You are on page 1of 163

DEFORMATION MICROSTRUCTURES AND MECHANISMS IN MINERALS AND ROCKS

This page intentionally left blank


Deformation Microstructures and Mechanisms
in Minerals and Rocks

by
Tom Blenkinsop
Department of Geology,
University of Zimbabwe, Harane Zimbabwe

KLUWER ACADEMIC PUBLISHERS


NEW YORK, BOSTON, DORDRECHT, LONDON, MOSCOW
eBook ISBN: 0-306-47543-X
Print ISBN: 0-412-73480-X

©2002 Kluwer Academic Publishers


New York, Boston, Dordrecht, London, Moscow

Print ©2000 Kluwer Academic Publishers


Dordrecht

All rights reserved

No part of this eBook may be reproduced or transmitted in any form or by any means, electronic,
mechanical, recording, or otherwise, without written consent from the Publisher

Created in the United States of America

Visit Kluwer Online at: http://kluweronline.com


and Kluwer's eBookstore at: http://ebooks.kluweronline.com
Contents

Acknowledgements ix

Symbols, Abbreviations and Units xi

1 Introduction and Terminology 1


1.1 Introduction 1
1.2 Classifications of deformation microstructures and mechanisms 1
1.3 Deformation microstructures and mechanisms in the earth: Brittle-semibrittle-plastic transitions 3
1.4 The description of deformation: Scale, continuity, distribution, mechanism and mode 4
1.5 Ductility and the “brittle-ductile transition” 4
1.6 Character and classification of deformation zone rocks 5
1.7 Format and use of this book 5

2 Cataclasis 7
2.1 Introduction 7
2.2 Fundamental cataclastic deformation mechanisms 7
2.2.1 Microcracking 7
2.2.2 Frictional sliding 10
2.3 Microcracks 10
2.3.1 Classification, characteristics and observation 10
2.3.2 Microstructures and mechanisms 12
2.3.3 Impingement microcracks 12
2.3.4 Flaw-induced microcracks 13
2.3.5 Microfracturing of pre-existing flaws 13
2.3.6 Cleavage microcracks 13
2.3.7 Elastic mismatch microcracks 13
2.3.8 Plastic mismatch microcracks 14
2.3.9 Microfault-induced microcracks: Microscopic feather fractures (mffs) 14
2.3.10 Thermally-induced microcracks 15
2.3.11 Phase transformation-induced microcracks 16
2.4 Microfaults 17
2.4.1 Characteristics 17
2.4.2 Mechanisms 17
2.5 Deformation bands 18
2.5.1 Characteristics and classification 18
2.5.2 Mechanisms 18
2.6 Distributed cataclasis and cataclastic flow 18
2.7 Gouge zone microstructures 19
2.8 Microfracture surface features 19
2.9 Crystallographic fabrics 22
2.10 Pre-lithification deformation microstructures and mechanisms 22
2.11 Pseudotachylites and frictional melting 22
2.11.1 Characteristics 22
2.11.2 Origin 22
2.11.3 Misidentification 23

v
vi CONTENTS

3 Diffusive Mass Transfer by Solution 24


3.1 Introduction 24
3.2 Fundamental deformation mechanisms of diffusive mass transfer by solution 24
3.3 Grain surface solution textures 25
3.4 Indenting, truncating and interpenetrating grain contacts 25
3.5 Strain caps 27
3.6 Microstylolites 27
3.6.1 Characteristics 27
3.6.2 Formation and propagation 27
3.7 Diffusive mass transfer and cleavage 28
3.7.1 Classification 28
3.7.2 Spaced cleavages 29
3.7.3 Continuous cleavage 30
3.8 Grain surface deposition textures 30
3.9 Overgrowths, porosity reduction, pressure shadows and fringes, and mica beards 32
3.9.1 Characteristics 32
3.9.2 Mechanisms 33
3.10 Grain shape fabrics 33
3.11 Fluid inclusion planes 33
3.12 Microveins 35

4 Intracrystalline Plasticity 39
4.1 Introduction 39
4.2 Fundamental mechanisms of intracrystalline plasticity 39
4.3 Deformation twins 39
4.4 Undulatory extinction 41
4.5 Intracrystalline deformation bands,
kink bands and subgrains: Recovery 41
4.6 Deformation lamellae 47
4.7 Grain shape fabrics and ribbon grains 47
4.8 New grains, core and mantle structure: Dynamic recrystallization 47
4.9 Crystallographic fabrics 50

5 Diffusive Mass Transfer and Phase Transformations in the Solid State 52


5.1 Introduction 52
5.2 Fundamental deformation mechanisms of solid state diffusive mass transfer and phase transformations 52
5.3 Grain shape fabrics and ribbon grains 52
5.4 Foam texture, static and secondary recrystallization 54
5.5 Decussate texture 54
5.6 Porphyroblasts and inclusion trails 54
5.6.1 Characteristics 54
5.6.2 Growth mechanisms 55
5.6.3 Relationship to deformation 55
5.7 Reaction rims, relict minerals, coronas and symplectites 57
5.8 Chemical zoning 57
5.9 Solid state phase transformation microstructures 57
5.10 Superplasticity 57

6 Magmatic and Sub-magmatic Deformation 59


6.1 Introduction 59
6.2 Fundamental deformation mechanisms and microstructures in rocks containing melt 59
6.2.1 Magmatic flow 59
6.2.2 Sub-magmatic flow 60
6.2.3 Magmatic and sub-magmatic flow and rheology 60
6.3 Mesoscopic evidence for magmatic and sub-magmatic flow 60
6.4 Magmatic microstructures 62
6.4.1 Grain shape fabrics 62
6.4.2 Crystallographic fabrics 62
6.5 Sub-magmatic microstructures 62
6.5.1 Grain shape fabrics 62
CONTENTS vii

6.5.2 Intracrystalline plasticity 62


6.5.3 Diffusive mass transfer 63
6.5.4 Cataclasis 63
6.6 Other microstructures 63
6.7 Non-magmatic deformation 63

7 Microstructural Shear Sense Criteria 65


7.1 Introduction 65
7.2 Curved foliation 66
7.3 Oblique foliations and shape preferred orientations 66
7.4 Porphyroclast systems 67
7.4.1 Characteristics and classification 67
7.4.2 Mechanisms of formation 68
7.4.3 Stair-step direction: and tails 69
7.4.4 Faces of a tail 69
7.4.5 Deflection and embayments of tails 69
7.5 S-, C- and 70
7.5.1 Characteristics and classification 70
7.5.2 Formation and evolution 72
7.5.3 Curvature of S-foliation 73
7.5.4 Shear on C- or 73
7.6 Pressure shadows and fringes 73
7.6.1 Kinematics of pressure shadows and fringes in shear zones 73
7.6.2 Geometry of the last increment of growth 73
7.6.3 Shape 73
7.7 Mica fish 74
7.8 Porphyroblast internal foliations 75
7.9 Crystallographic fabrics 75
7.10 Asymmetric microboudins 76
7.11 Asymmetric microfolds and rolling structures 77
7.12 Shear sense criteria in rocks containing melt 77
7.12.1 Magmatic shear zones 77
7.12.2 Oblique grain shape fabrics 78
7.12.3 Tiling and imbrication 78
7.12.4 S-C fabrics 78
7.12.5 Sub-magmatic microfractures 78
7.13 Shear sense criteria for faults 79
7.13.1 Shear sense observations on faults 79
7.13.2 Displaced grain fragments 79
7.13.3 Risers and slickenfibres 79
7.13.4 Gouges 79
7.13.5 Jogs and bends 79

8 Shock-induced microstructures and shock metamorphism 80


8.1 Introduction 80
8.2 Shock mechanisms 80
8.3 Microfractures 80
8.4 Planar Deformation Features (PDFs) 81
8.5 Mosaicism 82
8.6 Diaplectic glass 83
8.7 High pressure polymorphs of quartz - Coesite and stishovite 83
8.8 Lechatelierite 83
8.9 Tectites, microtectites and spherules 84
8.10 Shock barometry and thermometry 85
8.11 Calibration of shock pressures from microstructures 85
8.11.1 Calibration of shock pressures from optical properties of quartz 87
8.11.2 Problems of shock barometry 87
8.12 Diagnostic impact microstructures 88
viii CONTENTS

9 From Microstructures to Mountains: Deformation Microstructures, Mechanisms and Tetonics 90


9.1 Introduction 90
9.2 Failure criteria 90
9.2.1 Coulomb and Mohr failure criteria 90
9.2.2 Griffith failure criteria 90
9.3 Pore fluid pressure and faulting 91
9.4 Fracture mechanics and failure criteria 91
9.5 Frictional sliding laws 91
9.5.1 Byerlee’s law 91
9.5.2 Rate and state dependent frictional sliding 92
9.6 Flow laws 92
9.6.1 Diffusive mass transfer: Grain size sensitive creep 92
9.6.2 Intracrystalline plasticity 93
9.6.3 Empirical flow laws from experimental data 94
9.7 Polymineralic deformation 94
9.8 Deformation mechanism maps 97
9.9 Lithospheric strength envelopes 97
9.10 Palaeopiezometry 98
9.10.1 Methods and calibration 98
9.10.2 Recrystallized grain size 98
9.10.3 Subgrain size 100
9.10.4 Dislocation density 100
9.10.5 Twinning - differential stress 100
9.10.6 Deformation lamellae 101
9.10.7 Principal stress orientations from deformation lamellae 103
9.10.8 Principal stress orientations and strains from twins 103
9.10.9 General problems with palaeopiezometers 104
9.11 Geothermobarometry 104
9.11.1 Methods and calibration 104
9.11.2 Calcite twin morphology 104
9.11.3 Sutured quartz grain boundaries 105
9.11.4 Subgrain boundary orientation in quartz 105

References 107

Index 127

Color Plate Section 133


Acknowledgements
Most of the photomicrographs were developed and printed by Cuthbert Banda, whose assistance, with that of other members
of the staff of the Geology Department, University of Zimbabwe, was invaluable. James Preen guided the preparation of the
TEX version of the text. Faith Samkange and Maxwell Matongo were able research assistants, supported by the University of
Zimbabwe Research Board. The following are thanked for contributing photomicrographs or thin sections: P. Dirks (Plates
11, 30, 31, 32, 33, 35, 44), R. Fernandes (Fig. 2.9), H. Frimmel (Plate 18), S. Kamo (Figs. 8.4 - 8.6), H. Leroux (Figs. 8.2,
8.3), J.E.J. Martini (Figs. 8.7, 8.8), U. Reimold (Plate 46), J. Stowe (Plate 22), D. Van Der Wal and M. Drury (Fig. 4.2).
Plates 1 - 4 and Figs. 2.5, 2.18, 2.19, 3.18 of core material from the Cajon Pass drillhole were made at the Institute for Crustal
Studies, University of California, Santa Barbara, as part of research on deformation mechanisms with R. Sibson, supported
by the National Science Foundation, U.S.A., under grant DAR-84-10924. The assistance of the technical staff at U.C.S.B. is
gratefully acknowledged. Some research for this book was supported by the IUGS Commission on Tectonics, COMTEC.
Detailed reviews of chapters from the following are greatly appreciated: P. Dirks, R. Hanson, H. Jelsma, W. Means, A.
Ord, M. Paterson, U. Reimold, E.H. Rutter, A. Schmid Mumm.
Fig. 8.2 was reproduced from Leroux et al. (1994), and Fig. 8.7, 8.8 from Martini (1991), all with kind permission of
Elsevier Science - NL Sara Burgerhartstraat 25, 1055 KV Amsterdam, The Netherlands. Fig. 8.4 was reproduced from Krogh
et al. (1996) with kind permission from the American Geophysical Union. Fig. 9.5 was reproduced from Burkhard (1993),
and Fig. 7.20 from Goldstein (1988), both with kind permission from Elsevier Science Ltd. The Boulevard, Langford Lane,
Kidlington OX5 1GB, U.K. Figs. 9.6, 9.7 were reproduced from Kruhl and Nega (1996), with kind permission from Springer
Verlag. Full details of these publications are given in the references.

ix
This page intentionally left blank
Symbols, Abbreviations and Units
Numbers in brackets give the chapters and sections where the symbols are used. Units are given if they are referred to in the
text.
a Microcrack long axis (2.2.1), Rate and state variable friction law constant (9.3.2)
A Flow law constant, (9.4.1), or (9.4.3)
S to C-surface angle (7.5), Flow law constant (9.4.3), Effective stress coefficient (9.2.3)
b Microcrack short axis (2.2.1), Flow law constant (9.4.3)
Burger’s vector (4.2, 9.8)
B Particle velocity/unit potential gradient (3.2), Flow law constant (9.4.3)
c Microcrack or flaw length (2.2, 2.3, 2.4.2), concentration of particles (3.2)
Cohesion, MPa (9.2, 9.3), Reference state solubility, mole fraction (9.4.1)
Angles defining external asymmetry of an LPO (7.9)
CL Cathodoluminescence (2, 3, 8)
CMF Critical melt fraction (6.2)
d Particle size (2.2.2), Flaw spacing (2.4.2), Grain size, m (9.4), mm (9.8)
Subgrain size, mm (9.8.3)
D Fractal dimension (2, 9.9.3)
Reference state diffusion coefficient for Pressure solution creep, (9.4.1)
Reference state diffusion coefficient for Nabarro-Herring, Coble creep, (9.4.1)
S to C- or angle (7.5), Grain boundary width, m (9.4)
DMT Diffusive mass transfer (1, 3, 5, 9)
Twinning density, (9.8.5)
E Young’s modulus (2.2)
Strain rate, strain rate at 0 K in the Dorn Law, (9.4)
f Volume fraction of phase in polymineralic flow law (9.5)
Crystal fraction, fraction at critical packing (6.2.1)
Microcrack extension force, Critical value (2.2)
GBM Grain boundary migration (4.8, 9.4, 9.8)
Surface tension force (2.2), Angle between flaw and (2.3.4)
G Fracture toughness (2.2)
Viscosity of suspension, Viscosity of pure melt (6.2.1)
ISA Instantaneous stretching axes (7.1, 7.7)
Twinning incidence (9.8.5)
J Diffusive flux (3.2)
Angle between flaw array and (2.4.2)
k Dislocation density palaeopiezometer constant (9.8.4)
Stress intensity factors for mode I, II, and III opening (2.2)
Threshold, Critical stress intensity factors (2.2)
kC Velocity of deforming crystal face, (9.4.1)
l Model microcrack length (2.3, 2.4.2), Dislocation density palaeopiezometer constant (9.8.4)
L Critical slip distance (9.3.2), Length of grain boundary (9.9)
LPO Lattice preferred orientation (4.9, 5.10)
LSE Lithospheric strength envelope (9.7)
m Grain size exponent in flow law (9.4, 9.5), Grain size palaeopiezometer constant (9.8.2)
Coefficient of friction (2, 9.2), Chemical potential (3.2), Shear modulus, GPa (9.8.3, 9.8.4)
Coefficient of internal friction (9.2)
n Power law exponent for flow laws (9.4, 9.5)
P Mean stress, Pa (3.2, 9.4)
PDF Planar deformation feature (7)

xi
xii SYMBOLS, ABBREVIATIONS AND UNITS

Pore fluid pressure (3.2, 9.2.3)


PPL Plane polarized light
Flaw to microcrack angle (2.3.4, 2.4.2)
Activation enthalpy, pressure solution and grain boundary diffusion, (9.4)
Activation enthalpy for volume diffusion, and diffusion creep, (9.4)
r Length of stride in divider method (9.9.3)
R Gas constant, (9.4)
RF Reflected light
Density, (9.4.1), Dislocation density, (9.8.4)
S Deformation lamellae spacing, mm (9.8.6)
S Molar entropy (3.2)
Cohesion, MPa (9.2)
Oblique foliation, Bands parallel to shear plane (7.3)
External, Internal foliations (5.6, 7.8)
SEM Scanning electron microscope
ST Sensitive tint plate
Differential stress, Pa or MPa (9.4)
Maximum, intermediate, minimum principal stresses, Compression positive (2,3,5)
Remote stress for microcrack closure (9.2.2), Remote applied stress (2.2.1)
Normal stress (2,3, 9.2), Stress at 0 K in Dorn Law (9.4)
Temperature, 0C and K, Melting temperature
Uniaxial tensile strength, MPa (9.2.2)
TEM Transmission electron microscope
Shear stress, MPa (2.2.2, 9.2, 9.3)
u Dislocation density exponent in dislocation density palaeopiezometer (9.8.4)
U Molar internal energy (3.2)
Subgrain size exponent in subgrain size palaeopiezometer (9.8.3)
V Microcrack velocity (2.2.1), Sliding velocity (9.3.2), Activation volume, (9.4.3)
Molar volume, (3.2, 9.4)
Maximum twin volume, % (9.8.5)
Thickness of a grain boundary fluid (9.4.1), Grain size palaeopiezometer exponent (9.8.2)
Angles defining internal asymmetry of an LPO (7.9)
XP Crossed polarized light
State variable in dynamic frictional sliding law (9.3.2)
() Miller-Bravais indices of a crystal face
{} Miller-Bravais indices of a representative face of a form
<> Miller-Bravais indices of a crystal direction
Chapter 1

Introduction and Terminology

1.1 Introduction changes that remains after stress is removed, as opposed to


elastic (recoverable) deformation which is not seen directly
Deformation microstructures in rocks and minerals are mi- in the geological record. Deformation microstructures can be
croscopic features created by deformation. A deformation divided into three major categories:
mechanism is a process on one scale that accommodates an
imposed deformation on a larger scale. This book describes 1. Microfractures, displacement, and/or rotation of rigid
microstructures and mechanisms at the scale of a thin section, particles with no permanent lattice distortion. The typ-
the scale that most geologists use for detailed petrography, ical microstructures seen in thin section are microfrac-
based on the premise that many of the fundamental mechan- tures and fragments surrounded by a fine-grained mat-
isms can be inferred from microstructures at this scale. The rix.
book aims to be a guide to the recognition and interpretation
of deformation microstructures and mechanisms, and should 2. Microstructures showing material removal, transport
be used in conjunction with a petrographic microscope. and deposition without fracturing, permanent lattice dis-
Why is the study of deformation microstructures and tortion or melting. Examples of microstructures at sites
of material removal are distinctive types of grain con-
mechanisms useful ? Deformation mechanisms are determ-
tacts and microstylolites. Microstructures indicating
ined by temperature, stress (both hydrostatic and deviatoric
material deposition include microveins, overgrowths,
components), strain rate, pore fluids, mineralogy, and the tex-
pressure shadows and porphyroblasts. Many meta-
ture of the deforming rock (especially grain size and poros-
ity). Recognition of deformation mechanisms from micro- morphic textures associated with deformation fall into
structures allows limits to be placed on these variables. For this category.
example, microstructures called subgrains are formed by the 3. Permanent lattice distortion without fracturing. Ex-
deformation mechanism of intracrystalline plasticity, which amples of typical microstructures are undulatory ex-
indicates deformation at temperatures above 250°C in quartz. tinction, subgrains, recrystallized grains, and crystallo-
The size of the subgrains can be used to gauge the deviat- graphic fabrics.
oric stress during deformation. These are essential pieces of
information for tectonic analysis and for understanding the This simple classification scheme can be applied on the basis
behaviour of the lithosphere by mathematical modelling (e.g. of optical microscope observations. Table 1.1 summarizes
Kusznir and Park 1987, Molnar 1992, Beaumont et al. 1996). the scheme, gives examples of specific microstructures, and
Microstructures and deformation mechanisms are a grow- refers to the relevant chapters and sections of the book.
ing field of interest in the earth sciences. The application of Table 1.1 also shows the relation between this scheme and
materials science to minerals and rocks has provided much of a classification of deformation mechanisms, which has three
the new impetus. However, the literature is scattered through similar categories (e.g. Knipe 1989):
journals in a large number of disciplines, and most structural
geology textbooks have limited coverage of the field. At least 1. Cataclasis - deformation by microfracturing, sliding
some of this diverse literature is reviewed in this book, which and/or rotation of rigid particles (Chapter 2). Brittle de-
contains a comprehensive reference list. Hopefully the book formation is often used as a synonym for cataclasis, but
is written in sufficient depth to be useful at advanced under- is more accurately defined as deformation by fracture or
graduate level and above. Familiarity with elementary con- microfracture.
cepts and terms in structural geology is assumed.
2. Diffusive mass transfer (DMT) - deformation by diffu-
sion, the movement of lattice defects, ions, atoms or
1.2 Classifications of deformation mi- molecules in response to gradients of chemical potential
(Chapters 3 and 5).
crostructures and mechanisms
3. Intracrystalline plasticity - deformation by the move-
Deformation microstructures in minerals or rocks are the re- ments of extra half-planes of atoms (dislocations) in a
cord of permanent deformation, i.e. shape and/or volume crystal lattice (Chapter 4).

1
2 CHAPTER 1. INTRODUCTION AND TERMINOLOGY
CHAPTER 1. INTRODUCTION AND TERMINOLOGY 3

A general term for both categories 2 and 3 is useful because events, each of which may be associated with different mech-
they are often closely associated with each other: crystal anisms. One of the applications of this book should be to
plasticity or simply plasticity means deformation by either allow the unravelling of multiple deformations from their as-
or both DMT and intracrystalline plasticity. DMT can be sociated microstructures. Secondly, even within a single de-
split into two major sub-divisions: diffusion via a solution formation, mechanisms and microstructures vary from min-
eral to mineral within a polymineralic rock: a common ex-
(pressure solution, Chapter 3), and diffusion in the solid state
(Chapter 5). Solid state phase transformations may occur ample is the intracrystalline plasticity of quartz in a shear
during deformation, some involving DMT: these are included zone at greenschist facies, that contrasts with cataclastic de-
in Chapter 5. formation of feldspar in the same conditions.
Two or more mechanisms may act simultaneously together Thirdly, one mechanism may be incapable of accommod-
within a single mineral. An example is the combination of ating the imposed stress, strain or strain rate, so that other
sliding on grain boundaries, and mass transfer of material by
mechanisms are substituted or added during the same de-
diffusion to fill space created by sliding. This is is one type
formation: for example, faulting (cataclasis) may relieve high
of superplasticity, which is a composite deformation mech- stress levels in a shear zone otherwise deforming by intracrys-
anism; one in which two or more mechanisms are coupled talline plasticity. The record of both the cataclasis and plas-
together in the same mineral under the same conditions. An ticity may be preserved in the microstructures. Mechanisms
important composite deformation mechanism involves the in- that can only accommodate deformation in a single direction,
teraction of microfractures and intracrystalline plasticity: this
for example slip on a single fault set, are especially restrictive
is called semibrittle deformation. The different components and invariably require supplementary mechanisms.
of a composite mechanism may have variable strain rates: the
mechanism with the slowest rate determines the composite The variation in conditions in the earth, particularly of tem-
strain rate, and is said to be rate-limiting. perature and pressure, causes a corresponding variation in de-
formation mechanisms. Two systematic changes occur with
It is fortunate that the non-genetic classification of the mi-
increasing depth: temperature increases by the geothermal
crostructures matches the genetic classification of the mech-
gradient, and pressure increases due to the effect of gravity.
anisms so well, and this means that both classifications can be
referred to by similar names. The chapter titles of this bookCataclasis is relatively insensitive to the variation in temper-
ature, but is suppressed by pressure. The plastic deformation
use the names of the mechanisms for simplicity. The order of
mechanisms (intracrystalline plasticity and solid-state DMT)
the chapters follows the general change in deformation mech-
behave in the opposite way: they are strongly promoted by
anisms from low to high grade metamorphic conditions dur-
temperature but relatively insensitive to pressure. As a res-
ing deformation (see next section). Specific microstructures
that are diagnostic for each mechanism are shown in bold in ult, cataclasis is generally restricted to the upper crust (where
pressures are low), and crystal plasticity occurs in the lower
Table 1.1.
crust and the rest of the lithosphere (where temperatures are
Three chapters deal with relatively new developments in
high). The transition from cataclasis to plasticity is some-
structural geology, which can involve all three categories
times called the brittle-plastic transition. Experiements have
of deformation microstructures and mechanisms. Magmatic
identified an important intermediate regime of semibrittle be-
and sub-magmatic deformation (Chapter 6) describes the de-
haviour where microfractures interact with intracrystalline
formation of rocks that contain melt. Important deformation
plasticity (e.g. Carter and Kirby 1978, Kirby and Kronen-
mechanisms are flow of melt and crystals, with crystal de-
berg 1984), leading to the concept of two transitions in de-
formation (sub-magmatic flow) or without crystal deforma-
formation mechanism with increasing depth: firstly the brittle
tion (magmatic flow). This topic has great relevance to cur-
- semibrittle transition in the upper crust, and secondly the
rent debates about pluton emplacement mechanisms. Micro-
semibrittle - plastic transition in mechanisms at mid-lower
structural shear sense criteria (Chapter 7) provide clues to
crustal levels.
the displacement of rock masses during deformation on all
scales: this is one of the most important types of tectonic These generalizations need to be heavily qualified because
information. Shock-induced microstructures and shock meta- of the other variables that affect deformation mechanisms.
morphism are produced by meteorite impacts (Chapter 8), and Stress, strain rate and pore fluids are some of the most im-
are the focus of much current interest, because microstruc- portant additional variables to be considered. For example,
high stresses or strain rates may cause cataclasis at greater
tural studies have a central role to play in the debates about
depths or higher temperatures, which would otherwise be as-
mass extinction and other possible effects of large meteorite
impacts on the earth’s evolution. sociated with plasticity, as in the above example. Pore flu-
ids are essential for solution-assisted DMT, and may promote
cataclasis by mechanical and chemical effects. Furthermore,
1.3 Deformation microstructures and the various minerals within a polymineralic rock have differ-
mechanisms in the earth: Brittle- ent transitions from brittle to semibrittle to plastic behaviour,
and interactions between the minerals themselves also need to
semibrittle-plastic transitions be considered as an independent variable (Chapter 9). These
qualifications, particularly the variation in properties of dif-
Several different deformation microstructures commonly oc- ferent minerals, mean that no unique depth or temperature
cur together in rocks for three important reasons. Firstly, can be given for the brittle-semibrittle and semibrittle-plastic
deformation microstructures may record several deformation transitions. However, the transitions are important concepts
4 CHAPTER 1. INTRODUCTION AND TERMINOLOGY

macroscopic scale, these discontinuities are not seen and the


fold appears to be continuous.
Continuity should be described at the time of deformation,
but can easily be altered by subsequent events. Another prob-
lem with specifying deformation continuity is posed by fea-
tures such as overgrowths, pressure shadows, and pressure
fringes (Chapter 3.9), which may be continuous with the sur-
face from which they are grow, and discontinuous with the
surface towards which they grow. Boundaries between grains
are perhaps best regarded as continuous on a microscopic
scale during recrystallization, but after recrystallization they
appear as discontinuities. These examples show that a certain
amount of judgement may be necessary to describe continu-
ity, which is a necessary shortcoming of any description that
has to take into account the past history of deformation in
minerals and rocks.
Deformation distribution is also highly scale-dependent.
As scale of observation is increased, an localized deforma-
tion may appear pervasive: for example, microfractures are
highly localized deformation on a microscopic scale, but a
network of microfractures can have the effect of a pervasive
strain on the scale of an outcrop or a regional map.
The combination of deformation distribution (local-
ized/pervasive) and mechanism (cataclastic or plastic) was
described as a “mode of failure” by Rutter (1986). This
concept can be extended through the incorporation of deform-
ation continuity and scale. The combination of continuity,
mechanism and scale can be called a deformation mode, and
for understanding the behaviour of the earth, and the condi- represented on a diagram such as Fig. 1.1, where a mode
tions for the transitions can be predicted for specific models is specified by deformation mechanism (z-axis), continuity
of the earth, as described in Chapter 9. (which can be qualified by distribution, x-axis) and the scale
length (y-axis). The front of the diagram shows the field of
possible deformation modes described in this book (i.e. at the
1.4 The description of deformation: microscopic scale), with examples of some microstructures.
Scale, continuity, distribution, Figure 1.1 makes some important links between micro-
structures and mechanisms. Cataclastic deformation mi-
mechanism and mode crostructures are discontinuous, and intracrystalline plasti-
city microstructures are continuous, at the microscopic scale.
Deformation of minerals or rocks should be described in These observations point towards one of the major themes
terms of three fundamental attributes. Continuity is the con- of this book: deformation microstructures can be used, al-
nectivity of material points through the deforming body; de- beit with care and a certain amount of ambiguity, to identify
formation can be characterized as continuous if points remain deformation mechanisms. This link is possible by induction
connected, or discontinuous if not. The distribution of de- and deduction from the microstructures and theoretical un-
formation can be described as localized (e.g. a shear zone) or derstanding of the mechanisms, and by comparison with ex-
distributed. The deformation mechanism can be described in periments.
one of the three categories given in Section 1.2. All three at-
tributes depend on the scale of observation. Scales are loosely
defined in this book as macro (greater than outcrop, i.e. > 10
m), meso (outcrop, i.e. 1 cm - 10 m) or micro (microscopic,
1.5 Ductility and the “brittle-ductile
i.e. < 1 cm). transition”
Permanent deformation of minerals or rocks involves
breaking atomic bonds and is therefore is discontinuous at the Great confusion has been caused by the use of the terms
atomic scale. As the scale of observation is increased, these ductile and brittle-ductile transition. Much of this confu-
atomic discontinuities can not be discerned, and deformation sion can be traced to the dichotomy between laboratory (and
appears to be continuous. The top surface of Fig. 1.1 shows materials science)-based and field-based approaches to de-
schematically how deformation continuity is a function of the formation (the subject is well reviewed in Evans et al. 1990
scale of observation. An example of the discontinuous to con- and Williams et al. 1994). For example, rock mechanics
tinuous transition with increasing scale is the accommodation experiments use a maximum permanent strain before fail-
of a fold by cataclastic flow (Chapter 2.6). At micro to meso ure of more than 5% to define ductile behaviour (e.g. Pa-
scales, the deformation is by discontinuous fracture, but at a terson 1978). Similarly, Griggs and Handin (1960) defined
CHAPTER 1. INTRODUCTION AND TERMINOLOGY 5

relative ductility as: “the amount of permanent deformation the most widely used today. The scheme is based on the dis-
achievable prior to rupture”. These definitions of ductility tinction between random fabric and foliated deformation zone
based on stress-strain relationships are satisfactory and quan- rocks, as well as the cohesion of the rock (the degree to which
tifiable, but the widespread application of the term outside it behaves as a continuous body during deformation) and pro-
the rock mechanics laboratory necessitates alternative defin- portion of matrix to porphyroclasts. The original scheme is
itions, since stress-strain relationships are never known for slightly modified to allow for a range of foliation intensities in
rocks in the field (e.g. Griggs and Handin 1960). fault rocks in Table 1.2, by including the additional category
The use of deformation continuity or distribution in the of foliated cataclasites (e.g. Chester and Logan 1987), and
definition of ductile is implicit in the application of the terms by extending the gouge, breccia and pseudotachylite categor-
by most geologists in the field. Rutter (1986) suggested that ies to the foliated category. Random fabric has been changed
ductility is “the capacity for substantial, non-localized strain”, to Unfoliated to allow for deformation zone rocks that have
thus strictly excluding shear zones, and “is a concept which some order to their fabric, for example fragments with a jig-
must be defined on a macroscopic scale”. However, “ductile saw texture, and Foliated has been changed to Strongly fo-
shear zones” (e.g. Ramsay and Huber 1987) is a commonly liated in order to contrast with Unfoliated. Many deforma-
used phrase. It is proposed here that continuity rather than tion zone rocks have two or even three distinct foliations. All
distribution should be used to define ductility, and that a more foliations should be considered when assessing foliation in-
satisfactory definition of ductility is “macroscopically con- tensity, and the existence of different foliations can be used
tinuous deformation”. This definition has the merit of encom- to further classify deformation zone rocks (e.g. S-C mylon-
passing all features that the field geologist usually refers to as ites, Lister and Snoke 1984). The crush breccia series of the
ductile, including shear zones and macroscopic folds, and is original classification has been omitted for simplicity, and be-
preferred because continuity can be identified more precisely cause these terms have not found widespread use.
than distribution at any scale. The classification scheme in Table 1.2 is based on descrip-
Brittle and ductile are defined above by different cri- tion at the hand specimen scale, and recognizes that larger
teria. Therefore it is possible for a rock to be both brittle fragments in a finer grained matrix is a fundamental character
and ductile: for example, a type of deformation band of most deformation zone rocks. However, as noted by Sibson
(Chapter 2.5) forms by fracture (it is brittle) but has macro- (1977), any pidgeon-hole classification such as Table 1.2 suf-
scopic strain continuity (it is ductile). The concept of the term fers from the problem that the classification criteria may show
“brittle-ductile transition” is rendered questionable by these a continuous range of variation. This is especially problem-
definitions. Many terminological problems can be avoided atic in the assessment of foliation development, which may be
by using the concept of a deformation mode and the cat- difficult to quantify, or to judge objectively on a qualitative
egories of deformation continuity, distribution and mechan- basis. Another problem with the classification is the defini-
ism suggested in Section 1.4, and avoiding the use of the tion of cohesion, which is specified in the original classific-
term brittle-ductile transition, which has no meaning using ation as primary cohesion, i.e. cohesion during deformation.
the above definitions. Post-tectonic processes may decrease cohesion, for example
by weathering, or increase cohesion by cementation. These
processes may not be recognized easily or allowed for when
1.6 Character and classification of de- assessing primary cohesion.

formation zone rocks


1.7 Format and use of this book
Rocks within deformation zones are commonly highly
strained and may exhibit some of the best examples of de- Chapters 2 to 5 deal with the major categories of deform-
formation microstructures. Deformation reduces the grain ation microstructures and mechanisms. Each chapter begins
size of some of the protolith to produce a matrix of finer with a brief introduction to the fundamental mechanisms, and
grains surrounding remnant larger grains or grain aggregates, proceeds to descriptions of the characteristic microstructures,
known as porphyroclasts or clasts, which is the typical micro- identifying which are diagnostic, and illustrating them with
structure of deformation zone rocks. Deformation mechan- diagrams and black-and white photomicrographs embedded
isms and microstructures may differ between the matrix and in the text. Colour photomicrographs are collected separately
porphyroclasts. and referred to as plates in the text. Key words are emphas-
Some definitions and classifications of deformation zone ized where they occur for the first time in the chapter. Some
rocks have advocated the use of deformation mechanisms as mechanisms and microstructures, especially cataclasis, are
a classificatory tool. For example, Wise et al. (1984) pro- dealt with in more detail than others because comprehensive
posed to include crystal plasticity as an essential element descriptions are lacking in the literature. The mechanisms in-
of the definition of mylonites. As for any other classifica- volved in the formation of each microstructure are interpreted
tion of observational data, the use of mechanisms should be from experimental and theoretical backgrounds. The final
avoided because subjective interpretations are required, that chapter shows how deformation microstructures and mech-
may change in the light of new knowledge. Nomenclature anisms can be used to make quantitative inferences about de-
and classification of deformation zone rocks is extensively formation conditions for tectonic analysis. Throughout the
discussed in Snoke et al. (1998). text, relatively new developments in the subject, and those
The non-genetic classification scheme of Sibson (1977) is that have not been described in previous textbooks, have been
6 CHAPTER 1. INTRODUCTION AND TERMINOLOGY

more heavily referenced than other topics. The references are ST - Sensitive tint (gypsum) plate inserted
presented in two forms. The main list gives all references in
full. This is followed by a list of abbreviated references col- SEM - Scanning electron microscope optical image
lected by chapter, which shows important general sources for TEM - Transmission electron microscope image
the chapter topics in italics.
A general list of symbols, abbreviations and units prefaces CL - Cathodoluminescence image from the SEM
this chapter. Abbreviations and conventions for all photomic-
rographs are as follows: Figures below the captions give the horizontal dimension of
the image in mm. Shear sense (see Chapter 7) is given as
PPL - Plane polarized light sinistral or dextral (assuming that the shear or fault plane is
vertical); all illustrations with shear senses given are perpen-
XP - Crossed polarized light dicular to the shear plane and parallel to the shear direction,
RF - Reflected light and the shear plane is approximately parallel to the horizontal
edge of the image.
Chapter 2

Cataclasis
2.1 Introduction The second model treats the microcrack as flat with a sharp
tip (e.g. Lawn and Wilshaw 1975a), and gives a versatile
Microfractures, displacements and rotations of rigid particles solution for the stress field around the microcrack, in the polar
with no permanent lattice distortion are microstructures coordinates of Figure 2.1b:
formed by cataclasis. There are two fundamental cataclastic
deformation mechanisms: microcracking (Section 2.2.1) and
frictional sliding (Section 2.2.2). The single most distinct-
ive cataclastic microstructure is the microfracture, defined The stress is thus specified by K, the Stress intensity factor,
as a tabular or planar microscopic discontinuity. The term describing the intensity of the field around the microcrack,
thus includes microfaults, microscopic deformation bands, and the stress distribution, described by radial factor
microjoints, microcracks, microveins, and microscopic slip and a function of which depends on the propagation mode.
surfaces. Microfractures can be sub-divided into microfaults The three microcrack propagation modes shown in Figure 2.2
(Section 2.4), which contain a fragmental matrix, and mi- are tensile or opening mode, (mode I), in which displace-
crocracks (Section 2.3), which do not. Pseudotachylites are ments are perpendicular to the fracture plane, and two shear
briefly discussed in Section 2.11 because they are associated modes, in which displacements are parallel to the fracture
with cataclastic mechanisms. plane, sometimes called sliding and tearing modes (modes II
and III, Fig. 2.2). The stress intensity factor depends on the
propagation mode, the microcrack length and the applied
2.2 Fundamental cataclastic deforma- stress
tion mechanisms
2.2.1 Microcracking
Given these descriptions of the stress around the flat mi-
Microcracking involves microcrack nucleation followed by crocrack, it is possible to deduce the failure criteria for brittle
propagation. Nucleation is irrelevant in a geological context solids from Griffith’s postulate that a microcrack will extend
because of abundant heterogeneities in natural rocks and min- when the total energy change with the propagation of the mi-
erals. crocrack is negative or constant. The energy terms involved
in microcrack propagation are release of mechanical energy,
Dynamic microcrack propagation which must be equal to the energy required for creation of
surface area.In terms of force, the microcrack extension force,
Microcrack propagation from an initial flaw can be con- G, must be greater than or equal to twice the surface tension
sidered by two different models. The first model assumes force, (the doubling factor accounts for the two sides of
an elliptical microcrack (Inglis 1913, Griffith 1924). For a the crack). This leads to the classical Griffith result for fail-
uniaxial remote applied stress parallel to the microcrack ure under a tensile remote stress in a solid with Young’s
axis, the tangential stress, on the microcrack surface var- modulus E:
ies from a negative (tensile stress) equal to the value of at
the long axis of the microcrack, to a compressive stress at the
short axis (Fig. 2.la). In a biaxial stress field is equal to The well-known Griffith failure criterion and its derivatives
at the long axis of the microcrack (Fig. 2.1a; can be determined from this relation (Section 9.2). The Grif-
Jaeger and Cook 1979). The stress state around an elliptical fith failure criterion thus implicitly assumes the existence of
microcrack illustrates the essential concepts that large tensile microcracks of length Griffith observed that the theoretical
stresses can develop at the tip of a microcrack surface in com- strength of solids is much greater than their actual strength,
pression, and that the maximum tensile stress will develop and postulated that real solids contain microcracks which
in the direction of the maximum applied stress. These two concentrate stress leading to failure, thus explaining the dis-
concepts explain why extension microcracking perpendicular crepancy between theoretical and measured strengths. The
to the least principal stress is a widespread and fundamental existence of such microcracks or “Griffith flaws” is still ac-
cataclastic mechanism. cepted as the basis of most failure criteria.

7
8 CHAPTER 2. CATACLASIS
CHAPTER 2. CATACLASIS 9

The model so far assumes elastic behaviour, but an addi-


tional energy term must be incorporated to account for break-
ing of atomic bonds at the crack tip, which can be done by
postulating the existence of a small zone ahead of the micro-
crack tip in which non-linear forces are expended in breaking
bonds. The additional energy involved is incorporated in the
energy balance in the form of a new parameter the fracture
toughness, to replace the surface energy term The energy
balance condition is now:

This equation gives the condition for microcrack propagation


in terms of a value for G usually known as the critical frac-
ture toughness, which can be related to the critical stress
intensity factor, for a given geometry.
This analysis is valid provided that the size of the non-
linear zone is much smaller than the length of the microcrack
i.e. it does not affect the elastic stress system as a whole:
this assumption is called the small scale yielding (Rice 1968)
or the small scale zone approximation (Lawn and Wilshaw
1975a). More detailed analyses can relate to the dis-
placement on the microcrack if some function of the cohesive
forces ahead of the tip with distance is postulated: these are
the cohesive force models (e.g. Rudnicki 1980). The exist-
ence of such a non-linear zone ahead of the microcrack tip is
known experimentally from ceramics and metals, where it is
referred to as a process zone, and it has been detected from
acoustic emission and microcracking in geological materials
(e.g. Swanson 1981, Peck 1983, Labuz et al. 1987).

Sub-critical microcrack propagation

The above analysis is restricted to microcracks that propag- Five mechanisms of sub-critical microcrack growth have been
ate at speeds that are typically significant fractions of the ve- proposed, but stress corrosion, a general term for environ-
locity of elastic waves in solids, or dynamic microcracking. mentally influenced, stress driven, thermally activated chem-
However, microcracks may propagate under stress conditions ical reaction allowing breaking of bonds is considered to be
well below the critical stress intensity factor, at rates that the dominant mechanism for geological materials in crustal
depend on temperature and chemical environment as well as conditions (Atkinson 1982). Hydrolysis of the Si-O-Si bond
stress intensity. This phenomenon is known as sub-critical is likely to be responsible for the weakening. The microcrack
microcrack growth and is potentially of enormous geological velocity in Region 1 is controlled by the reaction rate of the
importance (e.g. Anderson and Grew 1977, Das and Scholz hydrolysis, and transport-rate control occurs in Region 2.
1981, Atkinson 1982, Meredith 1983). A diagram showing
Other types of chemical reaction occur with sub-critical
microcrack velocity (V) as a function of mode I stress intens-
microcracking. An example is the reaction of plagioclase to
ity factor illustrates the main features of sub-critical mi-
increasingly sodic compositions and ultimately to laumontite,
crocrack growth (Fig. 2.3), which have been demonstrated for
which creates a 60% volume increase (Blenkinsop and Sibson
a range of geological materials (e.g. quartz, granite, andesite,
1991). This alteration occurs along cleavage microcracks,
basalt, calcite, oil shale, sapphire and glass). The
resulting in a texture of expanded, matching fragments, de-
relationship falls into three parts:
rived from a single parent crystal (Plate 1). This texture sug-
Region 1. Velocity is highly sensitive to water concen- gests chemical alteration and sub-critical microcracking oc-
tration and temperature. There may be a threshold stress curred in a linked process, which can be called alteration-
intensity factor for microcrack growth to occur at enhanced microcracking.
all (Meredith 1983). Sub-critical microcrack growth has been incorporated into
some models of crustal processes, for example to explain the
Region 2. Velocity is dependent on water concentration and difference between creep (stress intensity factor between
temperature, but not and and seismic faulting (stress intensity factor equal to
Rudnicki 1980), the barrier theory for earthquake rupture
Region 3. Velocity increases extremely rapidly with until of Das and Scholz (1981), and features of magmatic intru-
microcrack growth becomes dynamic at sion and hydrofracture propagation (e.g. Anderson and Grew
10 CHAPTER 2. CATACLASIS

Adhesive strength, asperity interlocking and increasing


contact area have been combined in an elegant and simple
model by Wang and Scholz (1994, 1995), which accounts for
experimental results very well.
Frictional sliding leads to the production of gouge by as-
perity failure and ploughing. Once formed, fragmentation
within a gouge layer continues by mutual impingement of
particles, leading to particle size distributions (PSDs) that are
characteristically fractal (e.g. Blenkinsop 1991). A fractal
distribution of particle sizes can be described by the rela-
tion:

where N (d) is the number of particles greater than size d, and


D is the fractal dimension. D for particles in cataclastic rocks
ranges from 1.88 to 3.08 (e.g. Sammis et al. 1987, Sammis
and Biegel 1989, Olgaard and Brace 1983, Wang 1987), with
many results for natural and experimental gouges around the
value of 2.6 (Marone and Scholz 1989, Biegel et al. 1989).
Sammis et al. (1987) showed that a distribution of spheres
of unequal sizes reduces impingement stresses on individual
fragments, and proposed that microcracking in a gouge will
proceed in order to minimize the probability of neighbour-
ing grains having equal sizes. This constrained comminution
model predicts D = 2.58, which is very close to observed val-
1977, Atkinson 1982). ues (e.g. Sammis and Biegel 1989).
Gouge formation (and cataclasis generally) may occur by
at least two other processes of particle size reduction. Alter-
2.2.2 Frictional sliding ation, for example the laumontization of feldspar described
Amonton’s law that the steady state shear stress is propor- above, may lead to lower values of D (Blenkinsop and Sib-
tional to the normal stress on the sliding surface via the son 1991). At advanced stages of gouge formation, selective
coefficient of friction microfracture of larger particles takes place, creating PSDs
with fractal dimensions greater than 2.58 (Blenkinsop 1991).
Plates 1 to 4 show a sequence of cataclastic textures arranged
in order of increasing fractal dimension of PSDs, which is the
is predicted by a simple adhesion theory of friction. The the- evolutionary sequence of textures with strain.
ory assumes that the rough surfaces contact at asperities (pro- A simple theory of wear can predict that the volume
truding irregularities), which will yield under normal stress of gouge created by sliding, and hence the thickness of
until sufficient area of contact is established to support the the gouge, will increase linearly with displacement (Scholz
normal load (Fig. 2.4). The contact area is considered to have 1990), and some experimental results confirm this relation-
an adhesive strength which must be exceeded over the entire ship (e.g. Teufel 1981). This relationship has also been
area for sliding to occur. This model successfully predicts claimed for faults (e.g. Robertson 1983, Hull 1988). How-
the normal-stress dependence of friction, but it underestim- ever, there are at least two reasons the experimental data
ates the value of because there are other contributions to should not be applied directly to faults. Firstly, the roughness
the shear strength in addition to the adhesive strength of the of laboratory prepared surfaces is not comparable to natural
contacts (Scholz 1990). These include: fault surfaces (Brown and Scholz 1985), and secondly, once
a gouge layer has accumulated sufficient thickness to prevent
1. Interlocking of asperities. Oblique surfaces of contact interaction between the sliding surfaces, wear will no longer
are created between two asperities that come into contact occur according to the simple model of surfaces in contact
after sliding (Fig. 2.4). Interlocking may be relieved by with each other. The reported data on fault gouge thickness-
shearing through asperities (adhesive wear), or by slid- displacement relationships does not substantiate a proportion-
ing on the oblique contacts, which causes dilatancy. ality (Blenkinsop 1989).

2. Increasing area of contact due to asperity shearing. As


sliding continues, asperities fail and the contact area in- 2.3 Microcracks
creases.
2.3.1 Classification, characteristics and obser-
3. Asperity ploughing. An asperity with a greater strength vation
than the opposite surface will cut into the weaker ma-
terial, generating a groove and wear fragments. This is Microcracks can be classified as intragranular (within single
known as abrasive wear. grains), transgranular (across two or more grains) and cir-
CHAPTER 2. CATACLASIS 11
12 CHAPTER 2. CATACLASIS

cumgranular or grain boundary. The occurrence of these dif-


ferent types of microcrack depends on the microcrack mech-
anism (see below) and on the microstructure of the rock. In-
tragranular microcracks are characteristic of poorly cemen-
ted, highly porous rocks, whereas well-cemented, low poros-
ity rocks have transgranular microcracks. This classification
is useful in discriminating various microcrack mechanisms
described in Sections 2.3.3 to 2.3.11.
Tectonic trans- or intragranular microcracks commonly
link contact points between adjacent grains, and are kinked
or curved. Despite grain-scale irregularities in fracture geo-
metry, microcracks generally have a strong preferred orienta-
tion. Several sets of microcracks may exist within one rock.
Extension (mode I) microcracks are usually filled with ma-
terial in optical continuity with the host grains, so that their
importance may be underestimated or even completely over-
looked. The only manifestation of a fracture left in the rock
may be a line of fluid inclusions which are the trace of a fluid
inclusion plane (Fig. 2.5, Section 3.11), and careful obser-
vations of well-prepared thin sections at high magnifications
under the optical microscope are necessary to detect such mi-
crocracks if they contain small inclusions. Microcracks can
have a wide range of aspect ratios and densities. Extension
microcracks may have regionally systematic orientations be-
cause they form perpendicular to (Section 2.2.1.1), and
have been used very effectively to deduce regional stress sys-
tems (e.g. Lespinasse and Pêcher 1986).
Cryptic microcracks may be spectacularly revealed by
cathodoluminescence (CL) (e.g. Smith and Stenstrom 1965,
Sprunt et al. 1978, Sprunt and Nur 1979, Blenkinsop and
Rutter 1986). Luminescence depends on silica polymorph,
and chemical, thermal and mechanical histories (e.g. Seye-
dolali et al. 1997). Microfracture fillings which form under
different conditions from the host grain, and experience only
part of their tectonic history, therefore contrast in lumines-
cence with the rest of the grain and usually have very low
luminescence (Fig. 2.6). size, sorting, and grain shape. Impingement microfracturing
can be understood from an analysis of the stress field cre-
ated on loading a plane surface by a pointed (Boussinesq
2.3.2 Microstructures and mechanisms
configuration) or spherical (Hertzian configuration) object
Nine microcrack “mechanisms” can be distinguished, mainly (Fig. 2.8). The point load should produce radial microcracks
from experiments, where extension microcracks, usually sub-perpendicular to the indenter, and such microcracks have
known as axial microcracks, form from about half the peak been observed in indenter experiments (e.g. Lindquist et al.
strength through to post-failure (e.g. Tapponier and Brace 1984). The spherical indenter (possibly more geologically
1976). These are secondary mechanisms compared to the realistic) will contact the plane surface along a spherical sur-
fundamental physics of tensile microcracking described in face, inducing a region of compressive stresses immediately
Section 2.2.1, and they mainly describe specific geomet- below the indenter, surrounded by tensile stresses near the
ries that create microcrack tip tensile stresses (Krantz 1983, edge of the contact surface. An extension microcrack in the
Atkinson 1982). Table 2.1 summarizes the characteristic fea- form of a cone (cone microcrack) forms beneath the indenter
tures of the mechanisms, including the types of microcrack (Fig. 2.8) at a critical load. The critical indenter radius to gen-
(intra-, trans- or circumgranular) that can form by each mech- erate a tensile microcrack depends on the applied pressure via
anism. a square root (Lawn and Wilshaw 1975b): very modest pres-
sures for even quite large indenters can create microcracks.
2.3.3 Impingement microcracks A still more geologically realistic configuration is the case
of two spheres loading each other: in this case, an extension
Impingement microcracks link points of contact between ad- microcrack initiates at the edge of the contact plane between
jacent grains, and are usually intragranular. They may link the two spheres at a critical load that depends on porosity,
several pairs of contact points around grains. Four basic grain size, elastic moduli and fracture toughness, providing
patterns are shown in Fig. 2.7 (Gallagher et al. 1974), that there are pre-existing flaws present in the loaded grains
which depend on the boundary loads, packing arrangement, (Fig. 2.8, Zhang et al. 1990). The critical load measured in
CHAPTER 2. CATACLASIS 13

several rocks (Wong 1990) follows the theoretical relation- flaw (Fig. 2.6, Table 2.1). Such microcracks can be intra-,
ship and shows that the required pre-existing flaws have sub- trans- or circumgranular.
micron dimensions, comparable to the flaws invoked in Grif-
fith’s failure analysis (Section 2.2.1).
2.3.5 Microfracturing of pre-existing flaws
Photoelastic experiments on convexo-concave contact sur-
faces which model indenting grain contacts due to pressure Microfracturing of pre-existing flaws is known as an import-
solution (Section 3.4) show that tensile microcracks should ant and even dominant microcracking mechanism from ex-
form normal to the indenter contact, and shear failure is pre- periments (e.g Biegel et al. 1992, Menéndez et al. 1996).
dicted along curved trajectories that are approximately nor- It is considered to be at least as important as cleavage in
mal to the contact in its immediate vicinity, but deviate pro- controlling the overall microcracking process, and Atkinson
gressively with distance (McEwen 1981). (1982) asserts that it dominates the upper 20 km of the crust.
The diagnostic feature for recognizing the impingement The opening of grain boundaries (a type of pre-existing flaw)
mechanism is linking of contact points by intragranular mi- is well known to contribute to experimental cataclasis (e.g.
crocracks (Fig. 2.9, Table 2.1), although the contact points Dunn et al. 1973, Hadizadeh 1980, Tapponnier and Brace
may not be visible in the plane of the section. Impingement 1976). The weakness of some natural grain boundaries, even
microcracking is suppressed in well-cemented or low poros- when these are overgrown by optically continuous quartz ce-
ity crystalline rocks because impingement contacts are lack- ments, is apparent from the observation that overgrowths are
ing and tensile stress concentrations are dramatically reduced a major component in the matrix of faults (Pittman 1981).
by the cement (e.g. Wong and Wu 1995, Menéndez et al. Microfracturing of pre-existing flaws can be identified by mi-
1996). crofracture of a cement, and may involve all three types of
microcrack (Fig. 2.11, Table 2.1).
2.3.4 Flaw-induced microcracks
2.3.6 Cleavage microcracks
Flaw-induced microcracks are joined to flaws such as other
microcracks, pores and grain boundaries. They form because Microcracks in biotite are controlled by the basal (001) cleav-
of the tensile stresses that develop on the flaw surface when age (Plate 42, Wong and Biegel, 1985). In feldspars, the ma-
remote stresses are imposed. They are recognized in experi- jor (001) cleavage and also (010), and (110) planes ex-
ments on analogues and on rocks (e.g. Brace and Bombola- ert a strong influence on microcracks (Willaime et al. 1979,
kis 1963, Tapponnier and Brace 1976). Analytical solutions Brown and Macaudière 1984, Tullis and Yund 1992). Cleav-
show that the microcracks will grow along curved trajectories age microcracking of feldspars is important during deforma-
from both ends of a flaw to produce the well-known “wing tion of granitic rocks in the upper crust (Plate 1; Evans 1988).
crack” geometry (Fig. 2.10, Horii and Nemat-Nasser, 1985, The fracture toughness of quartz is least along the rhombo-
1986, Kemeny and Cook 1987, Baud et al. 1996). For isol- hedral planes, followed by the basal plane: these are prefer-
ated microcrack growth, a flaw length 2c, orientated at to entially exploited during fracture (e.g. Borg et al. 1960, Voll-
is assumed to have both cohesive and frictional resistance brecht et al. 1991). Cleavage microcracks can be recognized
to shear, and tensile microcracks grow from both ends with because they occur in crystallographically controlled sets par-
length l (Fig. 2.10), making an angle with the flaw. The res- allel to known cleavages within single grains (Table 2.1).
ults show that microcracks will grow by tensile failure from
the edges of the flaw along paths which fit experimental ob- 2.3.7 Elastic mismatch microcracks
servations very well (Horii and Nemat-Nasser 1985). After
a critical length (approximately 1.0) in even the slightest Microcracks have been noted in quartz and feldspar grains
tensile value of microcrack growth becomes unstable, but at contacts with micas in experimental studies (Tapponier
remains stable in compression. Flaw-induced microcracking and Brace 1976, Wong and Biegel 1985), and in a naturally
can be recognized by the connection of the microcrack to a deformed quartzite (Hippert 1994). These microcracks are
14 CHAPTER 2. CATACLASIS

thought to develop because of the difference in elastic strain microcracking is largely responsible for semibrittle behaviour
across the quartz-mica or feldspar-mica boundaries due to the (e.g. Carter and Kirby 1978). It can be recognized by
different elastic moduli of the two minerals in contact along the close association between intragranular microcracks and
a coherent interface (Plate 5, Fig. 2.11). They can be recog- areas or individual microstructures of intracrystalline plasti-
nized by the localization of intragranular microcracks around city, such as subgrains, kink bands, deformation lamellae or
contacts between grains of different mineralogy (Table 2.1), twins (Table 2.1).
but may be difficult to distinguish from thermally-induced
microcracks (see Section 2.3.10).
2.3.9 Microfault-induced microcracks: Micro-
scopic feather fractures (mffs)
2.3.8 Plastic mismatch microcracks
Mffs are intragranular microcracks found only adjacent to
Where intracrystalline plasticity is localized in one area (e.g. faults. They are characteristically wedge-shaped, opening to-
in twins, deformation lamellae and kinks, Chapter 4), mi- wards the fault plane (Fig. 2.12c). They were identified in ex-
crocracks may be initiated due to the strain incompatibility perimentally generated faults by Friedman and Logan (1970),
between the area of plastic deformation and the adjacent un- who found them exclusively within 5-10 grain diameters of
deformed area (e.g. Olson and Peng 1976, Wong and Biegel shear faults, and parallel to They did not occur adja-
1985; Fig. 2.12). Microcracks have been observed along cent to an incipient shear, and therefore formed in response
kink bands in naturally deformed enstatite, and normal to to shearing. Conrad and Friedman (1976) defined mffs as mi-
the kink bands in experimentally deformed quartz (Carter crocracks occurring only within grains adjacent to a fault, dy-
and Kirby 1978). Plastic mismatchs may account for the ing out rapidly away from the fault and statistically close to or
common microcracking of feldspar porphyroclasts surroun- parallel to the applied direction of Microcrack density and
ded by deformed quartz grains in quartzofeldspathic mylon- length increase with displacement and normal stress (confin-
ites (e.g. Evans and White 1984; Fig. 2.13). Plastic mis- ing pressure) (Conrad and Friedman 1976, Teufel 1981). Tec-
matches may also occur within single phases or grains due tonic analogues of mffs have been observed associated with
to stress concentrations created by intracrystalline plasticity shear surfaces between pebbles in contact with each other
(e.g. Lawn and Wilshaw 1975a). Plastic mismatch-induced (McEwen 1981). T fractures as described by Petit (1987)
CHAPTER 2. CATACLASIS 15

have the characteristics of mffs.


Mffs are created by tensile fracture at contact points along
the sliding surface (Teufel 1981). The relations between mi-
crocrack density, length, displacement and normal stress ob-
served in the experiments are all consistent with the formation
of mffs due to contact stresses on the sliding surface. Mffs are
intra- or transgranular microcracks that can be recognized by
their localization adjacent to the fault plane, their inclinations
of 20-50° to the fault plane, and their wedge-shape opening
towards the fault plane (Table 2.1). Mffs can be distinguished
from Riedel microfractures, which may also be associated
with fault planes (e.g. Petit 1987), by the shear offset along
the latter.

2.3.10 Thermally-induced microcracks


Microcracks can relieve stresses caused by differential
thermal expansion or contraction between adjacent minerals.
Such microcracks may form in grains of one mineral surroun-
ded by another during heating or cooling. If heating or cool-
ing are accompanied by pressure changes, elastic mismatch
microcracks may also form. Thermally-induced microcrack-
ing can only be distinguished from elastic mismatch-induced
microcracking if the P-T path is known. The case of cool-
ing granite has been considered in some detail by Bruner
16 CHAPTER 2. CATACLASIS

(1984) and Vollbrecht et al. (1991). Granite can be treated transformation involves a volume increase
as a composite of quartz surrounded by feldspar. Two ex- of 11%. Quartz inclusions in garnet or omphacite are sur-
treme cases can be considered during cooling and uplift. In rounded by radial extension microcracks (e.g. Chopin 1984,
isothermal decompression, greater elastic expansion of quartz Smith 1984, Wang et al. 1989), and indeed this texture has
may cause microcracking of feldspar, while isobaric cooling been used as evidence for the former presence of coesite
may lead to microcracking of quartz due to its greater thermal (e.g. Wang and Liou 1991). Although it is clear that the
contraction. The critical geothermal gradient for equilibrium transition has occurred in these rocks be-
between thermal and elastic strains in quartz and feldspar is cause relict coesite can be found in some inclusions, the ques-
10°C/km. Crystal anisotropy may be important on the grain tion arises whether such microcracks could be due to elastic
scale; quartz has a maximum coefficient of thermal expan- mismatches between the silica phase and the host. The key
sion perpendicular to the c-axis, favouring microcracks at low evidence for phase transition microcracking is the observation
angles to the c-axis. Regional stresses can also be important: that there is no microcracking around other types of inclusion,
microcracking will be favoured in those grains with appropri- including rutile. The extensional nature of the microcracks
ate orientations relative to the regional stress.(e.g. Vollbrecht and their origin from tips of inclusions are consistent with
et al. 1994). The general interaction between thermal and the mechanism. Fracture surface energy measurements in the
elastic stress around inclusions has been modelled by D’Arco conditions of the quartz phase transition imply that
and Wendt (1994), and specifically for garnet by Whitney microfractures could also be associated with this transforma-
(1996). tion, and other minerals undergoing similar phase transform-
Thermally-induced microcracking in granites can be re- ations (Kirby and Stern 1993). Radial microcracks have also
cognized by intragranular microcracks concentrated in quartz been observed around calcite inclusions which have replaced
surrounded by feldspar. Thermal microcracks in quartz may aragonite, a transformation that involves a 8.5% volume in-
have a preferred orientation parallel to the c-axis. crease (Wang and Liou 1991). The distinctive features of
phase transformation microcracking are the association of in-
tragranular microcracks with evidence for phase transform-
2.3.11 Phase transformation-induced micro- ation. In the case of the transition, radial
cracks microcracks around inclusions of quartz after coesite are dis-
tinctive.
The strain associated with solid state phase transforma-
tion can produce distinctive microcracks. For example, the
CHAPTER 2. CATACLASIS 17

2.4 Microfaults
2.4.1 Characteristics
Microfaults are shear microfractures that contain grain frag-
ments formed by cataclasis (Plate 3). Displacement paral-
lel to a microfault surface can be identified from displaced
grain boundaries or fragments. However, caution is needed
in positively identifying such shear displacements, because a
purely extensional microcrack can appear to have a shear dis-
placement when viewed in a section which is oblique to the
opening direction of the microcrack. A useful indication of
true shear displacement is the consistent sense of offset of a
number of grain boundaries at a variety of angles to the mi-
crocrack (Plate 45). axial microcracks, buckled at a critical fibre strain, allow-
Two components can comprise microfault matrix: frag- ing the first through-going fault plane to form (Fig. 2.14).
ments derived from the wall rock, and a precipitated cement. This linking mechanism can be contrasted with the beha-
The fragments are usually angular and very poorly sorted, viour of other siliclastic rocks, in which grain boundary mi-
ranging in size from micrometres up to the width of the mi- crocracks form first, to be linked by axial microcracks (Had-
crofault. It is sometimes possible to identify the parent grain izadeh 1980, Menéndez et al 1996). The detailed sequence
from which the fragments have been derived (the sensitive of microcracking and linkage probably depends on the ini-
tint plate is useful for this purpose; Boldt 1995), and frag- tial microstructure of the rock (Hadizadeh 1980, Blenkinsop
ment displacement or rotation may also be detected (Plates 2, and Rutter 1986). A third type of linking mechanism is the
4, 45). The proportion of cement may vary from relatively direct interaction of microcrack stress fields. Such interac-
little, in which the fragments may form a jigsaw texture, to a tions were classified into en-echelon and en passant types
majority, in which the fragments will be matrix-supported and by Krantz (1979) (Fig. 2.15). The former occur between
isolated from each other. Precipitated cements can be identi- two straight, sub-parallel microcracks, which are linked by
fied from euhedral crystal faces overgrowing primary grains a third straight microcrack (Fig. 2.15 a, b). En passant inter-
(e.g. Pittman 1981) or from CL studies (e.g. Stel 1981) be- actions involve curvature of microcrack paths because the mi-
cause the precipitated cement has a different luminescence crocrack tip stress fields influence each other (Fig. 2.15 b, c).
from the grains in the matrix (Section 3.9). A cyclic history The definitive theoretical model of microcrack interactions is
of cementation and shear can sometimes be deduced from the based on solutions for the behaviour of isolated microcracks
presence of resheared matrix. The edges of microfaults may in compression (Horii and Nemat-Nasser 1985). To model
be planar and truncate adjacent grains cleanly, or they may be shear failure, the microcracks interactions are considered in
irregular, perhaps comprising a large number of circumgran- an array of parallel flaws with individual angles and overall
ular microfractures. angle to and spacing d (Fig. 2.16). The results show that
axial compression at first increases with for all values of
i.e. microcrack growth is stable. However, at lower values
2.4.2 Mechanisms
of microcracks interact unstably at a certain stress. In gen-
Experiments identify the shear failure mechanism as the link- eral the model predicts that and are different, suggesting
ing of extension microcracks once a critical microcrack dens- that a failure plane will consist of oblique microcracks linked
ity is achieved (Krantz and Scholz 1977, Costin 1983). Sev- by axial microcracks. The model correctly predicts the effect
eral mechanisms of microcrack linkage have been sugges- of confining pressure on ultimate strength and is validated
ted. Peng and Johnson (1972) proposed that axial micro- by experiments on resin blocks. Natural microfaults seem to
cracks became linked when individual beams, bounded by evolve by microcrack linking at a critical microcrack density
18 CHAPTER 2. CATACLASIS

guished between deformation bands with no cataclasis, de-


formation bands with cataclasis, and deformation bands with
clay smearing.
Deformation bands tend to cluster together to form zones
of bands, often in an anastomosing pattern (e.g. Engelder
1974, Blenkinsop and Rutter 1986). The thickness of these
larger-order features depends on the number of individual
bands which comprise them, and their total displacement is
given by the sum of the individual components. Deforma-
tion zones may contain discrete surfaces with large discon-
tinuous displacements that cut through all other features, and
may have striations on their surfaces. These were described
as slip surfaces by Aydin and Johnson (1983), and are the
latest features to form in a sequence from deformation bands
to zones to slip surfaces.
The distinguishing characteristics of deformation bands are
shear displacements which preserve material continuity on a
mesoscopic scale, and their occurrence in porous, granular
rocks.

2.5.2 Mechanisms
Dilatancy caused by grain sliding, and impingement micro-
fracturing are important mechanisms in deformation band
formation. Experiments, theory and observations suggest
that the differences between the three categories of deform-
ation band are related to confining pressure and initial mi-
crostructure (Antonelli et al. 1994). Lack of grain micro-
cracking in the first category may indicate low confining pres-
sures during deformation compared to the bands in the second
category. Initial porosity determines whether a deformation
in similar ways to experiments and models (e.g. Blenkinsop bands dilates or compacts: low initial porosity leads to dila-
and Rutter 1986). tion, and vica versa. These relationships are predicted by crit-
ical state theory for granular materials (e.g. Schofield and
Wroth 1968). The third category forms as a result of the
2.5 Deformation bands higher clay content in these rocks.
The localization theory of cataclasis accounts for the form-
2.5.1 Characteristics and classification ation of microfaults, deformation bands, zones and slip sur-
faces (e.g. Rudnicki and Rice 1975, Aydin and Johnson
A distinctive type of localized deformation occurs in porous
1983). Localization is defined as a difference in strain rate
granular materials. Deformation bands are roughly planar
between a band and the matrix, and is a highly appropriate
features from millimetres to a hundred metres long, from
way to describe deformation bands since strain appears con-
fractions to a few millimetres wide, and with net slips from
tinuous across the bands. The theory predicts the formation
fractions to tens of millimetres. Displacement across deform-
of deformation bands at or before peak stress, and their sub-
ation bands is continuous at a mesoscopic scale, which distin-
sequent spread due to strain hardening. The development of
guishes deformation bands from faults. Porosity in deforma-
slip surfaces can be understood from the same theory as a pro-
tion bands may be greater or less than the wall rocks; in the
cess in which stress within the deformation band becomes too
more common case of porosity reduction, the bands contain
great to be accommodated with the result that a discontinuity
a layer of fine-grained matrix partly comprised of grain frag-
develops.
ments. Intragranular extension microcracks are found within
this type of deformation band.
There is a continuum between faults and deformation
bands, with many descriptions of faults having some charac-
2.6 Distributed cataclasis and cata-
teristics of deformation bands (e.g. Engelder 1974, House clastic flow
and Gray 1982, Jamison and Stearns 1982, Underhill and
Woodcock 1985, Narahara and Wiltschko 1986, Blenkinsop A number of experiments have produced distinctive cata-
and Rutter 1986, Cruikshank et al. 1991, Zhao and John- clastic microstructures in samples that have maintained
son 1991). Deformation bands were explicitly described by strength without localization of deformation. The micro-
Aydin (1978), Aydin and Johnson (1978, 1983), and most structures are characterized by distributed fracture and dis-
comprehensively by Antonellini et al. (1994), who distin- placement of fragments, but the deformation is macroscopic-
CHAPTER 2. CATACLASIS 19

ally continuous: this is called cataclastic flow, which can be 2.7 Gouge zone microstructures
defined as deformation by cataclastic mechanisms leading to
continuous flow at a given scale. Cataclastic flow occurs at Faults formed at low grade conditions commonly contain
higher confining pressures than faulting in equivalent rocks, gouge with a distinctive set of microstructures, which have
and is favoured by lower differential stresses. been closely replicated by experiments (e.g. Logan et al.
Extensively damaged grain boundaries, perhaps produced 1981, Rutter et al. 1986). Most of the features of gouge
by fracture under sliding indenters (microfault-induced mi- zones can be understood in the context of a zone of simple
crocracks, Section 2.3.9), and wear products accumulated in shear with boundaries parallel to the shear plane (Fig. 2.17).
the sliding zone and in pores are characteristic microstruc- P-foliation is formed by grains (most commonly phyllosilic-
tures of cataclastic flow in quartzites. (Rutter and Hadizadeh ates) aligned at an oblique angle to the gouge zone boundary
1991). Compaction and filling of pores by crushed material (Fig. 2.17, Plate 6). It may be inosculating and rather variably
seems to be typical of cataclastic flow of siliclastic rocks in developed. It is one of the first microstructures to form in ex-
general (e.g. Menéndez et al. 1996). Axially orientated mi- perimental gouges. Shears parallel to the P-foliation with the
crocracks are dominant. same sense of shear as the gouge zone are known as P-shears.
Riedel shears (R) form at an oblique angle in the opposite
Microstructures produced during experimental cataclastic
direction to the P-foliation (Fig. 2.17, Plate 6). They are dis-
flow of feldspar aggregates consist of intragranular shear mi-
crete shears with the same sense of shear as the gouge zone
crocracks mainly along feldspar cleavages producing blocky
and have geometrical similarities to extensional crenulation
fragments between the cleavage microcracks at 90° to each
cleavages and (Section 7.5, cf. Platt and Vissiers,
other (Tullis and Yund 1987, 1992). The spacing of the
1980). Conjugate Riedel shears form at a large angle
microcracks is as low as Deformed grains have a
to the gouge zone boundary with the opposite sense of shear
puckered appearance consisting of areas of patchy extinction,
(Fig. 2.17). They are generally much less prominent in gouge
caused by slight mismatches between adjacent blocks, which
zones than Riedel shears.
resemble subgrains or even recrystallized grains formed at
T fractures or extension fractures are inclined to the gouge
higher temperatures by intracrystalline plasticity. However,
zone boundary in the opposite direction to the P-foliation
TEM observations established that no dislocation processes
(Fig. 2.17). They may be localized in more competent rocks
contributed to the block rotation. Microcrush zones <
in the gouge zone. Competent rock may form boudins,
wide have sharp boundaries and strong grain rota-
which are commonly asymmetric (Section 7.10) and sep-
tion. A crystallographic preferred orientation and a strong
arated along Riedel shears (Fig. 2.17, Plate 6). Ductile
grain shape fabric developed by slip on the closely-spaced
stringers (Fig. 2.17) consist of relatively large, rigid, asym-
grain-scale faults that are geometrically similar to crystal
metric clasts elongated in the direction of the P-foliation, with
plastic slip systems, and mechanical twins were formed.
stepped tails that may become detached from the main clast
Similar features have been observed in cataclastic flow of
(Logan et al. 1981). Asymmetric folds are also found in
anorthosite, and lamellar features interpreted as bundles of
gouge zones: they may fold earlier boudins. The vergence
cleavage microcracks were also observed (Hadizadeh and
of the folds may be the same as the shear sense of the gouge
Tullis 1992). Twinning is evidence for the operation of
zone (Fig. 2.17), although opposite vergence can occur (Sec-
semibrittle deformation mechanisms.
tion 7.11). The orientation of fold axes within gouge zones
It is likely that quartz does not undergo the same cataclastic may define a girdle distribution parallel to the shear plane.
flow regime observed in feldspars, but requires a thermally- Y-shears are discrete, relatively long shears that cut through
activated mechanism to stabilize flow (Hirth and Tullis 1991). the gouge zone with the same sense of shear as the gouge zone
The difference between feldspar and quartz behaviour is due (Fig. 2.17, Plate 7). There is usually only one Y-shear in the
to the excellent cleavages in feldspar. Pyroxenes and am- central part of the gouge zone. In experiments, Y-shears are
phiboles may behave in a similar fashion to feldspar (Tullis the last feature to form.
and Yund 1992). Cataclastic flow has also been reported from The orientation of all features except Y-shears shown in
experiments on basalts (Mogi 1965; Shimada 1986), dunite Figure 2.17 is variable since their initial orientation is likely to
and gabbro (Byerlee 1968). change with finite strain, because material lines rotate during
A maximum of 30% shortening has been achieved in ex- simple shear. Features rotate clockwise in a dextral gouge
perimental cataclastic flow, after which faulting generally oc- zone such as illustrated in Figure 2.17. Components of pure
curs, and therefore it is unclear whether cataclastic flow can shear may complicate this general rule. The foliation in a
be stable to higher strains. The problem is complicated be- foliated cataclasite may originate as a fabric along any of the
cause faulting is artificially initiated by the experimental con- above shears, or as a P-foliation in a gouge zone (Plate 8).
figuration. If porosity reduction is necessary for cataclastic
flow in porous rocks (e.g. Rutter and Hadizadeh 1991), this
must limit the amount of strain that can accumulate by cata- 2.8 Microfracture surface features
clastic flow. At the opposite extreme of scale, cataclastic flow
has been suggested for continents (Gallagher 1981), and at Microfracture surfaces may show a variety of microstructures
intermediate scales such as kilometric scale folds (Stearns that convey useful information about cataclastic deformation.
1968, Hadizadeh and Rutter 1983, Blenkinsop and Rutter Intact specimens of microfracture surfaces can be easily ob-
1986). In all these cases, the deformation is discontinuous served under a binocular microscope, and at greater magni-
at a smaller scale than the scale of observations. fication in the SEM. It is also useful to examine the cross-
20 CHAPTER 2. CATACLASIS

sections by cutting thin sections perpendicular to the micro- category of slickenlines consists of equally developed ridges
fracture surface, preferably parallel and perpendicular to the and grooves with perfectly matching (or nested) profiles on
slip direction. both sides of a slickensided surface (Means 1987). The ridge
There is a rich field of microscopic studies on extension and groove structure can be formed by localized shear asso-
microfracture surfaces in the material science literature (e.g. ciated with a grain shape fabric, and the length of individual
Bansal 1977, Quakenbush and Frechette 1978, Michalske ridges and grooves may be greater than the displacement on
and Frechette 1980) which suggests that fracture surface fea- the shear zone (Wil and Wilson 1989). Ridge and groove
tures could be used to reveal rates of fracture propagation and structures can also be formed by inosculation of shear sur-
even fracture stresses (e.g. Scholz 1972, Martin and Durham faces (C-surfaces) in an S-C mylonite (Section 7.5, Lin and
1975, Swanson 1981, Norton and Atkinson 1981, Meredith Williams 1992). Another type of slickenside which may also
1983, Cox and Atkinson 1983). Surface microfracture fea- be formed by continuous deformation consists of very fine
tures such as mirror, mist, velocity hackle and forking, and grained quartz (0.01 to ) with a strong crystallographic
Wallner lines, which are known to be diagnostic of dynamic preferred orientation (Power and Tullis 1989). The preferred
microcrack propagation in glass, have not been found on frac- orientation may have resulted from orientated growth by dif-
ture surfaces in rocks. This has been taken to indicate that fusive mass transfer through a solution during the interseis-
the studied rock fractures were never dynamic (Kulander and mic, low strain rate part of the seismic cycle.
Dean 1995). This conclusion is supported by observations Microfracture surfaces may be offset along discontinuit-
such as irregular rib marks that imply stable propagation at ies approximately perpendicular to slickenlines, known as
low velocities, again by analogy with glass. risers or steps (Fig. 2.19). Risers are generally less dense
and coarser features than slickenlines, and are commonly
A shear microfracture with a smooth or shiny surface is
more irregular. They may be approximately linear or cres-
a slickenside, which is commonly marked by parallel lines
cent shaped. Risers are described as incongruous if the off-
known as slickenlines or slickenside striations. Many slick-
set opposes the direction of movement of the opposite block,
ensides are the surfaces of microfaults formed during tec-
and congruous if the opposite is true. Crystal fibres on fault
tonic deformation. The lines are parallel or tapering ridges or
surfaces are known as slickenfibres, and risers created by the
grooves parallel to the slip direction. There may be more than
crystal terminations are accretion steps (Norris and Barron
one set of slickenlines on a microfault surface in different dir-
1968). They are identified by the presence of fibres that are
ections. It is sometimes possible to distinguish overprinting
commonly monocrystalline. Risers may form either by ini-
relationships between different generations of slickenlines.
tial irregularities in the fault surface, or by the intersection of
Probably the most common mechanism of slickenline secondary fractures with the fault surface. Possible geomet-
formation is asperity ploughing (Section 2.2.2), which can ries of secondary fractures can be separated into P fractures,
be identified by the preservation of the asperity on one side R, T and fractures using the same terminology as intro-
of the microfracture and a matching groove on the opposite duced above for gouge zone features (Petit 1987). Slicken-
surface. The type of slickenline produced is called a wear fibres form by crystal growth in the dilatant gap adjacent to
groove (Fig. 2.18), tool track or mark, or prod mark (Tija congruous risers.
1967, Hancock 1985). Ridges of gouge may be formed by
wearing down asperities or by accumulating gouge around a
hard particle: this mechanism is debris streaking. Erosional
sheltering occurs where a solid ridge of microfracture sur-
face is preserved in the down-slip direction of a particle. A
CHAPTER 2. CATACLASIS 21
22 CHAPTER 2. CATACLASIS

2.9 Crystallographic fabrics Fine grained margins may occur at the contact with the wall
rock.
Experimental studies have shown that crystallographic fab- Pseudotachylite often occurs in planar veins a few mm to
rics can be produced by cataclasis (e.g. Borg and Maxwell 1 m thick from which veinlets may branch and penetrate the
1956, Borg et al. 1960). Crystallographic fabrics in feld- wall rock. Planar-convex lenses of pseudotachylite may oc-
spars have been produced by slip and rotation on multiple cur on either side of the vein. The pseudotachylite may occur
grain-scale faults (Tullis and Yund 1987, 1992). The faults in a volume between two parallel planar zones, and surround
are geometrically analogous to slip systems in intracrystal- large fragments of wall rock (e.g. Grocott 1981, Magloughlin
line plastic deformation (e.g. Allison and La Tour 1977, Sec- 1989). The wall rock fragments can form a jigsaw-textured
tion 4.9). Cataclastically-formed crystallographic fabrics can breccia with pseudotachylite matrix. Joints may form perpen-
be distinguished from intracrystalline plastic fabrics by the dicular to the vein walls. The pseudotachylite veins cross-cut
presence of multiple, grain-scale faults such as those illus- any earlier fabrics in the wall rock, and may contain frag-
trated by Tullis and Yund (1992). ments of previously-formed pseudotachylite, indicating cyc-
lical generation (e.g. Sibson 1980, Passchier et al. 1990). All
these features may occur on outcrop to thin section scale.
2.10 Pre-lithification deformation mi-
crostructures and mechanisms 2.11.2 Origin
The origin of pseudotachylite has usually been discussed in
The microstructures of pre-lithification deformation, and terms of the two mutually exclusive hypotheses of crushing
even the deformation itself, may be difficult to recognize. A (e.g. Wenk 1978) or frictional melting (e.g. Sibson 1977).
grain shape fabric can be produced by frictional grain sliding There is conclusive evidence that many pseudotachylites in-
and rotation, known as independent particulate flow (IPF) be- volved frictional melting during faulting. The intrusive tex-
cause the deformation is independent of fracture (Borradaile tures of pseudotachylite veins together with their flow band-
1981). This deformation mechanism is sometime regarded ing demonstrate that they formed in a fluid state. The spher-
as separate from cataclasis, because there is no fracturing; ulites and dendritic crystals and crystallites formed by cool-
here, however, cataclasis is used in a broader sense to en- ing from a melt or from devitrification of a glass formed
compass IPF. High pore fluid pressures may be essential in by quenching, by analogy with structures formed in rapidly
pre-lithification IPF (Knipe 1986, 1989, Maltman 1994). The cooled igneous rocks. The fine-grained margins of the veins
diagnostic feature of a grain shape fabric produced by IPF is can be interpreted as chilled margins, and the joints perpen-
the absence of microcracks and lack of internal deformation dicular to the veins walls can be interpreted as cooling frac-
of grains, but intracrystalline plasticity can also be involved tures (e.g. Camacho et al. 1995).
during pre-lithification deformation (e.g. kinking in biotite; Some pseudotachylites have the same bulk chemistry as
Morrit et al. 1982). The fabric of scaly clays (e.g. Agar et al. their hosts, and the pseudotachylite composition may change
1988) may also be characteristic of pre-lithification deforma- with that of the local wall rocks through which they pass (e.g.
tion. Maddock 1986, Maddock et al. 1987), demonstrating that the
melt was locally derived and not transported by an intrusive
dyke. In other pseudotachylites, there is a consistent differ-
2.11 Pseudotachylites and frictional ence in the bulk chemistry of the pseudotachylite and wall
melting rock which can be interpreted by fusion of the lowest melting
point fraction of the rock. Preferential melting of the hydrous
2.11.1 Characteristics mafic components of the host rock can generate a pseudot-
achylite matrix which is more mafic than the host rock, and
Pseudotachylite is a glassy or very fine-grained rock occur- explains why hydrous and mafic minerals are less common in
ring in veins and associated with deformation zones (Plates 9- the matrix than the host rock (e.g. Magloughlin 1989, Mad-
11). It generally consists of a dark matrix which encloses an- dock 1992, Camacho et al. 1995).
gular to rounded fragments of the wall rock (Plates 9, 11). The localization of pseudotachylite on and near planar sur-
The fragments are very poorly sorted and may have a high faces suggests that it is generated by sliding (hence these sur-
clast/matrix ratio. Clasts in the range to 10 mm in faces are called generation planes). The rounding of the frag-
some pseudotachylites have a fractal PSD, with a fractal di- ments may be due to thermal spalling. A link between slick-
mension of 2.5 (Shimamoto and Nagahama 1992). Quartz ensides, slickenlines and pseudotachylite has been suggested
and feldspar fragments may be more common in the pseudot- by Spray (1989) from observing mechanical excavation of
achylite than in the wall rock, while hydrous minerals such sandstone, which created thin layers of melt with shiny and
as biotite may be less common. Flow banding defined by striated surfaces.
colour variations may be seen in the matrix (Plates 10, 11), Pseudotachylite has also been successfully created in fric-
which may contain spherulites, dendritic crystals, crystal- tional sliding experiments (e.g. Spray 1987, 1995). Fila-
lites, and microphenocrysts, and may vary from opaque to ments of frozen melt drawn out across fractures “like warm
colourless and isotropic in thin section. Other microscopic mozarella” are a delicate feature of experimentally-produced
features include optically isotropic rims around clasts, em- pseudotachylite which has not so far been observed in tec-
bayments of clasts, sub-microscopic crystals, and amygdales. tonic pseudotachylite, but could be a useful diagnostic fea-
CHAPTER 2. CATACLASIS 23

ture. The experiments resolve the controversy about the ori- the Sudbury structure (Spray and Thompson 1994). Other
gin of pseudotachylites: both crushing and frictional melt- pseudotachylites have been described from the bases of large
ing are likely to be involved, depending on the strain rate landslides (e.g. Masch et al. 1985).
(Spray 1995). Crushing occurs in experiments at strain rates
of and melting at strain rates from to 2.11.3 Misidentification
The passage of seismic or shock waves, with a rapid in-
crease in strain rate, may involve three sequential processes: Veins of tourmaline, chlorite, ultracataclasite and fine-grained
initial fracture, comminution, and melting when heat can igneous dykes can easily be mistaken for pseudotachyl-
not be dissipated fast enough. Together with results from ite. Passchier and Trouw (1996) suggest some microstruc-
thermal modelling, experiments show that pseudotachylites tures that allow the distinction of pseudotachylite from other
should only form in dry rocks at relatively high temperatures, types of veins. The diagnostic microstructures that identify
because the presence of water decreases the effective stress pseudotachylite as a melt in distinction from hydrothermal
(Section 9.2.3) on the fault plane so that insufficient heat is veins are the quench and devitrification textures. The distinct-
generated to melt the rock (e.g. Spray 1987, 1988). On the ive chemistry, reflecting the wall rocks, is also diagnostic and
other hand, Magloughlin (1989) has argued that pseudota- distinguishes pseudotachylite from ultracataclasites. Distinc-
chylites were generated from a hydrous cataclasite with a low tion from igneous dykes relies on evidence for faulting associ-
melting temperature. The permeability of the wall rocks is ated with the generation of pseudotachylite, and the different
an important factor in determining whether excess fluid pres- chemistry of the dykes. Pseudotachylites should be distin-
sures can build up during faulting and suppress melting (Mase guishable from ultramylonites by their lack of evidence for
and Smith 1985). Experiments confirm the widely held view extensive intracrystalline plasticity.
that pseudotachylites can form during seismic slip at strain There are nevertheless many cases in which detailed evid-
rates greater than ence must be sought to prove the origin of a pseudotachyl-
Some pseudotachylites are associated with features such ite by frictional melting. It is a relatively uncommon rock
as craters, shatter cones and shock metamorphism which type in the field. The requirement of formation in hot anhyd-
show that they have been produced during meteorite impacts, rous conditions effectively limits the host rocks to igneous or
for example the Vredefort Dome in South Africa, from where metamorphic rocks.
Shand (1916) first coined the term pseudotachylite, and
Chapter 3

Diffusive Mass Transfer by Solution

3.1 Introduction The exact geometry of the diffusion path is still unclear.
Rutter (1976, 1983) has suggested that a water film on the
Microstructures that show evidence for material removal, order of nm thick must be present along grain boundaries to
transport and deposition without fracturing or lattice distor- allow DMT to occur. This concept is supported by observa-
tion form by diffusive mass transfer (DMT). It is often diffi- tions of arrays of fluid inclusions along grain boundaries in
cult to identify positively the nature of the diffusive flux, and naturally deformed halite, interpreted as evidence of a syn-
the evidence for the role of a solution may be indirect. Nev- tectonic continuous film of fluid along grain boundaries that
ertheless, it is a safe generalization that up to amphibolite fa- subsequently healed into discrete fluid inclusions (Urai et al.
cies, by far the majority of DMT occurs via solution because 1986), and by observations of grain boundary pores which
solid-state diffusion occurs very slowly at these temperatures probably formed in equilibrium with a fluid in quartz (Hip-
(e.g. Rutter 1983). pert 1994). Alternatively, islands of solid material may sup-
port normal stresses and allow diffusion to occur in a solution
within intervening channels (Raj and Chyung 1981, Raj 1982,
3.2 Fundamental deformation mech- Spiers and Schutjens 1990). It may be necessary for different
anisms of diffusive mass transfer phases to be adjacent to each other for a continuous fluid film
to be present (Hickman and Evans 1991).
by solution The undercutting mechanism proposed for pressure solu-
Diffusion of material occurs in response to gradients in chem- tion is completely different and does not require diffusion
ical potential, as summarized by Fick’s law which states that in a solute film between surfaces under high normal stress
the flux of the material (J) is proportional to the chemical gradients (e.g. Bathurst 1958, Tada and Siever 1986). In this
potential gradient mechanism, grain contact deformation occurs by intracrystal-
line plasticity or cataclasis, while dissolution of free surfaces
around the contacts maintains small contact areas and there-
fore the high stresses necessary for the solid-state grain de-
B is the particle velocity per unit potential gradient and c is formation. The relative importance of solute film diffusion
the concentration. The chemical potential of a component versus undercutting depends on temperature, grain size and
is given by: free grain surface area. Calculations by Tada et al. (1987)
suggest that undercutting will provide faster strain rates un-
der most circumstances, but that solute film diffusion will be-
where U is the molar internal energy, T the temperature, S come important for very small free grain surface areas, as in
the molar entropy, is the normal stress, is the pore the final stages of compaction, and in dense aggregates.
fluid pressure, and is the molar volume (e.g. Green 1980). The distribution of fluid in a rock, and therefore the effect-
Therefore variations in normal stress and pore fluid pressure iveness of pressure solution, is dependent on surface energy
can establish chemical potential gradients necessary for DMT (measured by the dihedral angle, ) and deformation. Fluids
to occur; internal strain energy may also play a role (Wintsch with greater than 60° will occur as isolated pores at equi-
and Dunning 1985, Bell and Cluff 1989). The diffusion will librium, as interconnected channels at triple grain junctions
be from sites of high to low normal stress; a gradient in for and along all grain surfaces for
normal stress is essential to cause material flux (Raj 1982). (e.g. Watson and Brennan 1987). The first two cases clearly
Hence deformation accommodated by DMT through an in- constrain the potential for pressure solution because the dif-
tergranular solute is often called pressure solution, despite fusion pathways are limited, and probably apply to fluids in
the fact that solubility itself is not significant to the thermo- most static geological situations. However, experiments and
dynamics of DMT. However, recent experiments that demon- theory show that fluids can wet grain boundaries completely
strate a crystallographic dependence of solubility may pose under deformation, leading to enhancement of diffusion creep
some problems for current pressure solution models based by orders of magnitude (e.g. Urai 1983, Copper et al. 1989,
on this thermodynamic approach (Becker 1995, Den Brock Heidug 1991, Tullis et al. 1996). Moreover, dihedral angles
1996). may be a function of pressure and temperature, as established

24
CHAPTER 3. DIFFUSIVE MASS TRANSFER BY SOLUTION 25

by experiments on brines in halite, which show that the di-


hedral angle reduces below the 60° threshold at elevated pres-
sures and temperatures (Lewis and Holness 1996). Unfortu-
nately experiments also suggest that evidence for fluids along
grain boundaries may be rapidly removed after deformation
ceases.
DMT is often inferred from microstructures that show a
systematic relation to surfaces which may have been under
higher normal stress (e.g. surfaces parallel to a shape fab-
ric). However, it is the principal strains which are apparent
from most microstructures, and thus there is always an ambi-
guity about interpreting DMT from the such microstructures
(cf. Groshong 1988). The above inference relies on an as-
sumption that the principal stress and strain axes are parallel,
which will be true for coaxial deformation. In this chapter,
finite shortening and extension directions will be emphas-
ized because they can be inferred from the microstructures.
Any links made between DMT microstructures and principal
stress orientations or gradients in normal stress must involve
some assumptions about the nature of the deformation, such
as coaxiality.

3.3 Grain surface solution textures


Solution creates distinctive grain surface textures which are
clearly observed in the scanning electron microscope. The
typical texture consists of pits on the grain surface, giving a
corroded appearance. Irregular pores with channel and cave
morphologies 1 to long and deep have been
described in a naturally deformed micaceous quartzite (Hip-
pert 1994); similar channel morphologies have been produced confidence in sediments with rounded grain shapes that al-
during experimental deformation of quartz by DMT (Den low the pre-solution grain shape to be accurately delineated.
Brock and Spiers 1991). Atomic-scale details of the solu- Truncated contacts may appear superficially like flat contacts
tion process can be seen using atomic force microscopy (e g. formed by mutual concordance between overgrowths (Sec-
Gratz et al. 1991).This technique shows that calcite dissolu- tion 3.9), but can be distinguished by the evidence for mater-
tion and precipitation occurs in layers, and that solution gen- ial removal.
erates crystallographically-controlled etch pinholes less than
The formation of indenting or truncating grain contacts can
5 nm deep, and rhombic etch cores more than 90 nm deep
be understood from theoretical and photoelastic treatments of
(Hillner et al. 1992). Quartz solution occurs along nm-sized
the stress distribution in contacting grains. The values of the
pits and ledges parallel to the and directions
principal stresses, and the mean and differential stresses, in-
(Gratz et al. 1991).
crease with the radius of curvature, suggesting that DMT will
preferentially occur in grains with larger radii of curvature,
3.4 Indenting, truncating and inter- creating indenting grain contacts (e.g. McEwen 1981). Trun-
cating grain contacts are expected from approximately equal
penetrating grain contacts rates of DMT between two grains of equal curvature.
Grain contacts affected by DMT are distinctive. Where two Interpenetrating grain contacts are mutually interlocking
grains of similar composition and orientation but with differ- protrusions of grains into each other, which have a similar
ent shapes are in contact, the grain with the smaller radius of morphology to microstylolites (Section 3.6), and indicate ma-
curvature typically penetrates into the grain with the larger ra- terial removal along the contacts (Figs. 3.1c, 3.4). Su-
dius of curvature (Figs. 3. la, 3.2), creating an indenting grain tured grain contacts may look superficially like interpenet-
contact. A flat or slightly curved contact is observed between rating grain contacts, but these are formed by intracrystal-
similar grains with approximately equal radii of curvature. line plasticity (Section 4.8). Sutured contacts can usually be
Such truncating contacts are common on longer edges of distinguished from interpenetrating contacts by the presence
grains which are parallel to any shape fabric (Fig. 3.3). Ma- of subgrains, which may form promontories along a sutured
terial removal is demonstrated by truncation of the original contact that are slightly misoriented with respect to the lattice
grain shape (Plate 12). The amount of material removed can of the rest of the grain (Fig. 3.1d). By contrast, interpenet-
be estimated by reconstructing the missing grain boundaries rating contacts formed by DMT usually occur between grains
(Fig. 3.1b; Onasch 1994). This can be done with reasonable with uniform lattice orientations.
26 CHAPTER3. DIFFUSIVE MASS TRANSFER BY SOLUTION
CHAPTER 3. DIFFUSIVE MASS TRANSFER BY SOLUTION 27

3.5 Strain caps crostylolite. Microslickolites are distinguished from micro-


stylolites by having teeth that are oblique to the plane. The
Strain Caps are strongly foliated domains enriched in micas oblique teeth create a lineation on the microslickolite surface.
or less soluble minerals around opposite surfaces of relatively Microstylolites may follow grain boundaries, especially of in-
rigid objects (Passchier and Trouw 1996). The foliation trace soluble grains. Microstylolites can be conspicuous because
in the strain cap is concordant with the margins of the object, of coatings or fillings of iron oxides, hydroxides, phyllosilic-
and grades into continuity with the bulk rock foliation with ates (which may have a strong grain shape fabric) or organic
increasing distance from the object, so that it appears to be matter, which distinguish them from the wall rock. The thick-
deflected around the object (Plate 13). The concentration of ness of the filling is commonly highly variable. The morpho-
less- soluble minerals such as micas in strain caps suggests logical classification of cross-sections through stylolites by
that they form by removal of the more soluble matrix around Guzetta (1984) into sharp peak, rectangular, wave, smooth
the rigid object, which can be predicted from the thermody- and composite types is useful for microstylolites (Fig. 3.6a,
namic considerations above, since the object will concentrate Plate 15). Parameters for describing microstylolites quant-
normal stress on some surfaces due to its greater rigidity. itavely are shown in Fig. 3.6b (from Andrews and Railsback
1997).

3.6 Microstylolites
3.6.2 Formation and propagation
3.6.1 Characteristics The geometrical similarities between stylolites and micro-
Microstylolites are microscale discontinuities with offset or stylolites suggests that they have the same genesis. The dis-
wave-like shapes that may truncate grains or other mark- cussion in this section refers to sylolites, because most previ-
ers such as bedding planes or fossils (Fig. 3.5, Plate 14). ous work has focussed on the mesoscale features.
Microstylolites are readily observed under the microscope, Truncation of markers is direct evidence that material has
and have most features in common with mesoscale stylolites. been removed along a stylolitic surface. The minimum thick-
Individual protrusions on a microstylolite are called teeth ness of material removed along a stylolite is approximately
or columns, which may have flat crowns at the top and twice the amplitude, or the height of the teeth (Tada and
steep walls along their sides (Fig. 3.5). The offset or wave- Siever 1989). The walls of the teeth are parallel to the direc-
like geometry is observed in all sections, showing that the tion of movement between the opposite sides of the stylolite.
three dimensional shape is a forest of columns and match- Railsback and Andrews (1995) have shown that the orienta-
ing pits approximately perpendicular to the plane of the mi- tion of teeth is more constant than the orientation of the plane
28 CHAPTER. 3 DIFFUSIVE MASS TRANSFER BY SOLUTION

of the stylolite, and is useful for kinematic analysis because ficient to cause a stylolite to develop and propagate under its
it shows the direction of maximum finite shortening. The ob- own stresses. Anticrack tip stresses may be dissipated in a
liquity of the teeth of slickolites shows that there has been zone at the stylolite tip which can be referred to as a process
a component of shortening parallel to the slickolite surface zone, analogous to the non-linear zone ahead of a microcrack
(Fig. 3.6c). tip (Section 2.2.1.1). Carrio-Schaffhauser et al. (1992) found
The coatings or fillings on stylolite surfaces are usually in- evidence for a process zone in the form of enhanced poros-
terpreted as insoluble residues that did not diffuse and were ity at stylolite tips in limestones, and suggested that material
concentrated at the sites of solution. The relative insolubil- removed from the process zone was deposited in a zone of
ity of these coatings is good evidence that diffusion was via reduced porosity adjacent to the stylolite.
solution. This is further supported by a correlation between Stylolites may form on previous fracture surfaces (e.g.
the thickness of the coating and the concentration of this ma- Petit and Matthauer 1995, Railsback and Andrews 1995).
terial in the adjacent layers of rock (e.g. Borradaile et al. Stylolites and pressure solution features may be localized
1982), and by the influence of heterogeneities on the stylolite along authigenic clay layers and around pre-existing phyllo-
trace, which show that solution was concentrated in areas silicates. The proposal that contacts between different phases
of higher inferred normal stress, and was impeded by insol- are necessary for continuous fluid films (Hickman and Evans
uble material. The concentration of insoluble material on a 1991) suggests a good reason for enhanced DMT around
stylolite surface compared to its concentration in the host rock phyllosilicates, and may also apply to the formation of strain
can be used to give an estimate of the amount of shortening caps.
(Railsback and Andrews 1995). However, in some cases the
stylolite filling material is not found in the host rock, and is
therefore the product of metamorphic reactions (Beach 1979). 3.7 Diffusive mass transfer and cleav-
Reaction products can also mark grain boundaries to indicate
where fluid transport has occurred (McCaig 1987).
age
A stylolite can be regarded as an anticrack, or an ellips-
3.7.1 Classification
oidal volume removed from the rock (Fletcher and Pollard
1981). If the resulting hole is closed up by elastic deform- Diffusive mass transfer via solution is an important and even
ation, stresses are induced around the anticrack tip, which dominant deformation mechanism in cleavage formation.
cause the stylolite to propagate in its own plane. Therefore Unfortunately there has been a strong tendency to invoke
a stress concentration under the appropriate conditions is suf- genetic nomenclature for cleavage formation (e.g. solution
CHAPTER 3. DIFFUSIVE MASS TRANSFER BY SOLUTION 29

the rock, forming a zonal cleavage. Another important as-


pect of description is the transition from cleavage domains
to microlithons, which may be sharp or gradational. Many
cleavages referred to as solution cleavages belong in this cat-
egory. Disjunctive cleavage domains are often composition-
ally distinct from microlithons, having a higher proportion of
opaque minerals and phyllosilicates. Zonal cleavages with
marked compositional contrasts between cleavage domains
and microlithons constitute a new compositional layering,
which has been called solution striping (e.g. Williams 1972).
Truncation of markers in cleavage domains suggests that ma-
terial loss by DMT has occurred (Fig. 3.8), although shear
displacements on fractures may create a similar effect. Ma-
terial removal can be distinguished from the latter effect in
some cases by inconsistent offsets of a folded layer on sev-
eral cleavage surfaces, and large variations in thickness of a
marker in adjacent microlithons (Fig. 3.8). The amount of
material removal can be estimated from truncation of markers
and other effects such as imbrication of chert pebbles and off-
set of bedding. The volume loss may vary from a few percent
for a weak cleavage to over 35% for very strong cleavage in
cherts (e.g. Alvarez et al. 1976), and 30-40% for disjunctive
cleavage in coarse siltstones and sandstones (Murphy 1990).
Material transfer by diffusion can also be inferred from the
compositional contrasts between the cleavage domains and
the microlithons. A solution can be inferred as the diffusing
phase because of the relatively low grades of cleavage form-
ation.

Crenulation cleavage
This is a spaced cleavage in which microlithons contain an
earlier fabric that is systematically related to the fabric of
the cleavage domains. The cleavage domains are usually
axial planar to folds (crenulations) in the earlier fabric (Plates
cleavage), which hinders discussion about deformation mech- 16, 17). The crenulations may be symmetric or asymmet-
anisms, but Powell (1979) and Borradaile et al. (1982) give an ric, and rounded or angular, and the transition from cleav-
excellent alternative, non-genetic classification, which is fol- age to microlithon may be sharp or gradational. The crenu-
lowed here. The main subdivision is made between spaced lation wavelength may be determined by the thickness of fol-
cleavage, consisting of cleavage surfaces that are separated ded layers as predicted by buckling theory (cf. Price and
by tabular bodies of rock called microlithons, and continu- Cosgrove 1990), or the grain size may dictate a minimum
ous cleavage, in which cleavage is penetrative throughout the wavelength. The cleavage domains are commonly localized
whole body of the rock. Since most cleavages are concen- along the limbs of the folds, especially those associated with
trated into domains at some scale, this distinction is evidently asymmetrical folds, in which the cleavage domains are usu-
scale-dependent. Powell (1979) suggests that any cleavage ally along only one of the two unequal limbs. Most crenu-
with domains spaced less than 0.01 mm (i.e. at the limit of lation cleavages are zonal because the cleavage domains are
optical microscopic resolution) should be called continuous compositionally quite distinct from the microlithons. Phyl-
at a microscopic scale. losilicates and opaques are concentrated in cleavage domains
and quartz, calcite and feldspar are concentrated in hinges.
While the essential role of buckling in crenulation cleav-
3.7.2 Spaced cleavages age formation is evident, the importance of DMT is clear
Disjunctive cleavage from the compositional zoning. Cleavage domains in typ-
ical crenulation cleavages are depleted in Si, Ca, Na,
Disjunctive cleavage is a spaced cleavage in which the cleav- P, Mg, Fe, and F, and enriched in Al, K, Rb, Y, Ce, Sc, and
age domains cut across the previous fabric of the rock Ba, and there is a net volume loss from the cleavage domains
(Fig. 3.7). The cleavage surfaces may be smooth, rough (dis- that is approximately balanced by volume gain in the micro-
continuous) or wave-like, and in relation to each other they lithons (Manktelow 1994). The zoning can be explained in a
may be sub- parallel, anastomosing, or trapezoidal/conjugate. simplistic way by the normal-stress dependence of diffusion:
Their spacing may be controlled by the rock type. The width crenulation hinges, where the primary foliation lies at a high
of the cleavage domains may be a substantial proportion of angle to the maximum principal stress, may be sites of low
30 CHAPTER 3. DIFFUSIVE MASS TRANSFER BY SOLUTION

normal stress. It is also possible that anisotropy in diffusion which must have occurred by DMT via solution at these meta-
pathways may play an important role in the differentiation morphic grades. However, bulk volume loss is not permit-
during crenulation (Marlow and Etheridge 1977). ted by several geochemical studies on the same rocks, as dis-
Other lines of evidence that can be used to show the import- cussed below (e.g. Erslev and Ward 1994).
ance of DMT are the narrow widths of quartz and feldspar Slaty cleavage domains are dominated by phyllosilicates,
grains in cleavage domains relative to microlithons, trunca- resulting in increased K, Al, Ti, Ba, and U, Na, Ca, Mn, Mg,
tion of grains, the association of crenulation cleavage with Fe, and Si are often depleted due to the loss of albite, carbon-
overgrowths and mica beards, and the lack of other deforma- ate, quartz and feldspar (Borradaile et al. 1982). The depleted
tion mechanisms (Gray 1977, Borradaile et al. 1992). elements may be redistributed by DMT into microlithons,
suggesting that there is little bulk volume change in the rock
(e.g. Groshong 1976). Large-scale volume fluxes are also un-
3.7.3 Continuous cleavage likely at low grades because average shale and slate compos-
itions are very similar (Erslev and Ward 1994). The reasons
The most characteristic continuous cleavage is slaty cleav- for the contradiction between this conclusion and the large
age, defined as a preferred orientation of phyllosilicates too volume losses suggested by the structural studies is presently
fine to be seen with the unaided eye (Fig. 3.9). Slaty cleavage unclear. Whether or not large scale volume changes accom-
commonly has cleavage domains with spacings of 0.1 mm pany cleavage formation, the pronounced local compositional
or less, which are easily visible under the optical microscope differentiation is clear evidence that DMT processes are es-
or the SEM. Mechanical rotation of grains (cataclasis), kink- sential to slaty cleavage development.
ing (intracrystalline plasticity) as well as DMT are involved
in the formation of slaty cleavage (e.g. Wood 1974). Elec-
tron microscope studies show that phyllosilicates and quartz 3.8 Grain surface deposition textures
are mobilized by DMT in the development of cleavage do-
mains (Knipe and White 1977,1979, White and Knipe 1978, Crystal growth from solution occurs by two different mech-
Knipe 1979). Structural studies using a variety of strain mark- anisms with different microstructures (Bennema and van der
ers including reduction spots, trace fossils and grain bound- Eerden 1987). Atomically flat crystal faces first nucleate
aries suggest that slaty cleavage formation is accompanied steps to which atoms can attach, and growth then occurs by
by volume loss (e.g. Ramsay and Wood 1973: 20%, Wright accretion along the steps in crystallographically controlled
and Platt 1982: 50%, Wright and Henderson 1992: 40-60%), planes. Dislocations (Section 4.2) may provide important
CHAPTER 3. DIFFUSIVE MASS TRANSFER BY SOLUTION 31
32 CHAPTER 3. DIFFUSIVE MASS TRANSFER BY SOLUTION

nucleation sites for both growth and solution on crystal sur- Porosity reduction may occur by mechanical compaction,
faces (e.g. Casey 1995). Above a critical temperature known solution and precipitation: the latter two are DMT processes,
as the roughening transition, growth rates are isotropic, and and they can be recognized by the loss of material at strain
sub-spherical crystals may be produced. Deposition by the markers and the precipitation of cements in pores in the form
first mechanism can be recognized from the euhedral nature of overgrowths (e.g. Carrio-Schaffhauser and Gaviglio 1990).
of grain surfaces seen under the microscope; the second may Crystal growth in pore spaces typically results in a convoluted
be responsible for botryoidal textures. pore-grain interface which is a fractal curve. Models suggest
that the fractal dimension of the curve increases with the ra-
tio between precipitation and dissolution rates, and also in-
3.9 Overgrowths, porosity reduction, creases from about 2.5 to about 2.75 in natural sandstones as
porosity reduces during diagenesis (Aharonov et al. 1997).
pressure shadows and fringes, and Pressure shadows and fringes are domains of secondary
mica beards mineral growth adjacent to grains (Plate 19). Pressure shad-
ows lack distinctive internal crystal forms, while pressure
3.9.1 Characteristics fringes have fibrous mineral fillings. Mica beards are a dis-
tinctive type of pressure fringe, consisting of fibrous mica de-
Overgrowths are mineral deposits surrounding a grain in a fining a good shape fabric. Passchier and Trouw (1996) prefer
rim. The rim and grain are usually the same mineral. The rim to use strain instead of “pressure” on the basis that it is non-
is commonly in crystallographic continuity with the grain, genetic; however, they point out that the term strain shadow is
and may therefore be difficult to detect. The overgrowth also potentially misleading as it incorrectly implies that strain
may, however, be separated from the original grain by a zone is low in the shadow zone. Analysis of steady-state pure shear
of inclusions (Fig. 3.10), and CL can distinguish subtle fea- around a rigid sphere shows that there are two volumes of low
tures of overgrowths (Plate 18). A distinctive microstructure pressure at the ends of the object perpendicular to the max-
formed by overgrowths in well cemented rocks are relatively imum applied stress, but that differential stress and strain are
straight grain boundaries and 120° triple junctions between high in these areas (Masuda and Mizuno 1995). With this in-
three grains (Fig. 3.11a). This texture arises from mutual im- sight, and since pressure shadow is so firmly entrenched in
pingement of overgrowths growing at equal rates from adja- the literature, it is probably the better choice of terminology,
cent grains. It can easily be confused with granoblastic poly- despite its genetic connotations.
gonal textures (Section 5.4) unless the overgrowth can be dis- The mineralogy in a pressure shadow or fringe may be the
tinguished from the grain. same as the grain, the same as the matrix, or different from
CHAPTER 3. DIFFUSIVE MASS TRANSFER BY SOLUTION 33

low for fibrous microveins (e.g. Durney and Ramsay 1973,


Elliot 1973, Ramsay and Huber 1983, Passchier and Trouw
1996). The use of fibres as shear sense indicators is discussed
in Section 7.7.

3.10 Grain shape fabrics


Grain shape fabrics are a common DMT microstructure.
They may form by material removal in the shortening dir-
ection (e.g. indenting, truncating or interpenetrating grain
boundaries), and by material addition (e.g. overgrowths)
in the extension direction (Fig. 3.11b). Grain shape fabrics
have been produced experimentally by DMT (e.g. Schutjens
1991). Lack of microfractures distinguish DMT grain shape
fabrics from cataclastic grain shape fabrics, and the lack
of undulatory extinction, sub-grains or recrystallized grains
(Chapter 4) distinguishes them from intracrystalline plastic
grain shape fabrics.

3.11 Fluid inclusion planes


Fluid inclusions consist of nm to mm sized cavities filled by
fluids, which may also contain solid material. They occur
as isolated inclusions, in clusters, and in planes. They are
common in thin section, where they may appear as dark in-
clusions which reveal their fluid contents on examination at
higher magnifications. The most common fluids are aqueous,
saline or with possible admixtures of
sulphur compounds, or more complex hydrocarbons.
An important distinction is made between primary fluid in-
both. Crystallographic continuity may be maintained with clusions that form during growth of the original minerals of
the grain, host, or neither. Fibres may be perpendicular to a rock, and secondary inclusions that form subsequently (e.g.
the grain boundary or oblique. They may be found on more Roedder 1984). Primary inclusions can be recognized be-
than two sides of polyhedral grains. Fibres may be straight, cause they are disposed on euhedral crystal forms, whereas
curved, deformed or undeformed. All of these basic aspects secondary inclusions cut across crystal growth features.
of pressure shadow and fringe description are important in Fluid inclusion planes (FIPs) are planes of fluid inclusions
interpreting the mechanism of formation. that often have a strong preferred orientation on a microscopic
scale, and may also have regionally consistent orientations
3.9.2 Mechanisms (Fig. 3.13, Fig. 2.5).
The fluids in any set of FIPs are generally composition-
Pressure shadows and fringes are considered to form by DMT ally homogeneous, and may be distinct from fluids in other
because they are a new mineral growth under low grade con- sets of FIPs that occur in the same rock. FIPs form by trap-
ditions. Diffusion towards sites of low mean stress is pre- ping of a fluid during precipitation of microcrack fillings.
dicted by the thermodynamic approach of Section 3.2, which The microcracks are usually extension microcracks formed
explains why pressure shadows and fringes form at the ends by the mechanisms described in Section 2.3. Healing of mi-
of grains perpendicular to the inferred maximum principal crocracks in quartz can occur as rapidly as micrometres/day
stress. (Smith and Evans 1984). The healing rate depends on tem-
However, the geometry of the pressure shadows and fringes perature, concentration of the fluid, and microcrack di-
is controlled by a number of other factors including strain, mensions (e.g. Brantley 1992). Healing leaves a plane of
grain shape, whether growth occurs at the grain interface (an- cylindrical or spherical fluid inclusions along the former mi-
titaxial) or the matrix interface (syntaxial), whether the ori- crocrack, firstly by forming cylindrical tubes of fluid paral-
entation of fibres is controlled by the grain surface (face con- lel to the microcrack tip (necking down), and pinching off or
trolled) or the incremental extension direction (displacement ovulation to form approximately spherical or negative crys-
controlled), and whether the shadows or fringes are deformed tal shapes (Fig. 3.14). Isolation of spheres occurs because
or not. Some of these factors are illustrated for a coaxial grain boundary migration rates depend on the thickness of
deformation in Fig. 3.12. Given due consideration of these the fluid phase (the microcrack can be considered as a type
factors, fibres can be used to deduce the extensional strain, of grain boundary): slower migration rates occur in thicker
the deformation path, and the type of flow, as discussed be- fluid films. A local thickening of the microcrack will slow
34 CHAPTER 3. DIFFUSIVE MASS TRANSFER BY SOLUTION
CHAPTER 3. DIFFUSIVE MASS TRANSFER BY SOLUTION 35

the grain boundary migration rate at that point, isolating a


fluid inclusion behind the rest of the more rapidly migrating
boundary (e.g. Urai 1983). FIPs are primary evidence for
the importance of fluids and DMT during deformation. They
form perpendicular to and can be very useful in kinematic
analysis, because they can constrain both the orientation of
and the fluid pressure (e.g. Lespinasse and Cathelineau
1995). They can also give evidence for fluctuating fluid pres-
sures associated with stress cycling and therefore probably
with earthquake faulting (e.g. Robert et al. 1995).

3.12 Microveins
Microveins are microscopic tabular zones of secondary min- orientation and grain size is characteristic of such “free-face”
eral growth. The rich variety of microvein textures observable growth.
under the microscope can allow very detailed interpretations In addition to the major filling phases, many microveins
of their formation. CL is valuable for analysis of microveins, also contain inclusions of the same mineralogy as the wall
because variations in fluid chemistry and temperature during rock. Lines of inclusions parallel to the microvein margins
microvein filling can be reflected in the luminescence, delin- are known as inclusion bands, while those at higher angles to
eating delicate growth features (e.g. Dietrich and Grant 1986, the margins are inclusion trails (Figs. 3.15, 3.16).
Urai et al. 1991). Inclusion bands and trails may form by two mechanisms:
One of the most important features of a microvein is the overgrowth of wallrock fragments, and fracturing of the wall
orientation of the opening vector, the vector which connects rock followed by incorporation of fragments into the filling.
points that were originally joined before microvein opening. The latter mechanism may occur where irregularities on mi-
The opening vector may be perpendicular to the wall (an crovein walls obstruct opening, and must be broken off for
extension microvein) or have components of displacement opening to occur (e.g. Urai et al. 1991). The presence of
both perpendicular and parallel to the walls (a shear micro- numerous inclusion bands is evidence for a cyclic process of
vein). The opening vector can be determined from displaced microcrack opening followed by filling, a process known as
markers, or from some features of microvein fillings, as de- crack-seal (Ramsay 1980). Each inclusion band represents
scribed below. The opening vector can be considered for the one cycle. The width between inclusion bands is
net opening history of the microvein (the cumulative open- for many rock types, and there may be up to thousands of
ing vector) or for individual increments of opening (the in- bands in a microvein (Ramsay and Huber 1987). Inclusion
cremental opening vector). trails are markers of the position of particular points on the
Common microvein fillings are carbonates, quartz, chlor- microvein margin at successive opening positions: they are
ite and epidote. The texture of the filling may be massive, therefore parallel to the opening vector (Fig. 3.15b).
equant (blocky; Plate 20), fibrous, laminated, euhedral (idio- Paradoxically, microstylolites sub-parallel to microvein
morphic), or botryoidal. Filling textures are a function of nuc- margins have also been described, particularly along inclu-
leation and growth kinetics, the rate of microvein opening, the sion bands (e.g. Cox 1987). The shortening demonstrated
geometry of the opening vector and the geometry of the walls. by such microstylolites could be part of the crack-seal cycle
Euhedral crystal terminations and botryoidal textures are dia- if fluid pressures decreased to less than lithostatic in part of
gnostic of growth into open space, and are useful in ore par- the cycle, causing the microvein to close and experience com-
agenetic studies for discriminating epithermal environments. pressional stress. These microstylolites may also be due to a
Euhedral crystal faces can be recognized by growth zoning in later deformation, unrelated to the microvein formation.
the crystal, which may be defined by fluid or solid inclusions. Laminated microveins have planar bands, often composed
Wilson (1984) suggests that considerable variation in crystal of phyllosilicates, sub- parallel to the margins. The bands
36 CHAPTER 3. DIFFUSIVE MASS TRANSFER BY SOLUTION
CHAPTER 3. DIFFUSIVE MASS TRANSFER BY SOLUTION 37

may form as inclusion trails where the opening vector has a metric antitaxial growth may result in a median line of inclu-
large component parallel to the margin (Cox 1987). In this sions along the centre of the microvein, unlike asymmetric
situation there is no distinction between inclusion trails and growth. Fibre widths can also be used to distinguish sym-
bands. metric growth (fibre widths increase symmetrically in two
opposite directions towards the wall rock) from asymmetric
Fibrous microveins are particularly rewarding for kin-
ematic interpretation. Fibrous fillings are common and form growth, in which the widths increase unidirectionally across
the microvein. Composite growth means both syntaxial and
by progressive microvein opening at a rate that can be
antitaxial protions in the same microvein (Fig. 3.17d). Non-
matched by crystallization of the filling. Fibres can grow
systematic growth (“ataxial” - Passchier and Trouw 1996)
by at least five different filling sequences: these need to be
histories may involve fracture at any point in the crystal
carefully established before kinematic interpretations can be
fibres, which are termed stretched crystal fibres by Durney
made. Syntaxial growth means growth from the wall rock
and Ramsay (1973). They exhibit none of the systematic fea-
towards the microvein centre on both sides of the fracture
tures described for the other four categories of growth above
(Fig. 3.17a, Plate 21). The diagnostic features of syntaxial
(Fig. 3.17e), and the sides of the fibres have distinctive in-
growth are two separate bands of fibres on either side of a
terlocking teeth, dividing then fibres into tablets (Plate 22).
central suture. The fibres are not continuous across the suture.
The lack of directional growth indicators (e.g. fibre widening
The microvein fill is similar to the wall rock and may be in
direction) in these microveins is diagnostic.
crystallographic continuity with it. Fibre widths generally in-
crease in the direction of growth due to competition between In many fibrous microveins, the fibres grow parallel to the
fibres, which ensures that the faster growing and therefore incremental opening vector: Such tracking or displacement-
larger fibres overgrow and eventually isolate the slower and controlled fibres can be used to deduce incremental strain
smaller fibres from the precipitating solution (e.g. Smith histories with great effect. They can be recognized because
1964). Therefore in syntaxial growth, fibre widths may in- fibres connect markers across the microvein, because they
crease towards the suture. By contrast, antitaxial growth oc- are parallel to inclusion trails (Urai et al. 1991), or because
curs from the microvein towards the wallrock, and may occur they have a constant orientation between microvein walls of
symmetrically on both sides of the microvein (Fig. 3.17b) or variable shape (Plate 23). The strain history deduced from
asymmetrically on only one side (Fig. 3.17c, 3.18). The dia- the fibres can be plotted on a diagram showing rotation of
gnostic feature of antitaxial growth is fibre continuity across the incremental strain on the horizontal axis against strain on
the microvein. Antitaxial growth is characterized by fillings the vertical axis (a cumulative incremental strain history or
of different material from the wall rock, and no crystallo- cish diagram, e.g. Fisher and Anastasio 1994, Hedlund et al.
graphic relationship between the filling and wall rock. Sym- 1994).
38 CHAPTER 3. DIFFUSIVE MASS TRANSFER BY SOLUTION

occurred. Inclusion trails may be used to distinguish track-


ing and non-tracking fibres in this case. Non-tracking growth
occurs because fibre growth directions are determined by the
orientation of the growth surface, like face-controlled pres-
sure fringes. Fibre boundaries are either perpendicular to the
growth surface (Fig. 3.19), or along the bisector of two ad-
jacent growth surfaces. The fibre boundaries have no fixed
relationship to the opening vector. The tracking efficiency,
or degree of match between the opening vector and the fibre,
is determined by the angle between the opening vector and
the growth surface and the shape of the growth surface
(Fig. 3.19, Urai et al. 1991). These fibres can only be par-
allel to the opening vector when when the tracking
efficiency is 1.
Fibres in microveins with any of the above growth histor-
ies are often curved. The curvature can be primary (formed
during crystal growth), or secondary (due to subsequent de-
formation). Primary curvatures can form in both tracking and
non-tracking fibres due to rotation of the incremental open-
ing direction with respect to the previously formed part of the
fibre, and can be recognized because the curved fibres show
no evidence of strain or recovery.
All low-temperature microvein fillings form by precipita-
tion from solution, and are therefore excellent evidence for
DMT via solution. Microveins are a conspicuous feature of
greenschist-facies and lower grade deformation because both
Non-tracking fibres do not track the incremental opening
cataclasis (fracture) and DMT via solution are required for
vector (e.g. Durney and Ramsay 1973). Non-tracking
their formation. At higher grades, other deformation mech-
fibres can be recognized because they do not connect
anisms operate, and microvein textures have low preservation
markers across the microvein (Fig. 3.19). However, markers
potential because of recrystallization.
may not be connected even by tracking fibres if a micro-
vein has an early history of shear displacement before filling
Chapter 4

Intracrystalline Plasticity

4.1 Introduction Dislocations move along planes in the crystal called slip
planes, in a direction within the plane known as the slip dir-
Permanent distortion of a crystal lattice without fracture oc- ection; this process is called dislocation glide. The combin-
curs by intracrystalline plasticity. The definitive feature of ation of the slip plane and slip direction is known as the slip
all intracrystalline plastic deformation mechanisms is the in- system, which is specified by the crystallographic orientations
volvement of dislocation motion (Section 4.2), but solid state of the slip planes and directions. For example, slip along the
diffusion (Chapter 5) is an integral part of some mechanisms prism planes of quartz in the < a > direction is annotated by
considered in this chapter. Intracrystalline plasticity has a Breaking bonds during glide requires en-
long history of research and a vast literature, particularly in ergy, which can be provided by heat, so the stress necessary to
the materials science field, which is necessarily rather con- cause glide decreases with increasing temperature. Slip sys-
densed in this chapter. Useful references that provide more tems in some directions are more active than in others because
explanation and detail are Hull (1975) and Nicolas and Poir- of anisotropy in crystal properties. The relative ease of glide
ier (1976). between different slip systems changes with temperature.
Glide is impeded by impurities in the crystal lattice, and
by the stress fields associated with other dislocations. These
4.2 Fundamental mechanisms of in- obstacles may be overcome by dislocations changing their
slip plane. Screw dislocations can accomplish this by a pro-
tracrystalline plasticity cess called cross-slip (Fig. 4.4). Edge dislocations can change
their slip plane if a lattice vacancy replaces the last atom in the
A dislocation is one type of imperfection or defect in a crystal half-plane of the dislocation (Fig. 4.4). This process of dis-
lattice. A useful way to understand dislocations is to imagine location climb therefore involves diffusion, although it is an
cutting into the lattice, stretching the lattice apart along the intracrystalline plastic deformation mechanism because dis-
cut, and inserting an extra plane of atoms into the cut. The location motion occurs. Deformation by a combination of
dislocation is the line along the edge of the extra plane of dislocation glide and climb is called dislocation creep, and
atoms (Fig. 4.1). The stretching of the lattice to accommod- occurs at higher temperatures, and lower strain rates than dis-
ate the extra atoms causes a distortion, the orientation and location glide (Nicolas and Poirier 1976).
size of which is measured by the Burgers vector which
can be perpendicular to the dislocation line (an edge disloca-
tion), parallel to the line (a screw dislocation) or oblique (a 4.3 Deformation twins
mixed dislocation) (Fig. 4.1). If has a magnitude of one
lattice unit, the dislocation is described as perfect, but par- A twin is a region of a crystal that is rotated or reflected with
tial dislocations have with magnitudes of fractions of lat- respect to the rest of the crystal (the host). Twins may form
tice units. The lattice distortion causes a stress field around during crystal growth or deformation: the latter are known as
the dislocation, and is one way of storing strain energy in the deformation or mechanical twins. Growth twins are generally
crystal. A direct image of dislocations can be obtained in the straight and of constant thickness, but deformation twins have
transmission electron microscope (TEM) where they usually variable thickness (thinning and branching towards the edge
appear as dark lines (Fig. 4.2). They can also be revealed of a crystal), and are commonly bent (Plate 24, Fig. 4.8). De-
by the technique of etching, in which a flat crystal surface formation twins are common in carbonates (where their mor-
is exposed to acid, creating a small depression (etch pit) at phology is temperature-dependent; Section 9.9.2), and feld-
the intersection of a dislocation with the surface of the crys- spars.
tal. The etching occurs due to dissolution which is enhanced Deformation twins form by shear of the crystal lattice with
by the lattice distortion and strain energy of the dislocation. respect to the host lattice along the twin plane, together with
The density of dislocations can be used to measure past stress minor rearrangements of the twinned lattice points. The
fields (Section 9.8.4). Movement of a dislocation is accom- twin plane is a mirror plane comprising an array of par-
plished by breaking bonds in the intact lattice ahead of the tial or twinning dislocations. The strain due to twinning
dislocation and re-forming the bonds behind the dislocation, can be measured, and deformation twins can also be used
causing it to advance through the lattice (Fig. 4.3). to measure stress and temperature during their formation

39
40 CHAPTER 4. INTRACRYSTALLINE PLASTICITY
CHAPTER 4. INTRACRYSTALLINE PLASTICITY 41

ilar to intracrystalline deformation bands, but have sharper,


more planar boundaries. The change in lattice orientation
across an intracrystalline deformation or kink band may be
visible from changes in extinction position (Fig. 4.7) or
changes in the orientation of twin or exsolution lamellae in
feldspars or calcite (Fig. 4.8). Kinks are different from twins
because the change in orientation across a kink plane is not a
fixed amount, and the kink plane is not a mirror plane.
Subgrains are seen under the microscope as areas within
grains of slightly different crystallographic orientation, separ-
ated by boundaries subparallel to crystal planes. The variation
(Sections 9.8.5, 9.8.9, 9.9.2).
in crystallographic orientation is visible in quartz as slight
changes in extinction position which affect discrete areas,
4.4 Undulatory extinction in contrast to the progressive change in extinction position
characteristic of undulatory extinction. Subgrains in quartz
Undulatory extinction is visible in cross-polarized light when are usually elongate parallel to the prism planes, which form
a single crystal has variable extinction positions (Fig. 4.5). the subgrain boundaries (Fig. 4.9). Tabular subgrains thus
The extinction position may change consistently from one have a similar geometry to kink bands. An alternative pat-
end of a crystal to the other, producing a continuous sweep tern in quartz consists of approximately square subgrains with
of extinction as the stage is rotated. Irregular patches of dis- boundaries parallel to both prism and basal planes, known as
tinct extinction positions can occur, and sectors of different a chessboard pattern (Fig. 4.10). Basal subgrain boundaries
extinction positions may radiate from the centre of the crystal. are only visible in grains with their c-axes subparallel to the
Undulatory extinction is caused by distortion of the crystal plane of the section, in contrast to prismatic subgrain bound-
lattice by dislocations of a consistent orientation (Fig. 4.6a). aries. Lattices with high dislocation densities possess a large
It forms at low strains during intracrystalline plastic deform- internal strain energy (Section 4.2), which can be lowered
ation, and as such it is a sensitive indicator of intracrystalline by dislocation movement into surfaces surrounding relatively
plastic deformation. dislocation-free volumes (Fig. 4.6). This process of recovery
results in a microstructure of low-energy volumes (subgrains)
surrounded by walls across which there is a slightly different
4.5 Intracrystalline deformation lattice orientation: hence the walls are sometimes called low-
angle boundaries. Undulose extinction, deformation and kink
bands, kink bands and subgrains: bands, and subgrains are a sequence of microstructures that
Recovery form with progressive strain, by generation and movement
of dislocations into low-angle boundaries (Fig. 4.6; White
Intracrystalline deformation bands are tabular low-strain do- 1976). The change in lattice orientation across the bound-
mains within crystals, separated from other parts of the crys- aries, and the dislocation density in the boundaries, increases
tal along approximately planar boundaries across which there with strain.
is a slight change in lattice orientation. “Intracrystalline” is Subgrain boundaries are orientated approximately perpen-
used to distinguish these smaller scale features from cata- dicular to the glide direction of the moving dislocations
clastic deformation bands (Section 2.5). Kink bands are sim- and therefore the orientations of the subgrain boundaries
42 CHAPTER 4. INTRACRYSTALLINE PLASTICITY
CHAPTER 4. INTRACRYSTALLINE PLASTICITY 43
44 CHAPTER 4. INTRACRYSTALLINE PLASTICITY
CHAPTER 4. INTRACRYSTALLINE PLASTICITY 45
46 CHAPTER 4. INTRACRYSTALLINE PLASTICITY
CHAPTER 4. INTRACRYSTALLINE PLASTICITY 47

walls made up of well-ordered arrays of two to three sets of


dislocations (Blenkinsop and Drury 1988, McLaren 1991).
The change in orientation across the subgrain walls is less
than 2°. The lamellae may have high fluid inclusion densit-
ies, and variable dislocation densities which suggest a highly
recovered structure (Blenkinsop and Drury 1988). These de-
formation lamellae probably formed by recovery of disloca-
tion slip bands, which leads to a variable decrease in dislo-
cation density and precipitation of water in fluid inclusions
(Drury 1993). Deformation lamellae can be used as a crude
paleopiezometer (Section 9.8.6). Planar deformation features
(PDFs) are different type of planar feature caused by shock
metamorphism (Section 8.4, 8.11).

4.7 Grain shape fabrics and ribbon


grains
One of the most characteristic microstructures formed by in-
tracrystalline plasticity are flattened or elongated grains with
a preferred orientation, resulting in a grain shape fabric, or
shape preferred orientation (Fig. 4.12). The effect of move-
ment of dislocations through a crystal is to change its shape
towards that of the strain ellipsoid. Apart from the distortion
may change with temperature. Most natural quartz subgrain of individual grains, intracrystalline plastic grain shape fab-
boundaries formed in the stability field are pris- rics can also form by coalescence of grains of similar orient-
matic, while in the stability field, both prismatic and ation during dynamic recrystallization (Section 4.8; Means
basal subgrain boundaries (chessboard pattern) form (Kruhl and Dhong 1982). Extreme strain can result in monocrys-
1996). These observations do not fully agree with exper- talline grains with very large aspect ratios known as ribbon
imental results, or with the suggestion that basal subgrains grains (Fig. 4.13); they may have intracrystalline plastic de-
are indicative of water-present deformation (Mainprice et al. formation features such as undulatory extinction or subgrains.
1986), but the observations are robust empirical evidence, al- Grain shape fabrics can also be produced by both cataclasis
lowing the occurrence of chessboard patterns to be used as a and DMT. Intracrystalline plastic grain shape fabrics can be
geothermobarometer (Section 9.9.4). The size of subgrains distinguished by their association with other intracrystalline
can be used as a palaeopiezometer (Section 9.8.3). plastic microstructures.

4.8 New grains, core and mantle struc-


4.6 Deformation lamellae
ture: Dynamic recrystallization
Deformation lamellae are crystallographically orientated
planar features wide (Drury 1993). Deforma- New grains, usually equant and strain free, are common
tion lamellae can be seen under the optical microscope us- around and within larger original grains in moderately or
ing a high magnification and a narrow diaphragm to en- highly strained rocks. The new grains may have a strong crys-
hance relief contrasts. In quartz, they have a slight extinc- tallographic fabric which is often systematically related to ad-
tion or refractive index contrast with the adjacent host grain jacent older grains. The new grains may be concentrated in
(Fig. 4.11), and they commonly have a sub-basal orienta- a mantle that partly or completely surrounds the older grains,
tion, sub-perpendicular to prismatic deformation bands (e.g. described as core-and-mantle structure (Fig. 4.14), and they
Spang and Van der Lee 1975, Drury 1993). Deformation may have similar sizes and orientations to adjacent subgrains.
lamellae are commonly slightly curved, and have a single, These features are characteristic of dynamic recrystallization,
consistent orientation within a grain. They are known from or recrystallization that occurs syntectonically. The lack of
both experimentally and naturally deformed rocks, especially strain in the new grains shows that they are at an advanced
in quartz but also in olivine and plagioclase (e.g. Den Brock stage of recovery. The new grains form by two main mech-
and Spiers 1991). The nature of deformation lamellae is enig- anisms: subgrain rotation (SGR), and grain boundary migra-
matic: observations include slip bands, walls of tangled dis- tion (GBM). SGR and GBM are structural transformations
locations, twin boundaries and planes of glass (e.g. McLaren which do not per se accommodate deformation, and therefore
et al. 1967, Christie and Ardell 1974, Twiss 1974). However, they are not strictly speaking deformation mechanisms (Urai
TEM studies have shown that typical deformation lamellae et al. 1986), but merely structural rearrangements.
in quartz consist of elongate subgrains wide with a Movement of dislocations into subgrain walls during re-
sub-basal orientation bounded by curved dislocation covery causes progressive rotation of the subgrains (see
48 CHAPTER 4. INTRACRYSTALLINE PLASTICITY

previous section) leading to the formation of a new grain


(Fig. 4.15). The collection of dislocations into the sub-
grain walls may occur by subgrain walls sweeping through
the crystal, or by movement of the dislocations towards the
walls, and some GBM probably accompanies SGR (Urai et
al. 1986). There is a continuum between subgrains and new
grains formed by this process, so that the definition of a new
grain as distinct from a subgrain is somewhat arbitrary. A
generally used criterion is that the new grain lattice orient-
ation differs from its host by more than 10°, but other val-
ues have been used (e.g. Urai et al. 1986). SGR is read-
ily inferred from the coexistence of new, smaller grains with
subgrains of similar size within the older grains (Fig. 4.16).
Other diagnostic microstructures for SGR include clusters of
new grains with similar orientations, which they inherit from
a single parent grain (Urai et al. 1986), and the coincidence
of one or more lattice directions in adjacent subgrains. GBM
occurs by the movement of a grain boundary from one grain
into another, and finally by the closure of the boundary to
isolate a new grain in a different lattice orientation from the
host (Fig. 4.17).
The grain boundary migrates in response to an internal
strain energy gradient from a less-deformed grain with a
lower dislocation density to a more highly strained one. The
final closure of the boundary may be achieved by an in-
termediate stage consisting of a bridging subgrain bound-
ary that accommodates progressively more misorientation
(Means 1981). GBM can be recognized by the presence of
CHAPTER 4. INTRACRYSTALLINE PLASTICITY 49
50 CHAPTER 4. INTRACRYSTALLINE PLASTICITY

new grains along grain boundaries, together with the serrated namic recrystallization in quartz is given in Table 4.1, and is
and lobate shapes of grain boundaries (Fig. 4.18, Plate 25). commonly applied to naturally deformed rocks.
The characteristic shape of these grain boundaries, also called
sutured grain boundaries, is caused by the fact that the
boundaries migrate away from their centre of curvature (e.g. 4.9 Crystallographic fabrics
Hirth and Tullis 1992). A bimodal distribution of grain sizes
may develop between the old and new grains (e.g. Lloyd A crystallographic fabric, or lattice or crystal preferred ori-
and Freeman 1994). New grains often develop where there entation (L or CPO) is a concentration of lattice orientations
is large lattice distortion and changes in lattice orientation, in one or a limited number of directions. Crystallographic
such as deformation bands and kink planes (Fig. 4.5). This fabrics can be recognized quickly under a optical microscope
occurs because grain boundaries migrate more rapidly at a by observing that many grains have the same extinction pos-
misorientation of about 5°. Drury et al. (1985) emphasize ition or interference colour. A useful way to detect the pres-
the importance of these special sites by proposing a third cat- ence of a crystallographic fabric in more detail is to insert
egory of dynamic recrystallization: sub-boundary migration, the sensitive tint plate or a quarter-wavelength plate, which
which is defined as subgrain growth in areas with large gradi- distinguish the different optical axes and therefore provides a
ents of strain and orientation. New grains also form in high more precise picture of the lattice orientations than extinction
stress sites around porphyroclasts. The size of recrystallized position or interference colour. This is particularly useful for
grains can be used as a paleopiezometer (Section 9.8.2). quartz because of its low order interference colours. Detailed
Dynamic recrystallization in naturally-deformed quartz oc- measurements of crystallographic orientation are made with
curs by either a mixture of 50% SGR and GBM (Fig. 4.19), the universal stage under the optical microscope, by electron
or by SGR alone (Drury et al. 1985). Both mechanisms may diffraction, channelling, and backscattering techniques under
be observed because they occur sequentially and cyclically the electron microscope, and by X-ray and neutron diffrac-
(Lloyd and Freeman 199la, b, 1994). Three regimes of dy- tion. Crystallographic fabrics can be produced in rocks by
namic recrystallization in quartz, depending on deformation at least four distinct mechanisms (e.g. Hobbs et al. 1976,
conditions, have been indentified from experiments (Hirth Mainprice and Nicolas 1989). The first two mechanisms, an-
and Tullis 1992). The Hirth and Tullis classification of dy- isotropic crystal growth (Chapter 5.3) and rigid-body rotation
CHAPTER 4. INTRACRYSTALLINE PLASTICITY 51

of grains in a flowing matrix, can be recognised because in- be understood through relatively complex models that allow
dividual grains are not deformed. A simple way in which for multiple slip systems and make additional assumptions,
intracrystalline plasticity can produce a crystallographic fab- such as strain compatibility between grains, uniform stress,
ric is shown in Fig. 4.20, in which a grain is deformed by the or viscoplastic self consistent theory, which minimizes stress
operation of a single slip system. As the grain rotates, the and strain differences from an average value (e.g. Wenk and
slip plane becomes orientated normal to the finite shortening Christie 1991). The work of Jessell (1988a, b, Jessell and
direction and the slip direction rotates towards a bulk shear Lister 1990) for quartz incorporates effects of both the intra-
plane (cf. Allison and LaTour 1977). Although this is a sim- crystalline plastic mechanisms referred to above, lattice rota-
plified case, it illustrates that lattice reorientation can occur tion and recrystallization. These models show that the dom-
by dislocation glide. Fourthly, crystallographic fabrics may inant slip plane does not necessarily align with the bulk shear
be created by recrystallization. The exact mechanism is un- plane, and the dominant slip direction is not necessarily paral-
clear, but experiments demonstrate that the mechanism of dy- lel to the shear direction, in contrast to the single slip system
namic recrystallization determines the type of LPO (Gleason model above. The simulations also show that the LPO is af-
et al. 1993). fected by the type of flow during deformation, the finite strain
A great deal of work has gone into theoretical attempts to magnitude and type, and the temperature, which determines
predict crystallographic fabrics, and to compare them with what slip systems are active. Crystallographic fabrics carry
experimental results and natural LPOs (see reviews by Law large amounts of structural information, and have been used
1990, Wenk and Christie 1991). The general development to analyze shear sense (Section 7.9), finite strain, and grade
of a crystallographic fabric by intracrystalline plasticity can of deformation.
Chapter 5

Diffusive Mass Transfer and Phase


Transformations in the Solid State
5.1 Introduction ture without bond breaking and with minor shape
change. An example is the phase transition in
Material removal, transport and deposition without fractur- quartz.
ing, lattice distortion or melting at metamorphic grades at
and above amphibolite facies suggest diffusive mass trans- 2. Martensitic-like transformations are coherent transform-
fer (DMT) in the solid state. The major evidence for solid ations involving dominantly shear strain. Important
state as opposed to fluid assisted DMT is provided by phe- geological examples include:
nomena that can be explained in terms of known solid state and
diffusion coefficients, including many metamorphic textures.
Recent experimental evidence suggests that a variety of trans-
formations, some involving DMT, may also be important de- 3. Coherent exsolution may involve both dislocation move-
formation mechanisms in the solid state (Section 5.9). There ment and DMT to effect crystallographic and chemical
rearrangements. Exsolution of clinopyroxene from or-
has been considerable recent discussion about the potential
thoenstatite is one of the best studied examples.
tectonic importance of superplasticity, which involves a com-
posite solid-state deformation mechanism: these microstruc- 4. Order-disorder transformations occur by disordering of
tures and mechanisms are described in the last section, 5.10. cation site occupancies, for example in Mg spinels.

5.2 Fundamental deformation mech- 5.3 Grain shape fabrics and ribbon
anisms of solid state diffusive mass grains
transfer and phase transforma-
A grain shape fabric of unstrained grains is an important mi-
tions crostructure in metamorphic rocks at higher grades (Plate 26).
The grain shape fabric can be produced by anisotropic grain
DMT in the solid state may occur though the crystal lat-
boundary migration recrystallization, although the details of
tice by the movement of lattice defects. This is known as
the mechanism are not clear (e.g. Jessell 1987), and by aniso-
volume diffusion, and the resulting deformation is Nabarro-
tropic crystal growth (e.g. Shelley 1989a, b). Both of these
Herring creep (e.g. Nicolas and Poirier 1976). Coble creep
processes are important DMT mechanisms.
is a different type of deformation resulting from solid state
Quart-mica rocks at higher metamorphic grades commonly
DMT through grain boundaries, which have different diffu-
have a distinctive microstructure of flat grain boundaries
sion characteristics from the grain interiors (e.g. White and
between quartz and mica parallel to the mica basal plane,
White 1981). Deformation by either or both types of creep
and quartz-quartz grain boundaries approximately perpendic-
is known by the general term diffusion creep. Diffusion oc-
ular to the mica flakes (Plate 27). A related effect is pin-
curs in response to gradients in chemical potential (Fick’s
ning of quartz-quartz grain boundaries at the end of mica
law) which may be created by variations in normal stress,
flakes. These microstructures attest to impeded grain bound-
internal strain energy, and grain boundary configuration (cf.
ary movement, and can be attributed to a greater surface en-
Section 3.2).
ergy between quartz and mica than between quartz and quartz
Solid state transformations can be subdivided into four
grains. Where one phase has a strong preferred orientation,
types (Green 1985, Kirby and Stern 1993), which may be im-
the impeding effect on grain boundary growth may create a
portant in deformation because they are sensitive to deviatoric
grain shape fabric in the other phase, which grows parallel to
stress. The transformed polymorph is related to the original
the shape fabric (Plates 26, 27).
by crystallographic rules in the first three types, which are
Monocrystalline quartz ribbon grains (Fig. 5.1) in high
known as coherent transformations.
grade gneisses may form by the above mechanism, and
1. Displacive transformations are changes in crystal struc- are characteristically free of internal structures (by contrast

52
CHAPTER 5. DIFFUSIVE MASS TRANSFER AND PHASE TRANSFORMATIONS IN THE SOLID STATE 53
54 CHAPTER 5. DIFFUSIVE MASS TRANSFER AND PHASE TRANSFORMATIONS IN THE SOLID STATE

to lower grade ribbons; Section 4.7, Fig. 4.13). Gower than pinning the crystal boundaries as described in Sec-
and Simpson (1992) proposed that the geometry of quartz- tion 5.3, which are then free to grow at a faster rate. This
feldspar grain boundaries in ribbon grains is largely con- has been distinguished as a separate type of grain boundary
trolled by a combination of dislocation creep and diffusion migration mechanism called fast or free grain boundary mi-
that results in a microstructure of straight quartz-feldspar gration by Urai et al. (1986).
boundaries perpendicular to the shortening, and cusps point-
ing in the extension direction. However, recently MacKinnon
et al. (1997) have proposed that the textures of some ribbon 5.5 Decussate texture
grains may form by filling of microfractures, based partly on
the evidence that grains adjacent to the ribbons appear to be Decussate texture consists of randomly orientated, interlock-
truncated by the ribbon boundaries, and that the ribbons con- ing elongate crystals (Plate 28). It is especially common in
tain wall rock particles that have very similar geometries to amphiboles and micas, and arises because of unequal growth
inclusion bands (Section 3.12). Some of these features are rates in different crystallographic directions. The growth
illustrated in Fig. 5.1. Both models account for many of the anisotropy modifies the ideal foam texture to favour grain
observed features of high-temperature ribbon grains. A pos- boundaries parallel to the faster growth directions, and may
sible approach to distinguishing the two mechansims may lie lead to randomly orientated acicular crystals for particularly
in careful examination of the ribbon tips, which could yield high anisotropies.
evidence for incipient microcracking or diffusion processes.

5.6 Porphyroblasts and inclusion


5.4 Foam texture, static and second- trails
ary recrystallization 5.6.1 Characteristics
Monomineralic metamorphic rocks, particularly at higher Porphyroblasts are single crystals grown during metamorph-
metamorphic grades, may have a distinctive microstructure ism with a larger size than the adjacent grains in the matrix.
consisting of approximately hexagonal grain sections with They are only widespread in rocks that have been at upper
straight or curved grain boundaries and strain-free grain in- greenschist facies or higher, and are most common in meta-
teriors, known as foam or granoblastic polygonal texture pelites or metabasites. Chlorite, chloritoid, biotite, garnet,
(Fig. 5.2). This microstructure is readily interpreted in terms cordierite, sillimanite, kyanite, andalusite, and staurolite are
of grain boundary migration and DMT in the solid state common porphyroblastic minerals.
driven by surface energy. The process of achieving a min- Porphyroblasts commonly contain inclusions, the most
imum energy configuration under hydrostatic stress is static common of which are opaque minerals, aluminium-rich
recrystallization. The lowest surface energy will occur for the phases such as sillimanite or spinel, quartz, zircon, apatite, ru-
smallest surface area to volume ratio. The minimum surface tile, and sphene. A porphyroblast with a very high density of
energy configuration for equal-sized, space-filling polyhedra inclusions is a poikiloblast. Inclusions may have a variety of
are rhomb dodecahedra (truncated octahedra). However, this textures that contain important microstructural information,
figure does not have an isotropic distribution of surface en- and need to be described carefully. The shape of individual
ergies, because angles between grain edges around four grain inclusions may be euhedral, platy, linear, or rounded. Inclu-
junctions are unequal. Grain edges and boundaries may curve sion trails are aligned inclusions that define a fabric which is
to allow all four grain edges to intersect in the same angle given the symbol where the refers to an internal fabric
of 109.5°, resulting in a structure of polyhedra with curved (compared to for fabric external to the porphyroblast).
faces, similar to liquid films and as observed in annealed al- usually has a two-fold rotational symmetry about an axis par-
loys and chromitites in ultramafic intrusions (Smith 1964). In allel to the length of the porphyroblast. may be straight
two dimensions, 120° triple junctions are common between or curved. It may be continuous and curve smoothly from
aggregates of the same phase. The surface area to volume the core of the porphyroblast to the rim; it may be sharply
ratio can also be reduced by an increase in grain size, which deflected or cut off along deflection and truncation surfaces
often accompanies static recrystallization, and is known as respectively, which are sub-parallel to more external parts of
Ostwald ripening or exaggerated grain growth. This micro- the inclusion trail.
structure can be useful in revealing a post-tectonic period of Special types of inclusion trails include snowball textures
relatively high temperatures. (Plate 29), which are spiral-shaped inclusion trails that curve
Grain shape fabrics can be produced even under static re- through more than 180°, with the total curvature decreasing
crystallization by at least two processes. An older fabric from the centre towards either end of the rotation axis (e.g.
can control grain boundary mobility by surface energy effects Powell and Treagus 1970, Busa and Gray 1992). Millipede
(Section 5.3) or a new mineral growth may overgrow a preex- texture consists of a straight, parallel which is deflected
isting fabric: this is known as mimetic crystallization. into at the porphyroblast margin. Helicitic texture consists
Relatively large, irregular crystals which commonly con- of folds defined by Inclusions may be crystallographically
tain inclusions of secondary phases are formed by the pro- controlled by the porphyroblast: this can give rise to a number
cess of secondary recrystallization. This occurs when second of distinctive textural zoning patterns such as sector zoning
phases become incorporated into the growing crystal, rather and re-entrant zones. The relation between and is very
CHAPTER 5. DIFFUSIVE MASS TRANSFER AND PHASE TRANSFORMATIONS IN THE SOLID STATE 55

5.6.3 Relationship to deformation


Porphyroblast textures can yield detailed information about
the relation between porphyroblast growth (and thus P-T con-
ditions) and deformation. This is best analyzed by looking
at the relationship between inclusions and which can be
classified into four non-interpretive categories (Fig. 5.3):

1. Inclusions are randomly orientated, is commonly also


curved around the porphyroblast (Fig. 5.3a).

2. is discontinuous with is often also curved


around the porphyroblast (Fig. 5.3b).

3. is continuous with but and have different


shapes or orientations (Fig. 5.3c).

4. is continuous with and and have similar


shapes and orientations (Fig. 5.3d).

These categories are objective descriptions, but also allow


straightforward interpretations based on the original ideas of
Zwart (1960, 1962), and the recent comprehensive treatment
of Passchier and Trouw (1996). The first category suggests
growth of a porphyroblast before deformation i.e. pretectonic
(e.g. Borradaile et al 1982, p. 438-441), but exceptions are
noted by Passchier and Trouw (1996). The curvature of is
due to deformation of the matrix around the more rigid por-
phyroblast (Plate 30).
The second category is called intertectonic by Passchier
and Trouw (1996), and indicates porphyroblast growth fol-
lowing the deformation event that generated but preced-
ing another event that formed (Fig. 5.4). The two deform-
ations that created and may be separated by other de-
formations that are not recorded by or (e.g. Johnson
and Vernon 1995). An intertectonic interpretation is reliably
demonstrated when is folded and is straight (Plate 31,
important: can be continuous with or truncated by e.g. Borradaile et al. 1982, p. 441, 453).
at the porphyroblast boundary. This forms the basis of the The third category indicates syntectonic porphyroblast
interpretation of inclusion trail patterns (Section 5.6.3). growth (Plate 32). A common example of different geomet-
ries assumed by and is a coarser grain size in than
which can be interpreted to show that the matrix was over-
grown by the porphyroblast relatively early and subsequently
underwent grain size reduction. Coarsening of towards the
5.6.2 Growth mechanisms margin of a porphyroblast indicates growth during prograde
conditions. A particularly complex pattern may be produced
Porphyroblasts are a mineral growth and therefore form by by overgrowth of one type of in a pressure shadow and an-
DMT. In many cases there is no evidence for a fluid phase other type of in a strain cap, both Si being formed in the
and it is assumed that diffusion was in a solid state. This same event (Shoneveld 1977).
can sometimes be demonstrated by the observation that the The fourth category indicates post-tectonic growth (Plate
most mobile components are those with the highest solid state 33). However, even in the case of concordant and with
diffusion coefficients. Porphyroblasts grow in response to no obvious differences in geometry, may be curved around
thermodynamic considerations including composition, tem- the porphyroblast, suggesting either that the last increment
perature, internal strain, and surface energy. Unusually large of deformation outlasted porphyroblast growth, or that low
porphyroblasts, especially those which contain high densities strain deformation occurred subsequently.
of inclusions, may have grown by secondary recrystallization Detailed analyses shows that some porphyroblasts can not
(Section 5.4). The size and distribution of porphyroblasts re- be accommodated into the above simplified scheme. Trunca-
flects a balance between the energy required for nucleation tion of within the porphyroblast may indicate two phases
and that for growth. High ratios of nucleation to growth ener- of porphyroblast growth and foliation development. Trunca-
gies will favour the formation of fewer and larger porphyro- tion surfaces can also be formed by post-tectonic overgrowth
blasts. of an early foliation which has been truncated by a strain cap
56 CHAPTER 5. DIFFUSIVE MASS TRANSFER AND PHASE TRANSFORMATIONS IN THE SOLID STATE
CHAPTER 5. DIFFUSIVE MASS TRANSFER AND PHASE TRANSFORMATIONS IN THE SOLID STATE 57

during a later deformation, or a later stage of a progressive de- ural examples are so far virtually undescribed. The
formation (Passchier et al. 1992). Bell et al. (1992) have ar- quartz phase transition can be viewed as Dauphiné twinning
gued that truncation surfaces form in the strain shadow at the on the scale of the unit cell, so that relict Dauphiné twins as
end of a porphyroblast during progressive deformation. The well as microfractures (Section 2.3.11) could constitute mi-
same interpretations may apply to deflection surfaces. Prob- crostructural evidence of this deformation mechanism (Kirby
lems of interpreting curved inclusion trails are discussed in and Stern 1993).
Section 7.8. Martensitic-like transformations can occur during cool-
ing under hydrostatic stress, but in the case of the
transformation, deformation-induced trans-
5.7 Reaction rims, relict minerals, formation can be distinguished by lack of twinning in the
coronas and symplectites clinoenstatite (e.g. Kirby and Stern 1993). The best known
microstructures of Martensitic-like transformation are those
Reaction rims are rims of altered mineralogy around grains. of the olivine-spinel system. “Fingers” or lobes of spinel
This common metamorphic texture testifies to DMT since grow into olivine parallel to in experiments on
metamorphic mineral growth requires mobility of compon- (Vaughan et al. 1984), and a similar phase transformation mi-
ents. Relic minerals are mineral grains that have been mostly crostructure of aragonite crystals has been observed growing
replaced by reaction rims. Reaction rims that completely sur- into calcite parallel to (Hacker and Kirby 1993). These
round a grain are coronas (Plate 34), which are more common microstructures are consistent with the theoretical analysis
in high grade rocks. Symplectites are lamellar or vermicular of Green (1985), which suggest that ellipsoids of the stable
intergrowths that are common in reaction rims (Plate 35). The phase should grow with long axes parallel to possibly co-
intimate mixture of the phases involved, the fine grain size alescing into ellipsoids separated by cusps. However, later
and the disequilibrium grain shape shows that diffusion was experiments on olivine showed lens shaped spinel inclusions
not possible over long distances. perpendicular to These can be accounted for by stress re-
distribution at a microscopic scale between stronger olivine
and weaker spinel (Green and Burnley 1989).
5.8 Chemical zoning Exsolution of Ca-clinopyroxene from orthoenstatite is sim-
ilar to the transformation described
Metamorphic minerals, especially garnets, amphiboles, above, with the addition of Mg, Fe and Ca diffusion. Cooling
pyroxenes and feldspars, commonly show systematic com- exsolution may be distinguished from deformation exsolution
positional variations from core to rim, or a relatively con- by the orientation of the exsolved phase (Champness and Lor-
stant core composition and a quite different rim composition. imer 1974).
These variations can be seen in the colours of some min-
erals (e.g. blue-green colour variations in amphiboles), but
are most accurately revealed by electron microprobe studies. 5.10 Superplasticity
Zoning is clear evidence for DMT, and may form in two dif-
ferent ways. Superplasticity was first used to describe the behaviour of
metals deformed in extension to large strains without failure
Growth zoning. Preferential partitioning of an element into
(e.g. Langdon 1982). The dominant deformation mechan-
a mineral during growth can cause growth zoning by de-
ism in these experiments is grain boundary sliding, with in-
pletion of that element within an equilibrium volume.
compatibilities between grains mainly relieved by solid state
This describes how growth zoning may occur at con-
DMT around grain boundaries. Unfortunately the term has
stant temperature and pressure: however, the P-T con-
been used in at least three different ways in a geological con-
ditions may change during mineral growth, which will text (Gilotti and Hull 1990):
also change the distribution coefficients for elements
between mineral and matrix. This is probably the most 1. As a deformation mechanism consisting of grain bound-
common way in which growth zoning is produced. ary sliding accommodated by DMT and/or intracrystal-
line plasticity.
Reaction zoning. After a crystal has formed, DMT may oc-
cur with neighbouring minerals in response to a change 2. As a set of deformation conditions and mechanical re-
in P-T conditions from the formation of the original sponses in rock. In particular, dependence of strain rate
metamorphic rock. Commonly the rims of minerals have on an inverse power of grain size, high temperatures, and
compositions that reflect lower metamorphic grades than proportion between strain rate and a low power of stress
the cores: the rims have equilibriated during the cooling have been regarded as characteristic (Section 9.4.3).
of a rock. Such zoning is known as retrograde zoning.
3. As a description of strain e.g. “continuous, homogen-
eous deformation to very large strain”. This “phenomen-
5.9 Solid state phase transformation ological” definition proposed by Gilotti and Hull (1990)
is not in widespread use, and the more common first geo-
microstructures logical definition will be followed here.
Experiments suggest that phase transformations under devi- Microstructures characteristic of rocks considered to have
atoric stress could have distinctive microstructures, but nat- been deformed by Superplasticity have been summarized by
58 CHAPTER 5. DIFFUSIVE MASS TRANSFER AND PHASE TRANSFORMATIONS IN THE SOLID STATE

Fliervoet and White (1995). They include: Several of these microstructures are ambivalent in their in-
terpretation. For example, Fliervoet and White (1996) de-
1. A very fine grain size (less than in quartzites, scribe an exceedingly fine-grained quartz mylonite which de-
more in carbonates) (e.g. Behrmann 1985). formed exclusively by dislocation creep with no grain bound-
2. Equiaxed grains with diamond or blocky shapes (e.g. ary sliding. A fine grain size is apparently a necessary but
Drury and Humphreys 1988). not sufficient condition for superplasticity. Crystallographic
fabrics may be developed during superplastic flow of fine-
3. Alignment of grain boundaries over several grains (e.g grained calcite rocks, with the implication that they could in
White 1977). principle become strong after sufficient strain (Schmid et al.
1987, Rutter et al. 1994). Fig. 5.5 shows an ultramylonite
4. An inverse correlation between finite strain and grain that shows some of the characteristic features of superplas-
size (e.g. Evans et al. 1980). ticity. The clear definition and recognition of superplasti-
5. A weak crystallographic fabric (Rutter et al. 1994). city in rocks is still a difficult matter. An important example
of superplasticity related to phase transformation may occur
6. An inverse correlation between dislocation density and when olivine transforms to spinel. This is known as the anti-
grain size (e.g. Behrmann 1985). crack theory of phase transformation faulting, which suggests
that olivine transforms into spinel in anticracks which local-
7. High dislocation densities and voids at grain triple junc- ize into faults within which grain boundary sliding occurs on
tions (e.g. White 1977). fine grained spinel (Green and Burnley 1989, Burnley et al.
8. Large mismatches between the crystallographic orienta- 1991). The theory accounts well for earthquakes that occur at
tions of adjacent grains (e.g. Fliervoet and White 1995). depths too great for frictional behaviour, and is well suppor-
ted by experimental evidence (e.g. Tingle et al. 1993).
9. Rotations between grains that can not be explained by
rotation axes perpendicular to known Burgers vectors
(e.g. Fliervoet and White 1995).
Chapter 6

Magmatic and Sub-magmatic Deformation

6.1 Introduction 40 to 60% (Lejeune and Richet 1995). The importance of the
CMF is further suggested by the observation that the max-
Identification of deformation microstructures and mechan- imum proportion of phenocrysts in volcanic rocks is 55-65%:
isms in rocks containing melt has several important tec- volcanic rocks with larger proportions of phenocrysts may not
tonic implications, notably for the problem of melt extraction be able to erupt because their viscosities are too high (Marsh
and in the interpretation of pluton ascent and emplacement 1981, Wickham l987).
mechanisms. However, microstructural criteria to distinguish A minimum melt proportion for magmatic flow can also
magmatic, sub-magmatic and non-magmatic deformation mi- be deduced from the critical packing density of crystals to
crostructures and mechanisms are not well established, and bring them into a coherent mass. The critical packing density
the distinction is best made using a combination of meso- depends on crystal shape, size, size distribution, packing ar-
scopic and microscopic evidence. The mesoscopic evidence rangement, and amount of compaction. Table 6.1 summarizes
is briefly described in Section 6.4. porosity at critical packing density, which is equal to the min-
imum melt proportion, for some combinations of these factors
in geometrical models, and estimates for magmas. Surface
6.2 Fundamental deformation mech- energies of the crystal and liquid phases may be important in
flow because they can affect melt distribution (e.g. Jarewicz
anisms and microstructures in and Watson 1984, 1985), and indeed an explicit relationship
rocks containing melt between the CMF and surface energy can be formulated (Ri-
ley 1990). A more fundamental parameter than melt fraction
6.2.1 Magmatic flow for determining the rheology of melt-laden systems may be
the contiguity (the fraction of grain surface area in contact
Flow of magma (i.e. melt and crystal phases) by transport with other grains), because contiguity affects surface energy
of rigid crystals is often regarded as the typical deformation
and the resistance to shearing at the average surface. A load
mechanism in melt-bearing rocks. Magmatic flow has been bearing framework of crystals breaks down at contiguities of
defined as flow by displacement of melt and rigid-body rota- less than 0.15-0.2 (Miller et al. 1988), which correspond to
tion of crystals without sufficient interaction to cause crystal variable equilibrium melt fractions depending on surface en-
plastic deformation (Paterson et al. 1989). As shown be- ergies and grain size and shape distributions (German 1985).
low, crystal interaction in melts may lead to deformation by These factors may explain the variation in estimates of the
cataclasis and diffusive mass transfer (DMT) as well as in- CMF.
tracrystalline plasticity. A more useful, general definition for Viscosities of suspensions depend on fluid (melt) com-
magmatic flow is flow of melt and crystals without crystal de- position, pressure, temperature, and proportion, size and
formation; this definition allows for the possibility that crys- shape distribution of solids. The viscosity of melts contain-
tals may be deformed by processes other than crystal plas- ing spherical crystals is often approximated by the Einstein-
ticity, and also describes the flow of a suspension. It leads Roscoe equation (Roscoe 1952):
naturally to a definition of magmatic microstructures as those
which indicate melt-present deformation without crystal de-
formation.
Some experimental studies suggest that viscosities of melt- where is the viscosity of the pure melt, and is the
laden systems reduce abruptly by orders of magnitude when crystal fraction for coherent packing. Values of 0.6 for
the proportion of melt increases beyond a value known as the seem to represent silicate systems well. n has a theoretical
critical melt fraction, CMF (Arzi 1978, Van der Molen and value of 2.5 for variably-sized spheres, which is also a good
Paterson 1979). The CMF is commonly taken to be 30%, but fit to data for silicate systems, but the viscosity-melt fraction
may be as much as 50% (Vernon et al. 1988) or as little as relationship has a slight temperature dependence which can
10-20% for gabbroic rocks (Nicolas et al. 1988). Experi- be allowed for by letting n vary between 2.0 and 2.5 with
ments on two silicate melts reported an increase in viscosity temperature (Lejeune and Richet 1995). Another relationship
of three orders of magnitude, and a change from Newtonian between viscosity and crystal fraction includes the effect of
to non-Newtonian behaviour, as melt fraction increased from grain size (Sherman 1968). At lower melt fractions, the rhe-

59
60 CHAPTER 6. MAGMATIC AND SUB-MAGMATIC DEFORMATION

ology of the magma may be at least partly controlled by the 6.2.3 Magmatic and sub-magmatic flow and
mechanical properties of the solid phases or by the factors af- rheology
fecting diffusive mass transfer (DMT) through the melt phase
(see below). It appears that two fundamental changes in deformation
mechanisms occur as melt fraction is increased from the
While some experiments and observations suggest a signi-
solid to liquid states, and these correspond to reductions in
ficant mechanical change at the CMF, other experiments cast
strength or viscosity. An order of magnitude decrease in
doubt on the existence of a large change in viscosity over a
strength occurs as melt fraction rises from zero to only a
narrow range of melt fraction (e.g. Rushmer 1995), and it
few percent melt, and solid-state deformation mechanisms
has been argued that the abrupt change in viscosity observed
change to melt-enhanced diffusion creep, with other deform-
in some previous experiments was due to a combination of
ation mechanisms (cataclasis, grain boundary sliding, and in-
temperature effects and increase in water contents of melts
tracrystalline plasticity) also possibly accommodating crys-
(Rutter and Neumann 1995). There may be a continuous de-
tal deformation. As melt fraction increases still further, sus-
crease in magma viscosity as melt fraction increases, and no
pension flow can occur, reducing strength by three orders
CMF, if the change in water content is allowed for. Yet other
of magnitude, and crystal deformation ceases. Figure 6.1
experiments suggest that the change in strength at the CMF
shows these two transitions in mechanism schematically, but
is due to dilatancy hardening (cf. Brace and Martin 1968) at
it should be emphasized that the values and range of melt frac-
low melt fractions (Renner et al. 1999).
tions over which the transitions occur, and the magnitudes of
the changes in viscosity, vary with magma composition, and
are presently the subject of some debate.
6.2.2 Sub-magmatic flow
6.3 Mesoscopic evidence for magmatic
Sub-magmatic flow can be defined as deformation involving
flow of melt and crystals with crystal deformation. Sub- and sub-magmatic flow
magmatic microstructures are accordingly those indicating
melt-present deformation with crystal deformation, and cor- It is important to use mesoscopic evidence in conjunction
respond to the pre-full crystallization fabric of Hutton (1988). with microscopic observations to distinguish magmatic, sub-
Crystal deformation by DMT through the melt phase is prob- magmatic and non-magmatic flow. Fabrics in magmatic rocks
ably the dominant deformation mechanism at low melt frac- can be defined by phenocryst alignment, matrix mineral shape
tions and lower strain rates. The thermodynamic consider- fabrics, compositional banding, enclave or schlieren shape
ations for DMT through melt are the similar to those de- fabrics, or magnetic anisotropy. Phenocrysts may be tiled or
scribed for DMT through solution in Section 3.2. Experi- imbricated (e.g. Blumenfeld 1983, Blumenfeld and Bouchez
ments show that strength decreases by an order of magnitude, 1988). The fabrics can have any shape from prolate to ob-
and intracrystalline plasticity changes to melt-enhanced diffu- late. The mere existence of any of these fabrics is not dia-
sion creep, when melt proportion increases from zero to only gnostic of melt-present flow, but detailed examination on the
3-5% (Cooper and Kohlstedt 1984, Dell’Angelo and Tullis mesoscopic scale may provide some strong indicators to dis-
1988, Dell’Angelo et al. 1987). The weakening occurs due tinguish magmatic/sub-magmatic from non-magmatic flow.
to enhanced grain boundary diffusion though the melt. Ex- The following criteria are proposed as diagnostic of magmatic
periments on rock analogues by Park and Means (1996) also or sub-magmatic flow because they suggest extremely large,
demonstrate the importance of this process in some systems at non-systematic strain gradients in an homogeneous rock on
low melt fractions. Surface energy considerations may be im- a meter scale, which are mechanically unlikely in the non-
portant in sub-magmatic deformation. Most diffusion creep magmatic state. Microscopic studies should be used to back
models assume isotropic surface energy, but measurements up these lines of evidence.
suggest an order of magnitude anisotropy in olivine, for ex- 1. Abrupt and non-systematic changes in fabric orientation.
ample (Cooper and Kohlstedt 1982). This may affect strength
by allowing a continuous film of melt along two-grain bound- 2. Abrupt and non-systematic changes in fabric intensity.
aries instead of limiting melt to three or more grain junctions,
as usually assumed (Hirth and Kohlstedt 1995). 3. Irregular, polyclinal, disharmonic or rootless folds in
CHAPTER 6. MAGMATIC AND SUB-MAGMATIC DEFORMATION 61

massive rocks that have no sign of mechanical aniso- compositions. These foliations may have formed in a short
tropy or shear surfaces (e.g. McLellan 1984). interval of cooling after the melt has been emplaced but be-
fore final solidification.
The relationship between igneous layering and fabrics can Phenocryst density and size distributions may be affected
also be used to suggest magmatic/sub-magmatic flow. For by magmatic flow. Interaction between phenocrysts in con-
example, Pons et al. (1995) describe centimetre-scale cycles centrations greater than 8% in a moving fluid creates a “grain
with sharp bases consisting of a layer of fine-grained am- dispersive pressure” which is proportional to the velocity
phiboles to a medium grained amphibole-plagioclase layer to gradient in the magma and is therefore greatest at the margins
a slightly porphyritic plagioclase-K-feldspar-quartz layer in of an intrusion (Bagnold 1954, Komar 1972a, b). Phenocrysts
alkali granites. The layers are parallel to the preferred orient- are concentrated into the centre of the intrusion, and may also
ation in any of these minerals, and the fabric never cross-cuts coarsen in this direction. The velocity distribution and pheno-
the layering. The layering and fabric can be explained by cryst concentration should be plug-shaped with a central re-
trapping of a mafic cumulate layer from a slowly convecting gion of high density and an abrupt decrease in density towards
magma, leaving an increasingly felsic rich magma to crys- the sides of an intrusion, due to the effect of phenocryst con-
tallize as the upper layers. By contrast, Paterson and Ver- centration on viscosity, and this is indeed observed in many
non (1995) discuss several examples of an apparently mag- natural examples (e.g. Ryan 1995).
matic foliation that cross-cuts both gradational compositional Discordance between fabrics in a xenolith and a surround-
changes and contacts between phases of different magmatic ing igneous rock is often interpreted as evidence for mag-
62 CHAPTER 6. MAGMATIC AND SUB-MAGMATIC DEFORMATION

matic flow, but this criteria is not diagnostic: fabrics may formed inclusions in annealed minerals, and by a granoblastic
be preserved in xenoliths during non-magmatic deformation texture in mineral aggregates (especially quartz) which them-
due to competence contrasts between the xenolith and sub- selves define a shape fabric. Primary igneous fabrics can also
solidus matrix. Sub- or non-magmatic fabrics on this scale be recognized by random spatial relationships between dif-
are demonstrated by deformation of individual crystals such ferent phases. By contrast, solid state deformation produces
as feldspar megacrysts. However, concordance between fab- higher frequencies of contacts between like grains because
rics within enclaves and the host rock can be evidence for new grains preferentially nucleate on the same phase (e.g.
magmatic/sub-magmatic flow if the enclaves can be inter- Ashworth and McLellan 1985).
preted as partially molten during deformation (e.g. Vernon
and Paterson 1993).
6.4.2 Crystallographic fabrics
Shear zones may develop during intrusion by magmatic,
sub-magmatic and non-magmatic mechanisms (e.g. Guine- Magmatic crystallographic fabrics are due to orientation of
berteau et al. 1987, Pons et al. 1995). Magmatic shear euhedral, inequant phenocrysts during flow. In mafic rocks,
zones can be identified by fabrics defined by unstrained ig- (010) crystal faces are parallel to the magmatic foliation in
neous minerals (e.g. Miller and Paterson 1994) or reorient- olivines, pyroxenes and feldspars (Benn and Allard 1989).
ation of a magmatic fabric in the shear zones (e.g. Pons et The [001] direction of olivine and clinopyroxene is parallel to
al. 1995). Non-magmatic shear zones in the latter study had the magmatic lineation, and [100] has been observed parallel
pressure shadows filled by epidote around megacrysts, and to the magmatic lineation in feldspars within gabbros and ton-
ribbon grains formed by intracrystalline plasticity in quartz. alites (Benn and Allard 1989). Such crystallographic fabrics
Magmatic shear zones were identified in the experiments of can be distinguished from fabrics due to intracrystalline plas-
Park and Means (1996) by analyzing movements of solid in- ticity by complete lack of evidence for intracrystalline plastic
clusions, but there was very little direct microstructural evid- microstructures (Chapter 4), and in the case of olivine, by the
ence to distinguish the shear zones from the unsheared walls. fact that high-temperature non-magmatic deformation leads
Criteria which are sometimes misused to demonstrate mag- to [100] parallel to the lineation rather than [001], as observed
matic flow include imbrication, which may occur in mylon- for magmatic fabrics (Benn and Allard 1989).
ites (Section 7.11.3), magnetic anisotropy, which is well
known in metamorphic rocks, and S-C fabrics, which may
form in either sub-magmatic or non-magmatic deformation 6.5 Sub-magmatic microstructures
(Blumenfeld and Bouchez 1988).
6.5.1 Grain shape fabrics
6.4 Magmatic microstructures Grain shape fabrics may be expected to form in sub-
magmatic flow by both rigid body rotation and crystal de-
formation by DMT or intracrystalline plasticity. The crystal
6.4.1 Grain shape fabrics
deformation combined with evidence of melt such as matrix
The typical microstructure of magmatic flow consists of a with igneous textures is diagnostic of sub-magmatic flow (e.g.
preferred orientation of euhedral phenocrysts, in an isotropic Quick et al. 1992). Localization of melt in structures that are
matrix that shows igneous textures (Plate 36). Grain shape coeval with the foliation constitutes good evidence for sub-
fabrics are commonly defined by feldspars and micas in felsic magmatic deformation: this sort of evidence is more readily
rocks, and may be defined by feldspar, olivine and pyroxene seen at outcrop scale. As for magmatic flow, static recrystal-
in mafic rocks (e.g. Benn and Allard 1989). This is diagnostic lization may give the matrix the false appearance of a primary
of magmatic flow if there is no sign of any other deforma- igneous texture.
tion mechanism. Since intracrystalline plastic deformation
microstructures are readily visible in quartz after only small
6.5.2 Intracrystalline plasticity
strains, the lack of deformation in quartz (no undulose ex-
tinction, subgrains, or kink bands) is a reliable criterion for Intracrystalline plasticity in quartz may be expected in the
magmatic flow. Lack of sub-solidus deformation can also be presence of melt under appropriate differential stresses (>1
checked from inclusions (e.g. rutile) in quartz: the inclusions MPa) and temperatures (700 to 800°C) on the basis of granite
should be undeformed themselves and randomly orientated and quartz flow laws (Rutter and Neumann 1995). The very
(e.g. Mitra 1976, Stel 1991). Other primary, undeformed common appearance of undulose extinction in quartz within
igneous features such as strongly euhedral crystals, igneous granites suggests that intracrystalline plasticity occurs during
(e.g. idiomorphic) zoning, growth twins, and ophitic texture sub-magmatic deformation (Plate 37). Subgrain formation or
can be used to demonstrate the absence of intracrystalline de- recrystallization of quartz has been taken as an indicator of
formation. A fabric in igneous rocks can only be confirmed as sub-magmatic deformation where an overall igneous texture
a magmatic microstructure by lack of any evidence for other is preserved and no other non-magmatic deformation event is
deformation microstructures or mechanisms, including any of known (e.g. Bouchez and Gleizes 1995). Deformation twins
the microstructures described in previous Chapters. and bent twins in feldspar may also hint at sub-magmatic
It may be difficult to distinguish magmatic deformation intracrystalline plasticity (Plate 38). However, these micro-
from non-magmatic deformation followed by static recrystal- structures on their own do not demonstrate that melt was
lization. Post-deformational annealing can be revealed by de- present during deformation.
CHAPTER 6. MAGMATIC AND SUB-MAGMATIC DEFORMATION 63

6.5.3 Diffusive mass transfer of a competent body sub-divided by fractures that are filled by
melt. Melt may form in pressure shadows at the ends of boud-
Despite the experimental evidence for the importance of melt-
ins, and fill faults (e.g. Quick et al. 1992). Quartz-feldspar
aided diffusion creep, microstructural evidence has remained
neosomes have been described accumulated under imper-
elusive, probably because melt films have almost no poten-
meable refractory layers such as amphibolite sheets which are
tial for preservation in the geological environment. However,
boudinaged, allowing the neosomes to rise into boudin necks,
experiments suggest some features that could be used: trun-
and forming a geopetal structure (Burg 1991). Segregations
cated, embayed, scalloped or overgrown grain boundaries in
in shear zones and along axial planes of folds in migmatites
igneous rocks may be analogous to some of the features dis-
are common. Localization of the melt in these structures in-
cussed in Chapter 3 that indicate DMT through a fluid phase
dicates that the melt was syntectonic. This sort of evidence
(e.g. Dell’Angelo et al. 1987, Park and Means 1996). The
has great relevance to the problem of extracting melt to form
potential importance of surface energy anisotropy in melt dis-
plutonic bodies (e.g. Wickham 1987).
tribution needs to be investigated and may have some micro-
scopic expression in the form of differences between crystal
faces.
6.6 Other microstructures
6.5.4 Cataclasis Park and Means (1996) introduced the term “contact melting”
to describe melting at contact points between grains observed
Paradoxically, some of the clearest sub-magmatic microstruc- in their experiments, and suggested that indented boundaries
tures involve cataclasis of the crystals, such as microfractures and truncation of growth zoning might be indicative of the
which are healed by melt (e.g. Hibbard 1987, Bouchez et al process in nature. Another suggestion from their experiments
1992, Karlstrom et al. 1993). Microfractures in plagioclase is that filter pressing and expulsion of melt might be recog-
can be demonstrated to have been filled by melt from the fol- nizable from layers characterized by intracrystalline plasti-
lowing criteria: city adjacent to layers of less-deformed or undeformed rock
1. The microfractures are intragranular. This allows that formed by crystallization of the expressed melt. Overgrowths
the plagioclase crystals were in contact with melt. of feldspar along low-stress boundaries have been considered
as indicators of magmatic or sub-magmatic deformation (Hi-
2. The microfracture filling is compositionally and crystal- bbard 1987), but they may equally be be formed by sub-
lographically continuous with the same phase in the ig- solidus DMT processes (Paterson et al. 1989).
neous groundmass of the rock.

3. The composition of the microfracture filling is compat- 6.7 Non-magmatic deformation


ible with the later stages of the igneous petrographic his-
tory of the rock. Plagioclase microfracture fillings may Non-magmatic microstructures reflect deformation without
have lower anorthite contents than their host crystals, any melt present, which has also previously been known as
consistent with progressive evolution towards a min- sub-solidus or solid state deformation. Non-magmatic is pre-
imum melt composition (Bouchez et al. 1992). The rela- ferred to “solid state” because it allows for the presence of
tion between quartz and plagioclase in the microfracture fluids other than melt. Two sort of evidence can be used to
fillings also suggests a residual melt: feldspars are on suggest that non-magmatic deformation has continued as part
the walls or tips of the microfractures. of the same deformation event shown by magmatic or sub-
4. Early crystals (e.g. biotite, sphene) are trapped within magmatic features. The first sort is evidence for high temper-
the microfracture fillings. ature deformation, e.g. prism <c> slip in quartz (recognized
by quartz c-axes with an orientation close to the lineation e.g.
Cataclastic microstructures observed in experiments con- Law et al. 1992, Lagarde et al. 1994), basal subgrain bound-
firm the potential importance of cataclasis in sub-magmatic aries (Section 4.6), or albite exsolution lamellae indicating
deformation in the experiments of Rutter and Neumann exsolution above the alkali feldspar solidus. A distinctive
(1995). Up to 10% melt, axially-orientated cracks formed grain boundary shape in quartz consisting of reticular grain
and filled with melt, and the sample was faulted. Between boundaries, resulting in a mosaic pattern, indicates crystallo-
10 and 45% melt, cataclastic flow occurred with pore col- graphic control of grain boundaries, extreme grain boundary
lapse. Axially orientated microcracks formed by high melt mobility and therefore high temperatures (Gapais and Bar-
pressures have been observed in granitic aggregates contain- barin 1986). However, a difficulty with this sort of evidence
ing 2-15% melt (Dell’Angelo and Tullis 1988). Connolly is that the high temperature deformation may be a later event
et al. (1997) have demonstrated that microcracking caused which is entirely unrelated to the magmatic deformation (Pa-
by volume increase during melting is a viable way to create terson et al. 1989).
permeable fracture and melt-pool networks in a muscovite- The second line of evidence is kinematic; if magmatic/sub-
bearing quartzite. The syn-kinematic experiments by Park magmatic microstructures are kinematically compatible with
and Means (1996) also recorded fracture, localized along a non-magmatic microstructures, they can plausibly be related
kink band boundary. to the same event. This type of evidence is probably more
On a larger scale, several cataclastic features are commonly reliable. One example is non-magmatic S-C fabrics record-
associated with melt in migmatites. Metatextites often consist ing the same shear direction and sense as magmatic/sub-
64 CHAPTER 6. MAGMATIC AND SUB-MAGMATIC DEFORMATION

magmatic flow (Blumenfeld and Bouchez 1988, Miller and Myrmekite is sometimes taken as evidence for non-
Paterson 1994). Another case is the formation of quartz rods, magmatic deformation, but this is unreliable as myrmekite
folds and boudins in a non-magmatic state along an intrus- can form by direct crystallization (Paterson et al. 1989). Hi-
ive interface, while the bulk of the intrusion was still mag- bbard (1987) has suggested that magmatic myrmekite can be
matic, as suggested by the fact that all these structures die distinguished by growth in dilational sites (“pressure shad-
out away from the interface (Stel 1991). The non-magmatic ows”) around phenocrysts, in contrast with tectonically in-
simple shear zones described by Ramsay (1989) have a tan- duced myrmekite which forms in volumes perpendicular to
gential extension direction around an intrusion: these can be the shortening direction (e.g. Simpson 1985).
related to ballooning strains during successive phases of ex-
pansion of the pluton.
Chapter 7

Microstructural Shear Sense Criteria


7.1 Introduction nite stretching axis, and the vorticity profile plane must be
perpendicular to the lineation. The vorticity vector in gen-
Shear sense can be defined formally as the rotation sense eral may lie at any angle to the maximum ISA, depending on
(sign) of the average angular velocity of material lines with the amount of transpression, and this relationship can change
respect to the directions of maximum and minimum stretch- across a single shear zone (Robin and Cruden 1994, Tikoff
ing rate, the instantaneous stretching axes, ISA (e.g. Hanmer and Greene 1997). In these circumstances, the shear direc-
and Passchier 1991). This rotation is also known as the shear- tion can only be established reliably by using shear sense in-
induced vorticity or internal vorticity (Means et al. 1980, Wil- dicators other than displaced markers, observed in a variety of
liams et al. 1994). Features such as pressure shadows, por- sections. The shear direction is within a plane across which
phyroclast inclusion trails and fibrous vein fillings may pre- the shear sense appears to reverse.
serve some information about the orientation of the ISA, and These complications do not apply to any of the examples
can be used to determine internal vorticity. However, shear used in this chapter, in which all illustrations are in the plane
sense is commonly used in a much looser way to indicate the perpendicular to the foliation and parallel to the lineation,
displacement sense of a fault or shear zone, usually specified which is the correct shear sense observation plane in these
by one or a combination of the usual terms for fault or shear cases. A convention used in many studies is to take the shear
zone movement i.e. normal, reverse, sinistral, dextral. In this plane as perpendicular to the plane of the illustration, and the
wider usage, shear sense is defined with respect to a shear shear direction as horizontal within the plane of the illustra-
plane, which is a useful external reference frame. Rotation tion, so that the shear sense can be specified by dextral (clock-
relative to this frame of reference is external vorticity, con- wise) or sinistral (anticlockwise). Unfortunately this conven-
sisting of the sum of internal vorticity and any rotation of the tion is not always explicitly recognized, sometimes creating
ISA with respect to the external frame of reference, or spin, the misleading impression of horizontal movement on dip-
which should be accounted for in a rigorous shear sense eval- slip faults or shear zones. The convention is used throughout
uation. this book, so that dextral or sinistral can be used to describe
Shear sense should only be analyzed in a section that is shear sense. Microscopic observations such as shown in the
perpendicular to the shear plane and parallel to the shear dir- plates in this chapter are often invaluable aids to shear sense
ection, which is known as the shear sense observation plane determination, especially in fine-grained rocks. This is be-
or the vorticity profile plane (Robin and Cruden 1994). The cause of the finer detail visible in thin sections, and because
shear direction is often assumed to be parallel to mineral the sections can be cut accurately in the shear sense observa-
stretching lineations or slickenlines, and the shear plane to tion plane, which may not be visible in outcrop.
foliations or fault planes; hence the shear sense observation Contradictory shear sense indicators are common in out-
plane would be parallel to the lineation and perpendicular crop or thin section. There are good theoretical reasons for
to the foliation (Fig. 7.1). Other sections can give incor- this, including the fact that material lines rotate in oppos-
rect shear senses when using marker displacements as shear ite directions under general shear (e.g. Passchier and Trouw
sense indicators. However, the assumption that mineral foli- 1996). Therefore it is often necessary to evaluate a large num-
ation and lineation are parallel to the shear plane and shear ber of shear sense indicators for reliability. This may mean
direction breaks down in two circumstances: when the total examining many thin sections, but a useful technique can be
shear strain is low, and in complex types of three-dimensional to cut a number of hand specimens on a rock saw in the shear
strain. Transpression with a high ratio of pure shear to simple sense observation plane. The shear sense may become visible
shear is one example of the latter. In this case, the maximum due to the smooth surface of the saw cut and because it is the
finite stretching axis (tracked by the mineral lineation) may correct observation plane, and the procedure allows rapid ex-
come to lie perpendicular to the shear direction (Fig. 7.2; amination of a large number of specimens. Specimens must
Tikoff and Fossen 1993). This may explain some puzzling be orientated correctly in the field as discussed by Prior et al.
cases where shear sense indicators are not seen in the plane (1987) or Passchier and Trouw (1996). Orientation errors can
parallel to the lineation and perpendicular to the foliation, but occur in the thin section laboratory, but errors can be checked
in planes perpendicular to the lineation (e g. the Larder Lake- if all rock fragments created during sectioning are carefully
Cadillac deformation zone, Canada; Robert 1989). The vor- preserved, so that they can be reconstructed to the pre-sawing
ticity vector in these cases must be near to the maximum fi- configuration.

65
66 CHAPTER 7. MICROSTRUCTURAL SHEAR SENSE CRITERIA

ing, as suggested by the observation that the rock has no fab-


ric outside the shear zone. A pre-shear foliation may not ex-
hibit this characteristic curvature, and is a less reliable shear
sense indicator.

7.3 Oblique foliations and shape pre-


ferred orientations
Oblique foliations are grain shape fabrics that maintain an ap-
proximately constant, acute angle to the shear plane across a
shear zone. The shear plane in this case is often represen-
ted by a compositional banding, so that the fabric consists of
compositional bands within which there is an oblique foli-
ation. The shear sense is given by the acute rotation from the
oblique foliation to the compositional banding (clockwise or
dextral in Fig. 7.3). Oblique foliations and the bands paral-
lel to the shear plane have been called and respectively
(Law et al. 1984). The geometry of oblique fabrics has strong
similarities with some S-C fabrics (see Section 7.5). The ob-
7.2 Curved foliation lique foliations described above are examples of shape pre-
ferred orientations, which can be used as shear sense indicat-
Many shear zones have a foliation that curves smoothly ors by a different method (Shelley 1995). The aspect ratio of
across the shear zone as shown in Fig. 7.1. The angle between the grain long axes is measured and plotted against their angle
the foliation trace and the shear plane decreases from about with respect to an arbitrary reference orientation. Frequency
45° or less to approach zero at the centre of the shear zone curves can be plotted for the highest and the mean aspect ra-
(Plate 39). The rotation of the foliation from the outside to tios, in class intervals of 5-10°. The angular distribution of
the centre of the shear zone (clockwise in Fig. 7.1) is the same aspect ratios is typically skewed, and the shear sense is given
as the shear sense of the shear zone (dextral). This criterion by two criteria: the rotation sense from the longer part of the
is one of the most common and reliable for deducing shear curve towards the highest aspect ratio orientation, and from
sense, and can be applied to shear zones on any scale. the mean orientation of all grains towards the orientation of
This sort of foliation follows the orientation of the XY the highest aspect ratio orientation.
plane of the finite strain ellipsoid, and the curvature is due Shape preferred orientations can result from passive be-
to increased intensity of finite strain towards the centre of haviour of grains that behave as strain markers (i.e. they
the shear zone (Ramsay and Graham 1970). Such foliations flatten and rotate towards the shear plane as described in
are referred to as strain-sensitive by Hanmer and Passchier Section 7.2), from anisotropic mineral growth (Chapter 5.3),
(1991) and constitute one of the most common and reliable or from rotation of grains behaving as rigid objects in a
shear sense indicators. An important caveat to the use of such fluid (e.g. Shelley 1989b). Development of shape preferred
foliations is that the foliation should be formed during shear- orientations may be counteracted by dynamic recrystalliza-
CHAPTER 7. MICROSTRUCTURAL SHEAR SENSE CRITERIA 67

tion of new equant grains; oblique foliations probably re-


flect a balance between shaping and homogenizing processes
which can lead to an equilibrium state or steady-state foliation
(Means 1981). These foliations are distinguished as strain-
insensitive because they do not track the finite strain ellipsoid
closely (Hanmer and Passchier 1991).

7.4 Porphyroclast systems


7.4.1 Characteristics and classification
A porphyroclast system comprises a porphyroclast with a
mantle and tails (or wings) of grains attached and derived
from the porphyroclast (Passchier and Simpson 1986). The
mineralogy of the mantle and tails may be the same as the por-
phyroclast (a “mantled porphyroclast”; Passchier and Trouw
1996), or may be related to the porphyroclast by a meta-
morphic reaction, such as a feldspar porphyroclast breaking
down to tails of quartz and mica.
Two tails are usually developed around a porphyroclast,
symmetrically related to each other by a 180° rotation about
an axis parallel to the shear plane and perpendicular to
the shear direction. Fig. 7.4 shows four categories of por-
phyroclast system geometry, three of which are named from
their similarity to the Greek letters: and (Passchier
and Simpson 1986, Passchier 1994). and have
monoclinic symmetry (Figs. 7.5-7.7), while have
orthorhombic symmetry. Complex types (Fig. 7.8) consist
of more than one tail on each side of the porphyroclast
(Passchier and Simpson 1986). The geometry of porphyro-
68 CHAPTER 7. MICROSTRUCTURAL SHEAR SENSE CRITERIA

clast systems can be described with respect to a reference sidered to develop by dynamic recrystallization of the por-
plane that passes through the centre of the porphyroclast par- phyroclast to form a mantle of finer grains around the clast,
allel to the planar part of the tail far from the porphyroclast, which subsequently become entrained in the flow of the mat-
and that contains a symmetry axis (Fig. 7.4; Passchier and rix. The geometry of tails depends on the rate of recrystalliz-
Simpson 1986). An important distinction is made between ation relative to the rate of rotation of the porphyroclast, the
in-plane tails that both lie on the reference plane, and stair- shape of the porphyroclast, the rheology of the tails and mat-
stepping tails that lie on opposite sides of the reference plane rix, the shear strain, and the degree to which the tails and mat-
(Fig. 7.3). types have been subdivided into which rix adhere to the porphyroclast (the coupling). Many of these
are found around isolated porphyroclasts in an homogeneous factors can be simplified by analyzing the relation between
matrix, and which are are found in S-C or the recrystallizing mantle and the “separatrix”, which is a
mylonites, with short tails that curve into C- or surface around the porphyroclast that separates closed from
(Passchier and Simpson 1986). tails were origin- open flow lines. The separatrix is either eye-shaped or bow-
ally defined by the criterion that an individual tail crossed tie shaped in simple shear, depending on a variety of factors
the reference plane near the porphyroclast, and therefore the including the rheology of the matrix. The shape of the tails
two tails stair-stepped (Passchier and Simpson 1986). How- is governed by the shape of the separatrix and the location of
ever,the characteristic feature of a is commonly taken the separatrix relative to the mantle. A few generalizations
to be a deflection near the porphyroclast which produces a can be made from the experimental work of Ten Brink and
feature known as an embayment (Fig. 7.9), so that systems Passchier (1995):
do not necessarily stair step. There is a great deal of variation
in the detailed morphology of porphyroclast systems, and all A mantle entirely within the separatrix will not develop tails.
gradations exist between and (Fig. 7.10).
A mantle that intersects the separatrix will first form
tails which then develop into tails, thus explaining
7.4.2 Mechanisms of formation why there are transitional types between the two end-
members.
The formation of porphyroclast systems is now understood
to some extent from experiments and numerical simulations A mantle that encloses the separatrix will first form
(e.g. Passchier and Simpson 1986, Passchier et al. 1993, tails, which develop into type tails for an eye-
Passchier 1994, Passchier and Sokoutis 1993, Bjørnerud and shaped separatrix, but remain as tails for a bow-
Zhang 1995, Ten Brink and Passchier 1995). Tails are con- tie shaped separatrix.
CHAPTER 7. MICROSTRUCTURAL SHEAR SENSE CRITERIA 69

tails will stair step in a bow-tie shaped separatrix, but 7.4.4 Faces of a tail
not in an eye-shaped one.
The asymmetry of an individual tail can be used to
define shear sense (Simpson and Schmid 1983). An indi-
tails only develop with high coupling. vidual tail has two faces in contact with the matrix
around the porphyroclast, one of which is approximately par-
allel to the shear plane, and the other which is curved and
The evolution of tails is incompletely understood at present,
oblique (Fig. 7.9a). Provided the face parallel to the shear
but these approaches offer the possibility that analysis of tails
plane can be identified, the curvature and obliquity of the
will yield considerable information on rheology and deform-
other face can be used to give the shear sense: it is convex
ation conditions.
in the direction of movement of the adjacent matrix, and the
There are three simple guidelines for using porphyroclast acute rotation of the curved face to the shear plane is in the
systems as shear sense indicators. It should be emphasized same sense as an oblique foliation. It is usually possible to
that shear sense can only be determined from tails with mono- check this criterion by stair-step direction, but it is sometimes
clinic symmetry, i.e. and tails. useful if only one tail is well-developed (Fig. 7.5).

7.4.5 Deflection and embayments of


7.4.3 Stair-step direction: and tails tails
The stair-step direction can be defined by taking an imagin- The deflection of a tail from the distal end towards
ary walk along one tail towards the porphyroclast in the shear the porphyroclast can be used to give the rotation sense of
sense observation plane: the stair-step direction is the dir- the porphyroclast (Passchier and Simpson 1986). The de-
ection of the step that needs to be taken over the porphyro- flection of the tail is in the opposite sense to the rotation of
clast to reach the opposite tail. Thus the tails on the left of the porphyroclast. For example, the deflection of the tails in
Fig. 7.4 step to the left. The stair-step direction defines the Fig. 7.4 is to the left as the porphyroclast is approached, giv-
shear sense unambiguously: the shear sense is the opposite to ing a clockwise or dextral rotation and sense of shear. An em-
the stair-step direction i.e. for a left stair-step direction, the bayment is the term for the approximately triangular region
shear sense is right-handed (dextral) and vice versa (Figs. 7.5 of matrix between a tail and the adjacent porphyro-
to 7.10). This is probably the most reliable way in which por- clast (Fig. 7.9b; Passchier and Simpson 1986). The interface
phyroclast systems can be used for shear sense determination. between the porphyroclast system and the matrix in the em-
70 CHAPTER 7. MICROSTRUCTURAL SHEAR SENSE CRITERIA

bayment has the shape of a fold: the direction of fold clos- nis and Secor 1987, 1990). Figure 7.11 shows the essen-
ure is the same direction as the rotation of the porphyroclast tial geometry of these fabrics and the angular relationships
(Figs. 7.7, 7.9b). between them, which are useful for description and classific-
Ideal conditions for the use of porphyroclast systems as ation. S-surfaces intersect C- or in a line perpen-
shear sense indicators are that the grain size of the matrix is dicular to, or at a high angle to, the shear direction, as defined
small with respect to the porphyroclast, the fabric of the mat- by the lineation on C- or surfaces. Angle is defined
rix is homogeneous, the porphyroclast systems were formed as that between the shear plane and the local orientation of
in one phase of deformation, the tails are long enough to S-surfaces between C or and angle is that
define a reference plane, and observations are made in the between the shear plane and which are inclined
shear sense observation plane (Passchier and Simpson 1986). to the shear plane in the opposite direction to the S-surfaces.
Porphyroclasts should be spaced widely enough not to inter- C-surfaces are parallel to the shear plane. In addition an angle
fere with one another. Structures similar to tails can develop can be defined, which is the angle between S- and C- or
in a passive matrix around a rotating porphyroclast, but these is equal to between S- and C-surfaces, and
have formed in a different manner from the porphyroclast sys- between S- and S-, C- and have
tems described above and should not be used in the same way a wide variation in morphology (Fig. 7.12, Plates 40, 41).
for shear sense determination (Section 7.10). Two major subdivisions of S-C mylonites were proposed by
Lister and Snoke (1984), but problems with this classifica-
tion have been pointed out (Passchier and Trouw 1996). An
7.5 S-, C- and alternative approach to description and classification (Blen-
kinsop and Treloar 1985) uses:
7.5.1 Characteristics and classification 1. The angles and

Many shear zones contain two or three planar fabrics that 2. The structures that define the S-surfaces,
can be divided into a sigmoidal penetrative foliation (an S- 3. The spacing of the C- or and
foliation), and discrete planar shears that displace the fo-
liation, which are known as C- or (Berthé et 4. The relative strengths of S- and C- or
al. 1979), extensional crenulation cleavage or ecc (Platt This classification is descriptive, yet uses features that have
and Vissers 1980), shear bands or shear planes (White et possible genetic significance. Three end members (porphyro-
al. 1980, Simpson 1986), or normal-slip crenulations (Den- clastic, megacrystic and banded types) identified using the
CHAPTER 7. MICROSTRUCTURAL SHEAR SENSE CRITERIA 71
72 CHAPTER 7. MICROSTRUCTURAL SHEAR SENSE CRITERIA

localized shear analogous to Y-shears and Riedel shears re-


spectively, as observed in simple shear experiments and fault
gouges (e.g. Rutter et al. 1986, Evans and Dresden 1991)
and in strike-slip fault systems (e.g. Sylvester 1988). Some
may form in the orientation of a Coulomb failure
surface. It is likely that anisotropy controls the formation and
the orientation of i.e. the value of (Platt and
Vissers 1980, White et al. 1980, Platt 1984, Passchier 1984).
If S-surfaces are parallel to the axes of the local finite strain
ellipsoid, then the value of should decrease with progress-
ive strain in these domains. However, C- and
probably do not not rotate with increasing strain. This is
suggested by the planar nature of C- and if sig-
nificant rotations occurred in a material with discontinuous
C- and strain incompatibilities would distort ini-
tially planar surfaces (Shimamoto 1989). If decreases and
scheme are shown in Fig. 7.13 and the relations between remains constant with progressive strain, then should also
and for the different fabric types are summarized in decrease, which has been reported in many studies (e.g. Zee
Fig. 7.14. This scheme seems to have wide applicability, et al. 1985, Burg 1987, Ghisetti 1987, Bal and Brun 1989,
based on other published descriptions. Scheuber and Andriessen 1990, Rykkelid and Fossen 1992),
but is not compatible with steady-state models (e.g. Dennis
and Secor 1987, 1990). Volume changes and “stretching”
7.5.2 Formation and evolution shear zones that extend parallel to their length may be import-
Variations in the values of and are controlled by strain ant in the formation of some S-, C-, and (Behrmann
and microstructure, and can be used to distinguish between 1984, Passchier 1991).
models for the formation of S-, C- and A com- Complications in the use of S-, C-, and as kin-
mon feature of several models is that S-surfaces are parallel ematic indicators have been pointed out by Behrmann (1987),
to the XY plane of the local finite strain ellipsoid (Ramsay but the general rules given below (Sections 7.5.3 and 7.5.4)
1967, 1980, Berthé et al. 1979, Lister and Snoke 1984, Blen- can be used for all fabrics. These relationships may be seen
kinsop and Treloar 1995). C- and are zones of particularly clearly in thin section.
CHAPTER 7. MICROSTRUCTURAL SHEAR SENSE CRITERIA 73

7.5.3 Curvature of S-foliation geometry that can be used in the same way as porphyroclasts
(Section 7.4.3). It may be difficult to distinguish pressure
Many S-foliations curve into like the curved foliation de- shadows from porphyroclasts (Plate 44). Rotation of the por-
scribed in Section 7.2, and the rotation sense from the S- phyroclast relative to the ISA and deformation of the pressure
foliation to the C- or gives the sense of shear re- shadow or fringe complicate this simple geometry. Despite
liably (Fig. 7.12, Plates 39, 40). the large variety of intricate structures that may be produced
(e.g. Etchecopar and Malavieille 1987, Aerden 1996) there
7.5.4 Shear on C- or are two reliable methods of determining shear sense from
more complex pressure shadows and fringes.
The sense of shear on C- or is always the same
as the overall sense of shear. It can sometimes be determined
7.6.2 Geometry of the last increment of growth
independently from the curvature of S-foliations by identify-
ing features such as veins that are offset by shear on C- or The last increment of growth may be virtually undeformed
and can therefore indicate the direction of the last maximum
ISA. In the case of antitaxial growth, the last increment will
be closest to the inclusion, while the opposite will be true for
7.6 Pressure shadows and fringes syntaxial growth. The sense of rotation from the last max-
imum ISA to the shear plane give the shear sense. Fibre ori-
7.6.1 Kinematics of pressure shadows and entation in pressure fringes can be used to indicate the ISA
fringes in shear zones only if the fibres are displacement-controlled.

The formation and geometry of pressure shadows and fringes


7.6.3 Shape
has been discussed in Chapter 3.9, and here it needs to be
re-emphasized that pressure shadows and fringes grow in the Fortunately a simple empirical rule for shear sense determ-
volume of low mean stress around the inclusion, which ap- ination is clear from computer simulations and natural ex-
proximately coincides with the direction of the ISA. There- amples, even without detailed considerations of the mechan-
fore in the simplest case of undeformed pressure shadows, ism of pressure shadow or fringe formation. The shape of the
the pressure shadow or fringe will be oblique to the shear whole pressure shadow or fringe, defined either by its median
plane, and the shear sense can be deduced from the acute line or by its enveloping surfaces, forms a doubly-inflected
rotation from the pressure shadow/fringe to the shear plane curve which is either an S or a Z shape. S shapes imply dex-
(Fig. 7.15). Some pressure shadows/fringes have a stair-step tral rotation of the core objects and therefore dextral shear
74 CHAPTER 7. MICROSTRUCTURAL SHEAR SENSE CRITERIA

sense, and vice versa (Figs. 7.16, 7.17). One complication


that may occur in the use of pressure shadows or fringes as
shear sense indicators is that the shadow/fringe may be ro-
tated with the inclusion (White and Wilson 1978). Pressure
shadows formed in the extension field may be rotated into the
shortening field and therefore give the incorrect shear sense.
Longer pressure shadows can come to resemble por-
phyroclast tails by this process (Hanmer and Passchier 1991).

7.7 Mica fish


Mica fish are a distinctive type of porphyroclast consisting
of elongated single crystals of mica, often associated with
tails (e.g. Lister and Snoke 1984, Hanmer 1986, Passchier
and Trouw 1996). Stair-stepping of the tails can be used as a
shear sense indicator following Section 7.4, and the sense of
the acute rotation from the long axis of the fish to the shear
plane also gives the shear sense directly (Plate 42). The form-
ation of mica fish is not well understood, but may involve
development of tails by dynamic recrystallization as well as
cataclasis, and rotation of the fish. Other minerals (e.g. horn-
CHAPTER 7. MICROSTRUCTURAL SHEAR SENSE CRITERIA 75

blende, Plate 43) may show the same geometrical features as understand the porphyroblast growth mechanism, especially
mica fish. whether the porphyroblast is in a domain of coaxial or non-
coaxial flow.

7.8 Porphyroblast internal foliations


7.9 Crystallographic fabrics
Curved foliations within porphyroblasts (internal foliations,
usually defined by inclusion trails) have been generally Crystallographic fabrics are usually represented by lower
considered to form by rotation and growth of porphyroblasts hemisphere, equal area stereographic projections of Crystal-
with respect to the matrix foliation and the flow axes dur- lographic directions of individual grains or sub-grains. It is
ing non-coaxial flow (Fig. 7.18a). However, Bell and cowork- standard practice to use the shear sense observation plane
ers (e.g. Bell 1985, Bell and Johnson 1989, Bell et al. 1992, as the projection plane, with the foliation as a vertical east-
Aerden 1995) have argued that many form by porphyro- west plane, and the lineation as a horizontal east-west line.
blasts overgrowing a new foliation that formed in a differ- This section focuses on the use of quartz c-axis Crystallo-
ent orientation relative to an external frame of reference, and graphic fabrics, which are readily obtained from the universal
that the porphyroblasts themselves do not have an external stage following the procedures in Turner and Weiss (1963) or
vorticity. The latter interpretation leads to the opposite sense Passchier and Trouw (1996).
of shear to the former (Fig. 7.18). A third possibility for inter- Quartz c-axes in non-coaxial shear are known to have an
preting curved is that a pre-existing foliation rotates during asymmetrical distribution relative to foliation and lineation
coaxial flow (Fig. 7.18b; Ramsay 1962). A fourth possibility from experiments, numerical models and natural examples.
for generating curved inclusions with continuous and is The asymmetry has been demonstrated as a reliable shear
by post-tectonic overgrowth of folds (helicitic structure). The sense criterion (e.g. Behrmann and Platt 1982, Bouchez et
last case can be distinguished by the presence of folds with al. 1983, Simpson and Schmid 1983, Law 1986, 1987, 1990,
a similar geometry to the inclusion trail in the matrix away Law et al. 1994) but important exceptions are known (e.g.
from the porphyroblast. Passchier 1983). Crystallographic fabrics should only be used
The above ambiguities are sufficient to preclude the use for shear sense determination when deformation is the res-
of porphyroblast inclusion trails as a shear sense indicator ult of intracrystalline plasticity, the foliation and lineation are
unless the mechanism by which they formed is well known. clearly defined and related to the finite strain, the deformation
It is important to know the regional kinematic framework to is homogeneous, and the mineral used for the determination is
76 CHAPTER 7. MICROSTRUCTURAL SHEAR SENSE CRITERIA

Quartz a-axes can also be used for shear sense determin-


ation, but they require X-ray measurement. Calcite c-axes
develop a girdle in non-coaxial shear that is inclined to the
foliation in the opposite direction to the quartz c-axis girdle
(e.g. Schmid et al. 1987, Wenk et al. 1987). Asymmet-
ric girdles in calcite determined by X-ray diffraction
have proven to be reliable shear sense indicators (Erskine et
al. 1993). Plagioclase, olivine and orthopyroxene asymmetric
dominant in volume (Bouchez et al. 1983). Crystallographic fabrics have also been used to determine shear sense (Mercier
fabrics should always be used in conjunction with other shear 1985, Gapais and Brun 1981, Passchier and Trouw 1996).
sense indicators.
The crystallographic fabric asymmetry is best defined by
the fabric skeleton, which are the lines that connect the max- 7.10 Asymmetric microboudins
imum concentrations of c-axes (Behrmann and Platt 1982,
Law 1987). Two types of asymmetry are distinguished: ex- Microboudins can be observed in thin section and share the
ternal asymmetry is defined by the obliquity of the central same characteristics as the mesoscale examples described in
girdle to the foliation (angle in Fig. 7.19; angle nomen- outcrop or experiments. Boudins in non-coaxial deformation
clature after Law 1987), and by the difference between the are asymmetric, due either to modification of initially sym-
angles and (Fig. 7.19). Internal asymmetry is defined metrical boudins (types I and II in Fig. 7.20; Hanmer 1986),
by the angles and between the arms of the crossed or to development of asymmetric boudins ab initio (type III
girdle and the central girdle (Fig. 7.19). The most important in Fig. 7.19, Goldstein 1988). The three types of asymmetric
shear sense criterion is the obliquity of the central girdle to boudin can be used as shear sense criteria according to the
the foliation (Fig. 7.19). The shear sense is given by the dir- following rules: Type 1 asymmetric boudins have two faces
ection of the acute rotation from the girdle to the shear plane. parallel to the shear plane, and two inclined to the shear plane
The difference between the values of and or between such that an acute rotation from the inclined faces to the shear
and can also be used (Fig. 7.19, cf. Law 1987, 1990, plane gives the shear sense. Types II and III boudins are sep-
Law et al. 1994). At high temperatures where prism <c> slip arated on shears that are inclined to the shear plane, and dis-
operates, a completely different type of fabric forms, consist- place the boudins in the same sense as the overall shear zone
ing of a point maximum of c-axes near the lineation. The (analogous to and Riedel shears, Section 7.5, Plate 6).
acute rotation from the point maximum of the c-axes to the However, the geometry of boudins in non-coaxial shear de-
lineation is in the opposite direction to the shear sense (Main- pends on the initial shape and orientation of the boudinaged
price et al. 1986). layer (Goldstein 1988). Boudins can be separated by shears
CHAPTER 7. MICROSTRUCTURAL SHEAR SENSE CRITERIA 77

antithetic to the dominant sense of shear, even when layers


lie in the extensional field (Fig. 7.21). “Bookshelf sliding”
also involves antithetic shearing. Therefore microboudins can
only be reliably used for shear sense determination if their
initial shape and orientation are known: this may be possible
from observations outside shear zones.

7.11 Asymmetric microfolds and


rolling structures
The vergence of asymmetric folds in shear zones is often the
same as the dominant shear sense (e.g. Fossen and Hoist
1995), but fold vergence opposed to the shear sense is also
well documented (e.g. Krabbendam and Leslie 1996). Op-
positely verging folds can be generated by shear of buckle
folds in a layer inclined in the shortening direction of non-
coaxial flow (e.g. Ramsay et al. 1983, Little et al. 1994) or
by heterogeneous simple shear (Fig. 7.22; Krabbendam and
Leslie 1996). Folds in simple shear zones may be highly non-
cylindrical, and this also complicates the use of fold vergence
for shear sense determination. Asymmetric microfolds, like
asymmetric microboudins, should not be used for shear sense
determination unless the way in which they formed is known.
Reliability can be enhanced when the vergence of several lay-
ers in different orientations can be observed, or when the ini-
tial orientation of the folded layer can be determined, for ex-
ample by observations outside the shear zone.
A special type of asymmetric fold is developed by the in-
teraction between rigid inclusions and layering. This type of
fold is localized around the inclusion, and has the same ver-
gence as the sense of shear. Such structures developed at high
shear strains by interaction between porphyroclast and matrix
are referred to as rolling structures (Van Den Driessche and
Brun 1987).

7.12 Shear sense criteria in rocks con-


taining melt
7.12.1 Magmatic shear zones
Shear planes and shear directions that existed during deform-
ation of melt-bearing rocks may be difficult to identify be-
cause they leave no microstructural imprint (e.g. Park and
Means 1996). Compositional layering may function as a
shear plane because of its rheological anisotropy (Benn and
Allard 1989). It has been suggested that magmatic flow
planes and directions may correspond to the average ori-
entation of planar and linear shape fabrics respectively (e.g.
Guineberteau et al. 1987). However, magmatic foliation
defined by megacrysts is not generally parallel to the shear
plane (e.g. Paterson and Vernon 1995, Yoshinobu and Pa-
terson 1996) because megacrysts rotate at variable rates de-
pending on their aspect ratios and the rheology of the matrix
(e.g. Tikoff and Teyssier 1995). Indeed the first shear sense
criterion suggested below is based on the obliquity between
grain shape fabrics and the shear plane. Possible dangers of
the interpretation of magmatic or sub-magmatic foliations as
78 CHAPTER 7. MICROSTRUCTURAL SHEAR SENSE CRITERIA

shear planes are pointed out by Tikoff and Teyssier (1995). 3. Clasts are passive markers, but trains do not rotate
(“March-fixed train" model).
7.12.2 Oblique grain shape fabrics In all cases, the imbrication of clast trains gives the shear
sense according to the rule that the acute rotation from the
Grain shape fabrics of olivine, pyroxene and feldspar in gab-
train long axes to the shear plane is the same as the shear
bros have been used to determine sense of shear (Benn and
sense. Individual clast long axes can not be used reliably as
Allard 1989) on the basis that they form a fabric at an acute
shear sense indicators because they may rotate past the shear
angle to the shear plane. The shear sense is given by the acute
plane in the first two models, even in simple shear. The results
rotation from the fabric to the shear plane. However, since
of numerical modelling show that clast interaction depends
elongate grains may rotate past the shear plane, this criterion
strongly on clast density, and that the March models (2 and
needs to be used with caution, and a large number of obser-
3), possibly corresponding to sub-magmatic and solid state
vations should be collected.
deformation, are more effective at imbricating clasts. There-
fore clast imbrication is not diagnostic of melt-present de-
7.12.3 Tiling and imbrication formation.
Tiled and imbricated clasts have been proposed as shear
sense indicators in rocks containing melt (e.g. Blumenfeld 7.12.4 S-C fabrics
1983, Paterson 1989), and as diagnostic of melt-present, non-
S-C fabrics have been described in rocks that have no evid-
coaxial deformation. Imbrication has been comprehensively
ence of solid state flow (Blumenfeld and Bouchez 1988, Benn
modelled by Tikoff and Teyssier (1994), who suggested three
and Allard 1989), and these fabrics can be used as a shear
possibilities corresponding to magmatic, sub-magmatic and
sense indicator as described in Section 7.5. Euhedral, little
solid state deformation respectively:
deformed plagioclase grains in S- and C-orientations have
1. Rigid clasts in a fluid, which form imbricated trains of been described by Miller and Paterson (1994), suggesting that
clasts that rotate as a unit when they have coalesced the S-C fabric formed in a sub-magmatic state.
(“Jeffrey rotating train” model, after Jeffrey’s (1922)
analysis of rigid objects in a fluid). 7.12.5 Sub-magmatic microfractures
2. Clasts behave as passive markers in a fluid; trains rotate Sub-magmatic microfractures can be used as a shear sense in-
as a unit (“March-rotating train” model, after March’s dicator. They are inclined to the shear plane in the same ori-
(1932) description of the rotation of passive markers in a entation as T fractures (Fig. 2.17); the acute rotation from the
fluid). March-like behaviour of rigid clasts is suggested fractures to the shear plane opposes the shear sense (Bouchez
by the experiments of Ildefonse and Manktelow (1993). et al. 1992).
CHAPTER 7. MICROSTRUCTURAL SHEAR SENSE CRITERIA 79

7.13 Shear sense criteria for faults been suggested as shear sense indicators by observations on
faults with known shear senses in specific conditions (e.g.
7.13.1 Shear sense observations on faults Petit 1987, Doblas et al. 1997). It is not clear how widespread
these features are, and many of them could have ambiguous
Determining the shear sense (sensu lato) of faults requires interpretations on faults with unknown shear sense. Until fur-
observations in the shear sense observation plane which can ther work is reported, the only generally reliable fault surface
only be identified if the net slip vector of the fault is known. kinematic indicator at present are risers created by slicken-
In practice this almost always requires the presence of slick- fibre terminations (or accretion steps), which are always con-
enlines to identify the shear direction within the fault plane. gruous (Fig. 7.23). When using this criterion, it is useful to
Incohesive fault rocks are particularly difficult to sample and examine fault surfaces in more detail under a binocular mi-
section. It may be possible to collect intact and orientated croscope such as those in common use for micropalaeonto-
specimens by removing them in a metal casing, and to pre- logy, or under the SEM. This technique is particularly good
pare them for thin sectioning by impregnation with adhesive. for dealing with incohesive specimens.

7.13.2 Displaced grain fragments 7.13.4 Gouges


Small displacements on faults can be measured by matching The characteristic structures of gouges shown in Fig. 2.17
fragments of distinctive grains in the cataclastic fault matrix can be used to identify shear sense on a microscopic scale,
to their parents in the walls of the fault, or by matching indi- including P-foliation, Riedel and conjugate Riedel shears, T-
vidual grains on either side of the fault (Fig. 7.23, Plate 45). fractures and ductile stringers.
The sensitive tint plate is useful for this task. However, the
sense of shear determined from fracturing of porphyroclasts
or boudins is ambiguous, because both synthetic and anti- 7.13.5 Jogs and bends
thetic fractures can form (Section 7.10). Compressional or dilational volumes are created at bends or
jogs in faults. On a microscopic scale, compression can often
7.13.3 Risers and slickenfibres be recognized by pressure solution, and dilation by creation
of open space or mineral deposition (e.g. Gamond 1987). The
Risers (Section 2.9) on fault surfaces have been used histor- shear sense can be determined from the sense of step and the
ically to determine shear sense, making the assumption that type of deformation: Right-handed bends or jogs are com-
they are congruous, but the occurrence of incongruous steps pressional for sinistral faults and dilational for dextral faults.
is well known. A variety of other fault surface features have The opposite rules apply for left-handed bends or jogs.
Chapter 8

Shock-induced microstructures and shock


metamorphism

8.1 Introduction
A number of distinctive natural microstructures have recently
been accepted as characteristic of shock metamorphism on
the basis of field observations, detailed microstructural stud-
ies, and comparisons with impact experiments. The products
and processes of shock deformation are distinct from other
geological microstructures and mechanisms, which justifies
a separate Chapter for their description. Shock effects have
well-established links with meteorite impact structures on
earth, which form under pressures ranging from <5 to > 100
GPa, temperatures up to 10 000 °C, and typical strain rates of
to (Fig. 8.1; Stöffler and Langenhorst 1994, Reimold
1995).
These pressures and strain rates are far greater than values
for tectonic crustal deformation (using the term tectonic to
imply endogenous processes). There has been vigorous de-
bate about which microstructures are diagnostic of meteorite
impact: this is discussed in Section 8.11.
a distinctive type of glass called diaplectic glass at moder-
ate pressures (25-50 GPa; 8.6), and a fused glass called le-
8.2 Shock mechanisms chatelierite at higher pressures (Section 8.8). Coesite and
stishovite, the high pressure polymorphs of silica, may be
Four fundamental shock mechanisms have been identified formed during shock metamorphism by solid state transform-
from experimental work that simulates shock conditions: ation or direct crystallization from melt (Section 8.7). These
cataclasis, intracrystalline plasticity, solid state phase trans- shock microstructures can be classified into high and low
formation, and melting (e.g. Huffman et al. 1993, Stöffler and pressure types (Table 8.1).
Langenhorst 1994). Cataclasis results in irregular, randomly
orientated microfractures, and sub-parallel sets of planar mi-
crofractures (Section 8.3). Intracrystalline plasticity is shown 8.3 Microfractures
by twinning and, possibly, shock mosaicism (Section 8.4), but
the formation of the latter is not well-understood. Non-planar, unorientated microfractures are common in
The most distinctive shock-induced microstructures and shocked materials. Planar, open microfractures (termed
mechanisms are associated with phase transformation at high planar fractures) parallel to low index crystallographic planes
compressive stress and with pressure-release melting. Quartz in quartz are a distinctive shock microstructure. They oc-
undergoes a phase transformation to an amorphous state in cur parallel to (0001) and planes, with a spacing of
which Si coordination with O changes from four-fold to six- (Stöffler and Langenhorst 1994). In experiment-
fold under pressures of approximately 5-35 GPa (Stöffler and ally shocked dunite, intragranular microfractures formed in
Langenhorst 1994). The phase transformation is heterogen- sets of variable orientation with separations of less than
eous at relatively low temperatures and high strain rates, and and more planar fractures with a spacing of
is localized on crystallographic planes leading to the form- formed in subregions of single crystals (Reimold and Stöffler
ation of one type of planar deformation feature (PDF, Sec- 1978). Their orientation is crystallographically controlled
tion 8.4). At higher temperatures and strain rates, the amorph- and a function of the direction of shock wave propagation,
ization is homogeneous. Quenching following melting forms and their density increases linearly with shock pressure in a

80
CHAPTER 8. SHOCK-INDUCED MICROSTRUCTURES AND SHOCK METAMORPHISM 81

given rock type. PDFs (next Section) do not cross the mi-
crofractures, which has been interpreted to indicate that the and PDFs have spacings from 1 to and thicknesses
fractures formed earlier than the PDFs during the shock event of under the TEM. Four different types of PDFs can
(Stöffler and Langenhorst 1994). be distinguished in the TEM (Goltrant et al. 1991):
1. Brazil twins in the basal plane.
8.4 Planar Deformation Features 2. Bands of high dislocation density, with concentrations
(PDFs) of voids or bubbles in rhombohedral planes.

Planar Deformation Features are intracrystalline zones of op- 3. Bands of variable proportions of glass and crystalline
tical and/or crystallographic contrast with the host crystal, quartz in the form of microcrystallites ~ 10 nm in size
typically on the order of micrometres wide and with spacings in rhombohedral planes.
of 2 to as seen under the optical microscope (Fig. 8.2,
4. Glass lamellae in rhombohedral planes.
Plate 46). Quartz is the mineral in which shock characteristic
deformation is best developed and has been most extensively The basal orientation of the Brazil twins (type 1, Fig. 8.3)
studied. PDFs in quartz occur in sets parallel to the crys- distinguishes them from Brazil growth twins which are on
tallographic directions given in Table 8.2 Other orientations rhombohedral planes, and their association with dislocations
include and As many as shows that they are mechanical twins (Leroux et al. 1994).
18 different sets may form in one grain. The orientations of Types 2 and 3 are thought to be secondary features due
PDFs in quartz can be related to the shock pressure as dis- to post-shock annealing of type 4 PDFs, which are recog-
cussed in Section 8.10. Planar features with the dimensions nized, together with basal Brazil twins, as characteristic of
of PDFs in shocked quartz are spectacularly revealed in SEM- fresh shock microstructures, and are the only type of lamellae
CL images (Seyedolali et al. 1997), though it is not known yet found in experimentally shocked material. The glass lamel-
certain that they correspond to PDFs. A fundamental distinc- lae can be subdivided into narrow transformation lamellae (<
tion can be made under the optical microscope between dec- 10 nm wide) which are only visible in the TEM, and wide
orated PDFs, which are marked by fluid inclusions or pores, transformation lamellae (50-500 nm wide) which are the op-
and non-decorated PDFs. tically visible lamellae, and which may contain a finer scale
It is necessary to turn to the TEM to study PDFs in more sub-lamellar structure of pillars at high angles to the lamellae
detail. About ten times more PDFs are visible in the TEM boundaries, which is clearly revealed by etching (Gratz et al.
than under the optical microscope (Goltrant et al. 1991), 1996).
82 CHAPTER 8. SHOCK-INDUCED MICROSTRUCTURES AND SHOCK METAMORPHISM

A number of models have been put forward for PDF form-


ation in quartz, which have largely been consolidated in the
work of Huffman et al. (1993), Langenhorst (1994), and
Huffman and Reimold (1996). Huffman et al. (1993) pro-
posed that PDFs formed by heterogeneous solid-state trans-
formation to amorphous silica along planes of progressively
lower compressibility, from basal to m, z, and a ori-
entations, with increasing pressure. Optically visible PDFs
are coalesced TEM-scale PDFs in domains between micro-
faults. Langenhorst (1994) considered the temperature pro-
files associated with shock fronts to show that, at relatively
low pressures (<25 GPa), melts were unlikely to form and
PDFs should consist of short-range order (quasi-diaplectic)
silica glass formed by solid-state transformation. At interme-
diate pressures (25-35 GPa), thin bands of melt may form in
addition to the solid state transformation, and at 35-50 GPa,
large-scale melting during compression may form diaplectic
glass. At >50 GPa, melting continues after shock compres-
sion to form lechatelierite by quenching under low pressure.
This model successfully accounts for the presence of both
amorphous silica in PDFs at temperatures too low for melting,
and diaplectic glass quenched from a melt. Fresh, impact- is uncertain whether these features are analogous to PDFs
induced PDFs are either Brazil twins or amorphous silica in quartz: recent TEM analyses of experimentally shocked
lamellae formed by solid state transformation or quenching zircon single crystals suggests that these features are micro-
of melt. cleavage, not PDF-type lamellar defects (Leroux et al. 1998).
Post-shock annealing may have several effects on PDFs. Zircon grains may also have a granular, polycrystalline tex-
The bands of high dislocation density in quartz (type 2, dec- ture of zircon crystals (“strawberry texture”), which has
orated PDFs) are likely to be due to post-shock annealing been observed in zircon from laboratory experiments and at
(Fig. 8.2). Annealing may be recognized by a distinct type several confirmed impact sites, including the Vredefort dome,
of fluid inclusion, which is typically much smaller (tens of but never in other circumstances: this texture may therefore
nm) than inclusions trapped during crystal growth (Goltrant be diagnostic of impact (Figs. 8.5, 8.6).
et al. 1991), although annealed PDFs may also contain in-
clusions up to diameter (Leroux et al. 1995, Leroux
and Doukham 1996). Fluid inclusions may form by diffusion 8.5 Mosaicism
along dislocations during recrystallization of PDFs, because
water is much more soluble in amorphous silica than in crys- Mosaicism is a pattern of domainal lattice misorientation seen
talline quartz. The tiny crystallites in the interior regions of as a mottled extinction pattern, which is distinctly different
PDFs (type 3 of Goltrant et al. 1991) are a characteristic an- from undulatory extinction, due to the sub-microscopic size
nealing texture (e.g. Langenhorst and Clymer 1996). of the domains (Plate 48, Stöffler and Langenhorst 1994).
Zircon, like quartz, is a favorable mineral to investigate Mosaicism can be described from spot or line broadening ob-
shock effects because of its transparency, lack of important served in X-ray diffraction (XRD) patterns from single crys-
twinning or cleavage, and uniaxial optical character. Fur- tal grains of quartz. As experimental shock pressures rise,
thermore, it is a very refractory mineral which is resistant the single crystal XRD pattern becomes increasingly similar
to alteration and thermal overprint. Shock features in zircon to a powder diffraction pattern. The technique can be used
can be seen after etching in the SEM (Bohor et al. 1993). to determine that domain size decreases from >3000 nm to
Planar features continuous across the whole grain occur with <200 nm with pressure. Single zircons with planar features
a spacing of or less (Fig. 8.4, Kamo et al. 1996). It can also show asterism (Bohor et al. 1993). The deforma-
CHAPTER 8. SHOCK-INDUCED MICROSTRUCTURES AND SHOCK METAMORPHISM 83

density (Table 8.3). Langenhorst’s (1994) model for the form-


ation of diaplectic glass suggests that it forms by quenching
of a melt before shock pressure is completely released. Its
distinctive properties are due to some long range order which
is inherited from the solid state. Diaplectic glass in natural
impact rocks recrystallizes during annealing to characteristic
textures which depend on the shock pressure reached. For
shock pressures less than 35 GPa, the glass becomes brown.
Microstructures of spherical single crystals, all in
the same orientation, called ballen, form at higher pressures.
At still higher pressures, the ballen have different crystallo-
graphic orientations (Grieve et al. 1996, Plates 47, 48).

8.7 High pressure polymorphs of


quartz - Coesite and stishovite
Coesite in some shocked samples appears as colourless-
brown, sized aggregates of individual grains less
than in size, typically twinned parallel to (010), often
enclosed by diaplectic glass or isotropic quartz with abundant
remnants of PDFs (Grieve et al. 1996). The aggregates may
be aligned on crystallographic planes. Coesite associated
with pseudotachylite veins from the Vredefort Dome is col-
ourless, with high relief, and very low birefringence (Fig. 8.7;
Martini 1991). It is either massive or forms radiating needles
up to long. Coesite has been observed at impact
craters and artificial explosion craters, but never in shock ex-
periments: this is attributed to the much shorter duration of
the pressure pulse in experiments compared to natural events
(Stöffler and Langenhorst 1994). Stishovite, the higher pres-
sure polymorph of silica, can be distinguished from coesite by
a higher birefringence and a brownish colour (Martini 1991).
It occurs in the same two habits described above for coesite
(Fig. 8.8). It appears in TEM as extremely fine-grained ag-
gregates parallel to PDFs in experimentally shocked quartz
(Ashworth and Schneider 1985). The difficulty of produ-
cing stishovite in shock experiments suggests that relatively
tion mechanism responsible for shock mosaicism is unclear.
long shock durations may be required, and that it may crys-
Although mosaicism appears similar to subgrain formation,
tallize from a high-pressure melt (Stöffler and Langenhorst
dislocations may not be mobile on the time scales of shock
1994). The preservation of stishovite appears to require rapid
events: therefore a direct comparison with subgrain forma-
quenching because it is transformed to amorphous quartz at
tion by recovery may not be valid (Stöffler and Langenhorst
relatively low temperatures.
1994). Remnants of planar features associated with mosa-
icism in quartz suggest that the mosaic domains may have
formed by the intersection of two sets of PDFs (Leroux and
Doukham 1996). Mosaicism has also been described from 8.8 Lechatelierite
other minerals, for example feldspars, olivine and pyroxene
(Hörz and Quaide 1973). A patchy optical extinction pattern Lechatelierite is a silica glass which may contain flow struc-
similar to mosaicism in feldspar can be formed by growth tures and vesicles, and may occur as veins within quartz with
zoning (Sharpton and Schuraytz 1989). PDFs. The optical properties of lechatelierite are summarized
in Table 8.3. The vesicles and flow structures demonstrate
that lechatelierite forms as a quench product from a melt un-
8.6 Diaplectic glass der low pressure, in distinction from diaplectic glass. Melting
to form lechatelierite requires pressures in excess of those for
Diaplectic glass is a distinctive type of silica glass that oc- whole rock melting, and therefore it is only preserved in veins
curs in homogeneous patches and associated with some types where pressures were locally increased by shock reverbera-
of PDF. It can be distinguished from either synthetic quartz tion, or as inclusions in whole rock melts or tectites (Stöffler
glass or lechatelierite (Section 8.8) by optical properties and and Langenhorst 1994).
84 CHAPTER 8. SHOCK-INDUCED MICROSTRUCTURES AND SHOCK METAMORPHISM

8.9 Tectites, microtectites and spher- MgO and ratios, and higher Cr, Ni, and Co (Koe-
berl 1990). These characteristics are similar to sediments.
ules
Tectites occur over large areas of the earth’s surface called
Tectites are rounded silicate glass bodies usually less than a strewn fields, demonstrating that they were deposited from
few centimetres in diameter (Glass 1990). They are usually the atmosphere. Terrestrial and lunar volcanism, and met-
black, but may be translucent, brown or green. Three major eorite impact, have been proposed as possible origins for
types are recognized: tectites, but a number of arguments favour the impact hypo-
thesis (Glass 1990). Flow structures demonstrate that tectites
1. Splash forms: spheres, disks, dumbell and teardrop were molten: the flanges of ablated forms were formed dur-
shapes. ing a second period of melting. Lechatelierite inclusions sug-
gest temperatures of 2000°C, far greater than volcanic tem-
2. Ablated forms, which are similar to splash forms but peratures. The chemical composition of tectites shows that
with an additional flange. they were derived by melting of sediments and not from ter-
restrial or lunar igneous rocks. Furthermore, the tectites in
3. Muong Nong types, which are larger, layered chunks of
two strewn fields can be linked to proven impact craters: for
tectite glass.
example, tectites in Czechoslovakia came from the Ries crater
Tectite glass has flow structures and abundant inclusions of in Germany, while the Ivory Coast strewn field is derived
from the Bosumtwi crater in Ghana.
lechatelierite. Baddeleyite, coesite, metallic spherules, and
corundum may occur as inclusions, as well as quartz, zir- Microscopic tectites (microtectites) have been described
con, rutile, chromite and monazite in the Muong Nong types from several sites at the Cretaceous-Tertiary boundary, and
(Glass 1990). The chemistry of tectites is similar to felsic vol- associated with impact at the Chicxulub site in Mexico (e.g.
canic glasses, with contents of >65%, but with higher Smit et al. 1992, Olsson et al. 1997). Many of these mi-
CHAPTER 8. SHOCK-INDUCED MICROSTRUCTURES AND SHOCK METAMORPHISM 85

crotectites are spherical, with diameters of 0.2 to 5 mm, and to mosaicism and PDFs, followed by diaplectic glass, coes-
composed of clays or calcite. Some of these spherules contain ite and stishovite, and lechatelierite at the highest pressure.
a core of glass, suggesting that the clays formed by devitrific- However, the accurate pressure calibration of this general se-
ation. Flow structures in the glass, and associated shocked quence is subject to the effects of a number of other variables
grains, demonstrate the impact origin of these spherule beds. (8. 10.3) which are not yet quantified with the exception of the
initial temperature of the target rocks and porosity.
Figure 8.9 shows onset pressures for shock microstructures
8.10 Shock barometry and thermo- in quartz and feldspar based on experiments on single crystals
of quartz or non-porous crystalline rocks. The lines showing
metry decrease in onset pressure with temperature were determined
by experiments on granite and quartzite (Reimold and Hörz
Establishing shock pressures has fundamental implications
1986, Huffman et al. 1993, Huffman and Reimold 1996).
for establishing the origin of shock microstructures and is
Microfracturing occurs at the lowest pressures: planar micro-
useful to estimate the rate of shock pressure attenuation for
fractures form at pressures greater than 5 to 10 GPa (Huffman
modelling impact processes. Shock pressures have been
et al. 1993). Mosaicism in quartz is first seen from pressures
calibrated experimentally using microstructures, the density
of 6 to 10 GPa. The onset pressures for PDFs in these experi-
of quartz, and the optical, X-ray diffraction, spectroscopic
ments fall between 15 and 18 GPa, but this is well above fig-
and thermoluminescent characteristics of quartz (Stöffler and
ures given by other workers, who suggest formation of non-
Langenhorst 1994, Grieve et al. 1996). Shock temperatures
basal PDFs (i.e. not Brazil twins) at pressures as low as 10
can be calculated in non-porous rocks from shock pressures
GPa. These lower pressures are shown on Fig. 8.2 by labels
(e.g. Langenhorst 1994). In the following Sections, two
giving the orientations of the dominant PDFs (after Stöffler
methods of shock wave barometry using optical microscopy
and Langenhorst 1994, Grieve et al. 1996). The nature and
(microstructures and optical properties of quartz) are briefly
orientations of PDFs in quartz changes as a function of pres-
described.
sure. Brazil twins form at the lowest pressures, followed by
PDFs. PDFs are generally considered to form at pressures
of ~20 GPa, and become the dominant orientation at ~25
8.11 Calibration of shock pressures GPa (e.g. Langenhorst and Deutsch 1994). Widths of experi-
from microstructures mentally created PDFs at the TEM scale change in thickness
from <100 nm at <25GPa to ~200 nm at >25 GPa. Ini-
Microstructures form with increasing shock pressure in a tial (i.e. target) temperature has an effect on PDF develop-
sequence that has been clearly established by experimental ment. Approximately equal numbers of grains show single
work. In quartz, from the lowest pressure, fracturing evolves and multiple sets of PDFs in experiments at 18 GPa and 250C
86 CHAPTER 8. SHOCK-INDUCED MICROSTRUCTURES AND SHOCK METAMORPH1SM
CHAPTER 8. SHOCK-INDUCED MICROSTRUCTURES AND SHOCK METAMORPHISM 87

(Huffman and Reimold 1996). The number of grains with


PDFs decreases as temperature increases, and multiple sets
are preferentially eliminated; at 750°C, there are very few
grains with single sets of PDFs in samples shocked at this
pressure. The shape of the PDFs also changes from narrow
with sharp boundaries at low temperatures to broad with wavy
boundaries at higher temperatures. There is some evidence
for a change from dominantly to orientations as temper-
ature is increased from 25°C to 440°C (Huffman et al. 1993).
Quartz phase may affect the character of PDFs: samples
shocked at 20-25 GPa and 540°C showed dominantly
PDFs, compared to a broad range of PDF orientation at
630°C, corresponding to the change from to sta-
bility in the target rocks (Langenhorst et al. 1992, Langen-
horst and Deutsch 1994). Experimentally shocked and
are also distinguished by the observation that positive
and negative planes of the same form never coexist in samples
shocked at less than 573°C, the transition tem-
perature, as expected from the trigonal symmetry of
This is in distinction from PDFs parallel to the hexagonal pyr-
amids of samples shocked in the stability field, and
has been used as a geothermometer for the pre-shock tem- from Fig. 8.7. The effect of temperature on the high pres-
perature of the target rocks: for example the temperature of sure phases is not well known, however. Microstructures in
the Ries crater and Eocene-Oligocene impact targets can be feldspar show a similar pattern with the notable difference
constrained to less than 573°C (Engelhardt and Bertsch 1969, that mosaicism occurs at considerably lower pressures. The
Langenhorst and Clymer 1996). best pressure estimates are given by the coexistence of two or
more microstructures.
A pressure calibration for porous rocks (Table 8.4) has
been established from work on the Barringer Crater in Ari-
zona (Kieffer 1975, 1981, Kieffer et al. 1976). In general, 8.11.1 Calibration of shock pressures from op-
the onset pressure for any microstructure decreases with in- tical properties of quartz
creasing porosity, because stress concentrations can occur on
The refractive indices and birefringence of quartz at room
the grain scale. PDFs seem to be less abundant in porous
temperature decrease continuously from samples shocked at
targets than crystalline ones, and there is some evidence that
25 GPa to those shocked at 35 GPa, which corresponds to the
different orientations occur in porous rocks:
transition to diaplectic glass (Fig. 8.10). The refractive index
and may be important (e.g. French et al. 1974). A
of diaplectic glass decreases slightly as pressure increases fur-
sequence of microstructures similar to those in quartz and
ther to 50 GPa. The effect of increasing pre-shock temperat-
feldspar with increasing shock pressures has been established
ure is to restrict the pressure interval over which the decreases
in olivine and pyroxene. Planar fractures and mosaicism de-
occur: at 630°C, the change occurs between 25 and 26 GPa
velop in olivine at 10 to 15 GPa, intergranular brecciation
(Stöffler and Langenhorst (1994).
at 30 GPa, and glass, annealing and melting occur above 50
GPa. In pyroxenes, twinning occurs at 5-10 GPa, mosaicism
and PDFs at 20-25 GPa, and glass, melting and recrystalliza- 8.11.2 Problems of shock barometry
tion above 75 GPa (Huffman and Reimold 1996).
Experiments have also shown that strain rate, pulse duration,
Details of the techniques for measuring PDF orientations and rate of pressure increase may also be important variables
are described in Engelhardt and Bertsch (1969), Stöffler and that should be considered in an accurate calibration of shock
Langenhorst (1994) and Grieve et al. (1996). There are vari- pressure (e.g. Huffman et al. 1993, Huffman and Reimold
ous methods of assigning an average shock pressure to ori- 1996). Different calibrations between experiments with dif-
entation measurements from a sample, which should consist ferent pulse lengths suggest that this is a factor which has not
of as many grains as possible. According to the method of been adequately explored, and raises the problem of extra-
Robertson (1975), individual grains are assigned the value of polation from experiment to nature (Huffman and Reimold
the highest pressure PDF observed within the grain accord- 1996). Grain size, grain size distribution, porosity and modal
ing to some experimental calibration. The shock pressure is mineralogy may also have effects (e.g. Robertson and Grieve
taken as the average of individual grain pressures. Variations 1977, Reimold and Stöffler 1978, Grieve et al. 1996).
on this method which account, respectively, for grains with Basal PDFs appear to be more resistant to post-shock an-
no PDFs and grains converted to diaplectic glass, have been nealing than other types of PDF (Leroux et al. 1994), render-
proposed (Fel’dman 1994). ing shock pressure calibrations from PDFs unreliable where
Mosaicism, the first appearance of diaplectic glass, 100% there has been post-shock annealing. Other barometers may
diaplectic glass, and lechatelierite, as well as the existence be questionable in this situation: for example, diaplectic glass
of coesite and stishovite can all be used as quartz barometers is a metastable phase and may not be preserved. Stishovite
88 CHAPTER 8. SHOCK-INDUCED MICROSTRUCTURES AND SHOCK METAMORPHISM

inverts in only 3 days at 250°C in the presence of water, and in impact rocks, so that the third criterion is not general. Mul-
thus has limited preservation potential (Gigl and Dachville tiple, intersecting sets of tectonic lamellae are also known
1968). (e.g. Lyons et al. 1993), but these lamellae are distinctly dif-
ferent to PDFs, mainly because they are non-planar (Reimold
1994). Several studies of PDF orientations in shocked targets
8.12 Diagnostic impact microstruc- show that criterion 4 is false: in particular, shock-induced
basal PDFs (Brazil twins) are common at low shock pres-
tures sures. Straightness (planarity), and sharpness or definition re-
main important characteristics for identifying PDFs, but have
The identification of microstructures that are diagnostic of
not been quantitavely defined: this lack of objectivity has
impact has profound implications for the understanding of the
evolution of the earth. The subject has been very controver- caused considerable controversy in the past. Optical studies
alone appear to be incapable of fully distinguishing impact
sial, with much debate on the extent to which shock-induced
from tectonic lamellae.
microstructures could be produced by volcanic eruptions or
other endogenous processes as alternatives to meteorite im- SEM and TEM-scale features do allow a clear distinction
pact. between fresh PDFs and tectonic deformation lamellae. Etch-
PDFs have been associated with shock metamorphism on ing with HF liquid or vapour followed by SEM examination
the basis of experiments and observations of their distribu- is a more widely available technique than the TEM (Gratz et
tion in suspected impact target rocks (e.g. McIntyre 1962, al. 1996). Relatively wide, glass-filled PDFs from impacts
Carter 1965, Hörz 1968, Robertson and Grieve 1977). How- and shock experiments are revealed as sharp, planar features
ever, PDFs can be superficially similar to deformation lamel- less than wide after etching, often containing sublamel-
lae formed by tectonic processes as described in Section 4.6, lar structure (see Section 8.4). By contrast, etched tectonic
and this raises the controversial questions of whether PDFs lamellae are much wider than and often sinuous, similar
can be diagnostic of shock metamorphism, and whether shock to their appearance under the optical microscope (e.g. Joreau
features can be produced in volcanic eruptions. et al. 1997; Fig. 4.11). Table 8.5 summarizes the differences
Commonly used criteria for identifying impact-induced between shock-induced PDFs and deformation lamellae in
PDFs include the following (e.g. Alexopoulos et al. 1988): quartz. Post-shock annealing reduces PDFs to a dislocation
structure similar to deformation lamellae, making the distinc-
1. Well-defined, sharp features. tion between the two difficult (cf. Goltrant et al. 1992). Basal
Brazil twins have only ever been observed in quartz from
2. Straight. impact structures and in experiments with deviatoric stresses
over ~4 GPa at room temperature (Leroux et al. 1994). These
3. Often occur in multiple sets per grain. observations were used by Leroux et al. (1994) to argue that
basal Brazil twins are strong evidence for shock.
4. 90% are in the and directions.
Stöffler (1972) suggested that mosaicism is the most com-
However, single sets of PDFs are developed in shock experi- mon manifestation of shock. However, mosaicism is not
ments under higher temperature conditions and are observed diagnostic of shock effects, and may also be caused by tec-
CHAPTER 8. SHOCK-INDUCED MICROSTRUCTURES AND SHOCK METAMORPHISM 89

tonic deformation or even chemical effects, especially in feld- within the field of crustal metamorphism (0.5 GPa) to values
spars (e.g. Sharpton and Schuraytz 1989, Officer and Carter as much as 10 GPa. Planar microstructures, mosaicism and
1991). The occurrence of coesite is known from inclusions Brazil twins in volcanic rocks have been claimed as shock
in high-pressure metamorphic rocks and kimberlites, but the microstructures (e.g. Carter et al. 1986, 1990, Huffman and
very fine-grained, polycrystalline nature of coesite in impact Reimold 1993), but much of this evidence has been strongly
related rocks is diagnostic (Grieve et al. 1996). Diaplectic refuted (Sharpton and Schuraytz 1989). The case for shock
glass can be formed by static compression at 25-35 GPa, but features related to volcanism or deep-seated explosive activ-
this is much higher than any pressures recorded by rocks at ity is not strong. In summary, Brazil twins, glass-filled PDFs,
the surface of the earth (Fig. 8.1): diaplectic glass remains a diaplectic glass, the distinctive very fine grained, polycrystal-
diagnostic indicator for shock pressures. line occurrence of coesite, and shock features in zircon are
Explosive volcanism and deep-seated explosive activity diagnostic of impact shock. As stressed by Reimold (1994),
are possible alternatives to meteorite impact as a source of a thorough evaluation of the case for shock metamorphism
shock waves. Controversy exists over what pressures may be should examine a variety of minerals and evidence on macro-
obtained in volcanic explosions; estimates range from values scopic to microscopic scales.
Chapter 9

From Microstructures to Mountains:


Deformation Microstructures, Mechanisms
and Tectonics
9.1 Introduction 9.2 Failure criteria
9.2.1 Coulomb and Mohr failure criteria
Deformation microstructures and mechanisms can be used to The simplest failure criterion, known as the Coulomb cri-
deduce conditions of deformation, because they imply partic- terion, is that failure occurs when the shear stress on a plane
ular relations between stress, strain rate (and possibly strain), equals a material property called the cohesion and
temperature, fluids, microstructure, and mineralogy. These the normal stress multiplied by the coefficient of internal
relations are derived from laboratory experiments and the- friction,
ories of material deformation. There are three types of re-
lationship: failure criteria (Section 9.2), which give stress
The criterion in terms of the principal stresses is:
conditions for faulting, frictional sliding laws (Section 9.3),
which give stress or stress-strain rate conditions for frictional
sliding, and flow laws (Section 9.4), which give stress-strain
where is the uniaxial compressive strength. The value of
rate relationships for flow accommodated by intracrystalline
is around 0.58 for many rocks, while varies widely for
plasticity, diffusive mass transfer (DMT), or composite de-
different rock types. A useful generalization of rock types
formation mechanisms. Relationships between flow laws are
in terms of their uniaxial compressive strengths is given by
usefully illustrated on a deformation mechanism map (Sec-
Paterson (1978):
tion 9.6) which shows fields of stress, temperature and grain
size in which particular mechanisms give the highest strain 1. Igneous and high-grade metamorphic rocks.
rate. The rheology of the lithosphere is often shown by a plot
of stress against depth, or a lithospheric strength envelope 2. Low-porosity sedimentary and low-medium grade meta-
(Section 9.7). morphic rocks.
Although failure criteria and flow laws are essential tools 3. High porosity sedimentary and some low grade meta-
in tectonic interpretation, they are based on laboratory exper- morphic rocks.
iments which have serious limitations when extrapolated to
4. Low porosity dolomites and quartzites.
nature (e.g. Carter and Tsenn 1987, Paterson 1987). Two of
the most important problems are the limited scale of triaxial However, the relationship between shear and normal stress
tests, which means that mesoscopic heterogeneities in rocks of real rocks is not linear as suggested by the Coulomb cri-
are not accounted for, and the fast strain rates of experiments terion. The Mohr criterion describes the relationship by an
(typically to ) compared to most tectonic de- empirical fit to data from rock mechanics experiments, and
formation ( to ), which mean that the deform- so is limited to materials that have been extensively tested.
ation mechanisms may differ between nature and experiment. The Griffith criterion (next Section) is a better approximation
Microstructures offer a potential solution to these problems: to rock behaviour in most situations.
if deformation microstructures and mechanisms can be iden-
tified in nature that correspond to experiments, then laborat- 9.2.2 Griffith failure criteria
ory results can be extrapolated with more confidence. This
Chapter provides some basic failure criteria, frictional sliding Griffith’s thermodynamic approach to failure outlined in Sec-
laws, and flow laws to be used in conjunction with micro- tion 2.2.1 leads to a failure criterion of the form:
structural observations to make tectonic interpretations. The
primary intention of the Chapter is to provide quantitative
data based on the concepts in the previous Chapters.

90
CHAPTER 9. FROM MICROSTRUCTURES TO MOUNTAINS 91

is the uniaxial tensile strength. This simple criterion is and may also depend on the elastic moduli of the solid and
widely applicable, but limitations are the difficulty of meas- pore fluid (Bernabé 1987, Warpinski and Teuffel 1993).
uring (which is dependent on the type of test used) and generally lies between the value of porosity and 1; typical
the observation that in most rocks, the uniaxial compressive values for reservoir rocks are 0.5-0.8 (e.g. Detournay et al.
strength is 10 to 12 times the value of (Jaeger and 1989, Kumpel 1991). is approximately 1 for many crystal-
Cook 1979). The Griffith criterion predicts that line rocks. Departures from this value occur when the pore
Values of for use in the Griffith criterion can be calculated fluid has a chemical effect, when permeability is low enough
from the generalized values of given in Section 9.2.1. to prevent uniform pore fluid pressures during experiments,
The simple Griffith criterion does not predict any effect of and when porosity is not interconnected (Paterson 1978).
on failure: this was accommodated in the Extended Grif-
fith criterion by Murrell (1963), giving the failure criterion:
9.4 Fracture mechanics and failure
criteria
This criterion implies that in better agreement
with observed values. A more rigorous treatment of triaxial Shear failure occurs by linking of axial tensile microcracks
stress acting on a body containing randomly orientated ellips- (Section 2.4). This fundamental observation has been incor-
oidal cavities leads to a failure criterion which is independent porated into failure criterion (e.g. Peng and Johnson 1972)
of and implies lower values of than the Griffith fail- through the application of fracture mechanics (e.g. Lawn
ure criterion (Murrell and Digby 1970). The effect of fric- and Wilshaw 1975a, Nemat-Nasser and Horii 1982, Horii and
tional sliding on closed Griffith flaws with a coefficient of Nemat Nasser 1985, 1986, Kemeny and Cook 1987, Hallam
friction and stress required for closure was dealt with and Ashby 1990, Baud et al. 1996). Unfortunately the res-
by McClintock and Walsh (1962) in the Modified Griffith cri- ultant expressions for failure are complex and depend on the
terion: geometry and size of the initial flaws, which are not readily
measured microstructural properties.

is a function of Poisson’s ratio and the axial ratios of the 9.5 Frictional sliding laws
flaws: for penny-shaped flaws and
for biaxial stress. The value of is (Murrell 1965); 9.5.1 Byerlee’s law
a value of 100 MPa was used for in lithospheric failure
Amonton’s Law (Section 2.2.2) describes frictional behaviour
modelling by Kusznir and Park (1984, 1987). If is taken
of rocks well at low normal stresses, but at higher normal
to be negligible, this criterion becomes equivalent to the Cou-
stress, a threshold shear stress must be applied before any
lomb criterion and implies that
movement occurs: this is called the cohesion, so that fric-
which generally has little relationship to the observed values
tional sliding in rocks is described by:
of (Jaeger and Cook 1976). However, this criterion does
provide a better fit to stresses measured at compressive failure
than either the original or extended Griffith criteria.
The form of this equation is the same as used for the Coulomb
criterion; however, the cohesion and coefficient of friction of
9.3 Pore fluid pressure and faulting intact rock (the coefficient of internal friction) are different
The principle of effective stress (Terzaghi 1943) describes the from those of frictional sliding. Byerlee (1978) discovered
important mechanical effect of pore fluid pressure on that values of in most rocks are in the range of 0.5-1, with
failure in many experiments (e.g. Paterson 1978). The prin- an average of approximately 0.75, and that frictional sliding
ciple states that normal stresses are reduced by the value of in most rocks could be described by two simple equations,
the pore fluid pressure, so that effective stresses can be sub- depending on the normal stress.
stituted in any of the failure criterion above. The effective
stresses (“conventional effective stresses” of Paterson 1978)
are calculated from:
These two equations have become known as Byerlee’s law,
and are approximately independent of rock type, roughness
More detailed experiments and poroelasticity theory indicate of the sliding surface and temperature. Byerlee’s Law is of-
that the normal stress is not reduced by the exact value of the ten assumed to describe the state of stress in the upper crust
pore fluid pressure. The effective stresses are given by (e.g. Burov and Diament 1996) which implies that failure in
the upper crust is governed by frictional sliding of previously
formed faults in ideal orientations for slip.
where is known as the effective stress coefficient and has Marone et al. (1992) have proposed that distributed de-
values less than or equal to 1 for elastic properties. de- formation in a gouge zone has a separate frictional sliding
creases with permeability, porosity and confining pressure, law, because stress is re-orientated within the sheared layer
92 CHAPTER 9. FROM MICROSTRUCTURES TO MOUNTAINS

so that the maximum possible shear stress is supported. This localization on shear surfaces, which always has a negative
results in a sliding law of the form: velocity dependence (Marone et al. 1992).
Seismological observations support the above picture, but
there is a lack of corresponding knowledge about natural
microstructures produced by seismic slip. Pseudotachylites
where is the coefficient of sliding on a discrete (Section 2.11) are the only direct evidence for seismic slip
surface. (e.g. Sibson 1977), but seismic slip can also be tentatively
suggested by textures such as crack-seal vein fillings which
indicate repetitive fracture followed by fracture sealing (Sec-
9.5.2 Rate and state dependent frictional slid-
tion 3.12). Particle size distributions of gouge may give an in-
ing dication of its mechanical behaviour: velocity strengthening
Dynamic effects during frictional sliding, such as stick slip behaviour changed to velocity weakening when the fractal di-
behaviour, are not described by Amonton’s and Byerlee’s mension of the particle size distribution of the gouge attained
Laws. Stick-slip behaviour involves a steady rise in stress a value of 2.6 in experiments (Biegel et al. 1989). Another
with no movement followed by a sudden rapid episode of potential microstructural indicator of seismic slip is the evid-
stress drop and sliding, and is analogous to earthquake fault- ence of continuous deformation alternating with discontinu-
ing. Stick slip behaviour is favoured over stable sliding in ous deformation in slickenside surface material (Power and
rocks with higher quartz contents, by sliding along rougher Tullis 1989).
surfaces, by the lack of gouge, by high normal stresses, and
low temperatures (e.g. Paterson 1978). Dynamic frictional
sliding can be modelled by making the coefficient of friction 9.6 Flow laws
in Amonton’s Law depend on the rate of frictional sliding
V: 9.6.1 Diffusive mass transfer: Grain size sens-
itive creep
The fundamental flow laws for deformation accommodated
where is the coefficient of friction at a reference velocity by diffusive mass transfer can be derived from theoretical
V*, a and b are empirically determined constants, and is a principles for three cases: diffusion through a grain bound-
state variable (e.g. Rice and Gu 1983). L is the critical slip ary fluid (pressure solution, Section 3.2), diffusion through
distance for sliding to return to a stable rate after a change in the grains (Nabarro-Herring creep, Section 5.2), and diffu-
sliding velocity. It is possible to expand the above expression sion through grain boundaries (Coble creep, Section 5.2). All
with more than one state variable. The value of a expresses flow laws for diffusion-accommodated deformation have one
the dependence of on a change in velocity, and b expresses distinguishing characteristic: strain rate is a function of grain
a change in that decays with time. This behaviour is found size.
to describe laboratory experiments very well. The stability
of a sliding process depends on the value of if this
is negative, increasing velocity will lead to a reduction in Pressure solution creep
and the fault is said to show velocity weakening, and hence be Diffusive mass transfer through a solution requires three
capable of nucleating earthquakes. On the other hand, if steps: solution, diffusion and precipitation. The strain rate
is positive, velocity changes will lead to increased friction and may be controlled by the rate of diffusion in the fluid, the rate
the sliding will be stable. of solution or precipitation at the source or sink, or the rate of
The rate dependent friction law can be applied to the nuc- diffusion at the source or sink (Paterson 1995), and the flow
leation of earthquakes in the crust (Tse and Rice 1986, Scholz law depends on which process is rate-limiting. The follow-
1990). is expected to be positive in the upper few km of ing flow law can be derived theoretically for diffusion control
well-developed faults due to the presence of a layer of gouge (e.g. Spiers et al. 1990, Lehner 1990):
(which is characterized by positive values of ). Therefore
earthquakes are not expected to nucleate in the top part of the
crust except in areas lacking well-developed faults. Below the
level of gouge, there is an upper stability transition to negative
values of which are the conditions for earthquake nucle- where is the strain rate A is a dimensionless para-
ation (Scholz 1990). becomes positive with increasing meter varying from 40 to 140 that depends on a geomet-
temperature, so that there is also a lower stability limit be- rical model for the arrangement of the grains, is the molar
low which earthquakes will not nucleate: this is probably ap- volume is the thickness of the grain boundary fluid
proximately at the 300°C isotherm in continental crust, which (m), is the reference state solubility of the solid in the fluid
explains the observation that earthquakes generally nucleate (mole fraction), is the reference state diffusion coefficient
in the top 15 km of the crust (e.g. Sibson 1982, Meisner in the grain boundary fluid is the activation en-
and Strehlau 1982). However, this picture is complicated by thalpy for pressure solution R is the gas constant
the relation between velocity dependence and displacement: T is the temperature (K), and is the
positive velocity dependence at small displacements changes differential stress (Pa). The reference state for and is
to negative dependence with increasing displacement, due to It is conventional to quote flow laws for the uniaxial
CHAPTER 9. FROM MICROSTRUCTURES TO MOUNTAINS 93

shortening strain rate under axisymmetric stress, and the dif- Nabarro-Herring creep
ferential stress relations between other components
The flow law for Nabarro-Herring creep is similar to the pre-
of the stress and strain rate tensors can be obtained from ap-
vious law, with the following important differences: the grain
propriate geometrical modifications.
size dependence is to the square of d there is no grain bound-
The terms are sometimes
ary thickness to take into account, and the diffusion coeffi-
gathered together as a phenomenological coefficient repres-
cient and activation enthalpies are different:
enting effective grain boundary diffusivity (e.g. Spiers et al.
1990). Diffusion control is relevant for salt in which solution
rates are very fast (e.g. Spiers et al. 1990), and has been
reported for experiments on quartz (e.g. Gratier and Guiget
1986). This flow law is only valid for dense polycrystals: where is the reference state volume diffusion coefficient,
porosity is a very important factor in pressure solution creep, and is the activation enthalpy for volume diffusion. Para-
which requires a specific geometrical model (e.g. Lemée and meters for the flow law are given in Table 9.1.
Guegen 1996). Paterson (1995) presents a model for pressure A flow law for diffusion creep that combines both Coble
solution creep in porous materials based on an analogy with and Nabarro-Herring creep (Ashby 1970, Raj and Ashby
particulate flow in soil mechanics, leading to equations for 1971) is:
each of the three possible rate limiting processes.
Diffusion-limited pressure solution creep in quartz has
been described by the following flow law (modified by An-
gevine and Turcotte 1983 after Rutter 1976):
The value of A may differ in all the above expressions since
it depends on the preferred geometrical model of the grains.

where A is in is the effective normal stress in Pa, 9.6.2 Intracrystalline plasticity


is the solubility of quartz at the reference state in and
is the density in Two different flow laws can be formulated from theory to de-
A flow law for the solution/deposition-controlled case is scribe flow controlled by dislocation glide (abbreviated to dis-
given by : location glide; Section 4.2) and flow controlled by dislocation
climb, known as dislocation creep (Section 4.2).

where kC is the velocity of the deforming crystal face Dislocation glide


e.g. Raj 1982). Solution/deposition controlled pressure solu-
Theoretical flow laws for dislocation glide are characterized
tion has been described in quartz (Gratier and Jenatton 1984),
by an exponential relationship between strain rate and stress,
but seems to be less common than diffusion control. Pa-
known as the Dorn Law, of the form:
terson’s (1995) model suggests that for quartz and rocksalt
aggregates, source/sink diffusion control is likely to be rate
limiting under geological conditions, but in high-temperature
laboratory conditions, precipitation control is effective for
quartz. Experimentally determined values for the relevant where and are the strain rate and stress at 0 K, and
parameters are given in Table 9.1. F is a constant (e.g. Dorn 1954). Dislocation glide is not
considered to be important in most minerals under geological
Coble creep conditions because of the relatively high stresses required for
the process.
The flow law for Coble creep has a very similar form to that
for pressure solution creep since both processes involve diffu-
sion around grain boundaries. The major differences are the Dislocation creep
lack of a term for solubility, and differences in the values of At lower stresses, higher temperatures and slower strain rates,
the diffusion coefficient and activation enthalpy: a power law relationship describes the strain rate-stress rela-
tionship:

is the reference state diffusion coefficient for grain where A is a pre-exponential constant, is the activation
boundary diffusion, is the grain boundary width, and enthalpy for dislocation creep, and n is the stress exponent
is the activation enthalpy for grain boundary diffusion. The (Weertman 1968). A depends on the volume diffusion coeffi-
specification of is difficult because it depends on the cient and the Burgers vector of the active dislocation system.
diffusing species. The strain rate will however be limited by Unfortunately there are severe problems in using this
the diffusion rate of the slowest species, which is normally simple flow law. One of the most important variables that is
quoted. In olivine, the values of the diffusion coefficients in- not explicitly taken into account is the effect of water. For ex-
crease in the order Si < O < Fe; the value for Si is rate limit- ample, “Dry” quartz, with H contents of less than
ing. Values for Coble creep flow laws are given in Table 9.1. Si, is an order of magnitude stronger than “wet” quartz (the
94 CHAPTER 9. FROM MICROSTRUCTURES TO MOUNTAINS

considerable amount of literature on this subject is summar- ation energy P is the mean stress (Pa), V is an ac-
ized in Paterson and Luan 1990). The effect is probably due in tivation volume often ignored due to its relatively
part to replacement of Si-O-Si bonds by weaker Si-OH-OH- small contribution to the exponential term), and m and n are
Si (hydroxyl) bonds, known as hydrolytic weakening (Griggs general dimensionless grain size and stress exponents. The
and Blacic 1965, Griggs 1967), which is similar to the mech- values of m and n suggest interpretations of the deformation
anism that controls sub-critical microcrack propagation (Sec- mechanism from experimental data by analogy with the the-
tion 2.2.1.1). Hydrolytic weakening enables easier disloca- oretical relationships: and indicates dislocation
tion glide and creep, although other mechanisms may also creep, and indicates pressure solution or Coble
play a role in the weakening effect of water. This suggests creep, and and indicates Nabarro-Herring creep.
that the flow law should incorporate a function of the water Superplasticity is characterized by and (e.g.
activity (e.g. Kronenberg and Tullis 1984, Ord and Hobbs Boullier and Guegen 1975, Schmid et al. 1977). Tables 9.2-
1985). Experiments show that the power law exponent de- 9.4 are a compilation of some results for minerals and rocks,
pends on water content, at least for low water contents (e.g. divided into grain-size sensitive Table 9.2), and
Blacic and Christie 1984). In any event, it is difficult to de- grain-size insensitive flow laws ( Table 9.3 for min-
termine water contents at the time of deformation, although erals and Table 9.4 for rocks). When using these tables or
Kronenberg et al. (1990) have shown how quartz water con- any such rock mechanics data, it is advisable to consult the
tents measured across a natural shear zone vary with shear original reference to find any important specific conditions of
strain, possibly reflecting a variation during deformation. the experiments which may limit the applicability of the flow
Mean stress is another variable that should appear in the laws.
flow law, which may have important rheological effects by At high stress levels, the power law relationship does not
increasing water diffusion rates, increasing water solubility, fit the experimental data (“power law breakdown”, Carter and
and increasing activation enthalpy. Quartz phase appears to Tsenn 1987). Two alternative empirical relationships may de-
affect both the activation enthalpy ( is higher for than scribe the data well:
and the power law exponent n which is lower for
(e.g. Linker and Kirby 1981). The quartz phase af-
fects diffusion coefficients in a remarkably similar way to ac- is an exponential law, with B and as empirical constants,
tivation enthalpies of power law creep, showing that the creep suggesting that flow controlled by dislocation glide is import-
enthalpies are essentially the same as the enthalpies of diffu- ant (Tsenn and Carter 1987). An expression that fits the in-
sion (Gilletti and Yund 1984, Dennis 1984). Trace amounts of termediate and high stress behaviour is :
impurities in quartz may also cause considerable weakening,
probably through their control on glide velocity (e.g. Jaoul et
al. 1984).
e.g. Heard (1963), with C, and n as the empirically-derived
constants.
9.6.3 Empirical flow laws from experimental
data
9.7 Polymineralic deformation
Experimental data can be fitted to empirical laws of the fol-
lowing general form for low stress levels: While flow laws such as those in Section 9.4.3 can describe
the bulk rheology of rocks, microstructural evidence shows
that strain is distributed in a complex manner between miner-
als with different rheologies, which should be accounted for
where A is a pre-exponential constant or in a detailed understanding of deformation. An empirical ap-
depending on the units of ), Q is an activ- proach was taken by Tullis et al. (1991), who derived a flow
CHAPTER 9. FROM MICROSTRUCTURES TO MOUNTAINS 95
96 CHAPTER 9. FROM MICROSTRUCTURES TO MOUNTAINS

law for a bimineralic rock that weighted the flow law para- The distinction between LBF and IWL behaviour occurs at
meters of each mineral by the volume fractions of the the conditions where both give the same bulk strength. On
weak and strong minerals as follows: either side of this equality, the more stable microstructure
is the one that dissipates less strain energy. For a quartz-
feldspar composite with dislocation creep flow laws, the LBF
microstructure is only stable for very low volume fractions of
quartz.
This model shows that strength of the composite always
lies between the limiting strengths of the strong and weak
where n, Q, and A are the parameters in the dislocation creep
phases, but the relative strengths of the composite compared
flow law (Section 9.4.2.2), and the subscripts and s are
to the end members are a complex function of their strength
for composite, weak, and strong minerals respectively. This
contrasts and their volume fractions. In a granite, 10-20%
gives a reasonable fit to experimental data.
volume fraction of quartz reduces the composite strength
A more fundamental way to treat polymineralic rocks is
by one half compared to the strength of pure feldspar, as-
to model the relationship between the behaviour of different
suming IWL behaviour, and that the deformation occurs in
minerals. Handy (1990, 1994) has suggested that for two
the conditions appropriate for dislocation creep flow laws.
minerals of contrasting strength, the microstructure may be
Quartz effectively governs the strength of the composite at
described as a load bearing framework (LBF) if the stronger
volume fractions over 10%. Feldspars in gabbros signific-
mineral supports stress around pockets of a weaker mineral,
antly weaken the composite strength from a pure pyroxene
or as an interconnected weak layer (IWL) if the weaker min-
rock up to 900°C: at higher temperatures, pyroxene is weaker
eral forms a matrix around microboudins of the stronger min-
than feldspar, and the composite has an intermediate strength.
eral. The behaviour of the composite depends on the contrast
Both the Tullis and Handy approaches outlined above are
on rheological properties of the minerals, and on their volume
limited to two phase rocks. A generalization to any number
fraction.
of phases is given by Ji and Zhao (1993) for two extreme
Quantitative observations can be used to distinguish LBF
assumptions. If all phases are loaded equally (isostress cri-
and IWL microstructures. In granite deformed under am-
terion), the flow law is given by taking the volume average of
phibolite facies, for example, the LBF case may be char-
the strain rates of individual phases:
acterized by similar grain shapes and sizes in all minerals,
and strain magnitudes, as estimated approximately by grain
shapes, are low (Schulmann et al. 1996). Quartz has large
recrystallized subgrains, and K-feldspars are fractured, while
plagioclase recrystallizes. There is a transition to an IWL
with increasing strain, which is manifest from higher strain where n, A, and Q are the flow law parameters, the subscripts
in quartz than feldspar, and the appearance of S-C fabrics. r and i refer to the isostressed aggregate and the individual
The ultimate stage of the process is a transition to another phases (total number m and volume fractions f) respectively.
type of IWL microstructure in which there is little contrast An alternative assumption is that all phases have the same
between the quartz and feldspar grains: all minerals have re- strain and strain rate. This isostrain criterion leads to the fol-
crystallized completely in a mylonite. lowing aggregate properties:
Two flow laws are necessary to describe the LBF and IWL
cases separately, which can be derived from considering that
the total rate of viscous strain energy dissipation is the sum
of the rates in the weak and strong minerals. In the LBF case,
both minerals deform at the same rate and the shear strength
of the composite is given by the sum of the shear An average flow law is considered to lie between the
strengths of each mineral multiplied by their volume frac- physically unrealisitic extremes of the isostress and isostrain
tions: cases. The average is determined by minimising the dif-
ference between two composite flow laws, that each allow
where and are the shear strengths and volume for variable, complementary proportions of the isostress and
fractions of the weak and strong minerals respectively. For isostrain flow law parameters. The results achieve reasonable
the IWL case, fits to some experimental data and the method is advantage-
ous when the effects of more than two phases need to be con-
sidered.
In another approach to the deformation of polymineralic
where a function x is introduced that describes the sensitivity rocks, Ji and Zhao (1994) have used a fibre loading model to
of the strain rate partitioning in the rock to the strength con- derive an expression for the elastic solution to the problem
trast of the minerals. The nature of this function can be sug- of rigid fibres in a weaker matrix, and generalized from the
gested by considering limiting values. One possibility sug- elastic solution to a viscoplastic case. The equation they de-
gested by Handy (1994) is: rive includes an expression for the aspect ratio of the fibres.
Their model describes some experimental results better than
the Tullis model, but they point out that it assumes uniformly
CHAPTER 9. FROM MICROSTRUCTURES TO MOUNTAINS 97

distributed fibres with a constant orientation, which may limit haviour of the lithosphere and have important applications to
its geological application. Flow laws for polymineralic rocks large-scale tectonic problems (e.g. Burov and Diament 1995,
require more theoretical and experimental research. 1996).
The upper part of the LSE is usually constructed using
Byerlee’s Law for frictional sliding to represent cataclastic
9.8 Deformation mechanism maps behaviour. Coulomb or Griffith failure criteria can also be
used to model the behaviour of intact lithosphere (cf. Sib-
The rheology predicted by the flow laws in Section 9.4 can son 1983, Kusznir and Park 1987). A power law for disloca-
be compared on a deformation mechanism map (Fig. 9.1), tion creep is usually assumed for the lower, plastic part of the
which shows the relationships between deformation mech- LSE, justified by observations that suggest that lithospheric
anism, stress, strain rate, temperature, and grain size (e.g. stress/temperature/grain size conditions are appropriate for
Stoker and Ashby 1973, Ashby and Verrall 1978). These this deformation mechanism. Dislocation glide can also be
maps may be shown in plots of stress normalized by the included (e.g. Burov and Diament 1995).
shearmodulus against temperature normalized by melt-
The LSE is based on a model of the compositional structure
ing temperature (homologous temperature, ). The shear
of the earth. Gross simplifications are necessary due to the
modulus normalization allows for the systematic dependence
limited availability of rock mechanics data, and the difficulty
of the flow law constant A on and for the temperature de-
of applying them to the earth. The upper crust is commonly
pendence of the modulus (cf. Tsenn and Carter 1987). The
modelled by wet quartzite, the mid to lower crust by feldspar
melting temperature normalization allows for better compar-
or diorite, and the mantle as olivine. A temperature-depth
ison between different materials because transitions between
profile is chosen for the crustal model. Appropriate steady
different rheological regimes depend on These nor-
state geotherms can be calculated from surface heat flux (e.g.
malizations are relevant for materials-science applications,
Kusznir and Park 1984, 1987), but Burov and Diament (1995)
but the direct values of stress and temperature are more useful
have modelled an evolving geotherm due to cooling after oro-
for tectonic studies.
geny.
The fundamental assumptions for constructing these maps
are that deformation will occur by the mechanism that The LSE is delineated by the lesser of the cataclastic or
provides the fastest strain rate, and that steady state conditions plastic strengths. The stresses can either be calculated for
are obtained. Deformation mechanism fields in stress and a given strain rate, or by applying a constant stress at the
temperature space are contoured for strain rate using the flow boundaries of a model that analyses the stress and strain with
law of the deformation mechanism with the fastest strain rate; time. The latter approach shows that stress applied uniformly
boundaries between fields correspond to loci of equal strain at the boundaries of the lithosphere is redistributed into the
rate. Different maps are constructed for different grain sizes; more competent layers which deform elastically, a process
alternatively grain size can be added as a third dimension. referred to as stress amplification (Kusznir 1982). Eventu-
The possibility of contributions to strain rate by more than ally the elastic strength of all layers is exceeded, leading to
one mechanism can be allowed for (e.g. Stoker and Ashby whole lithosphere failure, and deformation continues by fric-
1973). For example, both diffusional creep mechanisms can tional sliding and intracrystalline plasticity (Kusznir and Bott
be combined as in the last equation in Section 9.4.1.3, and dif- 1977, Kusznir and Park 1982, 1984, 1987). Hence steady
fusion and dislocation processes can also combine. However, state LSEs have no part that is controlled by elastic strength.
flow by dislocation glide and dislocation creep are mutually The shape of the simplest LSE (for a lithosphere of homo-
exclusive. geneous composition) consists of a rapid linear increase in
Deformation mechanism maps for geological materials strength with depth from the surface of the earth governed by
have been produced for only quartz, olivine and calcite (e.g. the high normal-stress dependence of Byerlee’s Law or a fail-
Rutter 1976, White 1976, Carter and Tsenn 1987), because a ure criterion (e.g. the top part of Fig. 9.2). The cataclastic
range of flow laws for different mechanisms are not available part of the LSE depends on the orientation of the failure sur-
for other minerals or rocks. Microstructures are a guiding face: differential stresses for faulting or frictional sliding at
principle to the construction and use of deformation mechan- any level of the crust increase in the order normal - strike-
ism maps: similarity in experimental and natural microstruc- slip - thrust fault. The strength of the lithosphere reaches
tures and mechanisms allows for the correct choice of flow a maximum where the cataclastic strength is equal to the
law parameters (e.g. Busch and Van der Pluijm 1995). flow stress for intracrystalline plasticity. Below this point the
flow stress is less than the cataclastic strength, and continues
to decrease with depth as temperature increases. LSEs for
9.9 Lithospheric strength envelopes layered models of the earth generally show the above pattern
repeated in each layer (Fig. 9.2). There are dramatic strength
A Lithospheric Strength Envelope (LSE) shows the relation- contrasts across layers, especially between the crust and the
ship of lithospheric strength with depth derived from elasti- mantle. This suggests that detachment could occur at these
city, failure criteria, frictional sliding laws and flow laws layer boundaries, which are sometimes misleadingly called
(Fig. 9.2). The term yield strength envelope is commonly “brittle-ductile transitions”, but are more accurately referred
used (e.g. Goetze and Evans 1979), but since the upper part to as brittle-plastic transitions (Section 1.3, 1.5).
of the envelope is governed by cataclastic failure, LSE is a However, saw-tooth LSEs such as those in Fig. 9. 2 can
better term. LSEs are a powerful way of describing the be- be criticized because the validity of Byerlee’s law has not
98 CHAPTER 9. FROM MICROSTRUCTURES TO MOUNTAINS

been established at mid to lower crustal pressures, and the entation of the stress tensor (Sections 9.10.7-9.10.8). Pa-
high stresses predicted by extrapolation of the law to these laeopiezometers are calibrated by experiments and theory,
conditions are unrealistic (Ord and Hobbs 1989). Moreover, but there are large variations in calibrations of the same pa-
the semibrittle deformation regime is not recognised, prob- laeopiezometer. Furthermore, different palaeopiezometers
ably because of the paucity of published flow laws. More yield considerably different values when applied to the same
realisitic possibilities for the shape of the LSE at mid crustal rocks. The latter variations can potentially lead to further tec-
depths are either a vertical line of constant stress with a value tonic insights (Section 9.10.9). Determination of the mean
equal to the stress at the base of the seismogenic zone (Hobbs stress is considered in the separate Section on geothermoba-
et al. 1986), or non-linear, pressure-dependent functions de- rometry (Section 9.11).
rived from Mohr-Coulomb constitutive laws for permanent
deformation (Ord 1991). The experimental basis for the lat-
ter laws is very limited, and they have the disadvantage of 9.10.2 Recrystallized grain size
predicting strain-dependent values of stress. The size of grains recrystallized during deformation (d, mm)
can be related to differential stress ( MPa) by an expression
of the form:
9.10 Palaeopiezometry
9.10.1 Methods and calibration where m and are constants, which can be determined ex-
perimentally or derived theoretically (e.g. White 1979). Val-
Palaeopiezometry is the determination of past stress fields, ues of m and are given in Table 9.5, and the calibrations
which can provide major constraints on tectonic models. Mi- for quartz are plotted in Fig. 9.3. The differential stresses
crostructural methods have concentrated mainly on measur- for quartz predicted by different calibrations differ by up to
ing differential stress (Sections 9.10.2-9.10.6), and the ori- four orders of magnitude, for at least seven possible reasons,
CHAPTER 9. FROM MICROSTRUCTURES TO MOUNTAINS 99
100 CHAPTER 9. FROM MICROSTRUCTURES TO MOUNTAINS

where and are calibration constants, is the shear mod-


ulus, and is the most common Burgers vector. Values of
1 and 2 for have been proposed on theoretical grounds
(Twiss 1977). Another possible relationship proposed by
Twiss (1986) is:

where C and are constants, implying that there is a


stable subgrain size at zero stress. However, the second rela-
tionship does not fit data for olivine and quartz better than the
first. This may be due to the lack of accurate, low stress meas-
urements (Twiss 1986). Until these are available, empirically
derived constants for the first relationship are probably the
most satisfactory calibration of the subgrain size palaeopiezo-
meter, as given in Table 9.6. Problems with the application
which are worth examining in detail because they highlight of this palaeopiezometer include a possible temperature de-
the problems of recrystallized grain size palaeopiezometry. pendence, the sensitivity of subgrain size to water known for
olivine (Twiss 1986, Van der Wal 1993) and the problem of
1. Recrystallized grain size appears to depend on the re- subgrain size measurement. Subgrain sizes may be measured
crystallization mechanism: subgrain rotation (regime 2) under the optical microscope in reflected light after etching
gives smaller grains than grain boundary migration (re- (see Ord and Christie 1984 for the method) or by electron mi-
gime 3) (e.g. Drury et al. 1985, Drury and Urai 1990, croscopy. Optically determined subgrain sizes are commonly
Hirth and Tullis 1992). an order of magnitude larger than subgrain sizes measured
by electron microscopy (Ord and Christie 1984). Available
2. The importance of the effect of water is obvious from
data at present may not allow reliable extrapolation outside
comparing the dry and wet calibrations of Ord and
the range of subgrain sizes used to calibrate the palaeopiezo-
Christie (1984). However, it does not appear to be im-
meter, since alternative calibrations fit the experimental data
portant in olivine (Van der Wal 1993).
almost equally well, yet give enormous differences when ex-
3. The possibility that the recrystallized grain size is sens- trapolated (Twiss 1986).
itive to temperature is suggested by some data on metals
and olivine (e.g. Mercier et al. 1977, Ross et al. 1980), 9.10.4 Dislocation density
and predicted by theory (Mercier 1980a).
The dislocation density relationship can be
4. The possibility that the recrystallized grain size is de- written as:
pendent on quartz phase is suggested by the flow law
dependence on phase (9.4.2.2).
where k and u are constants, and and are defined as above.
5. Kinetic effects may severely hamper determination of A theoretical value of u is 0.5 (Weathers et al. 1979), but
the calibration constants if equilibrium grain size is not measured values have a range from 0.45 to 3.33 (Twiss 1986).
attained during experiments (Twiss 1977). The lack of An alternative form of the relationship sometimes used is:
such equilibrium for grain sizes on the order of mm is
strongly suggested by the experiments of Kronenberg
and Tullis (1984); hence recent experiments have been
on much finer grain sizes (e.g. Post and Tullis 1999). where D and are constants, implying a steady state
free dislocation density. However, data for olivine and quartz
6. Equilibrium grain size may not be achieved if grain
are not fitted any better by the second relationship. Table 9.7
boundaries are pinned by impurities (e.g. Evans and
presents calibration constants for quartz and olivine using the
White 1984).
first expression. Problems with this palaeopiezometer are
7. Recrystallized grain size has been measured by various possible temperature dependence and the difficulty of meas-
techniques and may be specified by different parameters: uring dislocation density accurately. Invisibility of disloca-
a standard practice is to measure the mean linear inter- tions under the TEM may cause an underestimate in disloca-
cept grain size and multiply by 1.5 to allow for trunca- tion density of 30% (Ord and Christie 1984).
tion and sampling effects (e.g. Christie and Ord 1980).
However, this will not be correct for grains with a fabric.
9.10.5 Twinning - differential stress

9.10.3 Subgrain size Three different approaches have been proposed to relate twin-
ning in carbonates to differential stress. Two of these (Jam-
Subgrain size can be related to stress by an expression ison and Spang 1976 and Laurent et al. 1981, 1990) take
of the form: no account of the fact that grain size is known to have an
effect on the twinning, and are therefore less suitable than
CHAPTER 9. FROM MICROSTRUCTURES TO MOUNTAINS 101

the approach of Rowe and Rutter (1990), who have calib- a non-twinning deformation mechanism. In principle, grain
rated three twinning palaeopeizometers from laboratory ex- size distribution could have an effect on this palaeopiezo-
periments, which all appear to be independent of temperature, meter because stress distributions in grains are strongly in-
strain rate and strain. fluenced by the size of neighbouring grains (Newman 1994),
but this has not yet been evaluated or demonstrated. This
Twinning incidence, palaeopiezometer has been calibrated only for calcite (other
carbonates twin at different stress levels), and since the cal-
is the proportion of grains with optically visible twins in ibration experiments were carried out at 400°C and above, it
any grain size class. It can be related to differential stress is probably not appropriate for low temperature deformation
(MPa) and grain size (d, mm) by the following equation: (Burkhard 1993). The best results will be obtained in samples
that have experienced a single, coaxial strain event (the con-
ditions of the experimental calibration), otherwise the method
In practice, stresses are calculated from for each grain size will overestimate stress (Rowe and Rutter 1990).
class and plotted against grain size: they should lie on lines There is some experimental evidence that twinning in
of constant slope for different values of The standard error pyroxene could be used as a palaeopiezometer: Tullis (1980)
in the experimental calibration was 31 MPa. suggests that the minimum stress needed for twinning in
pyroxenes is 100 MPa.
Twin density,
is the number of twins per mm. This is calculated by the
slope of the relation between number of twins in a grain size 9.10.6 Deformation lamellae
interval and grain size. appears to be independent of grain
size, and is related to with a standard error of 43 MPa, by:
The appearance of deformation lamellae (Section 4.7) in ex-
periments is characteristic of flow laws with an exponential
relationship between stress and strain rate, which occur at
Maximum twin volume, relatively high stresses. This observation leads to the sug-
gestion that there may be a critical differential stress for the
is the maximum volume proportion (%) of twinned ma- formation of deformation lamellae, dependent on the type of
terial in a grain size class. In the experiments, the area frac- bonding and crystal structure of the material (Blenkinsop and
tion of twins was taken as equal to the volume fraction, and Drury 1988). Values of 100-200 MPa have been suggested
was measured from a two-dimensional thin section. is for quartz. However, it is possible that the critical stress
related to with a standard error of 41 MPa, by : may be inversely temperature-dependent, so that at present
deformation lamellae in quartz should only be taken as qual-
itative indicators of relatively high stress levels (Drury 1993).
Spacing of deformation lamellae (s, mm) may constitute a
As with the previous palaeopeizometers, there are a number
palaeopiezometer (Koch and Christie 1981, McLaren 1991).
of potential problems. Clearly if twinning is the only deform-
The relation derived by Koch and Christie is ( MPa):
ation mechanism, the amount of twinning measured by any of
the three parameters will be a function of strain. The method
thus assumes that after twinning, equilibrium is reached with
102 CHAPTER 9. FROM MICROSTRUCTURES TO MOUNTAINS
CHAPTER 9. FROM MICROSTRUCTURES TO MOUNTAINS 103

9.10.7 Principal stress orientations from de- (e.g. Turner 1953, Weiss 1954). Twinning in calcite occurs
formation lamellae on the e planes in the direction of the plane contain-
ing the pole to the twin and the c-axis (Fig. 9.4). and
Techniques of using deformation lamellae in quartz to de- are therefore at 45° to the twin plane in the plane of the c-axis
duce principal stress orientations were developed by Carter and the pole to the twin - a similar geometry to that used for
and Friedman (1965) and Carter and Rayleigh (1969). Of the quartz deformation lamellae method. However, in con-
the three methods proposed by Carter and coworkers, the “ar- trast to that method, is in the opposite quadrant from the
row” method is the easiest and yields consistent results. The c-axis (Fig. 9.4). A similar construction can be applied to f
orientations of the c-axis and deformation lamellae are first twins in dolomite, but here is in the same quadrant as the
measured from individual quartz grains. The plane containing c-axis (Fig. 9.4).
the pole to a lamella and the c-axis is constructed (Fig. 9.4). A development of this technique is to use the right dihedra
and are located within this plane at 45° to the lamella
method to determine the and orientations that are com-
plane. is distinguished from by the fact that the c-axis
patible with the largest numbers of twins. The value of the
lies closest to An arrow from the c-axis to the lamella maximum number of compatible twins (the MAX number)
pole will point from to (Fig. 9.4). Unique and gives a measure of the degree to which the data can be fitted
orientations can be derived by eigenvector analysis from a by a single stress orientation (Pfiffner and Burkhard 1987).
number of such measurements (e.g. Spang and Van der Lee This method has the advantage that it does not assume a fixed
1975). This construction implies that the deformation lamel- angle between and the twin plane. A further development
lae form in a plane of maximum resolved shear stress with a is the determination of principal stress orientations and ab-
maximum shear stress direction parallel to the projection of solute magnitudes by assuming homogeneous stress distribu-
the c-axis onto the lamellae plane. The evidence presented in tion and a critical resolved shear stress for twinning (Laurent
Section 4.7 shows that recovery is an integral part of lamellae
et al. 1981, 1990).
development, so that a simple analogy with a slip system for
It is possible to calculate a finite strain tensor by incorpor-
the formation of lamellae may not be entirely valid. However,
ating the shear strain necessary for twinning into the orient-
the strength of the method is based on its empirical success
ation analysis (e.g. Groshong 1972, 1974, 1984). A com-
(Pavlis and Bruhn 1988, Drury 1993), even if its theoretical
puter program to carry out the analyses is described by Evans
basis remains unclear. The principal stress axes from deform-
and Groshong (1994). The measurements necessary for the
ation lamellae are in good agreement with those calculated
determination include the c-axis orientation, twin set orienta-
from the carbonate twin method from the same samples (e.g.
tion, average thickness and number of twins, and grain width.
Spang and Van der Lee 1975). It is possible to consider other
25 grains in two perpendicular thin sections should be meas-
angular relationships between and the lamella plane, but
ured.
some results suggest that the stress axes do not change ap-
Twins which do not fit the calculated strain or stress fields
preciably from the 45° constraint (Spang and Van der Lee
are a problem in all these methods. Two basic philosophies
1975). It is also possible to analyze for the principal stresses
to deal with this problem have been adopted. The first is to
without using a fixed angular relationship between and the
recognize that some twins will inevitably develop in incom-
lamella plane, by using the right dihedra technique (cf. An-
patible orientations due to stress inhomogeneity. The propor-
gelier 1984).
tion of such twins (negative expected values, NEV, Groshong
1972) can be used to test the homogeneity of the sample, and
9.10.8 Principal stress orientations and strains the incompatible grains can be removed from the data set.
from twins Alternatively, the incompatible twins are presumed to reflect
different superimposed stress fields, and they can be used to
Twins have been used to deduce principal stress orientations separate out a number of different stress fields (e.g. Lacombe
by assuming that the twin plane is a shear plane and the twin et al. 1990). This method has been criticized by Burkhard
direction is parallel to the maximum resolved shear stress (1993) because it has yielded unrealistic results, and the as-
104 CHAPTER 9. FROM MICROSTRUCTURES TO MOUNTAINS

for subgrain sizes and perhaps 30% for recrystallized grain


sizes; dislocation densities may therefore record later tectonic
events with low strains such as stress during uplift (e.g. White
1979a). This problem can potentially be overcome by plot-
ting stress levels derived from one palaeopiezometer against
another. If the stresses are recorded by both palaeopiezomet-
ers from the same part of the stress-time path, they will lie
on a line with a unit slope, while deviations from this line
indicate records from different parts of the stress path.

9.11 Geothermobarometry
9.11.1 Methods and calibration
The determination of past temperature and mean stress (geo-
thermobarometry) is traditionally the domain of thermody-
namics and mineral chemistry. However, there are a num-
ber of promising new developments in the application of mi-
crostructures to geothermobarometry, which are particularly
robust because they are calibrated from naturally deformed
samples.
In addition to these new, quantitative approaches, experi-
ments and thermodynamic results allow some generalizations
about the minimum temperatures for plasticity. Plasticity in
the form of twinning can occur even at room temperature in
calcite, but appears to require a temperature of 300°C in dolo-
mite. Plasticity (sensu lato) in quartz and mica is generally
restricted to temperatures greater than 200°C. The onset of
plasticity in feldspars is commonly taken as 450°C, but this
value is not well constrained. The common mafic minerals
(clinopyroxene, amphibole, orthopyroxene and olivine) may
be plastic from 500°C or above. These numbers must how-
ever be treated with great caution, because there are many
other variables involved in the mechanism transitions (not-
ably stress/strain rate, see Section 1.3), and because of the
problems of extrapolating from experiments to nature (Sec-
tion 9.1).

9.11.2 Calcite twin morphology


Twin morphology in calcite appears to be strongly temperat-
sumption of stress homogeneity is clearly false in a polysized
ure dependent, and has been proposed as a geothermometer
grain aggregate.
(Burkhard 1993). Four types of twins can be distinguished
(Fig. 9.5), together with appropriate temperature ranges for
9.10.9 General problems with palaeopiezomet- deformation:
ers
Type I. Thin (1 mm), straight, rational twins form at temper-
A general problem with all palaeopiezometers is the interpret- atures of less than 200°C.
ation of the results. The significance of the stress level recor-
ded can only be evaluated in the context of an evolving stress Type II. Thicker, straight and rational twins (> 1 mm),
field, analogous to the problem of interpreting results from which are slightly lens shaped, form at 150-300°C.
geothermobarometry within a P-T path. Interpretations be-
come especially problematic when considering results from Type III. Thick, curved, irrational twins, which may contain
the same sample that differ according to the method used. For twins within twins, form at temperatures greater than
example, stresses measured by dislocation density, subgrain 200°C.
size and recrystallized grain size may differ because they have
variable dependence on strain magnitude. Dislocation dens- Type IV. Thick, irrational and patchy twins, breaking up into
ities may be reset after less than 1% strain, compared to 5% trails of small grains, form at temperatures over 250°C.
CHAPTER 9. FROM MICROSTRUCTURES TO MOUNTAINS 105

9.11.3 Sutured quartz grain boundaries were limited to to and maximum strains of
31%. Ultimately the validity of a temperature - only depend-
The geometry of sutured grain boundaries (Section 4.8)
ence of D rests on the empirical evidence.
formed by grain boundary migration depends on the temper-
ature during deformation. With increasing temperature, the
length of segments, serrations or lobes increases (Fig. 9.6, 9.11.4 Subgrain boundary orientation in
Kruhl and Nega 1996). The geometry of a grain boundary can quartz
be characterized by its fractal dimension, D, which is best
The orientation of subgrain boundaries in quartz appears to
measured by the “divider” method (Kruhl and Nega 1996).
be controlled by the phase present during deformation (Sec-
A grain boundary is divided into linear segments (“strides”)
tion 4.6). The presence of chessboard patterns, indicating
of individual length r. The length of the grain boundary L
both basal and prism-parallel sub-grain boundaries, is restric-
is equal to the product of the number of segments and their
ted to deformation in the field (Kruhl 1996). This
length. The process is repeated over a range of values of r.
means temperatures above 573°C at 0 MPa, and 825°C at
For a fractal grain boundary, L is related to r by:
1000 MPa (Gross and Van Heege 1973). Chessboard patterns
can only be seen in grains with c-axes subparallel to the plane
of the section (high interference colours) and must be distin-
D is derived from the gradient of a log-log plot of L against guished from three types of pseudochessboard pattern (Kruhl
r. Kruhl and Nega (1996) established the relation shown 1996).
in Fig. 9.7 empirically from measurements of natural su-
1. Two sets of prismatic subgrain boundaries visible in
tures with known temperatures of deformation. The relation
grains with c-axes at high angles to the section (i.e. low
between T (°C) and D is approximately:
birefringence).
2. Rectangular kink band boundaries that are not crystallo-
graphically orientated.
D values of sutures produced in experimental deformation
appear to be affected by the experimental strain rate as well 3. Two sets of rhombohedral subgrain boundaries in grains
as temperature (Takahashi et al. 1997). However, the extent with c-axes subparallel to the section, which can be dis-
to which strain rate may be important in natural deformation tinguished by their inclined extinction positions.
is unknown, especially because the experimental strain rates
This page intentionally left blank
References
Aerden, D. G. A. M. 1995. Porphyroblast non-rotation during the upper mantle. Phil. Trans. R. Soc. Lond. A 288, 59-95.
crustal extension in the Variscan Lys-Caillaouas Massif, Ashworth, J. R. & McLellan, E. L. 1985. Textures. In Ash-
Pyrenees. J. Struct. Geol. 17, 709-726. worth, J. R. (ed.) Migmatites. Blackie and Sons Ltd.,
Aerden, D. G. A. M. 1996. The pyrite-type strain fringes from Glasgow, 180-203.
Lourdes (France): indicators of Alpine thrust kinematics Ashworth, J. R. & Schneider, H. 1985. Deformation and
in the Pyrenees. J. Struct. Geol. 18, 75-92. transformation in experimentally shock-loaded quartz.
Agar, S. M., Prior, D. J. & Behrmann, J. H. 1988. Back- Phys. Chem. Min. 11, 241-249.
scattered electron imagery of the tectonic fabrics of some Atkinson, B. K. 1982. Subcritical crack propagation in rocks:
fine grained sediments: implications for fabric nomen- theory, experimental results and applications. J. Struct.
clature and deformation processes. Geology, 17, 901-904. Geol. 4, 41-56.
Aharonov, E., Rothman, D. H. & Thompson, A. H. 1997. Avé Lallement, H. G. 1978. Deformation of diopside and
Transport properties and diagenesis in sedimentary rocks: websterite. Tectonophysics 48, 1-27.
The role of micro-scale geometry. Geology, 25, 547-550. Aydin, A. 1978. Small faults formed as deformation bands in
Alexopoulos, J. S., Grieve, R. A. F. & Robertson, P. B. 1988. sandstones. Pure & Appl. Geophys. 116, 931-942.
Microscopic lamella deformation features in quartz: Dis- Aydin, A. & Johnson, A. M. 1978. Development of faults as
criminative characteristics of shock-generated varieties. zones of deformation bands and as slip surfaces in sand-
Geology 16, 796-799. stones. Pure & Appl. Geophys. 116, 931-942.
Allison, A. & La Tour, T. E. 1977. Brittle deformation of Aydin, A. & Johnson, A. M. 1983. Analysis of Faulting in
hornblende in a mylonite: a direct geometrical analogue porous sandstones. J. Struct. Geol. 5, 19-31.
of ductile deformation by translation gliding. Can. J. Bagnold, R. A. 1954. Experiments on a gravity free disper-
Earth Sci. 14, 1953-1958. sion of large solid spheres in a Newtonian fluid under
Alvarez, W., Engelder, T. & Lowrie, W. 1976. Formation of shear. Proc. R. Soc. Lond. A225, 49-63.
spaced cleavage and folds in brittle limestone by dissolu- Balé, P. & Brun, J-P. 1989. Late precambrian thrust and
tion. Geology 4, 698-701. wrench zones in northern Brittany. J. Struct. Geol. 11,
Alvarez, W., Engelder, T. & Geisser, P. A. 1978. Classifica- 391-405.
tion of solution cleavage in pelagic limestones. Geology Bansal, G. K. 1977. On fracture mirror formation in glass and
6, 263-266. polycrystalline ceramics. Philos. Mag. 35, 935-44.
Ameen, M. S. (editor) 1995. Fractography: fracture topo- Bathurst, R. G. C. 1958. Diagnetic fabrics in some British
graphy as a tool in fracture mechanics and stress analysis. Dinantian limestones. Liverpool and Manchester Geol. J.
Spec. Publs. geol. Soc. Lond. 92. 2, 11-36.
Anderson, O. L. & Grew, P. C. 1977. Stress Corrosion Theory Baud, P., Reuschlé T. & Charlez, P. 1996. An improved wing
of Crack Propagation with application to geophysics. Rev. crack model for deformation and failure of rocks in com-
Geophys. and Space Phys. 15, 77-104. pression. Int. J. Rock Mech. & Mining Sci. & Geomech.
Andrews, L. M. & Railsback, L. B. 1997. Controls on Abs. 33, 539-542.
stylolite development: Morphologic, lithologic and tem- Beach, A. 1979. Pressure solution as a metamorphic process
poral evidence from bedding-parallel and transverse in deformed terrigenous sedimentary rocks. Lithos. 51-
stylolites from the U.S. Appalachians. J. Geol. 105, 59- 58.
73. Beaumont, C, Ellis, S., Hamilton, J. & Fullsack, P. 1996.
Angevine, C. L. & Turcotte, D. L. 1983. Porosity reduction by Mechanical model for subduction-collision tectonics of
pressure solution: A theoretical model for quartz arenites. Alpine-type compressional orogens. Geology 24, 675-
Bull. geol. Soc. Am. 94, 1129-1134. 678.
Angelier, J. 1984. Tectonic analysis of fault slip data sets. J. Becker, A. 1995. Quartz pressure solution: influence of crys-
geophys. Res. 89, 5835-5848. tallographic orientation. J. Struct. Geol. 17, 1395-1406.
Antonellini, M. A. & Pollard, D. D. 1994. Distinct element Behrmann, J. H. 1984. A study of white mica microstructure
modeling of deformation bands in sandstone. J. Struct. and microchemistry in a low grade mylonite. J. Struct.
Geol. 16, 1165-1182. Geol. 6, 283-292.
Arzi, A. A. 1978. Critical phenomena in the rheology of par- Behrmann, J. H. 1985. Crystal plasticity and superplasticity
tially melted rocks. Tectonophysics 44, 173-184. in quartzite: a natural example. Tectonophysics 115,101-
Ashby, M. F. & Verrall, R. A. 1978. Micromechanisms of 129.
flow and fracture, and their relevance to the rheology of Behrmann, J. H. 1987. A precautionary note on shear bands

107
108 REFERENCES

as kinematic indicators. J. Struct. Geol. 9, 659-666. ing and cataclasis involving reaction softening within core
Behrmann, J. H. & Platt, J. H. 1982. Sense of nappe emplace- material from the Cajon Pass drillhole. J. geophys. Res.
ment from quartz c-axis fabrics. Earth Planet. Sci. Lett. 97.5135-5144.
59, 208-215. Blenkinsop, T. G. & Treolar, P. J. 1995. Geometry, classific-
Bell, T. H. 1985. Deformation partioning and porphyroblast ation and kinematics of S-C fabrics in the Mushandike
rotation in metamorphic rocks: a radical interpretation. J. area, Zimbabwe. J. Struct. Geol. 17, 397-408.
metam. Geol. 3, 109-118. Blumenfeld, P. 1983. Le “tuilage des mégacristaux”-un
Bell, T. H. & Johnson, S. E. 1989. Porphyroblast inclusion critère d’écoulement rotationnel pour les fluidalités des
trails: the key to orogenesis. J. metam. Geol. 3, 109-118. roches magmatiques-application au granite de Barbey.
Bell, T. H. & Cluff, C. 1989. Dissolution, solution transfer, Bull. Soc. géol. Fr. 25, 309-318.
diffusion versus fluid flow and volume loss during de- Blumenfeld, P. & Bouchez, J. L. 1988. Shear criteria in gran-
formation/metamorphism. J. metam. Geol. 7, 425-447. ite and migmatite deformed in the magmatic and solid
Bell, T. H., Forde, A. & Hayward, N. 1992. Do smoothly states. J. Struct. Geol. 10, 361-372.
curving, spiral-shaped inclusion trails signify porphyro- Bohor, B. F., Betterton, W. J. & Krogh, T. E. 1993. Impact
blast rotation? Geology 20, 59-62. shocked zircons: discovery of shock-induced textures re-
Bell, T. H., Johnson, S. E., Davis, B., Forde, A., Hayward, flecting increasing degrees of shock metamorphism. Earth
N. & Wilkins, C. 1992. Porphyroblast inclusion-trail ori- Planet. Sci. Lett. 119, 419-424.
entation data: eppure non son girate. J. metam. Geol. 10, Boland, J. N. & Tullis, T. E. 1986. Deformation behaviour of
295-307. wet and dry clinopyroxentite in the brittle to ductile trans-
Bennema & van der Eerden, 1987. Morphology of crystals. ition region. In: Mineral and rock deformation: Laborat-
In I. Sunangwa (ed.). Terra Publications, Tokyo. ory Studies-The Paterson Volume (edited by Hobbs, B. E.
Benn, K. & Allard, B. 1989. Preferred mineral orientation & Heard, H. C.). Am. Geophys. Un. Geophys. Monogr.
related to magmatic flow in ophiolite layered gabbros. J. 36, 35-50.
Petrol. 30, 925-946. Boldt, L. A. 1995. Gypsum plate enhancement of cataclastic
Berckhemer, H., Auer, F. & Drisler, J. 1979. High temperat- fabric in thin section. J. Struct. Geol. 17, 301-303.
ure and elasticity and elasticity of mantle peridotite. Phys. Borg, I. & Maxwell, J. C. 1956. Interpretation of fabrics of
Earth & Planet. Interiors 20, 48-79. experimentally defromed sands. Am. J. Sci. 254, 71-81.
Bernabé, Y. 1987. The effective pressure law for permeabil- Borg, I., Friedman, M., Handin, J. & Higgs, D. 1960. Exper-
ity during pore pressure and confining pressure cycling of imental Deformation of St. Peter Sand: A Study of Cata-
several crystalline rocks. J. geophys. Res. 92, 649-657. clastic Flow. In: Rock Deformation (edited by Griggs, D.
Berthé, D., Choukroune, D. & Jegouzo P. 1979. Orthogen- & Handin, J.) Mem. geol. Soc. Am. 79, 193-226.
esis, mylonite and non-coaxial deformation of granites: Borradaile, J. G. 1981. Particulate flow and the generation of
the example of the South Armorican Shear Zone. J. Struct. cleavage. Tectonophysics, 72, 306-321.
Geol. 1, 31-34. Borradaile, J. G., Bayly, M. B. & Powell, C. M. A. 1982.
Biegel, R. I., Sammis, C. G. & Dieterich, J. H. 1989. The Atlas of deformational and metamorphic rock fabrics.
frictional properties of a simulated gouge having a fractal Springer, Berlin Heidelberg New York.
particle distribution. J. Struct. Geol. 11, 827-846. Bouchez, J. L. & Gleizes, G. 1995. Two-stage deformation of
Biegel, R. I., Wang, C. H., Boitnott, G. N. & Yoshioka, N. the Mont-Louis-Andorra granite pluton (Variscan Pyren-
1992. Micromechanics of rock friction, 1, Effects of sur- ees) inferred from magnetic susceptibility anisotropy. J.
face roughness on initial friction and slip hardening in geol. Soc. 152, 669-680.
Westerly granite. J. geophys. Res. 97, 8965-8978. Boullier, A. M. & Gueguen, Y. 1975. SP-Mylonites: origin of
Bjørnerud, M. G. & Zhang, H. 1995. Flow mixing, object some mylonitesby superplastic flow. Contr. Miner. Petrol.
matrix coherence, mantle growth and the development of 50, 93-104.
porphyroclast tails. J. Struct. Geol. 17, 1347-1350. Bouchez, J. L., Lister, G. S. & Nicolas, A. 1983. Fabric asym-
Blacic, J. D. & Christie, J. M. 1984. Plasticity and hydrolytic metry and shear sense in movement zones. Geol. Rdsch.
weakening of quartz single crystals. J. geophys. Res. 89, 72, 401-419.
4223-4239. Bouchez, J. L., Delas, C., Gleizes, G., Nedelec, A. & Cuney,
Blenkinsop, T. G. 1989. Thickness-displacement relation- M. 1992. Submagmatic microfractures in granites. Geo-
ships for deformation zones: Discussion. J. Struct. Geol. logy 20, 35-38.
11, 1051-1054. Brace, W. E. & Bombolakis, E. G. 1963. A Note on Brittle
Blenkinsop, T. G. 1991. Cataclasis and processes of particle Crack Growth in Compression. J. geophys. Res. 68, 3709-
size reduction. Pure & Appl. Geophys. 136, 1-33. 13.
Blenkinsop, T. G. & Rutter, E. H. 1986. Cataclastic deforma- Brace, W. F. & Martin, R. J. 1968. A test of the law of effect-
tion of quartzite in the Moine thrust zone. J. Struct. Geol. ive stress for crystalline rocks of low porosity. Int. J. Rock
8, 669-681. Mech. &Min. Sci. 5, 415-426.
Blenkinsop, T. G. & Drury, M. R. 1988. Stress estimates and Brantley, S. L. 1992. The effect of fluid chemistry on quartz
fault history from quartz microstructures. J. Struct. Geol. microcrack lifetimes. Earth Planet. Sci. Lett. 113, 145-
10, 673-684. 156.
Blenkinsop, T. G. & Sibson, R. H. 1991. Aseismic fractur- Brown, W. L. & Macaudière, J. 1984. Microfracturing in re-
REFERENCES 109

lation to atomic structure of plagioclase from a deformed formed quartz and calcite from the Dry Creek Ridge An-
meta-anorthosite. J. Struct. Geol. 6, 579-586. ticline, Montana. Am. J. Sci. 263, 747-785.
Brown, W. L. & Scholz, C. H. 1985. Broad bandwidth study Carter, N. L. & Rayleigh, C. B. 1969. Principal stress direc-
of the topography of natural rock surfaces. J. geophys. tions from plastic flow in crystals. Bull. geol. Soc. Am.
Res. 90, 12575-12582. 80, 1231-1264.
Bruner, W. M. 1984. Crack growth during unroofing of crustal Carter, N. L. & Tsenn, M. C. 1987. Flow properties of con-
rocks: Effects of thermoelastic behaviour and near - tinental lithosphere. Tectonophysics 136, 27-63.
surface stresses. J. geophys. Res. 89, 4167-4184. Carter, N. L., Officer, C. B., Chesner, C. A. & Rose, W. I.
Burg, J-P. 1987. Regional shear variation in relation to diapir- 1986. Dynamic deformation of volcanic ejecta from the
ism and folding. J. Struct. Geol. 9, 925-934. Toba caldera: possible relevance to Cretaceous/Tertiary
Burg, J-P. 1991. Quartz shape fabric variations and c-axis fab- boundary phenomena. Geology 14, 380-383.
rics in a ribbon-mylonite: arguments for an oscillating fo- Carter, N. L., Officer, C. B. & Drake, C. L. 1990. Dynamic
liation. J. Struct. Geol. 8, 123-132. deformation of quartz and feldspar: clues to causes of
Burkhard, M. 1993. Calcite twins, their geometry, appearance some natural crises. Tectonophysics, 171, 373-391.
and significance as stress-strain markers and indicators of Carter, N. L, Horseman, S. T., Russell, J. E. & Handin, J.
tectonic regime: a review. J. Struct. Geol. 15, 351-368. 1993. Rheology of rocksalt. J. Struct. Geol. 15, 1257-
Burnley, P. C., Green, H. W. & Prior, D. J. 1991. Fault- 1271.
ing associated with the olivine-spinel transformation in Casey, W. H. 1995. Surface chemistry during the dissolution
and its implications for deep-focus earth- of oxide and silicate materials. In: Mineral Surfaces (ed-
quakes. J. geophys. Res. 96, 425-443. ited by Vaughan, D. J. & Pattrick, R. A. D.). Chapman
Burov, E. B. & Diament, M. 1995. The effective elastic thick- and Hall, London. 185-217.
ness (Te) of continental lithosphere: What does it really Champness, P. E. & Lorimer, G. W. 1974. A direct lat-
mean? J. geophys. Res. 100, 3905-3927. tice resolution study of precipitation (exsolution) in or-
Burov, E. B. & Diament, M. 1996. Isostacy, equivalent thopyroxene. Philos. Mag. 30, 357-365.
elastic thickness, and inelastic rheology of continents and Chester, F. M. & Logan, J. M. 1987. Composite planar fabric
oceans. Geology, 24, 385-480. of gouge from the Punchbowl Fault, California. J. Struct.
Busa, M. D. & Gray, N. H. 1992. Rotated staurolite porphyro- Geol. 9, 621-634.
blasts in Littleton Schist at Bolton, Connecticut, U.S.A. J. Christie, J. M. & Koch, P. S. 1982. Grain size-flow stress rela-
metam. Geol. 15, 351-368. tions for novaculite deformed in hydrous assemblies (ab-
Busch, J. P. & Van der Pluijm, B. A. 1995. Calcite textures, stract). Eos, Trans. Am. Geophys. Un. 63, 1095.
microstructures and rheological properties of marble Christie, J. M. & Ord, A. 1980. Flow stress from microstruc-
mylonites in the Bancroft shear zone, Ontario, Canada. ture of mylonites: example and current assessment. J.
J. Struct. Geol. 17, 677-688. geophys. Res. 85, 6253-6262.
Byerlee, J. D. 1968. Brittle-ductile transition in rocks. J. geo- Christie, J. M. & Ardell, A. J. 1974. Substructures of deform-
phys. Res. 73, 4741-4750. ation lamellae in quartz. Geology 2, 405-408.
Byerlee, J. D. 1978. Friction of rock. Pure & Appl. Geoph. Chopin, C. 1984. Coesite and pure pyrope in high grade
116, 807-839. blueschists of the Western Alps: A first record and some
Camacho, A., Vernon, R. H. & Fitz Gerald, J. D. 1995. Large consequences. Contr. Miner. Petrol. 86, 107-118.
volumes of anhydrous pseudotachylite in the Woodroffe Connolly, J. A. D., Holness, M. B., Rubie, D. C. & Rush-
Thrust, eastern Musgraves Ranges, Australia. J. Struct. mer, T. 1997. Reaction-induced microcracking: An ex-
Geol. 17, 371-383. perimental investigation of a mechanism for enhancing
Caristan, Y. 1982. The transition from higher temperature anatectic melt extraction. Geology, 25, 591-594.
creep to fracture in Maryland diabase. J. geophys. Res. Conrad, R. E. & Friedman, M. 1976. Microscopic Feather
87, 6781-6790. Fractures in the Faulting Process. Tectonophysics 33,
Carrio-Shaffhauser, E. & Gaviglio, P. 1990. Pressure solution 187-198.
and cementation stimulated by faulting in limestones. J. Cooper, R. F. & Kohlstedt, D. L. 1982. Interfacial energies
Struct. Geol. 12, 987-994. in the olivine-basalt system. In: High pressure research in
Carrio-Schaffhauser, E., Raynaud, S., Latière, H. J. & geophysics (edited by Akimata, S. & Manghnari, M. H.).
Mazorelle, F. 1992. Propagation and localization of Adv. Earth Planet Sci. 12, 217-228.
stylolites in limestones. In: Deformation Mechanisms, Cooper, R. F. & Kohlstedt, D. L. 1984. Solution-precipitation
Rheology and Tectonics (edited by Knipe, R. J. & Rut- enhanced diffusional creep of particles molten olivine-
ter, E. H.). Spec. Publs geol. Soc. Lond. 54, 193-199. basalt aggregates during hot- pressing. Tectonophysics
Carter, N. L. 1965. Basal quartz deformation lamellae - A cri- 107, 207-233.
terion for recogniton of impactites. Am. J. Sci. 263, 786- Cooper, R. F., Kohlstedt, D. L. & Chyung. 1989. Solution-
806. precipitation enhanced creep in solid-liquid aggregates
Carter, N. L. & Kirby, S. H. 1978. Transient Creep and that display a non-zero dihedral angle. Acta. Metall. 37,
Semibrittle behaviour of Crystalline Rocks. Pure & Appl. 1759-1771.
Geophys. 116, 807-839. Costin, L. S. 1983. A Microcrack Model for the Deformation
Carter, N. L. & Friedman, M. 1965. Dynamic analysis of de- and Failure of Brittle Rock. J. geophys. Res. 88, 9485-
110 REFERENCES

9492. temperature creep. J. Mech. Phys. Solids 3, 85-116.


Cox, S. J. D. & Atkinson, B. K. 1983. Fracture Mechanics Drury, M. R. 1993. Deformation lamellae in metals and
and acoustic emission of anti-plane shear cracks in rocks. minerals. In: Defects and processes in the solid state:
Earthquake Prediction and Res. 2, 1-24. Geoscience Applications - The McLaren Volume (edited
Cox, S. F. 1987. Antiaxial crack-seal vein microstructures and by Boland, J. N. & Fitzgerald, J. D.). Elsevier Science
their relationship to displacement paths. J. Struct. Geol. 9, Publishers B. V. 195-212.
779-788. Drury, M. R. & Humphreys, P. J. 1988. Microstructural
Cruikshank, K. M., Zhao, G. & Johnson, A. M. 1991. Ana- shear criteria associated with grain-boundary sliding dur-
lysis of minor fractures associated with joints and faulted ing ductile deformation. J. Struct. Geol. 10, 83-89.
joints. J. Struct. Geol. 13, 865-886. Drury, M. R. & Urai, J. L. 1990. Deformation-related recrys-
D’Arco, P. & Wendt, A. S. 1994. Radial cracks around in- tallisation processes. Tectonophysics 172, 235-253.
clusions: A program to calculate P-T paths with re- Drury, M. R., Humphreys, P. J. & White, S. H. 1985.
spect to elastic properties of minerals. Computers and Large strain deformation studies using polycrystalline
Geosciences, 20, 1275-1283. magnesium as a rock analogue Part II: dynamic recrys-
Das, S. & Scholz, C. H. 1981. Theory of Time-Dependent tallization mechanisms at high temperatures. Phys. Earth
Rupture in the Earth. J. geophys. Res. 86, 6039-6051. Planet. Interiors 40, 208-222.
De Bresser, J. H. P. 1988. Deformation of calcite crystals Dunn, D. E., La Fountain, L. J. & Jackson, R. E. 1973. Poros-
by r+ anf f+ slip: mechanical behaviour and dislocation ity dependence and mechanism of brittle fracture in sand-
density vs. stress relation. Eos, Trans. Am. Gephys. Un. stones. J. geophys. Res. 78, 2403-2417.
69, 1418. Durney, D. W. & Ramsay, J. G. 1973. Incremental strains
De Bresser, J. H. P. & Spiers, C. J. 1988. Deformation of measured by syntectonic crystal growth. In: Gravity and
calcite crystals by r+ and f+ slip. In: Deformation Mech- tectonics (edited by de Jong, K. A., Scholten, R. ). John
anisms, Rheology and Tectonics (edited by Knipe, R. J. & Wiley and Sons, New York, 67-96.
Rutter, E. H.). Spec. Publs geol. Soc. Lond. 54, 285-298. Elliot, D. 1973. Diffusion flow laws in metamorphic rocks.
Dietrich, D. & Grant, P. R. 1986. Cathodoluminescence pet- Bull. geol. Soc. Am. 84, 2645-2664.
rography of syntectonic quartz fibres. J. Struct. Geol. 7, Engelder, T. 1974. Cataclasis and the generation of fault
541-568. gouge. Bull. geol. Soc. Am. 85, 1515-1522.
Dell’Angelo, L. N. & Tullis, J. 1988. Experimental deform- Englehardt, W. & von Bertsch, W., 1969. Shock-induced
ation of partially melted granitic aggregates. J. metam. planar deformation structures in quartz from the Ries
Geol. 6, 495-516. crater, Germany. Contr. Miner. Petrol. 20, 203-234.
Dell’Angelo, L. N., Tullis, J. & Yund, R. A. 1987. Transition Erskine, B. G., Heidelbach, F. & Wenk, H-R. 1993. Lattice
from dislocation creep to melt-enhanced diffusion creep preferred orientations and microstructures of deformed
in fine-grained granitic aggregates. Tectonophysics, 139, Cordilleran marbles: correlation of shear indicators and
325-332. determination of strain path. J. Struct. Geol. 15, 1189-
Den Brock, B. 1996. The effect of crystallographic orienta- 1206.
tion on pressure solution in quartzite. J. Struct. Geol. 18, Erslev, E. A. & Ward, D. J. 1994. Non-volatile element and
859-860. volume flux in coalesced slaty cleavage. J. Struct. Geol.
Den Brock, S. W. J. & Spiers, C. J. 1991. Experimental evid- 16, 531-554.
ence for water weakening of quartzite by microcracking Etchecopar, A. & Malavielle, J. 1987. Computer models of
plus solution-precipitation creep. J. geol. Soc. London pressure shadows: a method of strain measurement and
148, 541-548. shear sense determination. J. Struct. Geol. 9, 667-677.
Dennis, P. F. 1984. Oxygen self-diffusion in quartz under hy- Evans, B. & Dresden, G. 1991. Deformation of Earth Mater-
drothermal conditions. J. geophys. Res. 89, 4047-4057. ials: Six Easy Pieces. Contributions in Tectonophysics,
Dennis, A. J. & Secor, D. T. 1987. A model for the devel- U.S. National Report to International Union of Geodesy
opment of crenulations in shear zones with applications and Geophysics, 1987-1990. Am. Geophys. Un., 823-
from the southern Appalachian piedmont. J. Struct. Geol. 843.
9, 809-817. Evans, B., Rowen, M. & Brace, W. F. 1980. Grain-size sens-
Dennis, A. J. & Secor, D. T. Jr. 1990. On resolving shear itive deformation of stretched conglomerate from Ply-
direction in foliated rocks deformed by simple shear. Bull, mouth, Vermount. J. Struct. Geol. 2. 411-424.
geol. Soc. Am. 102, 1257-1267. Evans, B., Frederich, J. T. & Wong, T-F. 1990. The Brittle-
Detournay, E., Cheng, A. H-D., Roegiers, J-C. & McLennan, Ductile Transition in rocks: Recent Experimental and
J. D. 1989. Poroelasticity considerations in in-situ stress Theoretical Progress. In: The Brittle-Ductile Transition
determination by hydraulic fracturing. Int. J. Rock Mech. in Rocks (edited by A. G. Duba, W. B. Durham, J. W.
& Mining Sci. & Geomech. Abs. 26, 507-513. Handin & H. F. Wang) Am. Geophys. Un. Geophys.
Doblas, M., Mahecha, V., Hoyos, M. & Lopez-Ruiz, J. 1997. Monogr. 56, 1-20.
Slickenside and fault surface kinematic indicators on act- Evans, D. J. & White, S. H. 1984. Microstructural and fabric
ive normal faults of the Alpine Betic Cordilleras, Granada, studies from the rocks of the Moine Nappe, Eriboll, N.W.
southern Spain. J. Struct. Geol. 19, 159-170. Scotland. J. Struct. Geol. 6, 369-389.
Dorn, J. E. 1954. Some fundamental experiments on high Evans, J. P. 1988. Deformation mechanisms in granitic rocks
REFERENCES 111

at shallow crustal levels. J. Struct. Geol. 10, 437-444. preferred orientations in experimentally deformed quartz
Evans, M. A. & Groshong, R. H. 1994. A computer pro- aggregates. J. Struct. Geol. 15, 1145-1168.
gramme for the calcite strain-gauge technique. J. Struct. Gigl, P. & Dachville, F. 1968. Effects of pressure and temper-
Geol, 16, 277-282. ature on the reversal transition of stishovite. Meteoritics
Farver, J. R. & Yund, R. A. 199la. Oxygen diffusion in 4, 123-136.
quartz: dependance on temperature and water fugacity. Giletti, B. J. & Yund, R. A. 1984. Oxygen diffusion in quartz.
Chem. Geol. 90, 55-70. J. geophys. Res. 89, 4039-4046.
Farver, J. R. & Yund, R. A. 1991b. Measurement of oxygen Gilotti, J. A. & Hull, J. M. 1990. Phenomenological super-
grain boundary diffusion in natural, fine-grained quartz plasticity in rocks. In: Deformation Mechanisms, Rhe-
aggregates. Geochim. Cosmochim. Acta 55, 1597-1607. ology and Tectonics (edited by Knipe, R. J. & Rutter, E.
Fel’dman, V. I.1994. The conditions of shock metamorphism. H.). Spec. Publs geol. Soc. Lond. 54, 29-240.
Geol. Soc. Am. Sp. Pap. 293, 121-132. Goetze, C. & Evans, B. 1979. Stress and temperature in the
Finney, J. L. 1970. Random packing and the structure of bending lithosphere as constrained by experimental rock
simple liquids 1: The geometry of random close packing. mechanics. Geophys. J. R. astr. Soc. 59, 463-478.
Proc. R. Soc. London A319, 479-493. Goldstein, A. G. 1988. Factors affecting the kinematic in-
Fisher D. M. & Anastasio, D. J. 1994. Kinematic analysis of terpretation of asymmetric boudinage in shear zones. J.
a large scale leading edge fold, Lost River Range, Idaho. Struct. Geol. 10, 707-716.
J. Struct. Geol. 16, 337-354. Goltrant, O., Cordier, P. & Doukhan, J-C. 1991. Planar de-
Fletcher, R. C. & Pollard, D. D. 1981. Anticrack model for formation features in shocked quartz: a transmission elec-
pressure solution surfaces. Geology 5, 185-187. tron microscopy investigation. Earth Planet. Sci. Lett.,
Fliervoet, T. F. & White, S. H. 1995. Quartz deformation in 106, 103-115.
a very fine grained quartzo-feldspathic mylonite: a lack Goltrant, O., Leroux, H., Doukhan, J-C. & Cordier, P. 1992.
of evidence for dominant grain boundary sliding deform- Formation mechanisms of planar deformation features in
ation. J. Struct. Geol. 17, 1095-1110. naturally shocked quartz. Phys. Earth Planet. Interiors,
Fossen, H. & Holst, T. B. 1995. Strain distribution in a fold 74, 219-240.
in the West Norwegian Caledonians. J. Struct. Geol. 9, Gower, R. J. W. & Simpson, C. 1992. Phase boundary mobil-
915-924. ity in naturally deformed, high-grade quartzofeldspathic
Fowler, M. 1990. The Solid Earth. Cambridge University rocks: evidence for diffusional creep. J. Struct. Geol. 14,
Press, Cambridge. 310-313.
French, B. M., Underwood, J. R. & Fisk, E. P. 1974. Shock Grader, J. P. & Jenatton, L. 1984. Deformation by solution-
metamorphic features in two meteorite impact strucures, deposition, and re-equilibration of fluid inclusions in
southeastern Libya. Bull. geol. Soc. Am. 85, 1425-1428. crystals depending on temperature, internal pressure and
Friedman, M. & Logan, J. M. 1970. Microscopic Feather stress. J. Struct. Geol. 6, 189-200.
Fractures. Bull. geol. Soc. Am. 81, 3417-3420. Gratier, J. P. & Guiguet, R. 1986. Experimental pressure
Gallagher, J. J. 1981. Tectonics of China: Continental Scale solution-deposition on quartz grains: the crucial effect of
Cataclastic Flow. In: Mechanical Behaviour of Crustal the nature of the fluid. J. Struct. Geol. 8, 845-856.
Rock (edited by Carter, N. L., Friedman, M., Logan, J. M. Gratz, A. J., Manne, S. & Hansma, P. K. 1991. Atomic force
& Stearns, D. W.). Am. Geophys. Un. Geophys. Monogr. microscopy of atomic scale ledges and etch pits formed
24, 259-273. during dissolution of quartz. Science 251, 1343-1346.
Gallagher, J. J., Friedman, M., Handin, J. & Sowers, G. M. Gratz, A. J., Hisler, O. K. & Bohor, B. E. 1996. Distinguish-
1974. Experimental studies relating to microfractures in ing shocked from tectonically deformed quartz by the use
sandstone. Tectonophysics 21, 203-247. of the SEM and chemical etching. Earth Planet Sci. Lett.
Gamond, J. F. 1987. Bridge structures as sense of displace- 142, 513-521.
ment criteria in brittle rocks. J. Struct. Geol. 9, 609-620. Gray, W. A. 1968. The packing of solid particles. Chapman
Gapais, D. & Brun, J-P. 1981. A comparison of mineral grain and Hall, London.
fabrics and finite strain in amphiboles from eastern Fin- Gray, D. R. 1977. Differentiation associated with crenulation
land. Can. J. Earth Sci. 18, 995-1003. cleavage. Lithos 10, 89-101.
Gapais, L. & Barbarin, B. 1986. Quartz fabric transition in a Green, H. W. 1980. On the thermodynamics of nonhydrostat-
cooling syntectonic granite. (Hermitage Massif, France). ically stressed solids. Philos. Mag. A 41, 637-647.
Tectonophysics 125, 357-370. Green, H. W. 1986. Phase transformation under stress and
German, R. M. 1985. The contiguity of liquid phase sintered volume transfer creep. In: Mineral and rock deforma-
microstructures. Metall. Trans. 16, 1247-1252. tion: Laboratory Studies-The Paterson Volume (edited by
Ghisetti, F. 1987. Mechanisms of thrust faulting in the Gran Hobbs, B. E. & Heard, H. C.). Am. Geophys. Un. Geo-
Sasso chain, central Appenines, Italy. J. Struct. Geol. 9, phys. Monogr. 36, 201-212.
955-967. Green, H. W. & Burnley, P. C. 1989. A new self-organizing
Glass, B. P. 1990. Tektites and microtektites: key facts and mechanism for deep focus earthquakes. Nature 341, 733-
inferences. Tectonophysics 171, 393-404. 737.
Gleason, G. C., Tullis, J. & Heildelbach, F. 1993. The role Grieve, R. A. F., Langenhorst, F. & Stöffler, D. 1996. Shock
of dynamic recrystallization in the development of lattice metamorphism of quartz in nature and experiment: II.
112 REFERENCES

Significance in geoscience. Meteoritics and Planetary Sci- (edited by Barber, D. J. & Meredith, P. G.) Unwin Hyman,
ence 31, 6-35. London, 84-108.
Griffith, A. A. 1924. The theory of rupture. In: Proc. 1st Int. Hancock, P. L. 1985. Brittle microtectonics: principles and
Congr. Appl. Mech. (edited by Biezeno, C. B. & Burgers, practice. J. Struct. Geol. 7, 437-451.
J. M.). Delft: Tech. Boekhandel en Drukkerij J. Waltman. Handin, J. W. & Heger, R. V. 1957. Experimental deforma-
Jr., 54-63. tion of sedimentary rocks under confining pressure: pore
Griggs, D. T. 1967. Hydrolytic weakening of quartz and other pressure tests. Bull. Am. Assoc. Petrol. Geol. 44, 1853 -
silicates. Geophys. J. R. astr. Soc. 14, 19-31. 1873.
Griggs, D. T. & Handin, J. 1960. Observations on fracture Handy, M. R. 1990. The solid state flow of poly mineralic
and a hypothesis of earthquakes. In: Rock Deformation rocks. J. geophys. Res. 95, 8647-8661.
(edited by Griggs, D. & Handin, J.) Mem. geol. Soc. Am. Handy, M. R. 1994. Flow laws for rocks containing two non-
79, 347-364. linear viscous phases: a phenomenological approach. J.
Griggs, D. T. & Blacic, J. D. 1965. Quartz; Anomalous weak- Struct. Geol. 16, 287-302.
ness of synthetic crystals. Science 147, 292 - 295. Hanmer, S. 1986. Asymmetrical pull-aparts and foliation fish
Grocott, J. 1981. Fracture geometry of pseudotachylite gen- as kinematic indicators. J. Struct. Geol. 8, 111-122.
eration zones: a study of shear fractures formed during Hanmer, S. & Passchier, C. 1991. Shear sense indicators: a
seisimic events. J. Struct. Geol. 3, 169-179. review. Geol. Soc. Canada Paper 90, 17.
Groshong, R. H. Jr. 1972. Strain calculated from twinning in Hansen, F. D. & Carter, N. L. 1982. Creep of selected crustal
calcite. Bull. geol. Soc. Am. 83, 2025-2048. rocks at 1000 MPa. Eos, Trans. Am. Geophys. Un. 63,
Groshong, R. H. Jr. 1974. Experimental test of the least 437.
squares strain gauge calculation using twinned calcite. Hansen, F. D. & Carter, N. L. 1983. Semibrittle creep of dry
Bull. geol. Soc. Am. 85, 1855-1864. and wet Westerly granite at 1000 MPa. U.S. Symp. on
Groshong, R. H. Jr. 1976. Strain and pressure solution in the Rock Mechanics, 24th, Texas A & M. Univ. College Sta-
Martinsburg Slate, Delaware Water Gap, New Jersey. Am. tion, Texas, 429-447.
J. Sci. 276, 1131-1146. Heard, H. C. 1963. Effect of large changes in strain rate in the
Groshong, R. H. Jr. 1988. Low-temperature deformation experimental deformation of Yule marble. J. Struct. Geol.
mechanisms and their interpretation. Bull. geol. Soc. Am. 71, 162-195.
100, 1329-1360. Hedlund, C. A., Anastasio, D. J. & Fisher, D. M. 1994. Kin-
Groshong, R. H. Jr., Pfiffner, O. A. & Pringle, L. R. 1984. ematics of fault-related folding in a duplex, Lost River
Strain partitioning in the Helvetic thrust belt of eastern Range, Idaho, U S A. J. Struct. Geol. 16, 571-584.
Switzerland from the leading edge to the internal zone. J. Heidug, W. K. 1991. A thermodynamic analysis of the con-
Struct. Geol. 6, 19-32. ditions of equilibrium at nonhydrostatically stressed and
Gross, A. F. K. & Van Heege, J. P. T. 1973. The high-low curved phase boundaries. J. geophys. Res. 96, 909-921.
quartz transition up to 10 kb pressure. J. Geol. 81., 717- Hibbard, M. J. 1987. Deformation of incompletely crystal-
724. lised magma systems: granitic gneisses and their tectonic
Guinenberteau, B., Bouchez, J-L. & Vigneresse, J-L. 1987. implications. J. Geol. 951, 543-561.
The Montagne granite pluton (France) emplaced by pull- Hickman, S. H. & Evans, B. 1991. Experimental pressure
apart along a shear zone: structural and gravimetric argu- solution in halite: the effect of grain/interphase boundary
ments and regional implication. Bull. geol. Soc. Am. 99, structure. J. geol. Soc. Lond. 148, 549-560.
763-770. Hillner, P. E., Gratz, A. J., Manne, S. & Hansma, P. K. 1992.
Guzzeta, G. 1984. Kinematics of stylolite formation and Atomic-scale imaging of calcite growth and dissolution in
physics of pressure solution process. Tectonophysics 101, real time. Geology 20, 359-362.
383-394. Hippert, J. F. M. 1994. Grain boundary microstructures in
Hacker, B. R. & Kirby. S. H. 1993. High-pressure deform- micaceous quartzite: significance for fluid movement and
ation of calcite marble and its transformation to aragon- deformation processes in low metamorphic grade shear
ite under non-hydrostatic conditions. J. Struct. Geol. 15, zones. J. Geol. 102, 331-348.
1207-1222. Hirth, G. & Tullis, J. 1991. Mechanisms responsible for
Hadizadeh, J. 1980. An experimental study of cataclastic de- the brittle-ductile transition in experimentally deformed
formation in a quartzite. Unpublished Ph.D. thesis, Im- quartz aggregates. Eos, Trans. Am. Geophys. Un. 72, 286.
perial College, University of London.79, 133-191. Hirth, G. & Tullis, J. 1992. Dislocation creep regimes in
Hadizadeh, J. & Rutter, E. H. 1983. The low temperature quartz aggregates. J. Struct. Geol. 14, 145-149.
brittle-ductile transition in quartzite and the occurrence Hirth, G. & Kohlstedt, D. L. 1995. Experimental constraints
of cataclastic flow in nature. Geol. Rdsch. 72, 493-509. on the dynamics of the partially molten upper mantle: De-
Hadizadeh, J. & Tullis, J. 1992. Cataclastic flow and semi- formation in the diffusion creep regime. J. geophys. Res.
brittle deformation of anorthosite. J. Struct. Geol. 14, 57- 100, 1981-2001.
64. Hobbs, B. E., Ord, A. & Teyssier, C. 1986. Earthquakes in the
Hallan, S. D. & Ashby, M. F. 1990. Compressive brittle fail- ductile regime? Pure & Appl. Geophys. 124, 309-336.
ure and the construction of multi-axial failure maps. In: Horii, H. & Nemat Naser, S. 1985. Compression-InducedMi-
Deformation processes in Minerals, Ceramics and Rocks, crocrack Growth in Brittle Solids: Axial Splitting and
REFERENCES 113

Shear Failure. J. geophys. Res. 90, 3105-3125. Mechanisms, Rheology and Tectonics (edited by Knipe,
Horii, H. & Nemat Naser, S. 1986. Brittle failure in compres- R. J. & Rutter, E. H.). Spec. Publs geol. Soc. Lond. 54,
sion: splitting, faulting and brittle-ductile transition. Phil. 353-362.
Trans. R. Soc. Lond. A319, 337-374. Ji, S. & Zhao, P. 1993. Flow laws of multiphase rocks cal-
Hörz, F. 1968. Statistical measurement of deformation struc- culated from experimental data on the constituent phases.
tures and refractive indices in experimentally shock Earth Planet. Sci. Lett., 117, 181-187.
loaded quartz. In: Shock metamorphism of natural mater- Ji, S. & Zhao, P. 1994. Strength of two-phase rocks: a model
ials (edited by French, B. M. & Short, N. M.), Baltimore, based on fibre-loading theory. J. Struct. Geol. 16, 253-
Mono book corporation, 243-254. 262.
Hörz, F. & Quaide, W. L. 1973. Debye-Scherrer investiga- Johnson, S. E. & Vernon, R. H. 1995. Inferring the tim-
tions of experimentally shocked silicates. The Moon 6, ing of porphyroblast growth in the absence of continuity
45-82. between inclusion trails and matrix foliations: can it be
House, W. M. & Gray, D. R. 1982. Cataclasites along the really done? J. Struct. Geol. 17, 1203-1206.
Salteville thrust U.S.A., and their implications for thrust Jarewicz, S. R. & Watson, E. B. 1984. Distribution of partial
sheet emplacement. J. Struct. Geol. 4, 257-269. melt in a felsic system: the importance of surface energy.
Huffman, A. R. & Reimold, W. U. 1996. Experimental con- Contr. Miner. Petrol. 85, 25-29.
straints on shock-induced microstructures in naturally de- Jarewicz, S. R. & Watson, E. B. 1985. The distribution of
formed silicates. Tectonophysics 256, 165-217. partial melt in a granitic system: the application of liquid
Huffman, A. R., Brown, M. J., Carter, N. L. & Reimold, W. U. phase sintering theory. Geochim. Cosmochim. Acta 49,
1993. The microstrucutural response of quartz and feld- 1109-1121.
spar under shock loading at variable temperatures. J. geo- Joreau, P., Reimold, W. U., Robb, L. J. & Doukham, J-C.
phys. Res. 98, 22171-22197. 1997. A TEM study of deformed quartz grains from vol-
Hull, D. 1975. An introduction to dislocations. Pergammon caniclastic sediments associated with the Bushveld com-
Press, Oxford. plex, South Africa. Eur. J. Mineral. 9, 393-401.
Hull, J. 1988. Thickness displacement relationships for de- Kamo, S. L., Reimold, W. U., Krogh, T. E. & Colliston, W. P.
formation zones. J. Struct. Geol. 10, 431-435. 1996. A 2.023 Ga age for the Vredefort impact event and
Hutton, D. H. W. 1988. Granite emplacement mechanism and a first report of shock metamorphosed zircons in pseudot-
tectonic controls: inferences from deformation studies. achylite breccias and granophyre. Earth Planet. Sci. Lett.
Trans. R. Soc. Edinb. Earth Sci. 79, 245-255. 144, 369-387.
Ildefonse, B. & Mancktelow, N. S. 1993. Deformation around Karatao, S. & Wu, P. 1993. Rheology of the upper mantle: a
rigid particles: the influence of slip at the particle/matrix synthesis. Science 260, 771-778.
interface. Tectonophysics 221, 345-359. Karlstrom, K. E., Miller, C. F., Kingsbury, J. A. & Wooden,
Inglis, C. E. 1913. Stresses in a plate due to the presence of J. L. 1993. Pluton emplacement along a ductile thrust
cracks and sharp corners. Trans. Inst. Nav. Archit. London zone, Pine mountains, southeastern California: Interac-
55, 219-241. tion between deformational and solidification processes.
Jaeger, J. C. & Cook, N. G. W. 1979. Fundamentals of Rock Bull. geol. Soc. Am. 105, 213-230.
Mechanics. 3rd ed. London, Chapman and Hall, 585pp. Kemeny, J. M. & Cook, N. G. W. 1987. Crack models for the
Jamison, W. R. & Spang, J. H. 1976. Use of calcite twin failure of rocks in compression. In 2nd Int. Conf. Con-
lamellae to infer differential stress. Bull. geol. Soc. Am. stitutive laws for engineering materials, Tucson, 879-887.
87, 868-872. Kieffer, S. W. 1975. From regolith to rock by shock. The
Jamison W. R. & Stearns, D. W. 1982. Tectonic Deform- Moon, 13, 301-320.
ation of Wingate Sandstone, Colorado National Monu- Kieffer, S. W., Phakey, P. R. & Christie, J. M. 1976. Shock
ment. Bull. Am. Ass. Petrol. Geol. 66, 2584-2608. processes in porous quartzite; transmission electron mi-
Jaoul, O., Tullis, J. & Kronenberg, A. 1984. The effect of croscope observations and theory. Contr. Miner. Petrol.
varying water contents on the creep behaviour of Heavit- 59, 41-93.
ree quartzite. Am. J. Sci. 266, 1-42. Kirby, S. H. & Kronenberg, A. K. 1984. Deformation of
Jeffrey, J. B. 1922. The motion of ellipsoidal particles in vis- clinopyroxenite: Evidence for a transition in Flow Mech-
cous fluid. Proc. R. Soc. Lond. A102, 161-179. anisms and Semibrittle behaviour. J. geophys. Res. 89,
Jessell, M. W. 1987. Grain boundary migration microstruc- 3177-3192.
tures in a naturally deformed quartzite. J. Struct. Geol. 9, Kirby, S. H. & Stern, L. A. 1993. Experimental dynamic
1007-1014. metamorphism of mineral single crystals. J. Struct. Geol.
Jessell, M. W. 1988a. A simulation of fabric development in 15, 1223-1240.
recrystallizing aggregates -I: description of the model. J. Knipe, R. J. 1979. Chemical changes during slaty cleavage
Struct. Geol. 10, 771-778. development. Bull. Min. 105, 206-209.
Jessell, M. W. 1988b. A simulation of fabric development Knipe, R. J. 1986. Microstructural evolution af vein arrays
in recrystallizing aggregates -II: example model runs. J. preserved in D.S.D.P. cores from Japan Trench Leg 57. In:
Struct. Geol. 10, 779-793. Structural fabrics in D.S.D.P. Cores from Forearcs (edited
Jessell, M. W. & Lister, G. S. 1990. A simulation of the tem- by Moores, J. C.). Mem. geol. Soc. Am. 166, 151-160.
perature dependence of quartz fabrics. In: Deformation Knipe, R. J. 1989. Deformation mechanisms - recognition
114 REFERENCES

from natural tectonites. J. Struct. Geol. 11, 127-146. Kusznir, N. J. 1982. Lithosphere response to externally and
Knipe, R. J. & White, S. H. 1977. Microstructural variation of internally derived forces: a viscoelastic stress guide with
an axial planar cleavage around a fold - H.V.E.M. study. amplification. Geophys. J. R. astr. Soc. 70, 399-414.
Tectonophysics 39, 355-380. Kusznir, N. J. & Bott, M. P. H. 1977. Stress concentration in
Knipe, R. J. & White, S. H. 1979. Deformation in low grade the upper lithosphere caused by underlying visco-elastic
shear zones in the Old Red Sandstone, S.W. Wales. J. creep. Tectonophysics 43, 247-256.
Struct. Geol. 1, 53-66. Kusznir, N. J. & Park, R. G. 1982. Intraplate lithosphere
Koch, P. S. & Christie, J. M. 1981. Spacing of deformation strength and heat flow. Nature 299, 540-542.
lamellae as a paleopiezometer. Abstr., Eos, Trans. Am. Kusznir, N. J. & Park, R. G. 1984. Intraplate lithosphere de-
Geophys. Un. 62, 1030. formation and the strength of the lithosphere. Geophys. J.
Koeberl, C. 1990. The geochemistry of tectites: an overview. R. astr. Soc.79, 513-538.
Tectonophysics 171, 405-422. Kusznir, N. J. & Park, R. G. 1987. The extensional strength of
Kollé J. J. & Blacic, J. D. 1983. Deformation of single crystal the continental lithosphere: its dependence on geothermal
clinopyroxenes. 1. Mechanical twinning in diopside and gradient, and crustal composition and thickness. In: Con-
hedenbergite. J. geophys. Res. B-88, 2381-2393. tinental extensional tectonics (edited by Coward, M. P.,
Komar, P. D. 1972a. Mechanical interactions of phenocrysts Dewey, J. F. & Hancock, P. L.). Spec. Publs geol. Soc. 28,
and flow differentiation of igneous dykes and sills. Bull, 35-52.
geol. Soc. Am. 83, 973-988. Labuz, J. F., Shah, S. P. & Dowding C. H. 1987. The fracture
Komar, P. D. 1972b. Flow differentiation in igneous dykes process zone in granite: evidence and effect. Int. J. Rock
and sills: Profiles of velocity and phenocryst concentra- Mech. & Mining Sci. & Geomech. Abs. 24, 235-246.
tion. Bull. geol. Soc. Am. 83, 3443-3448. Lachenbruch, A. H. & Sass, J. H. 1980. Heat flow and ener-
Krabbendam, M. & Leslie, A. G. 1996. Folds with vergence getics of the San Andreas fault zone. J. geophys. Res., 85,
opposite to the sense of shear. J. Struct. Geol. 18, 777-782. 6185-6222.
Krantz, R. L. 1979. Crack-crack and crack-pore interactions Lacombe, O., Angelier, J., Laurent, P., Bergerat, F. & Tourn-
in stressed granite. Int. J. Rock Mech. & Mining Sci. & eret, C. 1990. Joint analyses of calcite twins and fault slips
Geomech. Abstr. 16, 37-47. as a key for deciphering polyphase tectonics: Burgundy
Krantz, R. L. 1983. Review of microcracks. Tectonophysics as a case study. Tectonophysics 182, 279-300.
100, 449-480. Lagarde, J-L., Dallain, C., Ledru, P. & Courrioux, G. 1994.
Krantz, R. L. & Scholz, C. H. 1977. Critical dilatant volume Strain patterns within the late Variscan granitic dome of
of rocks at the onset of tertiary creep. J. geophys. Res. 82, Velay, French Massif Central. J. Struct. Geol. 16, 839-
4893-4897. 852.
Krogh, T. E., Kamo, S. L. & Bohor, B. F. 1996. Shock meta- Langdon, T. G. 1982. Fracture processes in superplastic flow.
morphosed zircons with correlated U-Pb discordance and Metal. Sci. 16, 175-183.
melt rocks with concordant ages inidcate an impact origin Langenhorst, F. 1994. Shock experiments on pre-heated
for the Sudbury structure. In: Earth processes: reading the and II. X-ray and TEM investigations. Earth
isotopic code, (edited by Basu, A. & Hart, S.) Am. Geo- Planet. Sci. Lett. 128, 683-698.
phys. Un. Geophys. Monogr. 95, 343-354. Langenhorst, F. & Deutsch, A. 1994. Shock experiments on
Kronenberg, A. K. & Tullis J. 1984. Flow strengths of Quartz pre-heated and I. Optical and density data.
Aggregates: Grain Size and Pressure Effects due to Hy- Earth Planet. Sci. Lett. 125, 407-420.
drolytic Weakening. J. geophys. Res. 89, 4281-4297. Langenhorst, F. & Clymer, A. K. 1996. Characteristics of
Kronenberg, A. K., Segall, P. & Wolf, G. H. 1990. Hydrolytic shocked quartz in late Eocene impact ejecta from Massig-
weakening and penetrative deformation within a natural nano (Ancona, Itlay): Clues to shock conditions and
shear zone. In: The Brittle-Ductile Transition in Rocks source crater. Geology, 24, 487-490.
(edited by A. G. Duba, W. B. Durham, J. W. Handin & H. Langenhorst, F., Deutsch, A., Hornemann, U. & Stöffler, D.
F. Wang). Am. Geophys. Un. Geophys. Monograph. 56, 1992. Effect of temperature on shock metamorphism of
21-36. quartz. Nature 356, 507-509.
Kruhl, J. H. 1996. Prism and basal-plane parallel subgrain Laurent, P., Bernard, Ph., Vasseur, G. & Etchecopar, A. 1981.
boundaries in quartz: a microstructural geothermobaro- Stress tensor determination from the study of e-twins in
meter. J. metam. Geol. 14, 581-589. calcite: a linear programming method. Tectonophysics
Kruhl, J. H. & Nega, M. 1996. The fractal shape of su- 78, 651-660.
tured quartz grain boundaries: application as a geother- Laurent, P., Tourneret, C. & Laborde, O. 1990. Determining
mometer. Geol. Rdsch. 85, 38-43. deviatoric stress tensors from calcite twins: applications
Kulander, B. R. & Dean, S. L. 1995. Observations on fracto- to monophased synthetic and natural polycrystals. Tec-
graphy with laboratory experiments for geologists. In: tonics 9, 379-389.
Fractography: fracture topography as a tool in fracture Law, R.D. 1986. Relationship between strain and quartz crys-
mechanics and stress analysis (edited by Ameen, M. S.) tallographic fabrics in the Roche Marie quartzites of Plou-
Spec. Publs geol. Soc. Lond. 92, 59-82. gastel, Western Brittany. J. Struct. Geol. 8, 493-515.
Kumpel, H-J. 1991. Poroelasticity: parameters reviewed. Law, R. D. 1987. Heterogeneous deformation and quartz
Geophys. J. Int. 105, 783-799. crystallographic fabric transitions: natural examples from
REFERENCES 115

the Moine thrust zone at the Stack of Glencoul, northern slikenside striae and associated steps in S-C mylonite. J.
Assynt. J. Struct. Geol. 9, 819-834. Struct. Geol. 14, 315-321.
Law, R. D. 1990. Crystallographic fabrics: a selective review Lindquist, P. A., Lai, H. H. & Alm, O.1984. Indentation Frac-
of their applications to research in structural geology. In: ture Development in rock continuously observed with
Deformation Mechanisms, Rheology and Tectonics (ed- a Scanning Electron Microscope. Int. J. Rock Mech. &
ited by Knipe, R. J. & Rutter, E. H.). Spec. Publs geol. Mining Sci. & Geomech. Abstr. 21, 165-184.
Soc. Lond. 54, 335-352. Linker, H. F. & Kirby, S. H. 1981. Anisotropy in the rheology
Law, R. D., Knipe, R. J. & Dayan, H. 1984. Strain path parti- of hydrolytically weakened synthetic quartz crystals. In:
tioning within thrust sheets: microstructural and petrofab- Mechanical Behaviour of Crustal Rock (edited by Carter,
ric evidence from the Moine thrust zone at Loch Eriboll, N. L., Friedman, M., Logan, J. M. & Stearns, D. W.). Am.
northwest Scotland. J. Struct. Geol. 6, 477-497. Geophys. Un. Geophys. Monogr. 24, 29-48.
Law, R. D., Morgan, S. S., Casey, M., Sylvester, A. G. & Ny- Lister, G. S. & Snoke, A. W. 1984. S-C Mylonites. J. Struct.
man, M. 1992. The Papoose Flat Pluton of eastern Cali- Geol. 6, 617-638.
fornia: a reassessment of its emplacement history in the Little, T. A., Miller, E. L., Lee, J. & Law, R. D. 1994. Exten-
light of new microstructural and Crystallographic fabric sional origin of ductile fabrics in the Schist Belt, Central
observations. Trans. Roy. Soc. Edinb. 83, 361-375. Brooks Range, Alaska-1. Geological and structural stud-
Law, R. D., Miller, E. L., Little, T. A. & Lee, J. 1994. Exten- ies. J. Struct. Geol. 16, 899-918.
sional origin of ductile fabrics in the Schist Belt, Central Llyod, G. E. & Freeman, B. 199la. SEM electron channelling
Brooks Range, Alaska-11. Microstructural and petrofab- analysis of dynamic recrystallisation in a quartz grain. J.
ric evidence. J. Struct. Geol. 16, 919-940. Struct. Geol. 13, 945-953.
Lawn, B. R. & Wilshaw, T. R. 1975a. Fracture of Brittle Lloyd, G. E. & Freeman, B. 1991b. Dynamic recrystallisa-
Solids. Cambridge University Press, Cambridge. tion of a quartz porphyroclast. Proceedings International
Lawn, B. R. & Wilshaw, T. R. 1975b. Review Indentation Conference on Textures of Materials 9, Avignon, France
Fracture: principles and applications. J. Mater. Sci. 10, (1990). Textures and Microstruct. 14-18, 751-756.
1049-1081. Llyod, G. E. & Freeman, B. 1994. Dyanamic recrystallization
Lehner, F. K. 1990. Thermodynamics of rock deformation by of quartz under greenschist conditions. J. Struct. Geol. 16,
pressure solution. In: Deformation Porcesses in Minerals, 867-882.
Ceramics and rocks (edited by Barber, D. J. & Meredith, Logan, J. M.,Higgs, N. G. & Friedman, M. 1981. Laboratory
P. G.) Unwin Hyman, London, 296-333. Studies on Natural Gouge from U.S. Geological Survey
Lejeune, A-M. & Richet, P. 1995. Rheology of crystal- Dry Lake Valley No.l Well, San Andreas Fault Zone. In:
bearing silicate melts: An experimental study at high vis- Mechanical Behaviour of Crustal Rock (edited by Carter,
cosities. J. geophys. Res. 100, 4215-4229. N. L., Friedman, M., Logan, J. M. & Stearns, D. W.). Am.
Lemèe, C. & Guegen, Y. 1996. Modeling of porosity loss dur- Geophys. Un. Geophys. Monogr. 24, 121-134.
ing compaction and cementation of sandstones. Geology, Lyons, J. B., Officer, C. B., Borella, P. E. & Lahodynsly,
24, 875-878. R. 1993. Planar lamellar substructures in quartz. Earth
Leroux, H. & Doukham, J-C. 1996. A transmission electron Planet. Sci. Lett., 119, 431-440.
microscope study of shocked quartz from the Manson im- Martin, R. J. & Durham, W. B. 1975. Mechanisms of Crack
pact structure. Geol. Soc. Am. Spec. Pap. 302, 267-274. Growth in Quartz. J. geophys. Res. 8, 4837-4843.
Leroux, H., Warne, J. E. & Doukham, J-C. 1995. Shocked McIntyre, D. B. 1962. Impact metamorphism at Clearwater
quartz in the Alamo breccia, southern Nevada: Evidence lake, Quebec. (Abstr). J. geophys Res. 67, 1647.
for a Devonian impact event. Geology, 23, 1003-1006. MacKinnon, P., Fueten, F. & Robin, P-Y. F. 1997. A fracture
Leroux, H., Reimold, W. U. & Doukhan, J-C. 1994. A TEM model for quartz ribbons in straight gneisses. J. Struct.
investigation of shock metamorphism in quartz from the Geol. 19, 1-14.
Vredefort dome, South Africa. Tectonophysics 230, 223- Maddock, R. H. 1992. Effects of lithology, cataclasis and
239. melting on the composition of fault-generated pseudot-
Leroux, H., Reimold, W. U., Koeberl, C. & Hornemann, achylytes in Lewisan gneiss, Scotland. Tectonophysics
U. 1998. Experimental shock metamorphism of zircon. 204, 261-268.
Lunar Planet. Sci. XXIX, in press. Maddock, R. H. 1986. Partial melting of lithic porphyroclasts
Lespinasse, M. & Cathelineau, M. 1995. Paleostress mag- in fault generated pseudotachylites. Neues Jahrb. Miner.
nitude determination by using fault slip and fluid inclu- Abh. 155, 1-14.
sions planes data. J. geophys. Res. 100, 3895-3904. Maddock, R. H., Grocott, J. & Van Nes, M. 1987. Ves-
Lespinasse, M. & Pecher, A. 1986. Microfracturing and re- icles, amygdales and similar structures in fault generated
gional stress field: a study of the preferred orientations of pseudotachylites. Lithos 20, 419-432.
fluid-inclusion planes in a granite from the Massif Cent- Magloughlin, J. F. 1989. The nature and significance of
ral, France. J. Struct. Geol. 8, 169-180. pseudotachylite from the Nason terrane, North Cascade
Lewis, S. & Holness, M. 1996. Equilibrium di- Mountains, Washington. J. Struct. Geol. 11, 907-917.
hedral angles: high rock-salt permeability in the shallow Mainprice, D., Bouchez, J-L., Blumenfeld, P. & Tubia, J. M.
crust?, Geology 24, 431-434. 1986. Dominant c slip in naturally deformed quartz: Im-
Lin, S. & Williams, P. F. 1992. The origin of ridge-in-groove plications for dramatic plastic softening at high temperat-
116 REFERENCES

ure. Geology 14, 819-822. McLellan, E. 1984. Deformational behaviour of migmatites


Mainprice, D. & Nicolas, A. 1989. Development of shape and and problems of structural analysis in migmatite terrains.
lattice preferred orientations: application to the seismic Geol. Mag. 121, 339-345.
anisotropy of the lower crust. J. Struct. Geol. 11, 175-190. Means, W. D. 1981. The concept of a steady state foliation.
Maltman, A. 1994. Prelithification deformation. In: Contin- Tectonophysics 78, 179-199.
ental deformation (edited by Hancock, P. L.) Pergammon Means, W. D. 1987. A newly recognised type of slickenside
Press, Oxford, 143-158. striation. J. Struct. Geol. 9, 585-590.
Mancktelow, N. S. 1994. On volume change and mass trans- Means, W. D. & Dhong, H. G. 1982. Some unexpected effects
port during the development of crenulation cleavage. J. of recrystallization on the microstrucures of materials de-
Struct. Geol. 16, 1217-1232. formed at high temperatures. Mitt. Geol. Inst. ETH., 239,
March, A. 1932. Mathematische theorie der regelung nach 205-207.
der korn gestalt bei affiner deformation. Z. Kristallogor Means, W. D., Hobbs, B. E., Lister, G. S. & Williams, P. F.
81, 285-297. 1980. Vorticity and non-coaxiality in progressive deform-
Marlow, P. C. & Etheridge, M. A. 1977. Development of a ations. J. Struct. Geol. 2, 371-378.
layered crenulation cleavage in mica schists of the Kan- Meissner, R. & Strehlau, J. 1982. Limits of stress in contin-
matoo Group near Macclesfield, South Australia. Bull, ental crust and their relation to the depth-frequency rela-
geol. Soc. Am. 88, 873-882. tion of shallow earthquakes. Tectonics 1, 73-89.
Marone, C. & Scholz, C. H. 1989. Particle-size distibu- Menéndez, B., Zhu, W. & Wong, T-F. 1996. Micromechanics
tion and microstructures within simulated fault gouge. J. of brittle faulting and cataclastic flow in Berea sandstone.
Struct. Geol. 11, 799-814. J. Struct. Geol. 18, 1-16.
Marone, C., Hobbs, B. E. & Ord, A. 1992. Coulomb con- Mercier, J-C. C. 1980. Magnitude of Continental Litho-
stitutive laws for friction: contrasts in frictional behaviour spheric Stresses inferred from Rheomorphic Petrology. J.
for distributed and localized shear. Pure & Appl. Geo- geophys. Res. 85, 6293-6303.
phys. 139,195-214. Mercier, J-C. C. 1985. Olivine and pyroxenes. In: Preferred
Marsh, B. D. 1981. On the crystallinity, probability of oc- Orientation in deformed metals and rocks: an introduc-
currence and rheology of lava and magma. Contr. Miner. tion to modern texture analysis (edited by H. R. Wenk).
Petrol. 78, 85-94. Academic Press, Orlando, 407-430.
Martini, J. E. J. 1991. The nature, distribution and genesis of Mercier, J-C. C., Anderson, D. A. & Carter, N. L. 1977. Stress
coesite and stishovite associated with the pseudotachylite in lithosphere: Inferences from steady state flow of rocks.
of the Vredefort dome, South Africa. Earth Planet. Sci. Pure & Appl. Geophys 115, 199-226.
Lett. 103, 285-300. Meredith, P. L. 1983. A fracture mechanics study of ex-
Mase, C. W. & Smith, L. 1985. Pore fluid pressure and fric- perimentally deformed crustal rocks. Unpublished Ph.D.
tional heating on a fault surface. Pure & Appl. Geophys. Thesis, University of London.
122, 583-607. Michalske, T. A. & Frechette, V. D. 1980. Dynamic effects of
Masch, L., Wenk, H. R. & Preuss, E. 1985. Electron micro- liquids on crack growth leading to catastrophic failure. J.
scopy study of hyalomylonites - Evidence for frictional Am. Ceram. Soc. 63, 603-609.
melting in landslides. Tectonophysics 115, 131-160. Miller, C. F., Watson, E. B. & Harrison, T. H. 1988. Per-
Masuda, T. & Mizuno, N. 1995. Deflection of pure shear vis- spectives on source, segregation and transport of granitoid
cous flow around a rigid spherical body. J. Struct. Geol. magmas. Trans. R. Soc. Edinb. 79. 135-156.
17, 1615-1620. Miller, R. B. & Paterson, S. R. 1994. The transition from
McBirney, A. R. & Murase, T. 1984. Rheological properties magmatic to high-temperature solid-state deformation:
of magmas. Ann. Rev. Earth and Planet. Science, 12, 337- implications from the Mount Stuart batholith, Washing-
357. ton. J. Struct. Geol. 16, 853-866.
McCaig, M. A. 1987. Deformation and fluid-rock interaction Mitra, S. 1976. A quantatitive study of deformation mechan-
in metasomatic dilatant shear bands. Tectonophysics 135, isms and finite strain in quartzites. Contr. Miner. Petrol.
121-132. 59, 203-226.
McClintock, F. A. & Walsh, J. B. 1962. Friction on Griffith Mogi, K. 1965. Deformation and fracture of rocks under con-
cracks in rocks under pressure. In: Proc. 4th U.S. Nat. fining pressure (2), Elasticity and plasticity of some rocks.
Congr. Appl. Mech., Vol. II. New York: Am. Soc. Mech. Bull. Earthquake Res. Inst. Tokyo Univ. 43, 349-374.
Eng. 1015-1021. Molnar, P. 1992. Brace-Goetze strength profiles, the partition-
McEwen, T. J. 1981. Brittle deformation in pitted pebble con- ing of strike-slip and thrust faulting at zones of oblique
glomerates. J. Struct. Geol. 3, 25-37. convergence, and the stress-heat flow paradox of the San
McLaren, A. C. 1991. Transmission electron microscopy of Andreas fault. In: Fault Mechanics and Transport Prop-
minerals and rocks. Cambridge University press, New erties of Rocks: A Festschrift in honour of W. F. Brace
York, 387pp. (edited by Evans, B. & Wong, T. F.). Academic Press,
McLaren, A. C., Retchford, A. J., Griggs, D. T. & Christie, London, 124-142.
J. M., 1967. Transmission electron microscope study of Morrit, R. F. C., McA.Powell, C. & Vernon, R. M. 1982. Bed-
Brazil twins and dislocations experimentally produced in ding plane foliation I and II. In: Atlas of deformational
natural quartz. Phys. Status Solidi. 19, 631-644. and metamorphic rock fabrics (edited by Borradaille, G.
REFERENCES 117

J., Bayly, M. B. & McA.Powell, C.). Springer Verlag, Ord, A. 1991. Deformation of rock: A pressure-senstive,
Berlin, 46-49. dilatant material. Pure & Appl. Geophys. 137, 337-366.
Muller, W. H. & Briegel, U. 1978. The rheological behaviour Ord, A. & Christie, J. M. 1984. Flow stresses from micro-
of polycrystalline anhydrite. Eclogae Geol. Helv. 71, 397- structures in mylonitic quartzites of the Moine Thrust
407. zone, Assynt area, Scotland. J. Struct. Geol. 6, 639-654.
Murphy, F. X. 1990. The role of pressure solution and Ord, A. & Hobbs, B. E. 1985. Experimental control of the
intermicrolithon-slip in the development of disjunctive water weakening effect in quartz. In: Mineral and Rock
cleavage domains: a study from Helvick Head in the Irish Deformation: Laboratory Studies-The Paterson Volume.
Variscides. J. Struct. Geol. 12, 69-82. (edited by Hobbs, B. E. & Heard, H. C.). Am. Geophys.
Murrell, S. A. F. 1963. A criterion for brittle fracture of rocks Un. Geophys. Monogr. 36, 51-72.
and concrete under triaxial stress, and the effect of pore Ord, A. & Hobbs, B. E. 1989. The strength of the continental
pressure on the criterion. In: Rock Mechanics. (edited by crust, detachment zones and the development of plastic
Fairhurst, C.). Proc. 5th Symp. Rock Mech. New York: instabilities. Tectonophysics 158, 269-289.
Pergamon Press, 563-577. Park, Y. & Means, W. D. 1996. Direct observation of deform-
Murrell, S. A. F. 1965. The effect of triaxial stress on the ation processes in crystal mushes. J. Struct. Geol. 18, 847-
strength of rocks at atmospheric temperatures. Geophys. 858.
J. 10, 231-281. Passchier, C. W. 1983. The reliability of asymmetric c-axis
Murrel, S. A. F. & Digby, P. J. 1970. The theory of brittle frac- fabrics of quartz to determine sense of vorticity. Tectono-
ture initiation under triaxial stress conditions I. Geophys. physics 99, T9-T18.
J. R. astr. Soc. 19, 309-334. II, 499-512. Passchier, C. W. 1984. The generation of ductile and brittle
Narahara, D. K. & Wiltschko, D. V. 1986. Deformation in shear bands in a low angle mylonite zone. J. Struct. Geol.
the hinge region of a chevron fold, Valley and Ridge 6, 273-282.
Province, central Pennsylvania. J. Struct. Geol. 8. 157- Passchier, C. W. 1991. Geometric constraints on the develop-
168. ment of shear bands in rocks. Geologie en Mijnbouw 70,
Nemat-Nasser, S. & Horii, H. 1982. Compression-induced 203-211.
nonplanar crack extension with application to splitting, Passchier, C. W. 1994. Mixing in flow perturbations: a model
exfoliation and rockburst. J. geophys. Res. 87, 6805- for development of mantled porphyroclasts in mylonites.
6821. J. Struct. Geol. 8, 831-844.
Newman, J. 1994. The influence of grain size distribution on Passchier, C. W. & Simpson, C. 1986. Porphyroclast systems
methods for estimating paleostress from twinning in car- as kinematic indicators. J. Struct. Geol. 8, 831-844.
bonates. J. Struct. Geol. 16, 1589-1602. Passchier, C. W. & Sokoutis, D. 1993. Experimental model-
Nicolas, A. & Poirier, J. P. 1976. Crystalline Plasticity and ling of mantled porphyroclasts. J. Struct. Geol. 15, 895-
Solid State Flow in Metamorphic Rocks. London: Wiley- 910.
Interscience. Passchier, C. W. & Trouw, R. A. J. 1996. Microtectonics.
Nicolas, A., Reuber, I. & Benn, K. 1988. A new magma Springer-Verlag, Berlin Heidelberg New York.
chamber model based on structural studies in the Oman Passchier, C. W., Hoek, J. D., Bekendam, R. F. & De Boorder,
ophiolite. Tectonophysics 151, 87-105. H. 1990. Ductile reactivation of Proterozoic brittle fault
Norris, D. K. & Barron, K. 1968. Structural analysis of fea- rocks; an example from the Vestfold Hills, East Antarc-
tures on natural and artificial faults. Proc. Conf. on Re- tica. Precamb. Res. 47, 3-16.
search in Tectonics, Ottawa, 136-174. Passchier, C. W., Trouw, R. A. J., Zwart, H. J. & Vissers, R.
Norton, M. G. & Atkinson, B. K. 1981. Stress dependent L. M. 1992. Porphyroblast rotation: eppur si muove?, J.
morphological features on fracture surfaces of quartz and metam. Geol. 10, 283-294.
glass. Tectonophysics 77, 283-295. Passchier, C. W., Bons, P. D. & Jessell, M. W. 1993. The
Officer, C. B. & Carter, N. L. 1991. A review of the struc- analysis of progressive deformation in rock analogues. J.
ture, petrology and dynamic deformation characteristics Struct. Geol. 15, 403-412.
of some enigmatic terrestrial structures. Earth Science Paterson, M. S. 1978. Experimental Rock Deformation - The
Reviews 30, 1-49. Brittle Field. Springer Verlag, Berlin.
Olgaard, D. L. & Brace, W. F. 1983. The microstructure of Paterson, M. S. 1987. Problems in the extrapolation of labor-
gouge from a mining induced seisimic shear zone. Int. J. atory rheological data. Tectonophysics 133, 33-43.
Rock Mech. & Mining Sci. & Geochem. Abstr. 20, 11-19. Paterson, M. S. 1995. A theory for granular flow accommod-
Olsson, R. K., Miller, K. G., Browning, J. V., Hibb, D. & Sug- ated by material transfer via an intergranular fluid. Tec-
arman, P. J. 1997. Ejecta layer at the Cretaceous-Tertiary tonophysics 245, 135-151.
boundary, Bass River, New Jersey (Ocean Drilling Leg Paterson, M. S. & Luan, F. C. 1990. Quartzite rheology un-
174 AX). Geology 25, 759-762. der geological conditions. In: Deformation Mechanisms,
Olsson, W. A. & Peng, S. S. 1976. Microcrack nucleation in Rheology and Tectonics (edited by Knipe, R. J. & Rutter,
marble. Int. J. Rock Mech. & Mining Sci. 13, 53 - 59. E. H.). Spec. Publs geol. Soc. Lond. 54, 299-307.
Onasch, C. M. 1994. Assessing brittle volume-gain and pres- Paterson, S. R. & Vernon, R. H. 1995. Bursting the bubble of
sure solution volume-loss processes in quartz arenite. J. ballooning plutons: A return to nested diapirs emplaced
Struct. Geol. 16, 519-530. by multiple processes. Bull. geol. Soc. Am. 107, 1356-
118 REFERENCES

1380. ter transport through a liquid phase. J. geophys. Res. 87,


Paterson, S. R., Vernon, R. H. & Tobisch, O. T. 1989. A re- 4731-4739.
view of criteria for the identification of magmatic and tec- Raj, R. & Ashby, M. F. 1971. On grain boundary sliding and
tonic foliations in granitoids. J. Struct. Geol. 11, 349-364. diffusional creep. Metall. Trans. A2, 1113-1127.
Pavlis, T. L. & Bruhn, R. L. 1988. Stress history during Raj, R. & Chyung, C. K. 1981. Solution precipitation creep
propagation of a lateral fold-tip and implications for the in glass ceramics. Acta Metall. 29, 159-166.
mechanics of fold-thrust belts. Tectonophysics 145, 113- Ramsay, J. G. 1962. The geometry and mechanics of forma-
127. tion of "similar" type folds. J. Geol. 70, 309-327.
Peck, L. 1983. Stress corrosion and crack propagation in Ramsay, J. G. 1967. Folding and Fracturing of Rocks. Mc-
Sioux Quartzite. J. geophys. Res. 88, 5037-5046. Graw Hill, New York.
Peng, S. & Johnson, A. M. 1972. Crack growth and faulting in Ramsay, J. G. 1980. Shear zone geometry: a review. J. Struct.
a cylindrical sample of Chelmsford Granite. Int. J. Rock Geol. 2, 83-89.
Mech. & Mining Sci. 9, 37-86. Ramsay, J. G. 1980. The crack-seal mechanism of rock de-
Petit, J. P. 1987. Criteria for the sense of movement on fault formation. Nature 284, 135-139.
surfaces in brittle rocks. J. Struct. Geol. 9, 597-608. Ramsay, J. G. 1989. Emplacement kinematics of a granite
Petit, J. P. & Matthauer, M. 1995. Paleostress superimposition diapir: the Chinamora Batholith, Zimbabwe. J. Struct.
deduced from mesoscale structures in limestone: the Ma- Geol. 11, 191-210.
telles exposure, Languedoc, France. J. Struct. Geol. 17, Ramsay, J. G. & Graham, R. H. 1970. Strain variation in shear
245-256. belts. Can. J. Earth Sci. 7, 786-813.
Pfiffner, O. A. & Burkhard, M. 1987. Determination of paleo- Ramsay, J. G. & Wood, D. S. 1973. The geometric effects of
stress axes orientations from fault, twin and earthquake volume change during deformation processes. Tectono-
data. Annales Tectonicae 1, 48-57. physics 16, 263-277.
Platt, J. P. 1984. Secondary cleavages in ductile shear zones. Ramsay, J. G. & Huber, M.I.1983. The techniques of modern
J. Struct. Geol. 6, 439-442. structural geology, 1: Strain analysis. Academic Press,
Platt, J. P. & Vissers, R. L. M. 1980. Extensional structures in London.
anisotropic rocks. J. Struct. Geol. 2, 397-410. Ramsay, J. G. & Huber, M. I. 1987. The techniques of mod-
Pittman, E. D. 1981. Effect of fault-related granulation on ern structural geology, 2: Folds and fractures. Academic
porosity and permeability of quartz sandstones, Simpson Press, London.
Group (Ordovician), Oklahoma Bull. Am. Ass. Petrol. Ramsay, J. G., Casey, M. & Kligfield, R. 1983. Role of shear
Geol. 65, 2381-2387. in development of the Helvetic fold and thrust belt of
Pons, J., Barbey, P., Dupuis, D. & Léger, J. M. 1995. Mech- Switzerland. Geology 11, 439-442.
anisms of pluton emplacement and structural evolution of Raterron, P. & Jaoul, O. 1991. High-temperature deformation
a 2.1Ga juvenile continental crust: the Birrimian of south- of diopside single crystal. 1, Mechanical data. J. geophys.
western Niger. Precamb. Res. 70, 281-301. Res. B-96, 14277-14286.
Post, A. & Tullis, J. 1999. A recrystallised grain size piezo- Renner, J., Hirth, G. & Evans, B. 1999. The effect of melt
meter for experimentally deformed feldspar aggregates. pressure on the deformation behaviour of partially mol-
Tectonophysics 303, 159-173. ten rocks. Abstract, Deformation Mechanisms, Rheology
Powell, C. M. A. 1979. A morphological classification of rock and Microstructures, Neustadt an der Weinstrasse, 22-
cleavage. Tectonophysics 58, 21-34. 26th March 1999.
Powell, D. & Treagus, J. E. 1970. Rotational fabrics in meta- Reimold, W. U. 1994. Comment on planar lamellar substruc-
morphic minerals. Mineral Mag. 37, 801-814. tures in quartz by Lyons, J. B., Officer, C. B., Borella, P.
Power, W. L. & Tullis, T. E. 1989. The relationship between E. & Lahodynsky, R. Earth Planet. Sci. Lett. 125, 473-
slickenside surfaces in fine grained quartz and the seismic 471.
cycle. J. Struct. Geol. 11, 879-894. Reimold, W. U. 1995. Impact cratering - A review, with
Price, N. J. & Cosgrove, J. W. 1990. Analysis of Geological special reference to the economic importance of impact
Structures. Cambridge University Press. structures and the southern African impact crater record.
Prior, D. J. 1987. Syntectonic porphyroblast growth in phyl- Earth, Moon and Planets 70, 21-45.
lites: textures and processes. J. metam. Geol. 5, 27-39. Reimold, W. U. 1995. Pseudotachylite in impact structures
Quackenbush, C. L. & Frechette, V. D. 1978. Crack front - generation by friction melting and shock brecciation?:
curvature and Glass Slow Fracture. J. Am. Ceram. Soc. A review and discussion. Earth-Science reviews 39, 247-
61, 402-406. 265.
Quick, J. E., Sinigoi, S., Negrini, L., Demarchi, G. & Mayer, Reimold, W. U. & Stöffler, D. 1978. Experimental shock
A. 1992. Synmagmatic deformation in the underplated ig- metamorphism of dunite. Proc. Lunar Planet. Sci. Conf.
neous complex of the Ivrea-Verbano zone. Geology 20, 9th, 2805-2824.
613-616. Reimold, W. U. & Hörz, F. 1986. Textures of experimentally-
Railsback, L. B. & Andrews, L. M. 1995. Tectonic stylolites shocked (5.1-35.5 GPa) Witwatersrand quartzite. Ab-
in the ‘undeformed’ Cumberland Plateau of southern Ten- stracts, Lunar and Planetary Science conference 17, 703-
nessee. J. Struct. Geol. 17, 911-922. 704.
Raj, R. 1982. Creep in polycrystalline aggregates by mat- Rice, J. R. 1968. Mathematical Analysis in the mechanics of
REFERENCES 119

fracture. In: Fracture.(edited by H. Leibowitz). Academic transitions in rocks. Tectonophysics 122, 381-387.
Press, New York. Rutter, E. H. & Neumann, D. H. K. 1995. Experimental
Rice, J. R. 1992. Fault stress states, pore pressure distribu- deformation of partially molten Westerly granite under
tions, and the weakness of the San Andreas fault. In: Fault fluid-absent conditions, with implications for the extrac-
Mechanics and Transport Properties of Rocks: A Fests- tion of granitic magmas. J. geophys. Res. 100, 15697-
chrift in honour of W. F. Brace (edited by Evans, B. & 15715.
Wong, T-F.). Academic Press, London, 475-504. Rutter, E. H., Maddock, R. H., Hall, S. H. & White, S. H.
Rice, J. R. & Gu, J. C. 1983. Earthquake after effects and 1986. Comparative Microstructures of Natural and Exper-
triggered seisimic phenomena. Pure & Appl. Geophys. imentally Produced Clay-Bearing Fault Gouges. Pure &
121,187-219. Appl. Geophys. 124, 1-30.
Riley, G. N. Jr. 1990. Liquid distribution and transport in silic- Rutter, E. H. & Hadizadeh, J. 1991. On the influence of
ate liquid-olivine materials. Ph. D. Thesis, Cornell Univ. porosity on the low-temperature brittle-ductile transition
New York. in silicate rocks. J. Struct. Geol. 13, 609-614.
Robert, F. 1989. The internal structure of the Cadillac tectonic Rutter, E. H., Casey, M. & Burlini, L. 1994. Preferred crys-
zone southeast of Val d’Or, Abitibi belt, Quebec. Can. J. tallographic orientation development during the plastic
Earth Sci. 26, 2661-2675. and superplastic flow of calcite rocks. J. Struct. Geol. 16,
Robert, F, Bouillier, A-M. & Firdaous, K. 1995. Gold-quartz 1431-1446.
veins in metamorphic terranes and their bearing on the Ryan, B. 1995. Morphological features of multigeneration
role of fluids in faulting. J. geophys. Res. 100, 12861- basic dykes near Nain, Labrador: clues to orginal em-
12879. placement mechanisms and subsequent deformation. Pre-
Robertson, P. B. 1975. Zones of shock metamorphism at camb. Res. 75, 91-118.
the Charlevoix impact structure, Quebec. Bull. geol. Soc. Rykkelid, E. & Fossen, H. 1992, Composite fabrics in mid-
Am. 86, 1630-1638. crustal gneisses: observations from the Oygarden Com-
Robertson, P. B. & Grieve, R. A. F. 1977. Shock attentuation plex, West Norway Caledonides. J. Struct. Geol. 14, 1-10.
of terrestrial impact structures. In: Impact and explosion Sammis, C. G. & Biegel, R. L. 1989. Fractals, fault-gouge
cratering (edited by Roddy, D. J., Pepin, R. O. & Merrill, and friction. Pure & Appl. Geophys. 131, 254-271.
R. B.). Pergammon Press, New York. Sammis, C. G., King, G. & Biegel, R. L. 1987. The kinemat-
Robertson, E. C. 1982. Continuous formation of gouge and ics of gouge deformation. Pure & Appl. Geophys. 125,
breccia during fault displacement. In: Proc. 23rd Symp. 77-812.
Rock Mech. (edited by Goodman, R. E. & Heuze, F. E.). Scheuber, E. & Andriessen, P. A. M. 1990. The kinematic
University California, Berkeley, 397-404. and geodynamic significance of the Atacama fault zone,
Robertson, E. C. 1983. Relationship of fault displacement to northern Chile. J. Struct. Geol. 12, 243-258.
gouge and breccia thickness. Min. Engng. 35, 1426-1432. Schmid, S. M., Boland, J. N. & Paterson, M. S. 1977. Super-
Robin, P. Y. F. & Cruden, A. R. 1994. Strain and vorticity plastic flow in fine grained limestone. Tectonophysics 43,
patterns in ideally ductile transpression zones. J. Struct. 263-286.
Geol. 16, 447-466. Schmid, S. M., Panozzo, R. & Bauer, S. 1987. Simple shear
Roedder, E. 1984. Fluid inclusions. Rev. Mineral Soc. Am. experiments on calcite rocks: rheology and microfabrics.
12. J. Struct. Geol. 9, 747-778.
Roscoe, R. 1952. The viscosity of suspensions of rigid Schmid, S. M., Paterson, M. S. & Boland, J. N. 1980. High
spheres. Br. J. Appl. Phys. 3, 267-269. temperature flow and dynamic recrystallization in Carrara
Ross, J. V., Avé Lallemant, H. G. & Carter, N. L. 1980. Stress Marble. Tectonophysics 65, 6174-6184.
dependence of recrystallized-grain and subgrain size in Schulmann, K., B. & Melka, R. 1996. High temper-
olivine. Tectonophysics 70, 39-61. ature microstructures and rheology of deformed granite,
Rowe, K. J. & Rutter, E. H. 1990. Paleostress estimation us- Erzebirge, Bohemian Massif. J. Stuct. Geol. 18, 717-734.
ing calcite twinning: experimental calibration and applic- Schofield, A. N. & Wroth, C. P. 1968. Critical state soil mech-
ation to nature. J. Struct. Geol. 12, 1-7. anics. McGraw-Hill, London.
Rudnicki, J. W. 1980. Fracture Mechanics Applied to the Scholz, C. H. 1972. Static Fatigue of Quartz. J. geophys. Res.
Earth’s Crust. Ann. Rev. Earth Planet. Sci. 8, 489-525. 77, 2104-2114.
Rushmer, T. 1995. An experimental deformation study of par- Scholz, C. H. 1990. The mechanics of earthquakes and fault-
tially molten amphibolite: application to low melt fraction ing. Camb. Univ. Press, New York.
segregation. J. geophys. Res. 100, 15681-15695. Schoneveld, C. 1977. A study of some typical inclusions pat-
Rudnicki, J. W. & Rice, J. R. 1975. Conditions for the local- terns in strongly paracystalline rotated garnets. Tectono-
ization of deformation in pressure-sensitive dilatant ma- physics 39, 453-471.
terials. J. Mech. Phys. Solids 23, 371-394s Schutjens, P. M. T. M. 1991. Experimental compaction of
Rutter, E. H. 1976. The kinetics of rock deformation by pres- quartz sand at low effective stress and temperature con-
sure Solution. Phil. Trans. R. Soc. A283, 203-219. ditions. J. geol. Soc. Lond. 148, 527-539.
Rutter, E. H. 1983. Pressure solution in nature, theory and Seyedolali, A., Krinsley, D. H., Boggs, S. Jr., O’Hara, P.
experiment. J. geol. Soc. Lond. 140, 725-740. F., Dypvik, H. & Goles, G. G. 1997. Provenance in-
Rutter, E. H. 1986. On the nomenclature of mode of failure terpretation of quartz by scanning electron microscope-
120 REFERENCES

cathodoluminescence fabric analysis. Geology, 25, 787- escence as a petrologic tool. J. Geol. 73, 627-635.
790. Snoke. A. W., Tullis, J. & Todd, V. R. 1998. Fault-related
Shand, S. J. 1916. The pseudotachylite of Parijs (Orange Free rocks - A photographic Atlas. Princeton University Press,
State), and its relation to "trop-shotten gneiss" and "flinty Princeton, New Jersey.
crush-rock". Q. J. geol. Soc. London, 72, 198-221. Spang, J. H. & Van der Lee, J. 1975. Numerical dynamic ana-
Sharpton, V. L. & Schuraytz, B. C. 1989. On reported oc- lysis of quartz deformation lamellae and calcite and dolo-
currences of shock deformed clasts in the volcanic ejecta mite twin lamellae. Bull. geol. Soc. Am. 86, 1266-1272.
from Toba caldera, Sumatra. Geology 7, 1040-1043. Spiers, C. J. & Schutjens, P. M. T. M. 1990. Densification of
Sherman, P.I. 1968. Emulsion science. New York, Academic. crystalline aggregates by fluid phase diffusional creep. In:
Shelley, D. 1989a. Plagioclase and quartz preferred orienta- Deformation processes in Minerals, Ceramics and Rocks
tions in a low grade schist: the roles of primary growth (edited by Barber, D. J. & Meredith, P. G.). Unwin Hy-
and plastic deformation. J. Struct. Geol. 11, 1029-1038. man, 334-353.
Shelley, D. 1989b. P, M, and G tectonics: a classification Spiers, C. J., Schutjens, P. M. T. M., Brzesowsky, R. H.,
based on origin of mineral preferred orientation. J. Struct Peach, C. J., Liezenburg, J. L. & Zwart, H. J. 1990. Ex-
Geol. 11, 1039-1044. perimental determination of constitutive parameters gov-
Shelley, D. 1995. Asymmetric shape preferred orientations as erning creep of rock salt by pressure solution. In: De-
shear-sense indicators. J. Struct. Geol. 17, 509-518. formation Mechanisms, Rheology and Tectonics (edited
Shelton, G. L., Tullis, J. & Tullis, T. E. 1981. Experimental by Knipe, R. J. & Rutter, E. H.). Spec. Publs geol. Soc.
high temperature and pressure faults. Geophys. Res. Lett. Lond. 54, 215-237.
8, 55-58. Spray, J. G. 1987. Artificial generation of pseudotachylite us-
Shimada, M. 1986. Mechanism of deformation in a dry por- ing friction welding apparatus: simulation of melting on
ous basalt at high pressure. Tectonophysics 121, 153-173. a fault plane. J. Struct. Geol. 9, 49-60.
Shimamoto, T. 1989. The origin of S-C mylonites and a new Spray, J. G. 1988. Crystallization of an amphibolite shear
fault-zone model. J. Struct. Geol. 11, 51-64. melt: an investigation using radial friction welding ap-
Shimamato, T. & Nagahama, H. 1992. An argument against paratus. Contr. Miner. Petrol. 99, 464-475.
the crush origin of pseudotachylites based on the analysis Spray, J. G. 1989. Slickenside formation by surface melting
of clast-size distribution. J. Struct. Geol. 14, 999-1006. during the mechanical excavation of rock. J. Struct. Geol.
Sibson, R. H. 1977. Fault Rocks and Fault Mechanisms. J. 11, 895-906.
geol. Soc. Lond. 133, 191-213. Spray, J. G. 1995. Pseudotachylite controversy: Fact or fric-
Sibson, R. H. 1980. Transient discontinuities in ductile shear tion? Geology 23, 1119-1122.
zones. J. Struct. Geol. 2, 165-171. Spray, J. G. & Thompson, L. M. 1994. Friction melt distribu-
Sibson, R. H. 1982. Fault zone models, heat flow and the tion in a multi-ring impact basin. Nature 373, 130-132.
depth distribution of earthquakes in the continental crust Sprunt, E. S. & Nur, A. 1979. Microcracking and healing in
of the United States. Bull. seism. Soc. Am. 72, 151-163. granites: New evidence from cathodoluminescence. Sci-
Sibson, R. H. 1983. Continental fault structure and the shal- ence 205, 495-497.
low earthquake source. J. geol. Soc. Lond. 140, 741-767. Sprunt, E. S., Dengler, L. A. & Sloan, D. 1978. Effects of
Sibson, R. H., Moore, J. McM. & Rankin, A. H. 1975. Seis- metamorphism on quartz cathodoluminescence. Geology
mic pumping - a hydrothermal fluid transport mechanism. 6, 305-308.
J. geol. Soc. Lond. 131, 653-659. Stearns, D. W. 1968. Fracture as a mechanism of flow in nat-
Simpson, C. 1985. Deformation of granitic rocks across the urally deformed layered rocks. Geol. Surv. Can. Paper 68-
brittle-ductile transition. J. Struct. Geol. 7, 503-511. 52, 68-80.
Simpson, C. 1986. Determination of Movement Sense in Stel, H. 1981. Crystal growth in cataclasites: diagnostic mi-
Mylonites. J. Geol. Education 34, 246-261. crostructures and implications. Tectonophysics 78, 585.
Simpson, C. & Schmid, S. 1983. An evaluation of criteria Stel, H. 1991. Linear dilatation structures and syn-magmatic
to deduce the sense of movement in sheared rocks. Bull. folding in granitoids. J. Struct. Geol. 12, 625-634.
geol. Soc. Am. 94,1281-1288. Stöffler, D. 1972. Deformation and transformation of rock-
Smit, J., Montanari, A., Swinburne, N. H. M., Alvarez, W., forming minerals by natural and experimental shock pro-
Hildebrand, A. R., Margolis, S. V., Claeys, P. Lowrie, W. cesses. I. The behaviour of minerals under shock com-
& Asaro, F. 1992. Tektite-bearing, deep-water clastic unit pression. Fortschr. Mineral. 49, 50-113.
at the Cretaceous-Tertiary boundary in northeastern Mex- Stöffler, D. 1984. Glasses formed by hypervelocity impact. J.
ico. Geology 20, 99-103. Non-crystalline Solids 67, 465-502.
Smith, C. S. 1964. Some elementary principles of polycrys- Stöffler, D. & Langenhorst, F. 1994. Shock metamorphism of
talline microstructure. Metall. Reviews 9, 10-48. quartz in nature and experiment: I Basic observation and
Smith, D. C. 1984. Coesite in clinopyroxene in the Cale- theory. Meteoritics 29, 155-181.
donides and its implications for geodynamics. Nature, Stoker, R. L. & Ashby, M. F. 1973. On the rheology of the
310, 641-644. upper mantle. Rev. Geophys. and Space Phys. II, 391-421.
Smith, D. L. & Evans, B. 1984. Diffusional crack healing in Swanson, P. L. 1981. Sub-critical crack propagation in West-
quartz. J. geophys. Res. 89, 4125-4135. erly Granite: An investigation into the Double Torsion
Smith, J. V. & Stenstrom, R. C. 1965. Electron-excited lumin- Method. Int. J. Rock Mech. & Mining Sci. & Geomech.
REFERENCES 121

Abstr. 18, 445-449. Wong, T-F.). Academic Press, London. 89-117.


Sylvester, A. G. 1988. Strike-slip faults. Bull. geol. Soc. Am. Tullis, J., Yund, R. & Farver, J. 1996. Deformation-enhanced
100, 1666-1703. fluid distribution in feldspar aggregates and implications
Tada, R. & Siever, R. 1986. Experimental knife-edge pressure for ductile shear zones. Geology 24, 63-66.
solution of halite. Geochim. Cosmochim. Acta 50, 29-36. Tullis, T. E., Horowitz, F. G. & Tullis, J. 1991. Flow laws
Tada, R. & Siever, R. 1989. Pressure solution during diagen- of polyphase aggregates from end member flow laws. J.
esis. An. Rev. Earth Planet. Science 17, 89-118. geophys. Res. 96, 8081-8096.
Tada, R., Maliva, R. & Siever, R. 1987. A new mechan- Turner, F. J. 1953. Nature and dynamic interpretation of de-
ism for pressure solution in porous quartzose sandstone. formation lamellae in calcite of three marbles. Am. J. Sci.
Geochim. Cosmochim. Acta 51, 2295-2301. 251, 276-298.
Takahashi, M., Nagahaam, H., Masuda, T. & Fujimora, A. Turner, F. J. & Weiss, L. E. 1963. Structural analysis of meta-
1997. Fractal grain boundaries of recrystallized quartz morphic tectonites. McGraw Hill, New York.
and strain rate meter. Digest, 3rd International Sym- Twiss, R. J. 1974. Structure and significance of planar de-
posium on Fractals and Dynamic Systems in Geoscience, formation features in synthetic quartz. Geology 2, 329-
Hotel Academia, Stara Lesna, Slovakia, June 18-29th, 87. 332.
Tapponier, P. & Brace, W. F. 1976. Development of stress in- Twiss, R. J. 1977. Theory and applicability of a recrystallized
duced microcracks in Westerly granite. Int. J. Rock Mech. grain size palaeopiezometer. Pure & Appl. Geophys. 115,
& Mining Sci. & Geomech. Abstr. 13, 103-112. 227-244.
Ten Brink, C. E. & Passchier, C. W. 1995. Modelling of Twiss, R. J. 1986. Variable sensitivity piezometric equations
mantled porphyroclasts using non-Newtonian rock ana- for dislocation density and sub-grain diameter and relev-
logue materials. J. Struct. Geol. 17, 131-146. ance to olivine and quartz. In: Mineral and rock deforma-
Terzaghi, K. Van. 1943. Theoretical Soil Mechanics. New tion: Laboratory Studies-The Paterson Volume (edited by
York, Wiley. Hobbs, B. E. & Heard, H. C.). Am. Geophys. Un. Geo-
Teufel, L. W. 1981. Pore volume changes during frictional phys. Monogr. 36, 161-169.
sliding of simulated faults. In: Mechanical Behaviour of Underhill, J. & Woodcock, N. 1985. Faulting mechanisms in
Crustal Rock (edited by Carter, N. L., Friedman, M., Lo- high porosity sandstones, Permo-Triassic, Arran, Scot-
gan, J. M. & Stearns, D. W.). Am. Geophys. Un. Geophys. land. Abstract, Deformation Mechanisms in sediments
Monogr. 24, 135-145. and sedimentary rocks, Tectonic Studies Group Meeting,
Tija, H. D. 1967. Sense of fault displacements. Geologie University of London, 19-20.
Mijnb. 46, 392-396. Urai, J. L. 1983. Water assisted dynamic recrystallisation and
Tikoff, B. & Teyssier, C. 1994. Strain and fabric analyses weakening in polycrystalline bischoffite. Tectonophysics
based on porphyroclast interaction. J. Struct. Geol. 16, 96, 125-157.
477-492. Urai, J. L., Spiers, C. J., Zwart, H. J. & Lister, G. S. 1986.
Tikoff, B. & Fossen, H. 1993. Simultaneous pure shear and Weakening of rock by water during long-term creep.
simple shear: the unifying deformation matrix. Tectono- Nature 324, 554-557.
physics 217, 267-283. Urai, J. L., Means, W. D. & Lister, G. S. 1986. Dynamic re-
Tikoff, B. & Greene, D. 1997. Stretching lineations in trans- crystallisation of minerals. In: Mineral and rock deforma-
pressional shear zones: an example from the Sierra tion: Laboratory Studies-The Paterson Volume (edited by
Nevada batholith, California. J. Struct. Geol. 19, 29-39. Hobbs, B. E. & Heard, H. C.). Am. Geophys. Un. Geo-
Tingle, T. N., Green, W. H., Scholz, C. H. & Koczynski, T. phys. Monogr. 36, 161-199.
A. 1993. The rheology of faults triggered by the olivine- Urai, J. L., Williams, P. F. & Van Roermund, H. L. M. 1991.
spinel transformation in and its implications Kinematics of crystal growth in syntectonic fibrous veins.
for the mechanism of deep focus earthquakes. J. Struct. J. Struct. Geol. 13, 823-836.
Geol. 15, 1249-1256. Van Den Driessche, J. & Brun, J-P. 1987. Rolling structures
Tse, S. & Rice, J. 1986. Crustal earthquake instability in re- at large strain. J. Struct. Geol. 9, 691-704.
lation to the depth variation of frictional slip properties. J. Van der Molen, I. & Paterson, M. S. 1979. Experimental de-
geophys. Res. 91, 9452-9472. formation of partially melted granite. Contrib. Miner. Pet-
Tsenn, M. C. & Carter, N. L. 1987. Upper limits of power law rol. 70, 299-318.
creep of rocks. Tectonophysics 136, 1-26. Van der Wal, D. 1993. Deformation processes in mantle
Tullis, T. E. 1980. The use of mechanical twinning in miner- peridotites with emphasis on the Ronda peridotite, S
als as a measure of shear stress magnitudes. J. geophys. Spain. Geologica Ultraiectina 102, PhD thesis, University
Res. 85, 6263-6268. of Utrecht, 180pp.
Tullis J. & Yund, R. A. 1987. Transition from cataclastic flow Vaughan, P. J., Green, H. W. & Coe, R. S. 1984. Aniso-
to dislocation creep of feldspar: Mechanisms and micro- tropic growth in the olivine-spinel phase transformation
structures. Geology 15, 606-609. of under hydrostatic stress. Tectonophysics
Tullis, J. & Yund, R. 1992. The Brittle-Ductile Transition 108, 299-322.
in feldspar aggregates: An experimental study. In Fault Vernon, R. H., Etheridge, M. A. & Wall, V. J. 1988. Shape
Mechanics and Transport Properties of Rocks: A Fests- and microstructure of microgranitoid enclaves: Indicators
chrift in honour of W. F. Brace (edited by Evans, B. & of magma mingling and flow. Lithos 22, 1-11.
122 REFERENCES

Vernon, R. H. & Paterson, S. R. 1993. The Ardara pluton, a mylonite zone. Contr. Miner. Petrol. 70, 193-202.
Ireland: deflating an expanded intrusion. Lithos 31, 17- White, S. H. 1976. The effects of strain on the microstruc-
32. tures, fabrics and deformation mechanisms in quartzites.
Vollbrecht, A., Rust, S. & Weber, K. 1991. Development of Phil. Trans. R. Soc. Lond. A283, 69-86.
microcracks in granites during cooling and uplift: ex- White, S. H. 1977. Geological significance of recovery and
amples from the Variscan basement in NE Bavaria, Ger- recrystallisation processes in quartz. Tectonophysics 39,
many. J. Struct. Geol. 13, 787-799. 143-170.
Vollbrecht, A., Durrast, H., Kraus, J. & Weber, K. 1994. White, S. H. & Knipe, R. J. 1978. Microstructure and cleav-
Paleostress directions deduced from microcrack fabrics age develeopment in selected slates. Contr. Miner. Petrol.
in KTB core samples and granites from the sorrounding 66, 165-174.
area. Scientific Drilling, Springer - Verlag, 233-241. White, J. C. & White, S. H. 1981. The structure of grain
Walker, A. N., Rutter, E. H. & Brodie, K. H. 1990. Exper- boundaries in tectonites. Tectonophysics 78, 613-628.
imental study of grain-size sensitive flow of hot-pressed White, S. H. & Wilson, C. J. L. 1978. Microstructures of
calcite rocks. In: Deformation Mechanisms, Rheology some quartz pressure fringes. Neues Jahrbuch fur Min-
and Tectonics (edited by Knipe, R. J. & Rutter, E. H.). eralogie, Abhandlungen 134, 33-51.
Spec. Publs geol. Soc. Lond. 54, 259-282. White, S. H., Burrows, S. E., Carreras, J., Shaw, N. D.
Wang, Y. 1987. A study of cataclastic deformation and ret- & Humpreys, F. J. 1980. On mylonites in ductile shear
rogressive metamorphism in fault zones on Guernsey. D. zones. J. Struct. Geol. 175-187.
Phil. Thesis, Univ. of Lond. Whitney, D. 1992. Garnets as open systems during regional
Wang, W. & Scholz, C. H. 1994. Wear processes during fric- metamorphism. Geology 24, 147-150.
tional sliding of a rock: A theoretical and experimental Wickham, S. M. 1987. The segregation and emplacement of
study. J. geophys. Res. 99, 6789-6799. granitic magmas. J. geol. Soc. Lond. 144, 281-297.
Wang, W. & Scholz, C. H. 1995. Micromechanics of rock Will, T. M. & Wilson, C. J. L. 1989. Experimentally produced
friction 3. Quantitative modeling of base friction. J. geo- slickenside lineations in pyrophyllitic clay. J. Struct.
phys. Res. 100, 4243-4247. Geol. 11, 657-668.
Wang, X. & Liou, J. G. Regional ultrahigh-pressure coesite- Willaime, C., Christie, J. M. & Kovacs, M. P. 1979. Exper-
bearing eclogitic terrane in central China: Evidence from imental deformation of K-feldspar single crystals. Bull.
country rocks, gneiss, marble and metapelite. Geology, Mineral. 102, 168-177.
19, 933-936. Williams, P. F. 1972. Development of metamorphic layer-
Wang, X., Liou, J. G. & Mao, H. K. 1989. Coesite-bearing ing and cleavage in low-grade metamorphic rocks at Ber-
eclogite from the Dabie Mountains in central China. Geo- magui, Australia. Am. J. Sci. 262, 1-47.
logy, 17, 1085-1088. Williams, P. F, Goodwin, L. B. & Ralser, S. 1994. Ductile de-
Warpinski, N. R. & Teuffel, L. W. 1993. Laboratory meas- formation processes. In: Continental Deformation (edited
urements of the effective stress law for carbonate rocks by P. L. Hancock), Pergammon Press, Oxford, 1-27.
under deformation. Int. J. Rock Mech. & Mining Sci. & Wilson, C. J. L. 1994. Crystal growth during a single stage
Geomech. Abstr.30, 1169-1172. opening event and its implications for syntectonic veins.
Watson, E. B. & Brenan, J. M. 1987. Fluids in the lithosphere, J. Struct. Geol. 16, 1283-1296.
1. Experimentally determined wetting characteristics of Wintsch, R. P. & Dunning, J. 1985. The effect of dislocation
CO2-H2O fluids and their implications for fluid transport, density on aqueous solubility of quartz and some geolo-
host- rock physical properties and fluid inclusion forma- gical implications. J. geophys. Res. 90, 3649-3657.
tion. Earth Planet. Sci. Lett. 85, 497-515. Wise, D. U., Dunn, D. E., Engelder, J. T., Geiser, P. A.,
Weertman, J. 1968. Dislocation climb theory of steady-state Hatcher, R. D., Kish, S. A., Odom, A. L. & Schamel, S.
creep. Trans. Am. Soc. Met. 61, 681-694. 1984. Fault-related rocks: some suggestions for termino-
Weertman, J. 1978. Creep laws for the mantle of the earth. logy. Geology 12, 391-394.
Phil. Trans R. Soc. Lond. A288, 59-96. Wong, T-F. 1990. Mechanical compaction and the brittle-
Weiss, L. E. 1954. A study of tectonic style; Structural invest- ductile transtion in porous sandstones. In: Deformation
igaton of a marble quartzite complex in southern Califor- Mechanisms, Rheology and Tectonics (edited by Knipe,
nia. Univ. Calif. Publ. Geol. Sci. 30, 1-102. R. J. & Rutter, E. H.). Spec. Publs geol. Soc. Lond. 54,
Wenk, H-R. & Christie, J. M. 1991. Review paper - comments 111-122.
on the interpretation of deformation textures in rocks. J. Wong, T-F. & Biegel, R. 1985. Effects of pressure on the mi-
Struct. Geol. 13, 1091-1110. cromechanics of faulting in San Marcos gabbro. J. Struct.
Wenk, H-R. 1978. Are pseudotachylites products of fracture Geol. 7, 737-749.
or fusion? Geology 6, 507-511. Wong, T-F. & Wu, L. 1995. Tensile stress concentration and
Wenk, H-R, O’Brien, D. K., Ratschbacher, L. & You, Z. compressive failure in cemented granular material. Geo-
1987. Preferred orientation of phyllosilicates in phyllon- phys. Res. Lett. 22, 1649-1652.
ites and ultramylonites. J. Struct. Geol. 9, 719-730 Wood, D. S. 1974. Current views of the development of slaty
White, S. H. 1979a. Difficulties associated with palaeostress cleavage. Ann. Rev. Earth Sci. 2, 1-35.
estimates. Bull. Mineral. 102, 210-215. Wright, T. O. & Platt, L. B. 1982. Pressure dissolution and
White, S. H. 1979b. Grain and sub-grain size variations across cleavage in the Martinsburg shale. Am. J. Sci. 282, 122-
REFERENCES 123

135. Krantz 1979, 1983, Krantz & Scholz 1977, Kulander &
Wright, T. O. & Henderson, J. R. 1992. Volume loss during Dean 1995, Labuz et al. 1987, Lawn & Wilshaw 1975a, b,
cleavage formation in the Meguma group, Nova Scotia, Lespinase & Pêcher 1986, Lin & Williams 1992, Lindquist et
Canada. J. Struct. Geol. 14, 281-290. al. 1984, Logan et al. 1981, Maddock 1986, 1992, Maddock
Yoshinubo, A. & Paterson, S. R. 1996. Fracture-controlled et al. 1987, Magloughlin 1989, Maltman 1994, Marone
magma conduits in an obliquely convergent continental & Scholz 1989, Masch et al. 1985, Mase & Smith 1984,
magmatic arc: Comment. Geology 24, 669-671. McEwen 1981, Means 1987, Menéndez et al. 1996, Meredith
Zee, R. Y. S., Teyssier, C., Hobbs, B. E. & Ord, A. 1985. De- 1983, Michalske & Frechette 1980, Mogi 1965, Morrit et al.
velopment of foliations in the Wyangala Gneiss, Central 1982, Narahara & Wiltschko 1986, Norris & Barron 1968,
New South Wales, Australia. Abstract, J. Struct. Geol. 7, Norton & Atkinson 1981, Olgaard & Brace 1983, Olsson &
501. Peng 1976, Passchier & Trouw 1996, Passchier et al. 1990,
Zhao, G. & Johnson, A. M. 1991. Sequential and incremental Peck 1983, Peng & Johnson 1972, Petit 1987, Pittman 1981,
formation of conjugate sets of faults. J. Struct. Geol. 13, Platt & Vissiers 1980, Power & Tullis 1989, Quakenbush &
887-896. Frechette 1978, Rice 1968, Robertson 1983, Rudnicki 1980,
Zhang, J., Wong, T-F., Yanagidani, T. & Davis, D. M. 1990. Rudnicki & Rice 1975, Rutter & Hadizadeh 1991, Rutter
Pressure-induced microcracking and grain crushing in et al. 1986, Sammis & Biegel 1989, Sammis et al. 1987,
porous rocks. J. geophys. Res. 95, 341-352. Schofield & Worth 1968, Scholz 1972,1990, Seyedolali et al.
Zwart, H. J. 1960. The chronological succession of folding 1997, Shand 1916, Shimada 1986, Shimamoto & Nagahama
and metamorphism in the central Pyrenees. Geol. Rdsch. 1992, Sibson 1980, Sibson et al. 1975, Smith 1984, Smith
50: 203-218. & Stenstrom 1965, Spray & Thompson 1994, Spray 1987,
Zwart, H. J. 1962. On the determination of polymetamorphic 1988, 1989, 1995, Sprunt & Nur 1979, Sprunt et al. 1978,
mineral associations, and its application to the Bosot area Stearns 1968, Stel 1981, Swanson 1980, 1981, Tapponier &
(Central Pyrenes). Geol. Rdsch. 52, 38-65. Brace 1976, Teufel 1981, Tija 1967, Tullis & Yund 1987,
1992, Underhill & Woodcock 1985, Vollbrecht et al. 1991,
1994, Wang 1987, Wang & Liou 1991, Wang & Scholz 1994,
Abbreviated references collected by 1995, Wang et al. 1989, Wenk 1978, Whitney 1996, Will &
Wilson 1989, Willaime et al. 1979, Wong 1990, Wong &
chapter Biegel 1985, Wong & Wu 1995, Zhang et al. 1990, Zhao &
Johnson 1991.
Italics indicates important general sources for the chapter
topic.
Chapter 3. Aharonov et al. 1997, Alvarez et al. 1976,
Chapter 1. Beaumont et al. 1996, Byerlee 1968, Carter Andrews & Railsback 1997, Bathurst 1958, Beach 1979,
& Kirby 1978, Chester & Logan 1987, Evans et al. 1990, Becker 1995, Bell & Cluff 1989, Bennema & Van Der
Griggs & Handin 1960, Kirby & Kronenberg 1984, Knipe Eerden 1987, Borradaile et al. 1982, Brantley 1992, Carrio-
1989, Kusznir & Park 1987, Lawn & Wilshaw 1975a, Lister Schaffhauser & Gaviglio 1990, Carrio-Schaffhauser et al.
& Snoke 1984, Molnar 1992, Patterson 1978, Ramsay & 1992, Casey 1995, Cox 1987, Den Brock 1996, Den Brock
Huber 1987, Rutter 1986, Sibson 1977, Snoke et al. 1998, & Spiers 1991, Dietrich & Grant 1986, Durney & Ramsay
Williams et al. 1994, Wise et al. 1984. 1973, Elliot 1973, Erslev & Ward 1994, Fisher & Anastasio
1994, Fletcher & Pollard 1981, Gratz et al. 1991, Gray
1977, Green 1980, Groshong 1976, 1988, Guzetta 1984,
Chapter 2. Agar et al. 1988, Allison & La Tour 1977, Ameen
Hedlund et al. 1994, Heidug 1991, Hickman & Evans 1991,
1992, Anderson & Grew 1977, Antonellini et al. 1994,
Hillner et al. 1992, Hippert 1994, Knipe 1979, Knipe &
Atkinson 1982, Aydin & Johnson 1978, 1983, Aydin 1978,
White 1977, 1979, Lespinasse & Cathelineau 1995, Lewis
Bansal 1977, Baud et al. 1996, Biegel et al. 1989, 1992,
& Holness 1996, Manktelow 1994, Marlow & Etheridge
Blenkinsop 1989, 1991, Blenkinsop & Rutter 1986, Blen-
1977, Masuda & Mizuno 1995, McEwen 1981, McCaig
kinsop & Sibson 1991, Boldt 1995, Borg & Maxwell 1956,
1987, Murphy 1990, Onasch 1994, Passchier & Trouw 1996,
Borg et al. 1960, Borradaile 1981, Brace & Bombolakis
Petit & Matthauser 1995, Powell 1979, Price & Cosgrove
1963, Brown & Macaudiere 1984, Brown & Scholz 1985,
1990, Railsback & Andrews 1995, Raj 1982, Raj & Chyung
Bruner 1984, Byerlee 1968, Camacho et al. 1995, Carter
1981, Ramsay 1980, Ramsay & Wood 1973, Ramsay &
& Kirby 1978, Chopin 1984, Conrad & Friedman 1976,
Huber 1983, 1987, Robert et al. 1995, Roedder 1984, Rutter
Cooper et al. 1989, Costin 1983, Cox & Atkinson 1983,
1976, 1983, Schutjens 1991, Smith 1964, Smith & Evans
Cruickshank et al. 1991, D’Arco & Wendt 1994, Das &
1984, Spiers & Schutjens 1990, Tada & Seiver 1986, 1989,
Scholz 1981, Dunn et al. 1973, Engelder 1974, Evans 1988,
Tada et al. 1987, Tullis et al. 1996, Urai 1983, Urai et al.
Evans & White 1984, Friedman & Logan 1970, Gallagher
1986, 1991, Watson & Brennan 1987, White & Knipe 1978,
1981, Gallagher et al. 1974, Griffith 1924, Grocott 1981,
Williams 1972, Wilson 1994, Wintsch & Dunning 1985,
Hadizadeh 1980, Hadizadeh & Rutter 1983, Hadizadeh &
Wood 1974, Wright & Platt 1982, Wright & Henderson 1992.
Tullis 1992, Hancock 1985, Hippert 1994, Hirth & Tullis
1991, Horii & Nemat-Nassar 1985, 1986, House& Gray
1982, Hull 1988, Inglis 1913, Jaeger & Cook 1979, Jamison Chapter 4. Allison & LaTour 1977, Blenkinsop & Drury
& Stearns 1982, Kemeny & Cook 1987, Knipe 1986, 1989, 1988, Christie & Ardell 1974, Den Brock & Spiers 1991,
124 REFERENCES

Drury et al. 1985, Drury 1993, Gleason et al. 1993, Hirth & Passchier & Trouw 1996, Passchier et al. 1993, Paterson &
Tullis 1992, Hobbs et al. 1976, Hull 1975, Jessel 1988a, b, Vernon 1995, Paterson et al. 1989, Petit 1987, Platt 1984,
Jessel & Lister 1990, Kruhl 1996, Lloyd & Freeman 1991a, Platt & Vissers 1980, Prior et al. 1987, Ramsay 1962, 1967,
b, 1994, Mainprice & Nicolas 1989, Mainprice et al. 1986, 1980, Ramsay & Graham 1970, Ramsay et al. 1983, Robert
McLaren 1991, McLaren et al. 1967, Means 1981, Means & 1989, Robin & Cruden 1994, Rutter et al. 1986, Rykkelid
Dhong 1982, Nicolas & Poirier 1976, Spang & Van der Lee & Fossen 1992, Scheuber & Andriessen 1990, Schmid et
1975, Twiss 1974, Urai et al. 1986, Wenk & Christie 1991, al. 1987, Shelley 1989b, 1995, Shimamoto 1989, Simpson
White 1976. 1986, Simpson & Schmid 1983, Sylvester 1988, Ten Brink
& Passchier 1995, Tikoff & Greene 1997, Tikoff & Fossen
Chapter 5. Behrmann 1985, Bell et al. 1992, Borradaile et 1993, Tikoff & Teyssier 1995, Turner & Wiess 1963, Van den
al. 1982, Burnley et al. 199l, Busa & Gray 1992, Champness Driesche & Brun 1987, Wenk et al. 1987, White & Wilson
& Lolimer 1974, Drury & Humphreys 1988, Evans et al. 1978, White et al. 1980, Williams et al. 1994, Yoshinobu &
1980, Fliervoet & White 1995, Gilotti & Hull 1990, Green Patterson 1996, Zee et al. 1985.
1986, Green & Burnley 1989, Gower & Simpson 1992,
Hacker & Kirby 1993, Jessell 1987, Johnson & Vernon 1995, Chapter 8. Alexopoulos et al. 1988, Ashworth & Schneider
Kirby & Stern 1993, Langdon 1982, MacKinnon et al. 1977, 1985, Bohor et al. 1993, Carter 1965, Carter et al. 1986,
Nicolas & Poirier 1976, Passchier & Trouw 1996, Passchier 1990, Engelhardt & Bertsch 1969, Fel’dman 1994, French
et al. 1992, Powell & Treagus 1970, Rutter et al. 1994, et al. 1974, Gigl & Dachville 1968, Glass 1990, Goltrant
Schmid et al. 1987, Shelly 1989a, b, Shoneveld 1977, Smith et al. 1991, 1992, Gratz et al. 1996, Grieve et al. 1996,
1964, Tingle et al. 1993, Urai et al. 1986, Vaughan et al. Hörz 1968, Hörz & Quaide 1973, Huffman & Reimold 1996,
1984, White 1977, White & White 1981, Zwart 1960, 1962. Huffman et al. 1993, Joreau et al. 1997, Kamo et al. 1996,
Kieffer 1975, Kieffer et al. 1976, Koeberl 1990, Langenhorst
Chapter 6. Arzi 1978, Ashworth & McLellan 1985, Bagnold 1994, Langenhorst & Clymer 1996, Langenhorst & Deutsch
1954, Brace & Martin 1968, Benn & Allard 1989, Blu- 1994, Langenhorst et al. 1992, Leroux & Doukham 1995,
menfield 1983, Blumenfield & Bouchez 1988, Bouchez & 1996, Leroux et al. 1994, Lyons et al. 1993, Martini 1991,
Gleizes 1995, Bouchez et al. 1992, Burg 1991, Conolly et McIntyre 1962, Officer & Carter 1991, Reimold 1994, 1995,
al. 1997, Copper & Kohlstedt 1982, 1984, Dell’Angelo & Reimold & Hörz 1986, Reimold & Stöffler 1978, Reimold et
Tullis 1988, Dell’Angelo et al. 1987, Finney 1970, Gapais al. 1998, Robertson 1975, Robertson & Grieve 1977, Seye-
& Barbarin 1986, German 1985, Gray 1968, Guineberteau dolali et al. 1997, Sharpton & Schuraytz 1989, Stöffler 1972,
et al. 1987, Hibbard 1987, Hirth & Kohlstedt 1995, Hutton 1984, Stöffler & Langenhorst 1994.
1988, Jarewicz & Watson 1984, 1985, Karlstrom et al. 1993, Chapter 9. Angelier 1984, Angevine & Turcotte 1983,
Komar 1972a, b, Lagarde et al. 1994, Law et al. 1992, Ashby & Verrall 1978, Avé Lallement 1978, Baud et al.
Lejeune & Richet 1995, Marsh 1981, McBirney & Murase 1996, Berckhemer et al. 1979, Bernabé 1987, Biegel et al.
1984, McLellan 1984, Miller & Paterson 1994, Miller et al. 1989, Blacic & Christie 1984, Blenkinsop & Drury 1988,
1988, Mitra 1976, Nicolas et al. 1988, Park & Means 1996, Boland & Tullis 1986, Boullier & Guegen 1975, Burkhard
Paterson et al. 1989, Paterson & Vernon 1995, Pons et al. 1993, Burov & Diament 1995, 1996, Busch & Van der pluijm
1995, Quick et al. 1992, Ramsay 1989, Renner et al. 1999, 1995, Byerlee 1978, Caristan 1982, Carter & Friedman 1965,
Riley 1990, Roscoe 1952, Rushmer 1995, Rutter & Neumann Carter & Rayleigh 1969, Carter & Tsenn 1987, Carter et
1995, Ryan 1995, Sherman 1968, Simpson 1985, Stel 1991, al. 1993, Christie & Koch 1982, Christie & Ord 1980, De
Van der Molen & Paterson 1979, Vernon & Paterson 1993, Bresser 1988, De Bresser & Spiers 1990, Dennis 1984,
Vernon et al. 1988, Wickham 1987. Detournay et al. 1989, Dorn 1954, Drury 1993, Drury & Urai
1990, Drury et al. 1985, Evans & Groshong 1994, Farver &
Chapter 7. Aerden 1995, 1996, Balé & Brun 1989, Behr- Yund 199la, b, Fowler 1990, Gilletti & Yund 1984, Goetze
mann 1984, 1987, Behrmann & Platt 1982, Bell 1985, Bell & & Evans 1979, Gratier & Guiget 1986, Gratier & Jenatton
Johnson 1989, Bell et al. 1992, Benn & Allard 1989, Berthé 1984, Griggs 1967, Griggs & Blacic 1965, Groshong 1972,
et al. 1979, Bjørnerud & Zhang 1995, Blenkinsop & Treloar 1974, 1984, Gross & Van Heege 1973, Hallam & Ashby
1995, Blumenfield 1983, Blumenfield & Bouchez 1988, 1990, Handy 1990, 1994, Hansen & Carter 1982, 1983,
Bouchez et al. 1983, 1992, Burg 1987, Dennis & Secor 1987, Heard 1963, Hirth & Kohlstedt 1995, Hobbs et al. 1986,
1990, Doblas et al. 1997, Erskine et al. 1993, Etchecopar & Horii & Nemat-Nasser 1985, 1986, Jaeger & Cook 1979,
Malavielle 1987, Evans & Dresden 1991, Fossen & Holst Jamison & Spang 1976, Jaoul et al. 1984, Ji & Zhao 1993,
1995, Gamond 1987, Gapais & Brun 1981, Ghisetti 1987, 1994, Karato & Wu 1993, Kemeny & Cook 1987, Kirby &
Goldstein 1988, Guineberteau et al. 1987, Hanmer 1986, Kronenburg 1984, Koch & Christie 1981, Kollé & Blacic
Hanmer & Passchier 1991, Ildefonse & Mancktelow 1993, 1983, Kronenberg & Tullis 1984, Kronenburg et al. 1990,
Jeffrey 1922, Krabbendam & Leslie 1996, Law 1986, 1987, Kruhl 1996, Kruhl & Nega 1996, Kumpel 1991, Kusznir
1990, Law et al. 1984, 1994, Lister & Snoke 1984, Little et 1982, Kusznir & Bott 1977, Kusznir & Park 1982, 1984,
al. 1994, Mainprice et al. 1986, March 1932, Means 1981, 1987, Lacombe et al. 1990, Laurent et al. 1981, 1990, Lawn
Means et al. 1980, Mercier 1985, Miller & Paterson 1994, & Wilshaw 1975a, Lehner 1990, Lemée & Guegen 1996,
Park & Means 1996, Passchier 1983, 1984, 1991, 1994, Linker & Kirby 1981, Marone et al. 1992, McClintock
Passchier & Simpson 1986, Passchier & Sokoutis 1993, & Walsh 1962, McLaren 1991, Meisner & Strehlau 1982,
REFERENCES 125

Mercier 1980a, Mercier et al. 1977, Muller & Briegel 1978, 1977, 1980, Scholz 1990, Schulmann et al. 1996, Shelton et
Murrell 1963, 1965, Murrell & Digby 1970, Nemat-Nasser al. 1981, Sibson 1982, 1983, Spang & Van der Lee 1975, Spi-
& Horii 1982, Newman 1994, Ord 1991, Ord & Christie ers et al. 1990, Stoker & Ashby 1973, Takahashi et al. 1997,
1984, Ord & Hobbs 1985, Ord & Hobbs 1989, Paterson Terzaghi 1943, Tse & Rice 1986, Tsenn & Carter 1987, Tullis
1978, 1987, 1995, Paterson & Luan 1990, Pavlis & Bruhn 1980, Tullis et al. 1991, Turner 1953, Twiss 1977, 1986, Van
1988, Peng & Johnson 1972, Pfiffner & Burkhard 1987, der Wal 1993, Walker et al. 1990, Warpinski & Teuffel 1993,
Post & Tullis 1999, Power & Tullis 1989, Raj 1982, Raj & Weertman 1968, 1978, Weiss 1954, White 1976, 1979a.
Ashby 1971, Raterron & Jaoul 1991, Rice & Gu 1983, Ross
et al. 1980, Rowe & Rutter 1990, Rutter 1976, Schmid et al.
This page intentionally left blank
Index
(Bold numbers refer to plates or figures)

abrasive wear 10 classification 28, 29


accretion steps 20 continuous 29
activation enthalpy 92-96 crenulation 29, 30, Plates 16, 17
activation volume 94 disjunctive 29, 3.7
adhesive strength, wear 10 domain 29, 30
alteration 9, 10, Plate 1 slaty 30, 3.9
Amonton’s law 10, 91 spaced 29, 30
amphibolite facies 24, 52, 96 zonal 29
amygdale 22 coefficient of friction, coefficient of internal friction 90, 91
annealing 62, 82, 88 coesite 16, 80, 83, 8.7, 85-88
anticrack 58 coherent exsolution 52
antitaxial fibre growth 33, 37, 3.17, 3.18 coherent transformation 52
arrow method 102, 103 cohesion 5, 90, 91
asperities, asperity ploughing 10, 2.4, 20 composite deformation mechanism 3, 52, 90
asterism 82 composite fibre growth 37, 3.17
constrained comminution 10
ballen 83, Plates 47, 48 contact melting 63
barrier theory 9 contiguity 59
bend 79 core-and-mantle structure 4.14, 42, 49
bookshelf sliding 77 corona 2, 57, Plate 34
botryoidal texture 32, 35 crack-seal 35, 92
boudin 19, 2.17, Plate 6 crater 23
Boussinesq configuration 12 creep 92-94
Brazil twin 81, 8.3, 85, 88, 89 Coble 52, 92-94
breccia 5, 6 diffusion 52, 54, 60, 97
brittle 1,3 dislocation 93-97
brittle-ductile transition 4, 5 grain size sensitive 92, 93
brittle-plastic transition 3, 97 Nabarro-Herring 52, 92, 94
Burger’s vector, 39, 4.1, 58 pressure solution 92, 93
Byerlee’s law 91, 97 critical melt fraction 59
critical packing, packing density 59, 60
calcite twin morphology 104 critical slip distance 92
cataclasis 1-3, 5, 7-24, 30, 38, 47,60,63,74 cross-slip 39, 4.4
cataclasite 6 crystal plasticity 4, 5, 19, 59, 104
foliated cataclasite 6, Plate 8 crystallites 22
protocataclasite 6 crystallographic fabric, preferred orientation 1, 2, 19, 22, 50,
ultracataclasite 6, 23 51, 4.20, 58, 62, 75, 76, 7.19
cataclastic flow 4, 18, 19, 63 curved foliation 66
cathodoluminescence 2.6, 12, 17, 32,35, 81
cement, cementation 17, 32 Dauphine twin 57
chemical potential 24, 52 debris streaking 20
chemical zoning 2, 57 decussate texture 2, 54, Plate 28
chessboard pattern 41, 4.10, 47 defect 52
Chixculub structure 84 deflection surface 54, 57
Cish diagram 37 deformation bands 2, 5, 7, 18, 41, 4.6, 47, 50
cleavage 2, 13, 28-30 deformation lamellae 2, 14, 4.11, 47, 49, 101-103, 9.4

127
128 INDEX

deformation mechanism 1, 2, 3, 5, 90 fibre strain 17


classification 1, 2 growth histories: antitaxial, ataxial, composite, non-
map 90, 97, 98 systematic, syntaxial 37, 3.17, 3.19
deformation microstructure 1-3, 5, 90 non-tracking 38, 3.19
classification 1, 2 tracking 37, 38, Plate 23
deformation (mechanical) twin 2, 39-41, 4.8, Plate 24, 62, tracking efficiency, 38, 3.19
81, 87, 102-104, 9.4, 9.5 Fick’s law 24, 52
deformation zone rocks 5 finite shortening, extension 25
deformation flow laws 90, 92-97
continuity 4, 5, 18, 19,57 fluid inclusions planes 2, 12, 24, 2.5, 33-35, 3.13, 3.14, 47
distribution 4, 5, 18 foam texture 2, 54
mechanism and mode 4, 5 fold 4, 19, 2.17, 60, 63, 70
scale 4 foliation 54-57, 65, 75
dendritic crystals 22 curved 66, Plate 39
diaplectic glass 80-83, 85, 87, 89 external foliation 54, 55, 5.3, Plates 29-33
diffusion 24, 29, 53, 57, 82, 94-94 internal foliation, 54, 55, 5.3, Plates 29, 31-33
coefficient 52, 92-94 oblique foliation, 66, 67, 7.3
creep - see creep steady-state, strain insensitive 61
grain boundary 93 strain sensitive 66
volume 93 fractal dimension 10, 22, 92, Plates 1-4
diffusive mass transfer 1-3, 47, 63, 90, 92 fracture 1
by solution 3, 24-38 fracture toughness 9
in melts 60 fragmentation 10
solid state 52-58 frictional sliding 7, 10, 2.4, 90-92, 97
dihedral angle 24
dilatancy hardening 60 generation plane 22
dislocation 1, 30, 39, 4.1, 4.2, 4.3, 4.4, 47 geotherm 97
climb 39, 93 geothermobarometry 104, 105
creep 39, 5.1, 54, 93-97 gouge and gouge zone microstructures 5, 6, 10, 17, 2.17,
density 39, 41, 47-49, 58, 81, 82 Plate 6, 72
glide 39, 51,93-95 grain boundary migration 4.12, 47, 4.17, 48-50, 4.18, 4.19,
displaced grain fragments 79, 7.23, Plate 45 95, 99, 100
displacement control 33 grain boundary width - fast or free 54
displacive transformation 52 grain dispersive pressure 61
divider method 105 grain shape fabric 2
Dorn law 93 cataclastic 19, 33
ductile stringer 19, 2.17 diffusive mass transfer 33, 3.11, 52, Plates 26, 27, 54
ductility 4, 5 intracrystalline plastic 33, 4.12, 47
dynamic recrystallization 47-50, 4.19 magmatic 62, Plate 36
sub-magmatic 62
effective stress 91 grain size sensitive creep 92, 93
Einstein-Roscoe equation 59 grain surface deposition textures 2, 30
enclave 60 grain surface solution textures 2, 25
erosional sheltering 20 granoblastic polygonal texture 32, 54, 62
etching, etch pit 39, 88 growth twins 39, 62, 81
exaggerated grain growth 54 growth zoning 57, 83, Plate 18
exponential law 93-97
extensional crenulation cleavage 70 helicitic texture 54
external asymmetry 76, 7.19 Hertzian configuration 12
homologous temperature 97
fabric skeleton 76 hydrolytic weakening 9, 94
face control 33, 38, Plate 19
failure criteria 90, 91, 97 igneous zoning 62
Coulomb and Mohr failure criteria, 90, 97, 98 imbrication 60, 62, 78
and fracture mechanics 91 inclusion 35, 3.15, 3.16, 62, 84
Griffith criteria 91,97 band 35, 3.15, 3.16
fault gouge - see gouge trail 35, 3.15, 3.16, 37, 38, 54, 65
fault breccia - see breccia indenting grain contacts 2, 25, 3.1, 3.2, 33, 63
fibre 32, 33, 37, 65 independent particulate flow 2, 22
INDEX 129

instantaneous stretching axis, axes 65, 73, 7.15, 7.17 sub-critical propagation 9, 94
interconnected weak layer 96 thermally-induced 15, 16
internal asymmetry 76, 7.19 transgranular 10, 2.5, 12, 13, 16
internal strain energy 4, 48, 52, 55 wing 13, 2.10
interpenetrating grain contacts 2, 25, 3.1, 3.4, 33 microcrystallite 81, 82
intracrystalline plasticity 1, 3, 4, 14, 19, 22-24, 30, 39-51, 57, microfault 2, 7, 17, 18, 2.14-2.16, 7.23, Plates 3, 45
60, 62, 63, 75, 90, 93, 97 microfold 77, 7.22
intracrystalline deformation bands 2, 41 microfracture 1, 2, 4, 7, 80, 85, 86, Plate 4
island and channels 24 mirror 20
isostrain criterion 96 mist 20
isostress criterion 96 surface features 2, 19, 20
velocity hackle 20
jog 79 Wallner line 20
microlithon 29, 30
kink, band 2, 14, 30, 41, 4.6, 4.7, 4.8, 50, 62 microphenocryst 22
microslickolite, 27, 28, 3.6
landslide 23 microstylolite 1, 2, 27, 28, 3.5, Plate 14, 3.6
lattice preferred orientation 58, see crystallographic fabric characteristics 27
lechatelierite 80-85 filling 24-28
lineation 66, 7.1, 7.2, 75 formation and propagation 27, 28
lithospheric strength envelope 90, 97, 98, 9.2 teeth, walls, crown 27, 3.5, 28
load bearing framework 96 microvein 1, 2, 7, 35-38
low angle boundaries 41 free-face growth 35
lower stability limit 92 fibrous 33, 35, 37, 38
laminated 35
magmatic flow 3, 59-63, Plate 36 opening vector 35
magmatic microstructures 59, 62 migmatite 63
magmatic shear zones 62, 77, 78 millipede texture 54
magnetic anisotropy 60, 62 mimetic crystallization 54
Martensitic-like transformation 52, 57 mode of failure 4
mean stress 94 molar entropy 24
megacryst 62 molar internal energy 24
metatextite 63 molar volume 24, 92-94
mica beard 32 mosaicism 81-83, Plate 48, 85-89
mica fish 74, Plate 42 mylonite 5, 6, 62
microboudins, asymmetric 76, 7.20, 7.21, Plate 6 protomylonite 6
microcrack 2, 7, 2.1, 2.5, 2.6, 10-16, 33 ultramylonite 6
axial 12, 17, 2.14, 19, 63,91
characteristics and observation 10-12 necking down 33, 35, 3.14
circumgranular 10-12 neosomes 63
classification 10-12 new grains 47, 50
cleavage 13, 19 non-magmatic deformation 63, 64
cone 12 non-systematic fibre growth 37, 3.17, Plate 22
dynamic propagation 7 normal slip crenulation 70
elastic mismatch 13, 14, 2.11, Plate 5
en-echelon 17, 2.15 oblique grain shape fabric 78
en passant 17, 2.15 ophitic texture 62
extension 12, Plate 2 order-disorder transformation 52
extension force 7 Ostwald ripening 54
flaw-induced 2.6, 13, 2.10 overgrowth 1, 2, 4, 13, 32, 33, 3.10, Plate 18, 3.11, 4.11
grain boundary 13
impingement 2.6, 12, 13, 2.7, 2.8, 2.9 P-foliation 19, Plate 6, 79
intragranular 10, 2.5, 12, 13, 2.9, 18, 19 P-shear, fracture 19, 20
microfault- induced 14, 19 paleopiezometry, palaeopeizometer, 98-101
microscopic feather fracture 14, 15, 2.12 deformation lamellae 101-103
modes I, II, III 7, 2.2, 12 dislocation density 100, 101
phase transformation-induced 16 general problems 104
plastic mismatch 14, 2.12, 2.13 maximum twin volume, 101
refracturing 13, 2.11 principal stress and strains 103, 104, 9.4
130 INDEX

recrystallized grain size 98-100, 9.3 pseudotachylites 2, 5-7, 22, 23, Plates 9-11, 83, 92
subgrain size 100, 101 characteristics 22
twinning - differential stress 100, 101 misidentification 23
twinning density, 101 origin 22, 23
twinning incidence, 101
particle size distributions 10, 22, 92 rate and state dependent frictional sliding 92
perlitic texture Plates 47, 48 reaction rims 2, 57
permeability 91 reaction zoning 57
phase transformation microstructures 2, 57, 58, 80-92 recovery 2, 41, 47, 82
phenocryst 60-62 recrystallized grain size 49, 50
phenomenological coefficient 93 reentrant zoning 54
phyllonite 6 relict minerals 2, 57
pinching off 33, 3.14 rheology 90
planar deformation feature 47, 8.2, 80-89 rhomb dodecahedra 54
decorated 81,82 ribbon grain 2, 4.13, 52-54, 5.1
non-decorated 81, 82 ridges and grooves 20
sub-lamellar structure 81, 88 Riedel, conjugate Riedel shear 2.17, 19, Plates 6, 7, 20, 72,
plastic, plasticity 3, 14 79
poikiloblast 54 fractures 15
Poisson’s ratio 97 risers 20, 2.19, 79
polymineralic deformation 94-97 congruous/incongruous 20, 2.19
pore fluid, pressure 3, 22-24, 91 rolling structures 77
porosity 12, 13, 18, 87, 88, 93 roughening transition 32
reduction 19, 32
porphyroblasts 1, 2, 54-57 S-, C- and 5, 20, 62, 63, 66, 70-73, Plates 40, 41
characteristics 54, 55 characteristic and classification 70-72, 7.11, 7.13, 7.14
growth mechanisms 55 curvature of S-foliation 73
internal, external foliations 54, 75, 7.18 formation and evolution 72
intertectonic 55-57, 5.3, 5.4, Plate 31 shear on C-or 19, 73
plate 33 shear sense from 73
posttectonic 55-57, 5.3 scaly clays 22
pretectonic 55-57, 5.3, Plate 30 scanning electron microscope 2.6, 19, 2.18, 2.19, 25, 30, 50,
relationship to deformation 55-57, 5.3 79, 82, 88
shear sense indications 75, 7.18 schlieren 60
syntectonic 55-57, 5.3, Plate 32 secondary recrystallization 54, 55
porphyroclast, porphyroclast systems 5, 14, 67-70 sector zoning 54
characteristics and classification 67, 68, 7.4 semibrittle 3, 14, 19
complex type 67, 7.4, 7.8 sensitive tint plate 2.6, 12, 17, 50, Plates 1-4, 45, 79
67-70, 7.4, 7.7, 7.9, 7.10, 74 separatrix 68, 69
deflection, embayments 68-70, 7.7, 7.9 shape preferred orientation 47, 66
faces of a tail 69, 7.9 shatter cones 23
in-plane 67, 68 shear band 70
67, 7.4 shear modulus 97
mantle 67 shear sense 65-79
mechanisms of formation 68, 69 criteria 65-79
67-69, 7.4, 7.5, 7.6, 7.9 for faults 79, 7.23, Plate 45
stair-step 67, 68 in rocks containing melt 77-79
tail, wing 67 observation plane 65, 71
power law 93-97 shock-induced microstructures 3, 80-89, 8.1
breakdown 94 shock mechanisms and metamorphism 23, 80
pre-exponential constant 93-97 shock wave barometry and thermometry 85-88
pre-lithification deformation 2, 22 calibration 85-87
pressure shadows and fringes 1, 2, 4, 32, Plate 19, 33, 3.12, problems 87, 88
63, 65, 73,74 slickenfibre 20, 2.19, 79
last increment of growth 73, slickenline 20
kinematics in shear zones 73, 7.15 slickenside 20, 92
shape 73, 7.16, 7.17 slip 39
process zone 9, 28 direction 39, 51
prod mark 20, 2.18 plane 39, 4.1, 51
INDEX 131

system tiling 60, 78


snowball texture 54, Plate 29 tool track or mark 20, 2.18
solid state transformation 52-58, 80-82 transmission electron microscope 19, 39, 4.2, 81-83, 88, 101
solubility 24, 92-94 transpression 65, 7.2
spherulites 22, 83, 84 triple grain junction 54, 5.2
spin 65 truncating grain contact 2, 25, 3.1, 3.3, Plate 12, 30, 33
state variable in frictional sliding law 92 truncation surface 30, 3.7, 54, 55
static recrystallization 54, 62 twin, twinning 47, 57, see also deformation (mechanical)
stick-slip behaviour 92 twin, growth twin, paleopiezometry
stishovite 80, 83, 8.8, 85-87
strain cap 2, 27, Plate 13 undercutting mechanism 24
strain rate 90, 92-97 undulatory extinction 1, 2, 33, 41, 4.5, 4.6, 4.13, 49, 62, 82
strawberry texture 82, 8.5, 8.6 uniaxial compressive strength 90, 91
stress 1 uniaxial tensile strength 90, 91
amplification 97 upper stability transition 92
corosion 9
exponent 93-97 velocity weakening, strengthening 92
intensity factor 7, 9 vesicle 84
stretched crystal fibre 37, 3.17 viscosity 59-61
strewn field 84 volume loss 29, 3.8, 30
sub-boundary migration 50 vorticity 65
sub-magmatic flow 3 external 65
microstructures 3, 59-63, Plates 37, 38, 78, 79 internal 65
subgrain 1, 2, 14, 19, 32, 41, 47, 4.9, 4.13, 47, 4.16, 50, 62 profile plane 65, 7.1
boundary orientation in quartz 41, 63, 105 shear-induced 65
rotation 47, 4.15, 4.16, 48, 49, 4.19, 50, 99, 100 vector 65, 7.1
size 1, 47
superplasticity 3, 52, 57, 58, Fig. 5.5, 94 wear groove 20, 2.18
surface tension force, energy 7, Plate 27, 54, 55 whole lithosphere failure 97
sutured grain boundary 25, 50, 3.1, 4.19, 104, 105, 9.6, 9.7
symplectite 2, 57, Plate 35 Y-shear Plate 7, 19, 2.17, 72
syntaxial fibre growth 33, 37, 3.17, Plate 21 Young’s modulus 7

T fracture 19, 2.17, 20, 79 zircon, shocked 82, 8.4-8.6, 89


tectite, microtectite 83, 84
This page intentionally left blank
Color Plate Section
This page intentionally left blank
COLOR PLATE SECTION 135

Plate 1. Alteration-enhanced microcracking. The plagioclase crystal in the light blue colour is fragmented along cleavage planes without significant rotation
or displacement of the fragments. Alteration of the feldspar to more sodic compositions (and ultimately to laumontite) occurs along the microcracks, showing
that chemical reaction and deformation were linked in alteration-enhanced microcracking. The fractal dimension of the particle size distribution (PSD) is
~2.0, which is characteristic of this type of cataclasis. (Chapters 2.2.1, 2.3.6).
XPL, ST, 4.3 mm, Biotite granite, Cajon Pass drillhole, California, U.S.A.

Plate 2. Extension microcrack. Grain boundaries between the yellow and pink grains in the wallrock can be matched across the microcrack, demonstrating
extension in the plane of the section. The microcrack is filled by angular fragments of quartz, biotite and feldspar in random orientations which can not be
matched with adjacent grains in the wall-rock. These textures suggest that the fragments were transported in a single, chaotic manner from some distance,
probably in a fluid matrix. The fractal dimension of the PSD is approximately 2.8, slightly higher than typical for extension microcracks. (Chapter 2).
XPL, ST, 2.2 mm, Granodiorite, Cajon Pass drillhole, California, U.S.A.

Plate 3. Microfault. The great variety of colours of the fragments (mostly quartz) in the matrix under the sensitive tint plate shows that they have been
rotated and probably derived from grains at least outside the field of view. Many fragments are equant and sub-angular, suggesting some wear during shear.
The fractal dimension of the PSD is approximately 2.6, a common value for fragments in shear fractures, and also the value predicted by the constrained
comminution model. (Chapters 2.2.2, 2.4).
XPL, ST, 4.3 mm, Granite, Cajon Pass drillhole, California, U.S.A.
136 COLOR PLATE SECTION

Plate 4. Selective microfracture of larger fragments. Angular, randomly orientated fragments were produced by microfracture during cataclasis. A large
feldspar grain at lower right is separated into two fragments by a microfault. The fractal dimension of the PSD is approximately 2.9. (Chapter 2.2.2).
XPL, ST, 4.3 mm, Granite, Cajon Pass drillhole, California, U.S.A.

Plate 5. Elastic mismatch strains. Intracrystalline plastic strains around the ends of an inclusion of biotite are dramatically shown by the areas of quartz in
extinction, compared to the birefringent host grain elsewhere. The strained area has the same geometry as the area of high stress predicted from modelling
the elastic stress field around a weak inclusion, suggesting that the elastic stresses resulted in lattice distortion. See Fig. 2.11. (Chapter 2.3.7).
XPL, 0.5 mm, Chilimanzi granite, Great Zimbabwe, Zimbabwe.

Plate 6. Riedel shears, asymmetric boudin and P-foliation in a sinistral gouge zone. Riedel shears trend from top right to bottom left, and cut across a P-
foliation which trends in the opposite direction. The two features surround an asymmetrical boudin. All three features give a sinistral shear sense. Compare
with Figs. 2.17, 7.20. (Chapters 2.7, 7.13.4).
PPL, 2 mm, Tarskavaig thrust zone, Tarskavaig, Isle of Skye, Scotland, U.K.
COLOR PLATE SECTION 137

Plate 7. Y-Shear in a gouge zone. The central dark line parallel to the top/bottom edges of the photomicrograph is a Y-shear. Above and below are Riedel
shears similar to those in the previous plate, which give a sinistral sense of shear. Compare with Fig. 2.17. (Chapters 2.7, 7.13.4).
PPL, 2 mm, Tarskavaig thrust zone, Tarskavaig, Isle of Skye, Scotland, U.K.

Plate 8. Foliated cataclasite. The foliation (planar feature from upper right to lower left) is defined by bands of quartz and calcite. Calcite-filled microfractures
at high angles to the foliation demonstrate the brittle component of the deformation. (Chapter 1.3).
XP, 1 mm, Foliated cataclasite, Punchbowl fault zone, California, U.S.A.

Plate 9. Pseudotachylite. Dark, fine grained matrix containing clasts fills a triangular-shaped veinlet branching from a narrow vein of the same material
parallel to a fault surface. The veinlet cross-cuts an older foliation in the wall rock. Sub-angular fragments of wall rock in the veinlet have a large range of
sizes. All these features are typical of pseudotachylite. See Plates 10, 11. (Chapter 2.11).
PPL, 2 mm, Pseudotachylite in grey gneiss mylonite, Buhwa, Zimbabwe.
138 COLOR PLATE SECTION

Plate 10. Pseudotachylite. The same view as plate 9 in crossed polars. The opacity of the matrix suggests a glassy nature. Slight alteration to a very fine
grained, moderately birefringent phyllosilicate picks out a weak foliation in bands which are concave towards the apex of the veinlet. These bands suggest a
primary flow foliation. See Plates 9, 11. (Chapter 2.11).
XL, 2 mm, Pseudotachylite in grey gneiss mylonite, Buhwa, Zimbabwe.

Plate 11. Pseudotachylite. Branching veins contain an opaque, fine grained matrix suggesting glass. Poorly sorted, sub-rounded fragments define a weak
foliation parallel to a feint colour banding in the matrix. The foliation and banding are folded about an axis parallel to the margin of one vein, suggesting a
flow foliation. See Plates 9, 10. (Chapter 2.11).
PPL, 2 mm, Pseudotachylite in tonalitic granulite, Bollingen Islands, Antarctica.

Plate 12. Truncated Ooid. Concentric banding in the elliptical ooid is truncated at the upper and lower surfaces by quartz grains along approximately planar
surfaces. The length of the original ooid can be estimated as approximately 0.6 mm: its present length is 0.3 mm. (Chapter 3.4).
XPL, 1 mm, Bargate Stone, U.K.
COLOR PLATE SECTION 139

Plate 13. Strain cap. The porphyroclast of feldspar at centre has foliae of muscovite wrapped around the top and bottom surfaces. Flakes of muscovite
distributed sparsely through the matrix around the porphyroclast define a weak foliation. The presence of muscovite in the matrix suggests that the muscovite
was concentrated in the strain cap by diffusion of the other matrix components away from the strain cap. (Chapter 3.5).
XPL, 1 mm, Mylonite, Moine thrust zone, Scotland, U.K.

Plate 14. Microstylolite truncating crinoid fragment. The microstylolite has a filling of opaque iron oxides/hydroxides and truncates the crinoid fragment,
showing material removal. See Fig. 3.8. (Chapter 3.6).
PPL, 1 mm, Crinoidal limestone, Derbyshire, U.K.

Plate 15. Microstylolites. The wider stylolite is approximately planar, and the narrow stylolite is sinusoidal. Both are zones of concentrated opaque minerals,
which are found in low concentrations throughout the adjacent rock, showing that they have been concentrated by removal of the more soluble carbonates
along the stylolites. Shell fragments are truncated by the stylolites. The amplitude of the sinusoidal stylolite suggests approximately 0.3 mm of shortening
perpendicular to the stylolite trace. (Chapter 3.6).
PPL, 2 mm, Carboniferous limestone, Somerset, U.K.
140 COLOR PLATE SECTION

Plate 16. Incipient crenulation cleavage. A early fabric defined by muscovite and chlorite is folded into open folds. A weak second fabric is seen on the
fold limbs, defined by muscovite and chlorite alignment parallel to the axial surfaces of the folds, and by concentrations of opaque minerals which are found
elsewhere throughout the rock, suggesting that diffusive mass transfer has contributed to the formation of this cleavage by removal of the more soluble
phases from the fold limbs. (Chapter 3.7.2.2).
PPL, 2 mm, Albite chlorite schist, Perthshire, U.K.

Plate 17. Advanced crenulation cleavage and fabric transposition. Two distinct types of layers are visible in this rock. Layers consisting of strongly aligned
biotite also contain a few small elongate grains of quartz parallel to the biotite fabric (S2, not labelled). Layers between the S2 domains contain larger,
equant quartz grains and a higher proportion of quartz, which defines a compositional banding (S1, not labelled) of quartz layers alternating with biotite
layers, which also have a strong fabric parallel to S1. S1 is folded into tight to isoclinal asymmetric folds; small scale fold hinges can be seen in some of
the biotite grains and in the compositional layering. Depletion of quartz in the S2 layers shows that it has diffused away from the S2 domains. S1 has been
almost completely transposed into S2. (Chapter 3.7.2).
XP, 1 mm, Biotite schist, Cornwall, U.K.

Plate 18. Precipitation and solution textures in hydrothermal quartz. This SEM-CL false colour image shows a zoned quartz crystal with euhedral concentric
growth zones shown by different colours disrupted by solution. The complex zonation patterns show variations in fluid chemistry during growth. (Chapter
3.9).
SEM-CL, 0.9 mm. Hydrothermal quartz vein cross-cutting the Ventersdorp Contact Reef, Klerksdorp goldfield, Witwatersrand, South Africa.
COLOR PLATE SECTION 141

Plate 19. Pressure fringe. Fibrous quartz overgrows euhedral faces of pyrite crystals. The quartz fibres are perpendicular to several faces, showing a typical
face-controlled geometry. (Chapter 3.9).
XPL, 1 mm, Quartz-sericite schist, Sabi mine, Zvishavane, Zimbabwe.

Plate 20. Blocky microvein filling. Blocky grains of calcite filling this microvein suggest a single opening and filling event. (Chapter 3.12).
XPL, 2 mm, Mylonite, Freda Rebecca mine, Bindura, Zimbabwe.

Plate 21. Syntaxial vein filling. The photomicrograph shows one side of a syntaxial quartz and calcite vein filling. A narrow band of partly fibrous quartz
lines the vein adjacent to the dark wall rock. Some fibres overgrow quartz grains in the wall rock, and the quartz grain size increases away from the wall,
showing that the quartz grew syntaxially. The calcite consists of curved but unstrained calcite fibres, which also coarsen away from the wall rock. The
calcite fibres grow towards a median suture which is parallel to the vein wall. A second group of fibres growing from the opposite side of the vein meets
the first group at the median suture, but the two groups are not continuous. The continuity between the vein quartz and the wall rock, the increase in grain
size away from the wall, and the joining of two separate groups of fibres at the median suture are all distinctive features of syntaxial growth. The curved but
undeformed nature of the fibres shows that the curvature is a primary growth feature. Since the fibres grow towards a median suture, wall rock control of
their orientation is unlikely, and the fibres probably tracked the incremental opening vector, which rotated with respect to the vein. See Fig. 3.17a. (Chapter
3.12).
XPL, 4.3 mm, Quartz-calcite vein, locality unknown.
142 COLOR PLATE SECTION

Plate 22. Non-systematic (stretched) fibres. Individual quartz fibres stretch all the way across the microvein. There is no symmetry or suture, and the sides
of the fibres have interlocking teeth. They are subdivided into tablets perpendicular to their length, each of which represents a growth increment. These are
the typical features of non-systematic growth histories. See fig. 3.17e. (Chapter 3.12).
XPL, 4 mm, Semi-psammite, Anglesey, U.K.

Plate 23. Tracking fibres. Undeformed biotite fibres grow obliquely across the microvein, and remain parallel through a small jog in the margin (centre).
These observations suggest that the fibres were tracking an incremental opening direction that was oblique to the margin, since the change in the margin
orientation at the jog has no effect on the fibre orientation. (Chapter 3.12).
XPL, 1 mm, Serpentinite, Sabi mine, Zvishavane, Zimbabwe.

Plate 24. Deformation twins in calcite. Characteristic features of deformation twins are variable thickness, pinching out, and bending of individual twins.
These twins are type 2 of Burkhard’s classification, and therefore formed between 150 and 300°C. Also see Fig. 4.11. (Chapters 4.3, 9.11.2).
XPL, 0.25 mm, Limestone, Mushandike, Zimbabwe.
COLOR PLATE SECTION 143

Plate 25. Grain boundary migration. Advanced stages of grain boundary migration. As well as highly convoluted grain boundaries, blue grains can be seen
entirely surrounded (in the plane of the section) by the yellow grain. Also see Figs. 4.17, 4.18. (Chapter 4.8).
XP, ST, 0.5 mm, Deformed granitic rock, Unknown locality.

Plate 26. A strong grain shape fabric defined by hornblende and biotite shows no evidence of internal strain features, and formed by solid state diffusive
mass transfer. The fabric in the mafic minerals has controlled the shape of the quartz grains, imparting a similar fabric. Also see Plate 27. (Chapter 5.3).
PPL, 4 mm, Hornblende schist, Makaha greenstone belt, Zimbabwe.

Plate 27. Grain boundary control in a two-phase aggregate. Quartz grain boundaries are controlled by muscovite grains. Boundaries between two quartz
grains are perpendicular to muscovite basal planes, and often pinned at the ends of the muscovite grains (centre). Both these effects are due to the higher
surface energies between quartz and mica than between quartz grains. The resultant quartz grains are elongate parallel to the muscovite fabric. Also see
Plate 26. (Chapter 5.3).
XP, 2 mm, Mica-garnet schist, Miami district, Zimbabwe.
144 COLOR PLATE SECTION

Plate 28. Decussate texture. Randomly orientated, interlocking, elongate actinolite crystals show a good example of this texture, produced by diffusive mass
transfer processes in crystals with anisotropic growth rates. (Chapter 5.4).
XPL, 2 mm, Biotite-actinolite schist, Arcturus mine, Zimbabwe

Plate 29. Snowball texture in garnet. Inclusions trails of quartz define an internal foliation that curves through more than 180°. The external foliation is
defined by a strong grain shape fabric in adjacent muscovite. See Fig. 7.18. (Chapters 5.6, 7.8).
XPL, 4 mm, Muscovite-garnet schist, Makaha greenstone belt, Zimbabwe.

Plate 30. Pretectonic porphyroblast. The andalusite porphyroblast contains randomly orientated opaque inclusions. An external fabric defined by biotite
wraps around the porphyroblast. This is the typical texture of pre-tectonic porphyroblasts (Fig. 5.3c). Asymmetric quartz pressure shadows (light areas)
occur on either side of the porphyroblast. See Plate 44 for another example from this rock. (Chapter 5.6).
PPL, 2 mm, Biotite-andalusite schist, Sharriva greenstone belt, Zimbabwe.
COLOR PLATE SECTION 145

Plate 31. Intertectonic porphyroblasts. The two biotite porphyroblasts contain an internal fabric of relatively large quartz grains. The external fabric
is continuous with but has a smaller grain size The strong fabric localized on the limbs of folds in and the folds themselves, are not present in
the biotite porphyroblasts. The porphyroblasts grew after was formed, but before reached its present grain size and was folded, showing the diagnostic
features of intertectonic porphyroblast (Fig. 5.3b). (Chapter 5.6).
PPL, 2 mm, Metapelite, Longman Mountains, Sichuan province, China.

Plate 32. Syntectonic porphyroblast. The external foliation is continuous with the internal foliation, which is curved within the staurolite porphyroblast.
This is the typical texture of a syntectonic porphyroblast (Fig. 5.3c). (Chapter 5.6).
XPL, 4 mm, Staurolite schist, Dindi greenstone belt, Zimbabwe.

Plate 33. Post-tectonic porphyroblast. The kyanite porphyroblast at centre grows across a strong biotite-muscovite fabric. The internal fabric and the
external fabric are continuous, and is undeformed around the porphyroblast, indicating post-tectonic growth (See Fig. 5.3d). (Chapter 5.6).
PPL, 2 mm, Kyanite staurolite schist, Reynolds range, Arunta block, Australia.
146 COLOR PLATE SECTION

Plate 34. Corona. The central hornblende grain is surrounded by an intergrowth of orthopyroxene, plagioclase and magnetite, a typical texture produced by
prograde reaction of hornblende. (Chapter 5.7).
PPL, 1 mm, Mafic granulite, Northern Marginal Zone, Chief Bota, Zimbabwe.

Plate 35. Symplectite. Intergrowth between orthopyroxene (clear) and plagioclase (clouded appearance). Vermicular intergrowth of the two phases occurs
because diffusion distances during the reaction were too small to allow an equilibrium microstructure
to form. (Chapter 5.7).
PPL, 2 mm, Mafic granulite, Datong-Huai’an, China.

Plate 36. Magmatic grain shape fabric. The three large euhedral plagioclase phenocrysts and several smaller ones define a grain shape fabric trending from
upper right to lower left. The phenocrysts are completely unstrained and set in a matrix of pyroxene. Undeformed igneous zoning can be seen in the largest
phenocryst. No evidence for deformation is seen in the hand specimen, which also shows the alignment of euhedral phenocrysts. The complete lack of
microstructural evidence for strain demonstrates that the fabric is magmatic. (Chapter 6.4).
XPL, 2 mm, Basalt, Bembezi river, Zimbabwe.
COLOR PLATE SECTION 147

Plate 37. Magmatic/submagmatic grain shape fabric. Euhedral plagioclase and biotite crystals define a grain shape fabric parallel to the top/bottom of the
plate. The only other evidence for deformation is slight undulatory extinction in quartz. Primary, undeformed zoning is visible in the dark plagioclase crystal
on the right. The biotite and feldspar crystals define a moderate grain shape fabric in the hand specimen. There is no evidence for recrystallization (e.g.
a quartz aggregate shape fabric). The deformation was either magmatic, with a later non-magmatic deformation that created undulatory extinction in the
quartz, or sub-magmatic. In either case, the fabric was created with melt present. (Chapter 6.5).
XPL, 4 mm, Tonalite, Fort Rixon, Zimbabwe.

Plate 38. Sub-magmatic fabric. The euhedral outline of the feldspar megacryst is parallel to a biotite and hornblende grain shape fabric. A euhedral feldspar
megacryst alignment is a conspicuous feature of the outcrop from which this specimen was collected. The feldspar megacryst is recrystallized along its
margins, and slight bending of the feldspar twins is visible. This granite is part of a suite that is syntectonic with a major deformation event. The shape fabric
and the intracrystalline deformation features can therefore be interpreted as submagmatic. (Chapter 6.5).
XPL, 2 mm, Razi Granite, Mavizhu, Zimbabwe.

Plate 39. Curved foliation. Foliation defined by muscovite lies at angle of about 45° to the upper edge of the plate. It curves smoothly to become parallel to
the lower edge of the plate, which is the orientation of the shear plane. The clockwise rotation shows a dextral sense of shear (Fig. 7.1). (Chapter 7.2).
XPL, 1 mm, Granite, Mushandike, Zimbabwe.
148 COLOR PLATE SECTION

Plate 40. fabric or extensional crenulation cleavage. A penetrative S-foliation from upper left to lower right is defined by hornblende grains. S is
deflected into a narrow inclined in the opposite direction. The sinistral shear sense is clear from the curvature of S into (See Figs. 7.12, 7.13).
(Chapter 7.5).
XP, 1 mm, Hornblende Schist, Unknown locality.

Plate 41. S-, C- and Three fabrics are visible. The pervasive S-foliation inclined from upper right to lower left is defined by muscovite and quartz
grain shapes and compositional banding. Discrete shears parallel to the top and bottom edges are C-surfaces. A few discrete shear surfaces are inclined to
the shear plane from upper left to lower right; these are surfaces. Both S-C porphyroclastic fabrics and banded fabrics and
are visible (see Fig. 7.11 for definitions of and The dextral shear sense is clearly given by the curvature of S towards C- and (See
Figs. 7.12, 7.13). (Chapter 7.5).
PPL, 1 mm, Mylonite in grey gneiss, Buhwa, Zimbabwe.

Plate 42. Mica fish. The lozenge shape of this single crystal of biotite is typical of mica fish. Short tails of biotite extend from the end of the fish, and appear
to have formed by microfracture along basal cleavage planes. The basal cleavage is parallel to the long axis of the fish. Recrystallization occurs along the
left and right margins of the fish. The shear sense is clear from the left-stepping tails, and the clockwise rotation from the long axis of the fish to the shear
plane (parallel to the top/bottom edges of the plate). Also see Plate 43. (Chapter 7.7).
XPL, 0.5 mm, Mylonite, Shamva greenstone belt, Zimbabwe.
COLOR PLATE SECTION 149

Plate 43. Hornblende fish. The single hornblende crystal has a similar geometry to the mica fish in the previous plate. The dextral shear sense is clear from
the stair stepping of the tails and the obliquity of the fish long axis to the shear plane. Also see Plate 42. (Chapter 7.7).
XPL, 2 mm, Granite mylonite, Zivuku, Zimbabwe.

Plate 44. Asymmetrical pressure shadow. This andalusite porphyroclast gives a dextral shear sense from the left-stepping tails. The clear areas are
pressure shadows of quartz which also step up to the left. Plate 30 shows another example from this rock. (Chapter 7.6).
PPL, 2 mm, Biotite schist, Shamva greenstone belt, Zimbabwe.

Plate 45. Matching of grains across a microfault. The use of the sensitive tint plate allows grains of similar colour to be matched across the microfault,
showing a sinistral separation of ~ 0.1 mm in this section. The microfault matrix contains fragments of different colours from the adjacent grains, suggesting
rotation or transport from other grains, as well as a quartz cement which is continuous with the wall grains (Fig. 7.23). (Chapter 7.13).
XPL, ST, 1 mm, Barrios quartzite, Cantabrian zone, Villamanin, Spain
150 COLOR PLATE SECTION

Plate 46. Multiple sets of PDFs in quartz. The central grain shows at least three sets of PDFs, visible from slight contrasts in optical properties. The typical
planar, sharp and parallel (within individual sets) geometries of PDFs contrasts with the non-planar appearance of deformation lamellae, shown in Fig. 4.13.
(Chapter 8.4).
XPL, 0.45 mm, Aruonga Impact crater, Chad.

Plate 47. Suevite. Two types of clast can be distinguished. The large clear clast on the left has prominent circular features which are ballen structure. The
right hand clast is feldspar. The dark matrix anastomoses around the clasts, defining a weak fabric. (Chapter 8.6).
PPL, 1 mm, Suevite, Bosumtwi Impact crater, Ghana.

Plate 48. Suevite. The same view as the previous plate in cross polarized light. The clear clasts are non-birefringent. The boundaries between individual
ballen are filled by a birefringent phase, probably a clay mineral, suggesting that the ballen structure formed in a similar way to perlitic texture by hydration
of a glass. Traces of albite twins can be seen in the feldspar clast, which has the mottled extinction characteristic of mosaicism. However, in view of the
evidence for alteration, the mosaicism can not be unambiguously attributed to shock. (Chapter 8.6).
XPL, 1 mm, Suevite, Bosumtwi Impact crater, Ghana.

You might also like