress, 101 — fracture, 134
— axis, 7 — modulus, 154, 361
— deviatoric stress, 103 — strain, 4, 18, 20, 28, 34, 154
— planes of strain, 8 — stress, 60, 64, 72, 98, 356
— strain, 7 — stress vector, 357
384 INDEX

— zone, 18 — inversion, 281, 327


simple —, 11, 18–21, 28, 238, 239 — ratio, 103
Shear-rate dependent visocity, 237 — space, 239
Shear-thickening viscosity, 238 — tensor, 60
Shear-thinning viscosity, 238 — vector, 59
Simple crater, 251 axial —, 125
Simple shear rift, 41, 192 biaxial —, 109
Simple stretching model, 92 boundary condition of —, 113
Single layer folding, 217, 219–222, 224, 225, 268 classes of —, 108
Slip line, 250 couple —, 68
— field, 250 deviatoric —, 103, 239, 240
— theory, 248 differential —, 101
Slip tendency, 288 effective —, 139
Southwest Japan, 23, 118, 119 equivalent —, 249
Spatial hydrostatic —, 64
— coordinates, 6 hydrostatic state of —, 64
— description, 6 inverse —, 287
Spin tensor, 38 lithostatic —, 66
Stokes equation, 206 mean —, 103
Stokes’ theorem, 346 normal —, 60, 63
Stokesian fluid, 206 plane —, 68, 102, 109, 157
Strain, 3 principal —, 101
— ellipse, 8 principal deviatoric —, 103
— ellipsoid, 7 pure shear —, 109
— incremental theory, 248 shear —, 60, 63
— marker, 4 tectonic —, 103
— rate, 37 thermal —, 160
— rate tensor, 38 triaxial —, 109
definition of —, 3 yield —, 125
engineering shear —, 4, 5, 18, 20, 28, 34 Stretch, 3
infiniteismal —, 33, 34, 153, 154 Stretching
infiniteismal — tensor, 4, 33, 34, 38 — factor, 94
infinitesimal —, 31 — tensor, 38
logarithmic —, 3 Striae, 271
natural —, 3 Strike-slip regime, 138
permanent —, 126 Subsidence
plane —, 10, 157 — curve, 83
reciprocal — ellipsoid, 16 — due to mantle convection, 225
reciprocal — ellipsoid, 8 — due to sediment load, 78
shear —, 4, 18 — of foreland basin, 173
Stratigraphy, 23 initial —, 94
quantitative —, 80 post-rift —, 92, 95
Stream function, 208 rift-stage —, 92
Strength profile, 305 syn-rift, 95
Stress syn-rift —, 93
— axis, 101 thermal —, 94, 98
— concentration, 134 Suface force, 64
— difference, 286, 328 Symmetric
INDEX 385

— matrix, 334 — diffusivity, 88


— tensor, 337 — expansion coefficient, 89, 161
Syn-rift subsidence, 93, 95 — isostasy, 87
— stress, 160
Tangent-lineation diagram, 275 — subsidence, 94, 98
T-axis, 279 Thrust sheet, 104
Tectonic Tibetan Plateau, 86
— age, 199 Time constant for thermal conduction, 88
— inversion, 141 Topographic load, 172
— stress, 103 Transient crater, 252
Tectonics, v Transpression, 50
Tensile strength, 132 Transtension, 50
Tension fracture, 133 Tresca yield criterion, 243
Tensor, 337 Triaxial stress, 109
— of elastic constants, 156 Triple
Almansi’s finite strain —, 17, 32 — scalar product, 331
anisotropic —, 341 — vector product, 331
antisymmetric —, 35, 337, 348
asymmetric moment —, 53 Uncoupled thermoelastic theory, 161
basic invariants of —, 340 Uniaxial
contraction of —, 339 — compression, 109
deformation gradient —, 7, 9, 10, 12, 14, 31, — tension, 109
32, 40 Unit tensor, 337
deviatoric —, 237, 341, 342
Green’s finite strain —, 16, 17, 32 Vector
infinitesimal strain —, 4, 32–34, 38 — product, 329
left Cauchy-Green —, 14 axial —, 348
left stretch —, 13 scalar product of —, 329
orthogonal —, 337 stress —, 59
reduced stress —, 287 triple — product, 331
right Cauchy-Green —, 14, 16 Velocity gradient tensor, 38
right stretch —, 13 Venus, 163
rotation —, 33, 364 Viscoplastic body, 237
spin —, 38 Viscosity, 206
strain rate —, 38 dynamic —, 207
stress —, 60 effective —, 260, 261
stretching —, 38 kinematic —, 207
symmetric —, 337 shear-rate dependent —, 237
unit —, 337 shear-thickening —, 238
velocity gradient —, 38 shear-thinning —, 238
zero —, 337 von Mises yield criterion, 244
Terrain softening, 215 Vorticity, 44, 48, 56, 206, 207, 209, 235, 301
Theorem
Cayley-Hamilton —, 341 Wallace-Bott hypothesis, 273
Gauss’s divergence —, 345 Wavenumber, 189
polar decomposition —, 12, 343 — selection, 189
representation —, 129 Western Interior Basin, 173, 233
Stokes’ —, 346 Whole lithosphere failure, 248, 307
Thermal Wrinkle ridge, 22, 168, 200
386 INDEX

Yield
— condition, 242
— function, 242
— locus, 244
— point, 125
— stress, 125, 239
— surface, 242
Drucker-Prager — criterion, 246
extended — Tresca criterion, 246
extended — von Mises criterion, 246
Tresca — criterion, 243
von Mises — criterion, 244
Young’s modulus, 154, 156

Zero
— deformation, 10
— tensor, 337
References

[1] Agterberg, F. P. and Gradstein, F. M., 1988. Earth Sci. Rev., 25, 1–73.
[2] Allaby, A. and Allaby, M. (eds.), 1999. A dictionary of Earth sciences. Oxford Univ. Press, Oxford,
619p.
[3] Amadei, B. and Stephansson, O., 1997. Rock stress and its measurement. Chapman and Hall, London,
490p.
[4] Anderson, E. M., 1905. Trans. Edinburgh Geol. Soc., 8, 387–402.
[5] Anderson, F. C. and Smrekar, S. E., 1999. J. Geophys. Res., 104, 30743–30756.
[6] Angelier, J., 1979. Tectonophys., 56, T17–T26.
[7] Angelier, J., 1984. J. Geophys. Res., 89, 5835–5848.
[8] Angelier, J., 1989. J. Struct. Geol., 11, 37–50.
[9] Angelier, J., 1994. In Hancock, P. L., ed., Continental deformation, Pergamon Press, Oxford, p. 53–101.
[10] Angelier, J., Barrier, E., Hao, T. C., 1986. Tectonophys., 125, 161–178.
[11] Aono, H. and Masuda, F., 1989. In Taira, A. and Masuda, F., eds., Sedimentary facies in the active plate
margin, Terra Sci. Publ., Tokyo, p. 143–149.
[12] Armijo, R., Tapponnier, P., Mercier, J.L. and Han, T.L., 1986. J. Geophys. Res., 91, 13803–13872.
[13] Badley, M. E., Price, J. D. and Backshall, L. C., 1989. Geol. Soc. Lond. Spec. Publ., 44, 201–219.
[14] Baer, G., Beyth, M. and Reches, Z., 1994. J. Geophys. Res., 99, 24039–24050.
[15] Barrell, J., 1914. J. Geol., 22, 729–741.
[16] Bassi, G. and Bonnin, J., 1988. Geophys. J. Int., 93, 485–504.
[17] Bassi, G. and Bonnin, J., 1988. Geophys. J. Int., 94, 559–565.
[18] Bechtel, T. D., Forsyth, D. W., Sharpton, V. L. and Grieve, R. A. F., 1990. Nature, 343, 636–638.
[19] Beeman, M., Durham, W. B. and Kirby, S. H., 1988. J. Geophys. Res., 93, 7625–7633.
[20] Bott, M. H. P., 1959. Geol. Mag., 96, 109–117.
[21] Brace, W. F. and Kohlstedt, D. L., 1980. J. Geophys. Res., 85, 6248–6252.
[22] Brereton, R. and Müller, B., 1991. Phil. Trans. R. Soc. Lond. A, 337, 165–179.
[23] Brotchie, J. F., 1971. Modern Geol., 3, 15–23.
[24] Brown, E. T. and Hoek, E., 1978. Int. J. Rock Mech. Min. Sci. Geomech. Abstr., 15, 211–215.
[25] Buck, W. R., Lavier, L. L. and Poliakov, A. N., 1999. Phil. Trans. R. Soc. Lond. A, 357, 671–393.

369
370 REFERENCES

[26] Burgess, P. M., Gurnis, M. and Moresi, L., 1997. Geol. Soc. Am. Bull., 109, 1515–1535.
[27] Burov, E. B., Jaupart, C. and Mareschal, J. C., 1998. Earth Planet. Sci. Lett., 164, 205–219.
[28] Burov, E. and Diament, M., 1995. J. Geophys. Res., 100, 3905–3927.
[29] Buck, W. R. and Sokoutis, D., 1994. Nature, 369, 737–740.
[30] Bulin, N. K., 1981. Geotectonics, 3, 133–139.
[31] Bullard, D. H., 1981. Pattern Recognition, 13, 111–122.
[32] Butler, R.F., 1992. Paleomagnetism: magnetic domains to geologic terrains. Blackwell Scientific Pub-
lisher, Boston, 319p.
[33] Byerlee, J. D., 1978. Pure Appl. Geophys., 116, 615–626.
[34] Cande, S. C., Stock, J. M., Muller, R. D. and Ishihara, T., 2000. Nature, 404, 145–150.
[35] Carey, E. and Brunier, B., 1974. C. R. Acad. Sci. Paris, ser. D, 279, 891–894.
[36] Carslaw, H. S. and Jaeger, J. C., 1959. Conduction of heat in solids, 2nd ed. Clarendon Press, Oxford,
510p.
[37] Carter, N. L. and Tsenn, M. C., 1987. Tectonophys., 136, 27–63.
[38] Chadwick, P., 1999. Continuum mechanics, concise theory and problems. Dover, Mineola, 187p.
[39] Choi, P.-Y.,Kwon, S. K., Hwang, J. H., Lee, S. R. and An, G. O., 2001. block rotation. Geosci. J., 5,
1–18.
[40] Clow, G. D. and Carr, M. H., 1980. Icarus, 44, 268–279.
[41] Cobbold, P. R., Brun, J. P., Davy, P., Fiquet, G., Basile, C.and Gapais, D., 1989. In Kissel, C. and
Laj, C., eds., Paleomagnetic rotations and continental deformation, Kluwer Academic Publ., Dordrecht,
145–155.
[42] Coblentz, D. D., Richardson, R. M. and Sandiford, M., 1994. Tectonics, 13, 929–945.
[43] Cochran, J. A. et al., 1989. Proc. Ocean Drilling Program, Part A, Init. Rep., 116, 3–11.
[44] Collins, G. C., Head, J. W., III and Pappalardo, R. T., 1998. Geophys. Res. Lett., 25, 233–236.
[45] Cox, A. D., Eason, G. and Hopkins, H. G., 1961. Phil. Trans. Roy. Soc. Lond., 254A, 1–45.
[46] Cross, T. A. and Pilger, R. H., Jr., 1978. Nature, 274, 653–657.
[47] Dahlen, F. A., 1981. J. Geophys. Res., 86, 7801–7807.
[48] Dahlstrom, C.D.A., 1969. Can. J. Earth Sci., 6, 743–757.
[49] Davies, G. F., 1999. Dynamic Earth: plates, plumes and mantle convection. Cambridge Univ. Press, New
York, 458p.
[50] Doghri, I., 2000. Mechanics of deformable solids: linear, nonlinear, analytical and computational
saspects. Springer-Verlag, Berlin, 579p.
[51] Driscoll, N. W. and Karner, G. D., 1994. Geology, 22, 1015–1018.
[52] Duda, R. O., Hart, P. E. and Stork, D. G., 2001. Pattern classification, 2nd ed. Wiley, New York, 654p.
[53] Durham, W. B., Kirby, S. H. and Stern, L. A., 1997. J. Geophys. Res., 102, 16293–16302.
[54] Engelder, T., 1993. Stress regimes in the lithosphere. Princeton Univ. Press, Princeton, 457p.
[55] England, P., 1983. J. Geophys. Res., 88, 1145–1152.
[56] England, P. and McKenzie, D., 1982. Geophys. J. R. astr. Soc., 70, 295–321.
REFERENCES 371

[57] Fletcher, R. C., 1974. Am. J. Sci., 274, 1029–1043.


[58] Eringen, A. C., 1998. Microcontinuum field theories. 1: Foundations and solids. Springer-Verlag, New
York, 360p.
[59] Falvey, D. A., 1974. Austral. Petrol. Explor. Assoc. J., 14, 95–106.
[60] Fry, N., 1999. J. Struct. Geol., 21, 7–21.
[61] Fung, Y.C., 1965. Foundations of solid mechanics. Prentice-Hall, Englewood Cliffs, 525p.
[62] Fink, J. H. and Fletcher, R. C., 1978. J. Volcanol. Geotherm. Res., 4, 151–170.
[63] Fletcher, R. C. and Hallet, B., 1983. J. Geophys. Res., 88, 7457–7466.
[64] Forsyth, D. W., 1985. J. Geophys. Res., 90, 12623–12632.
[65] Furbish, D. J., 1997. Fluid physics in geology. An introduction to fluid motions on Earth’s surface and
within its crust. Oxford Univ. Press, New York, 476p.
[66] Garzione, C. N., Quade, J., DeCelles, P. G. and English, N. B., 2000. Earth Planet. Sci. Lett., 183,
215–229.
[67] Ghosh, S. K. and Ramberg, H., 1976. Tectonophys., 34, 1–70.
[68] Goff, D. F., Wiltschko, D. V. and Fletcher, R. C., 1996. J. Geophys. Res., 101, 11341–11352.
[69] Goldstein, H., Poole, C.P., Jr. and Safko, J.L., 2002. Classical mechanics, 3rd ed. Addison Wesley, San
Francisco, 638p.
[70] Grollimund, B. and Zoback, M. D., 2000. Tectonophys., 327, 61–81.
[71] Handin, J., Hager, R. V., Jr., Friedman, M. and Feather, J. N., 1963. Am. Assoc. Petrol. Geol. Bull., 47,
728–755.
[72] Hartley, R., Watts, A. B. and Fairhead, J. D., 1996. Earth Planet. Sci. Lett., 137, 1–18.
[73] Hayashida, A., 1994. Chikyu, 16, 135–138† .
[74] Hendrie, D. B., Kusznir, N. J. and Hunter, R. H., 1993. Earth Planet. Sci. Lett., 116, 113–127.
[75] Herget, G., 1986. Proc. Int. Symp. on Rock Stress and Rock Stress Measurements, p. 61–68.
[76] Herrick, D. L. and Stevenson, D. J., 1990. Icarus, 85, 191–204.
[77] Herrick, R. R., Sharpton, V. L., Malin, M. C., Lyons, S. N. and Feely, K., 1997. In Bougher, S. W.,
Hunten, D. M. and Phillips, R. J., eds., Venus II: geology, geophysics, atmosphere, and solar wind
environment, Arizona Univ. Press, Tucson, p. 1015–1046.
[78] Hess, H. H., 1962. Buddington Volume, Geol. Soc. Am., New York, p. 599–620.
[79] Hirooka, K. et al., 1986. and Takeuchi, A., 1986. J. Geomag. Geoelectr., 38, 311–323.
[80] Hodgson, J. H., 1957. Bull. Geol. Soc. Am., 68, 611–644.
[81] Honda, S., 1985. Tectonophys., 112, 69–102.
[82] Hopper, J. R. and Buck, W. R., 1993. J. Geophys. Res., 98, 16213–16221.
[83] Hornafius, J. S., Luyendyk, B. P., Teres, R. P. and Kamerling, M. J., 1986. Geol. Soc. Am. Bull., 97,
1476–1487.
[84] Hoshi, H. and Matsubara, T., 1998. Earth, Planets Space, 50, 23–33.
[85] Hoshi, H. and Takahashi, M., 1997. J. Geol. Soc. Japan, 103, 523–542‡ .
[86] Hubbert, M. K. and Rubey, W. W., 1959. Bull. Geol. Soc. Am., 70, 115–116.
[87] Hubbert, M. K. and Rubey, W. W., 1961. Bull. Geol. Soc. Am., 72, 1587–1594.
372 REFERENCES

[88] Ismail-Zadeh, A. T., 1998. J. Geodynamics, 26, 69–83.


[89] Jackson, J. and McKenzie, D., 1988. Geophys. J. Int., 93, 45–73.
[90] Jackson, M. D. and Pollard, D. P., 1988. J. Geol. Soc. Am., 100, 117–139.
[91] Jaeger, J. C. and Cook, N. G. W., 1976. Fundamentals of rock mechanics. Chapan and Hall, London,
585p.
[92] Jarvis, G.T. and McKenzie, D., 1980. Earth Planet. Sci. Lett., 48, 42–52.
[93] Johnson, C. L. and Sandwell, D. T., 1992. J. Geophys. Res., 97, 13601–13610.
[94] Kano, K. and Murata, A., 1998. Structural geology. Asakura, Tokyo† .
[95] Kakimi, T., 1974. J. Geol. Soc. Japan, 81, 39–51‡ .
[96] Kakimi, T. 1978. The analysis of geologic structures. Association for the Geological Collaboration in
Japan, Tokyo, 240p† .
[97] Karato, S.-I., 2003. The dynamic structure of the deep Earth: an interdisciplinary approach. Princeton
Univ. Press, Princeton, 241p.
[98] Karato, S. and Paterson, M. S. and Fitzgerald, J. D., 1986. J. Geophys. Res., 91, 8151–8176.
[99] Kirby, S. H., 1983. Rev. Geophys. Space Phys., 21, 1458–1487.
[100] Kitamura, N., 1963. Kaseki, 5, 124–137† .
[101] Klitgord, K. D., Hutchinson, D. R. and Schouten, H., 1988. In Sheridan, R. E. and Grow, J. A., eds., The
Geology of North America, v. I-2, The Atlantic continental margin, U.S., Geol. Soc. Am., Boulder, p.
19–55.
[102] Kobayashi, Y., 1979. Bull. Volcanol. Soc. Japan, 24, 203-212.
[103] Kominz, M. A., Miller, K. G. and Browning, J. V., 1998. Geology, 26, 311-314.
[104] Krantz, R. W., 1988. J. Struct. Geol., 10, 225–237.
[105] Kudo, T. et al., 2001. Island Arc, 10, 135–144.
[106] Kudo, T. and Yamaoka, K., 2003. Tectonophys., 364, 203–217.
[107] Kushiro, I., 1987. In Mysen, B. O., ed., Magmatic processes: physicochemical principles, Spec. Publ.
Geochem. Soc. University Park, p. 165–181.
[108] Kusuhashi, N. and Yamaji, A., 2001. J. Geol. Soc. Japan, 107, 26–40‡ .
[109] Kusznier, N. J. and Park, R. G., 1987. In Corward, M. P. et al., eds., Continental extensional tectonics,
Geol. Soc. London, p. 35–52.
[110] Kusznier, N. J. and Ziegler, P. A., 1992. Tectonophys., 215, 117–131.
[111] Lamb, S. H., 1987. Earth Planet. Sci. Lett., 84, 75–86.
[112] Lambeck, K. and Johnston, P., 1998. In Jackson, I. N. S., ed., The Earth’s mantle: compsotion, structrure
and evolution, Cambridge Univ. Press, Cambridge, p. 461–502.
[113] Lambeck, K., Johnston, P. and Nakada, M., 1990. Geophys. J. Int., 103, 451–468.
[114] Lambeck, K. and Nakada, M., 1990. Palaeogeogr. Palaeoclimantol. Palaeoecol., 89, 143–176.
[115] Lachenbruch, A. H. and Sass, J. H., 1980. J. Geophys. Res., 85, 6185-6222.
[116] Lama, R. D. and Vutukuri, V. S., 1978. Handbook on mechanical properties of rocks, Trans. Tech Publ.,
Clausthal, Germany, 481p.
REFERENCES 373

[117] Landau, D. L. and Lifshitz, E. M., 1976. Course in theoretical physics vol 1: mechanics, 3rd ed. Perga-
mon Press, Oxford, 170p.
[118] Lawn, B. R. and Wilshaw, T. R., 1975. Fracture of brittle solids, Cambridge Univ. Press, Cambridge,
204p.
[119] Lewis, J. S., 1995. Physics and chemistry of the Solar System. Academic Press, San Diego, 556p.
[120] Lisle, R.J., 1977b. Tectonophysics, 39, 381–395.
[121] Lisle, R.J., 1985. Geological strain analysis: a manual for the Rf /φ method. Pergamon Press, Oxford.
[122] Lisle, R. J., 1987. Ann. Tectonicae, 1, 155–158.
[123] Lliboutry, L. A., 1987. Very slow flows of solids: basics of modeling in geophysics and glaciology.
Martins Nijhoff Publ., Dordrecht, 510p.
[124] Lobkovsky, L. I. et al., 1996. Tectonophys., 266, 251–285.
[125] Lockner, D. A., 1995. In Ahrens, T. J., ed., Rock physics and phase relations: a handbook of physical
constants, AGU reference shelf 3, Am. Gephys. Union, Washington, D. C., p. 127–147.
[126] Lowry, A. R. and Smith, R. B., 1995. J. Geophys. Res., 100, 17947–17963.
[127] Lucchitta, B. K. and Watkins, J. A., 1978. Proc. Lunar Sci. Conf. 7th, p. 3459–3472.
[128] Łukaszewicz, G., 1999. Micropolar fluids: theory and applications. Brikhäuser, Boston, 253p.
[129] MacDonald, G. J. F., 1960. Planet. Space Sci., 2, 249–255.
[130] Mackwell, S. J., Zimmerman, M. E. and Kohlstedt, D. L., 1998. J. Geophys. Res., 103, 975–984.
[131] Malone, S. D., Rothe, G. H. and Smith, S. W., 1975. Seismol. Soc. Am. Bull., 65, 855–864.
[132] Malvern, L. E., 1969. Introduction to the mechanics of a continuous medium. Prentice-Hall, 713p.
[133] Marrett, R. and Allmendinger, R. W., 1990. J. Struct. Geol., 12, 973–986.
[134] Mase, G. E., 1970. Schaum’s outline of theory and problems of continuum mechanics. McGraw-Hill,
New York, 221p.
[135] Matson, D. L, Ransford, G. A. and Johnson, T. V., 1981. J. Geophys. Res., 86, 1664–1672.
[136] McAdoo, D. and Sandwell, D. T., 1985. J. Geophys. Res., 90, 8563–8569.
[137] McKenzie, D., 1969. Bull. Seismol. Soc. Am., 59, 591–601.
[138] McKenzie, D., 1978. Earth Planet. Sci. Lett., 40, 25–32,
[139] McKenzie, D. and Jackson, J., 1989. In Kissel, C. and Laj, C., eds., Paleomagnetic rotations and conti-
nental deformation, Kluwer Academic Press, Dordrecht, p. 17–31.
[140] McNutt, M. K., Diament, M. and Kogan, M. G., 1988. J. Geophys. Res., 93, 8825–8838.
[141] McNutt, M. K. and Menard, H. W., 1982. Geophys. J. Int., 71, 363–394.
[142] Means, W. D., 1989. J. Struct. Geol., 11, 625–627.
[143] Melosh, H. J., 1977. In Roddy, D. J., Pepin, R. O. and Merrill, R. B., eds., Impact and explosion cratering,
Pergamon Press, New York, p. 1245–1250.
[144] Melosh, H. J., 1982, J. Geophys. Res., 87, 371–380.
[145] Melosh, H. J., 1989. Impact cratering: a geologic process. Oxford Univ. Press, New York, 245p.
[146] Melosh, H. J. and Ivanov, B. A., 1999. Ann. Rev. Earth Planet. Sci., 27, 385–415.
[147] Melosh, H. J. and McKinnon, W. B., 1988. In Vilas, F., Chapman, C. R. and Matthews, M. S., eds.,
Mercury, Arizona Univ. Press, Tucson, p. 347–400.
374 REFERENCES

[148] Meyer, C. D., 2000. Matrix analysis and applied linear algebra. SIAM, Philadelphia, 718p.
[149] Menke, W., 1989. Geophysical data analysis: descrete inverse thoery, revised edition. Academic Press,
San Diego, 289p.
[150] Michael, A. J., 1984. J. Geophys. Res., 89, 11517–11526.
[151] Mino, K. and Yamaji, A., 1999. J. Geol. Soc. Japan, 105, 574–584‡ .
[152] Mitrovica, J. X., Beaumont, C. and Jarvis, G. T., 1989. Tectonics, 8, 1079–1094.
[153] Mitrovica, J. X. et al., 1996. J. Geodynamics, 22, 79–96.
[154] Mogi, K., 1972. Tectonophys., 13, 541–568.
[155] Molnar, P., 1983. J. Geophys. Res., 88, 6430–6432.
[156] Mörner, N.-A., 1980. In Mörner, N.-A., ed., Earth rheologhy, isostasy and eustasy, John Wiley and Sons,
New York, p. 251–284.
[157] Morris, A., Ferrill, D. A. and Henderson, D. B., 1996. Geology, 24, 275–278.
[158] Morrow, C., Moore, D. and Byerlee, J., 1983. In Evans, B. and Wong, T.-F., eds., Fault mechanics and
transport properties of rocks, Academic Press, London, p. 69–88.
[159] Murase, T. and McBirney, A. R., 1970, Science, 167, 1491–1493.
[160] Nakane, K., 1973. J. Geod. Soc. Japan, 19, 190–199, 200–208‡ .
[161] Nemcock, M. and Lisle, R. J., 1995. J. Struct. Geol., 17, 1445–1453.
[162] Oda, H., Torii, M. and Hayashida, A., 1989. Rock Mag. Paleogeophys., 16, 51–56.
[163] Oertel, G., 1965. Tectonophys., 2, 343–393.
[164] Orife, T., Lisle, R.J., 2003. J. Struct. Geol., 25, 949–957.
[165] Otofuji, Y., Matsuda, T. and Nohda, S., 1985. Earth Planet. Sci. Lett., 75, 265–277.
[166] Otofuji, Y. et al., 1994. Earth Planet. Sci. Lett., 121, 503-518.
[167] Otsubo, M., Sato, K. and Yamaji, A., 2006. J. Struct. Geol., 28, 991–997.
[168] Otsubo, M. and Yamaji, A., 2006. Compt. Geosci., 32, 1221–1227.
[169] Pang, M. and Nummedal, D., 1995. Geology, 23, 173–176.
[170] Passchier, C. W. and Trouw, R. A. J., 1996. Microtectonics, Springer-Verlag, Berlin, 289p.
[171] Pappalardo, R. T. et al., 1998. Icarus, 135, 276–302.
[172] Pappalardo, R. T., Reynolds, S. J. and Greeley, R., 1997. J. Geophys. Res., 102, 13369–13379.
[173] Parsons, B. and Daly, S., 1983. J. Geophys. Res., 88, 1129–1144.
[174] Parsons, B. and Sclater, J. G., 1977. J. Geophys. Res., 82, 803–827.
[175] Passchier, C. W. and Trouw, R. A. J., 1996. Microtectonics. Springer-Verlag, Berlin, 289p.
[176] Patterson, M.S., 1978. Experimental rock deformation: the brittle field. Springer-Verlag, Berlin, 254p.
[177] Pearce, S. J. and Melosh, H. J., 1986. Geophys. Res. Lett., 13, 1419–1422.
[178] Peltzer, G., Crampé, F. and King, G., 1999. Science, 286, 272–276.
[179] Percival, J. A. and Berry, M. J., 1987. In Fuchs, K. and Froidevaux, C., eds., Composition, structure and
dynamics of the lithosphere-asthenosphere system, Am. Geophys. Union, Washington, D. C., p. 33–59.
[180] Peska, P. and Zoback, M. D., 1995. J. Geophys. Res., 100, 12791–12811.
[181] Petit, J. P., 1987. J. Struct. Geol., 9, 597–608.
REFERENCES 375

[182] Pike, R. J., 1974. Geophys. Res. Lett., 1, 291–294.


[183] Pike, R. J., 1977. In Roddy, D. J., Pepin, R. O. and Merrill, R. B., eds., Impact and explosion cratering,
Pergamon Press, New York, p. 489–509.
[184] Pollard, D. D. and Fletcher, R. C., 2005. Fundamentals of structural geology. Cambridge University
Press, Cambridge.
[185] Pollard, D. D. and Johnson, A. M., 1973. Tectonophys., 18, 311–354.
[186] Price, N. J. and Cosgrove, J. W., 1980. Analysis of Geological Structures. Cambridge Univ. Press, Cam-
bridge, 502p.
[187] Pysklywec, R. N. and Mitrovica, J. X., 1997. Earth Planet. Sci. Lett., 148, 447–455.
[188] Ramsay, J. G., 1967. Folding and fracturing of rocks. McGraw-Hill, New York, 568p.
[189] Ramsay, J. G. and Huber, M. I., 1983. The techniques of modern structural geology, volume 1. Academic
Press, New York, 307p.
[190] Ramsay, J. G. and Lisle, R. J., 2000. The techniques of modern structural geology, volume 3. Academic
Press, New York, 608p.
[191] Ranalli, G., 1994. Tectonophys., 240, 107–114.
[192] Ranalli, G. and Murphy, D. C., 1987. Tectonophys., 132, 281–296.
[193] Reches, Z., 1978. Tectonophys., 47, 109–129.
[194] Reches, Z., 1983. Tectonophys., 95, 133–156.
[195] Reches, Z. and Deterich, J. H., 1983. Tectonophys., 95, 111–132.
[196] Richard, Y. and Froidevaux, C., 1986. J. Geophys. Res., 91, 8314–8324.
[197] Rice, J. R., 1992. In B. Evans and T.-F. Wong, eds., Fault mechanics and transport properties of rocks,
Academic Press, San Diego, p. 475–503.
[198] Robinson, E. M. and Parsons, B., 1988. J. Geophys. Res., 93, 3469–3479.
[199] Ron, H. and Nur, A., 1996. Geology, 24, 973–976.
[200] Sandwell, D. and Schubert, G., 1982. J. Geophys. Res., 87, 4657–4667.
[201] Sato, H., 1989. Mem. Geol. Soc. Japan, 32, 257–268‡ .
[202] Sato, H. and Amano, K., 1992. Sedimentary Geol., 74, 323–343.
[203] Sato, K. and Yamaji, A., 2006. J. Struct. Geol, 28, 957–971.
[204] Schelling, D. D., 1999. Geol. Soc. Am. Spec. Pap., n. 328, 287–302.
[205] Schenk, P. M., 1989. J. Geophys. Res., 94, 3813–3832.
[206] Schenk, P. M., 1992. Lunar Planet. Sci. 8th, p. 1215–1216.
[207] Scholtz, C. H., 2000. Geology, 28, 163–166.
[208] Schubert, G., Ross, M. N., Stevenson, D. J. and Spohn, T., 1988. In Vilas, F., Chapman, C. R. and
Matthews, M. S., eds., Mercury, Arizona Univ. Press, Tucson, p. 429–460.
[209] Shan, Y., Suen, H. and Lin, G., 2003. J. Struct. Geol., 25, 829–840.
[210] Shaw, H. R. and Swanson, D. A., 1970. In Gilmour, E. H. and Stradling, D. eds., Proc. 2nd Columbia
River Basalt Symposium, Eastern Washsinto State College, Cheney, p. 271–299.
[211] Sherwin, J. A. and Chapple, W. M., 1968. Am. J. Sci., 266, 167–179.
376 REFERENCES

[212] Shimamoto, T., 1977. Texas Eng. Exper. Station Rep., n. 3275-1, Texas A&M Univ., College Station,
122p.
[213] Shimamoto, T. and Logan, J. M., 1986. Geophys. Monogr., Am. Geophys. Union, 37, 49–63.
[214] Shimazu, Y., 1967. The evolution of the Earth. Iwanami, Tokyo, 359p† .
[215] Smith, R. B., 1975. Bull. Geol. Soc. Am., 86, 1601–1609.
[216] Smith, R. B., 1977. Bull. Geol. Soc. Am., 88, 312–320.
[217] Solomon, S. C., 1977. Phys. Earth Planet. Int., 15, 135–145.
[218] Solomon, S. C., 1986. In Hartman, W. K., Phillips, R. J. and Taylor, G. J., Origin of the Moon, Lunar
and Planetary Institute, Houston, p. 435–452.
[219] Solomon, S. C. and Chaiken, J., 1976. Proc. Lunar Sci. Conf. 7th, p. 3229–3243.
[220] Solomon, S. C. and Head, J. W., III, 1979. J. Geophys. Res., 84, 1667–1682.
[221] Solomon, S. C. and Head, J. W., III, 1979. Rev. Geophys. Space Phys., 18, 107–141.
[222] Sonder, L. J. et al., 1994. Tectonics, 13, 769–788.
[223] Spiegel, M. R., 1968. Mathematical handbook of formulas and tables. McGraw-Hill Inc., New York,
271p.
[224] Spudich, P. A., Guatteri, M. G., Otsuki, K. and Minagawa, J., 1998. Bull. Seismol. Soc. Am., 88, 413–427.
[225] Squyres, S. W., 1981. Icarus, 46, 156–168.
[226] Squyres, S. W., 1989. Icarus, 79, 229–288.
[227] Steckler, M. S., 1985. Nature, 317, 135–139.
[228] Stein, C. A. and Stein, S., 1992. Nature, 359, 123-129.
[229] Stephansson, O., 1988. Bull. Geol. Inst. Univ. Uppsla, 14, 39–48.
[230] Stern, T. A. and Ten Brink, U. S., 1989. J. Geophys. Res., 94, 10315–10330.
[231] Strehlau, J. and Meissner, R., 1987. In Fuch, S. and Froidevaux, C., eds., Composition, structure and
dynamics of the lithosphere-asthenosphere system, Am. Geophys. Union, Washinton, D. C., p. 69–87.
[232] Strom, R. G., Tras, N. J. and Guest, J. E., 1975. J. Geophys. Res., 80, 2478–2507.
[233] Stüwe, K., 2002. Geodynamics of the lithosphere: an introduction. Springer-Verlag, Berlin, 449p.
[234] Swarbrick, R. E. and Osborne, M. J., 1998. Am. Assoc. Petrol. Geol. Mem., 70, 13–34.
[235] Takada, A., 1987. Structure of a cauldron in the Otoge ring complex, Shitara District, Aichi Prefecture,
central Japan. J. Geol. Soc. Japan, 93, 107–120‡ .
[236] Takeshita, T. and Yamaji, A., 1990. Tectonophys., 181, 307–320.
[237] Tanaka, T. and Ogusa, K., 1981. J. Geol. Soc. Japan, 87, 725–736‡ .
[238] Tatsumi, Y. et al., 1983. J. Geophys. Res., 88, 5815–5825.
[239] Taylor, G. I., 1938. J. Inst. Metals., 62, 307–324.
[240] Taylor, M. et al., 2003. Tectonics, v, 22, doi:10.1029/2002TC001361.
[241] Thomas, P. G., Masson, P. and Lleitout, L., 1988. In Vilas, F., Chapman, C. R. and Matthews, M. S.,
eds., Mercury, Arizona Univ. Press, Tucson, p. 401–428.
[242] Tosha, T. and Hamano, Y., 1988. Tectonics, 7, 653–662.
[243] Truesdell, C., 1985. The elements of continuum mechanics. Springer-Verlag, Berlin.
REFERENCES 377

[244] Truesdell, C. and Noll, W., 1965. The non linear field theories of mechanics. Springer-Verlag, Berlin.
[245] Turcotte, D. L. and Schubert, G., 1982. Geodynamics: applications of continuum physics to geological
problems. John Wiley, New York, 450p.
[246] Timoshenko, S.P. and Goodier, J.N., 1970. Theory of elasticity. McGraw-Hill, New York, 567p.
[247] Tullis, T. E. and Tullis, J., 1986. In Geophys. Monogr., n. 36, Am. Geophys. Union, Washinton, D. C., p.
297–324.
[248] Twiss, R. J. and Moores, E. M., 1992. Structural Geology. W. H. Freeman and Co., New York.
[249] Twiss, R. J., Protzman, G. M. and Hurst, S. D., 1991. Tectonophys., 186, 215–239.
[250] Vandycke, S., Bergerat, F., 2001. J. Struct. Geol., 23, 393–406.
[251] Van Balen, R. T. and Skar, T., 2000. Tectonophys., 320, 331–345.
[252] Wallace, R. E., 1951. J. Geol., 59, 111–130.
[253] Wallis, S. R. and Behrmann, J. H., 1996. J. Struct. Geol., 18, 1455–1470.
[254] Wang, Y. and Mareschal, J.-C., 1999. Geophys. Res. Lett., 26, 3033–3035.
[255] Watts, A. B., 2001. Isostasy and flexure of the lithosphere. Cambridge Univ. Press, New York, 478p.
[256] Wessel, P., 1992. J. Geophys. Res., 97, 14177–14193.
[257] Weissel, J. K. and Karner, G. D., 1989. J. Geophys. Res., 94, 13919–13950.
[258] Wesnousky, S. G., Scholz, C. H. and Shimazaki, K., 1982. J. Geophys. Res., 87, 6829–6852.
[259] White, N., 1994. Earth Planet. Sci. Lett., 122, 351–371.
[260] Woodward, N. B., Boyer, S. E. and Suppe, J., 1989. Balanced geological cross-sections: an essential
technique in geological research and exploration. Am. Geophy. Union, Wasington, D.C., 132p.
[261] Wu, P., 1991. Geophys. Res. Lett., 18, 451–454.
[262] Yamaji, A., 1989. Mem. Geol. Soc. Japan, 32, 305–320‡ .
[263] Yamaji, A., 1990. Tectonics, 9, 365–378.
[264] Yamaji, A., 2000. J. Struct. Geol., 22, 441–452.
[265] Yamaji, A., 2003. J. Struct. Geol., 25, 241–252.
[266] Yamaji, A., 2003. Tectonophysics, 364, 9–24.
[267] Yamaji, A., 2005. J. Struct. Geol., 27, 2030–2042.
[268] Yamaji, A., Otsubo, M. and Sato, K., 2006. J. Struct. Geol., 28, 980–990.
[269] Yamaji, A. and Sato, K., 2006. Geophys. J. Int., 167, 913–942.
[270] Yamaji, A., Tomita, S. and Otsubo, M., 2005. J. Struct. Geol., 27, 161–170.
[271] Yamaji, A., Momose, H. and Torii, M., 1999. Earth, Planets Space, 52, 81–92.
[272] Yamaji, A. and Yoshida, T., 1998. J. Min. Petrol. Econ. Geol., 94, 389–408.
[273] Yamazaki, T., 1989. J. Geomag. Geoelectr., 41, 533–548.
[274] Zandt, G. and Ammon, C. J., 1995. Nature, 374, 152–154.
[275] Zapryanov, Z. and Tabakova, S., 1999. Dynamics of bubbles, drops and rigid particles, Kluwer Academic
Publ., Dordrecht, 514p.
[276] Ziegler, H., 1983. An introduction to thermomechanics. North-Holland, Amsterdam, 355p.
[277] Zoback, M. D. et al., 1987. Science, 238, 1105–1111.
378 REFERENCES

[278] Zoback, M. D. and Healy, J. H., 1984. Ann. Geophy., 2, 589–698.


[279] Zoback, M. L., 1992. J. Geophys. Res., 97, 11703–11728.
[280] Zhong, S. and Gurnis, M., 1994. J. Geophys. Res., 99, 15683–15695.
[281] Zuber, M. T., 1987. J. Geophys. Res., 92, 4817–4825.
[282] Zuber, M. T., Parmentier, E. M. and Fletcher, R. C., 1986. J. Geophys. Res., 91, 4826–4838.

, in Japanese

, in Japanese with English abstract

Вам также может понравиться