You are on page 1of 311
Ree ee Pee oe GAC | AGG Ra ee! A St ea Bieti: IEorccsl Se Be ee en ua A ‘The Geological Association of Canada ‘The GEOLOGICAL ASSOCIATION OF ‘CANADAis Canada's national society forthe ‘geosciences. It was established in 1947 10 ‘advance geology and its understanding ‘among both professionals and the general puptc. The GAC membership of 3.000 in- ‘ludes representatives of all geological cs ‘opines from across Canada and many pans Of the world employed in goverment. in- ‘ustry and academia. There we specialist ‘visions for environmental earth sciences, ‘geophysics, mineral deposits, paleontology. Precambran, Sedimentology, tectonics and vvolcanoiogy. Regional sections of GAC have ‘been set up in Victoria, Vancouver. Edmon- ton, Winnipeg and St John's, and there are ‘affiated oups in Toronto and the Mantmes. GAC actuities include the organization and ‘sponsorship of conferences, seminars, shor ‘courses. feld trips, lecture tours and student land professional awards and grants. The ‘Association publishes the quarterly journal Geoscience Canada and the quarterly news: leter Geolog, a Special Paper series. a Re- print Series, Short Course Notes. a membership lst: the sectons and civisions GAC also maintains liaison wih other C rnadian earth science organizations. names representatives to national and international scientific organizations, and provides advice to government and the public on geological Issues. The Association was incorporated tnder the Canada Corporation Act in Janu- ‘ay. 1984. For information contact: Geolog- ‘eal Associaton of Canada, Department of Earth Sciences, Memorial University of New- founiand, St. John's, Newfoundland AB 3X5, Canada. Telephone; (709) 737-7660 Association géologique du Canada ASSOCIATION GEOLOGIQUE DU CANADA est la société nationale canadienne pour les sciences Ge la Terre. Creee en 1947, elle a comme double ‘object de faire progresser a geologie et de sen- sSbiser les spécialisies et les membres du grand Dubie aux soences oe la Tere. Ses TOS mune membres repésentent utes les ciscplines gto- logques: is viennent de toutes les régions cu Caraca ‘et de nombreux autres pays: ils o@uvrent cans le ‘secteur public, dans le secteur industriel et dans le monde universitaire. LAssociation comprend des nisions de spécialistes en géophysiqus, en gise- ‘ments minéraux, en paléontologie, en Précaméren, ‘en sédimentologie, en tectonique. en volcancogie et en sciences de la Terre touchant & Femiron- rement. Des sections régionales existent a Vicon Vancouver, Edmonton, Winnipeg et St. Johns, et es groupes atTiés $e sont consttués a Toronto et Gans les provinces martimes, Les activites de Association géologique du Canada comprennent Torganisation et le parrainage de conferences, de colloques, de cours de formation de courte durée, de visites sur le lerrain et de tour- 1nées de conferences. Elle déceme des octois et des bourses aux étudiants et aux personnes qui ‘vavallent Gans le domaine des sciences dela Terre. LAssociaton pubie un joumal timestiel, Geo- ‘scence Canada, et un bulletin timestriel cinfor- ‘mation, Geolog, une série de mémoires, des notes de cours et diverses autres séries de publications. Ele assure en outrelalisison avec d autres socistes fen sciences de la Terre et fournt des conseils au {gouvernement et au grand pubic sur des questions: géologiques. L’Association a été constituée en cor- eration en janvier 1984, en vertu de la Lot sur les corporations canadiennes. Pour renseignements: Association géologique du Canada, Department of Earth Sciences, Menonal University of Newfoundland, St. John's, Terre-Newe Canada A1B 3X5. Telephone: (709) 737-7660. MINERALIZATION AND SHEAR ZONES Editor /Organizer: J.T. Bursnall Authors: J.T. Bursnall, C. Jay Hodgson, C. Hubert, R.W. Kerrich, P. Marquis, J.B. Murphy, I. Osmani, H. Poulsen, F. Robert, M. Sanborn-Bzrrie, G. Stott, and, H.R. Williams. Geological Association of Canada Short Course Notes Volume 6 Montréal, Québec May 12-14, 1989 © Copyright - Geological Association of Canada First printing — May, 1989 Reprinted — November, 1989 Reprinted — November, 1990 ‘The various chapters in this volume have been prepared primarily for short course participants at the 1989 Montréal GAC Annual Meeting. The volume has been reviewed within a panel of critical readers. All comments and requests should be forwarded to the authors, who retain responsibility for their specific contribution. Canadian Catalog Main entry under title: Mineralization and shear zones Publication Data (Geological Association of Canada short course notes; v.6) “The various chapters in this volume have been pre- pared primarily for short course participants at the 1989 Montréal GAC Annual Meeting.”’--cf. verso of t.p. Authorization to photocopy items for internal or personal use of specific clients, is granted by the Geological Association of Canada for libraries and other users registered with the Copyright Clearance Center (CCC) Transactional Reporting Service, provided that the base fee of $3.00 per copy is paid directly to CCC, 21 Congress Street, Salem MA 01970. US.A. $3.00 + 0.00 3 The above permission does not extend to other types of copying, such as copying for general distribution, for advertising or promotional purposes, for creating new collective works, or for resale. For such copying, arrangements must be made with the publisher. Additional copies may be obtained from: Geological Association of Canada Department of Earth Sciences, University of Newfoundland, St. John's Newfoundland Canada A1B 3X5 ISBN 0-919216-39-0 Printing: Love Printing Service, Ottawa. Includes bibliographical references. ISBN 0-919216-39-0 1, Ore deposits--Canada, 2. Geology, Economic-- Canada. 3. Mines and mineral resources--Canada. I Bursnall, J.T. (John Treharne), 1940- Il. Geological Association of Canada, Meeting (1989 : Montréal, Quebec). III. Series. QE35.M46 1989 553.4 €89-090461-8 L’Association géologique du Canada accorde Yautorisation de photocopier des documents pour un usage interne ou personnel, ou pour utilisation de clients particulier, aux biblioth@ques et autres utilisateurs inscrits au Copright Clearance Center (CCC) Transactional Reporting Service, la condition que le tarif de base 3 § par copie soit payé directement au Centre, au 21 Congress Street, Salem, MA 01970, US.A. 3,00 $ + 0.00. La présente permission ne s'applique pas & d'autres genres de reproduction, notamment la reproduction en vue d'une distribution générale, a des fins de publicité ou de promotion pour la création de nouveaux traveaux collectifs ou pour la revente. Dans ces cas, il faut prendre les dispositions qui simposent en communiquant avec l'éditeur. PREFACE Shear zones are commonly mineralized and have long been recognized as important sites for the precipitation of economically viable deposits of such mineral commodities as gold, silver, lead, tin, zinc, fluorite, copper, uranium and platinum (¢.g., McKinstry, 1948). Many of these structures are thought either to have pre-existed mineralization, subsequently acting as conduits for upward migrating fluids, or to be responsible for tectonically 'remobilizing’ a pre-existing ore deposit. Both of these processes are important, but it is only relatively recently that the full significance of the dynamic relationship between actively developing shear and the concomitant migration of fluids of appropriate composition has been recognized (e.g: Sibson, 1987, 1981, 1977; Kerrich et al., 1987); this is particularly true for vein mineralization complexes, for which there are abundant examples of a spatial-temporal relationship to shear zones. It is this aspect of mineralization that provided the motivation for this short course and is the central theme of this volume . One reason for the recognition of a genetic relationship between active deformation and mineralization has been the increase in interest in the dynamic (mechanical) and kinematic (translational) meaning of shear zone geomeiry and internal structure: see Ramsay (1980a) and Ramsay & Huber (1987) for reviews; another is the increased. realization of the importance of fluids in controlling rheological behaviour - from facilitating fracturing on the one hand to increasing rock ductility on the other (e.g: Fyfe et al., 1978; Rutter, 1972; Sibson, 1977). This research has produced a considerable increase in our understanding of why and how shear zones develop, which, in turn, has generated more detailed structural and other studies of shear zone hosted deposits (e.g., Robert et al., 1983); a critical component here, of course, is the renewed interest in gold deposit geology during the past decade - an interest that is strongly reflected in all the succeeding chapters of this volume. It has also generated an enormous literature. Roberts (1987) and Hodgson (in press) present recent reviews of lode gold deposits in a shear zone setting, and many deposit-specific descriptive studies are now available. With any major advance in science, as our knowledge improves so does the vocabulary, or ‘jargon’. There are good reasons for this (improved precision in communication and approach) and many times some bad results (confusion); the subject of shear zone studies is no exception. One of the subsidiary aims of iii the short-course, therefore, is to clarify some of the complex terminology that has resulted from the recent work - developed in our attempts to clarify what is a very complex subject: for example, "Cataclasites, hemiclasites, holoclasites, blastoditto and myloblastites - Cataclastic rocks" (Zeck, 1974). The primary aims of the course, however, are to introduce and to review fundamental concepts relating to the mineralization of shear zones, including: (®, the mechanical and chemical processes in the crust responsible for the development of the different types of shear zones, and (ii), the distribution, possible sources and mobility within the crust of the mineralizing fluids. These are related in studies of the structures within shear zones, their progressive development and patterns of mineralization (mainly Bursnall, Chapter 1; Murphy, Chapter 2; and Hodgson, Chapter 3); chemical and mineralogical changes associated with shear zones (mainly Murphy, Chapter 2; Kerrich, Chapters 4, and 5); sites of mineral precipitation, vein evolution and vein textures (mainly Hodgson, Chapter 3; Kerrich, Chapter 4; and Poulsen and Robert, Chapter 8); and reviews of current dynamic, kinematic and geochemical models for the localization of ore minerals in shear zones (e.g., Kerrich, Chapter 4). A number of deposit-scale examples are given and regional tectonic and structural appraisals of the Superior Province from a mineralization viewpoint are presented (e.g., Osmani et al., Chapter 6; Hubert and Marquis, Chapter 7, and Poulsen and Robert, Chapter 8), but the study is not restricted to that area. The emphasis is undoubtedly towards gold, resulting primarily from its over- whelming importance in the past few years (particularly within Archean terranes) and the interest of the authors, but the principles outlined in this volume are appropriate for other occurrences. The approach is, therefore, by no means comprehensive: the choices of examples of specific mine relationships are necessarily selective and have depended to a large extent on individual author's experience. Those deposits where the dynamic relationship between mineralization and shear zone development is somewhat ambiguous and where the temporal relationship of ore deposition and its concentration is unclear have for the most part been excluded: the emphasis, therefore, is on vein-hosted deposits rather than those where the ore has a granular and disseminated habit (e.g., Hemlo). It will be noticed that there is some overlap in the material contained in the first part of the volume (Chapters 1 through 4): this is in part accidental, resulting from the authors specific interests and the way the short course was set up (the editor accepts full responsibility), but also is in part intentional as a form of review as the text unwinds. References to common sections are incomplete, and so due scrutiny of the Table of Contents is recommended. Neither the short course nor this volume would have been possible without the help and encouragement of a large group of colleagues and friends. The editor and the authors would particularly like to thank the reviewers and others for their prompt response to urgent requests. The review panel included: M.Cherry, M. Downes, J. Fedorowich, J. Hamilton, H. Helmstaedt, R. King, AJ. Macdonald, G. McGill, R. McLeod, D. Nance, B.E. Nesbitt, E. G. Nisbet, C. Shrady, T. Stokes, C. Van Staal, J.Summers, and D. Wyman. Kathryn Baker, O. Grutiw, C.Hadubiak, Joan van Kralingen, and, Maria Maclsaac performed the valuable task of some of the drafting; and, K. Ansdell, B. Every,D. Wyman, R. Feng, A Hepner, and R. King are thanked for their assistance; the editor apologises for any omissions. Much of the research described in this volume was supported by NSERC operating grants. Grateful thanks go to a number of people for their good-humoured support during the preparation of this volume. The editor particularly wishes to thank Monica Easton of Geoscience Canada, B. Baragar of the GAC Publications Commitiee, and to Kevin Corrigan, Jamie O'Connel and staff of the Paper House Print Shop Amherst, Ma, a warm thanks for putting up with an amateur in their midst for a hectic four weeks. Finally, particular thanks to Catherine for her continued support throughout all stages of the project. (Note: References are within the combined references at the end of the volume) TABLE OF CONTENTS CHAPTER 1: REVIEW OF MECHANICAL PRINCIPLES, DEFORMATION: 11 12. 1.3. 13.1 132 Ess 14 141 14.2 14.2.1 143 by 15.1 15.1.1 152. 153 16 1.6.1 MECHANISMS AND SHEAR ZONE ROCKS Introduction Shear zones Crustal forces, deformation, and stress Introduction Stress Stresses acting across a plane Responses to stress I: examples from experimental work Introduction Short duration tests Structural anisotropy Long duration tests Responses to stress II: deformation mechanisms Brittle behaviour The role of pore fluid pressures Ductile behaviour The brittle-ductile transition Review of strain Introduction The strain ellipsoid The effects of simple shear Pure shear strains Shear zone rocks Introduction Brittle effects Ductile effects Discussion CHAPTER 2; TECTONIC ENVIRONMENT AND METAMORPHIC CHARACTERISTICS OF SHEAR ZONES 29 21 Introduction 29 22 Tectonic environment: extensional and compressional regimes 31 221 Introduction 31 2.2.2 Internal characteristics of ductile shear zones 31 pa eel Mineral lineations 31 2222 Related C and S fabrics 33 2235 Folds within ductile shear zones 33 2224 Rotation of rigid inclusions 34 22.25 Extension fracture in brittle-ductile shear zones 35 2226 Riedel shears 35 223 Extensional shear zones 35 223.1 Metamorphic core complexes 37 224 Compressional shear zones 39 23 Fluid chemistry and phase relationships 42 24 Metamorphic reactions, mineralization and brittle-ductile behaviour in shear zones 46 25 Summary and discussion: Metamorphic grade and style of deposit 48 CHAPTER 3: PATTERNS OF MINERALIZATION 51 31 Introduction 51 32 Stress and strain: some basic concepts 52 33 Shear zones and shear zone systems 54 331 Introduction 54 33.2 Classification of shear zones 54 33.3 Internal features of shear zones 56 3.3.4 External form of shear zones 57 335 Shear zone systems 57 34 Mineralization in, and related to, shear zones and folds 59 3.4.1 ‘Mineralization in shear and extension fractures 59 3.4.2 Mineralization in folds 63 3.43 Mineralization in systems of shear zones and related features 64 35 Form of mineralization 69 3.5.1 Classification 69 352 Layering 71 3. Breccia 75 354 Oriented mineral growth textures ve) 525 Replacement textures and disseminated mineralization 76 36 Spatial and temporal variations in mineralization 78 37 Mechanisms for generating dilatancy 79 371 Introduction 79 3.7.2 3.7.3 3.74 375 3.7.6 3.7.7 3.8 Opening of extension fractures Rotation of fractures formed at <45° to axis of maximum shortening Movement on non-planar shear surfaces 80 80 82 Variable external rotation of mechanically effective layering 82 Dilation on the convex side of fold hinges in competent layers 83 Dilation as a component of bulk inhomogeneous flattening summary CHAPTER 4, GEODYNAMIC SETTING AND HYDRAULIC REGIMES SHEAR ZONE HOSTED MESOTHERMAL GOLD DEPOSITS 89 41 42 421 422 423 424 425 425.1 4.2.5.2 426 4.2.6.1 42.6.2 43. 43.1 43.2 43.3 434 435 44 441 44.2 443 4.4.3.1 443.2 443.3 443.4 45 45.1 4.6. 47 Introduction and scope Geodynamic setting Monte Rosa lodes, Italian Alps Cordilleran mesothermal gold deposits Ballarat Slate Belt, S. Australia Phanerozoic geodynamic settings Superior Province Tectonic development Trondhjemitic and alkaline igneous rocks Summary of geodynamic setting Tectonic models Local controls on gold distribution The architecture of shear zones and Au distribution Introduction Distribution of gold: first and higher-order structures Controls on the brittle-ductile transition Evidence for transitions in rheological behaviour Implications of the brittle-ductile transition to gold precipitation Hydraulic conditions in shear zones “Introduction and boundary conditions Discussion of models Yellowknife shear zones type example Introduction Vein Set 1 Vein Set 2 Vein Set 3 Fluid generation and pressure in metamorphic belts Fluid volumes solute concentrations Real time analogs? Summary and conclusions 83 84 89 90 90 91 91 94 95 oF 97 100 100 102 103 103 103 107 108 110 112 112 113 117 117 117 118 118 119 120 122 123 CHAPTER 5: | GEOCHEMICAL EVIDENCE ON THE SOURCES OF 5.6.3 5.6.3.1 5.6.3.2 5.6.3.3 5.6.4.1 5.64.2 5.6.4.3 FLUIDS AND SOLUTES FOR SHEAR ZONE HOSTED MESOTHERMAL AU DEPOSITS Introduction and scope Geological characteristics of Au-Ag vein deposits Radiogenic isotope tracing Sr-isotope systematics Superior Province Norseman-Wiluna Belt Mother Lode Pb isotope systematics Superior Province Norseman-Wiluna Belt Summary Lithophile element and $/Se systematics Introduction Partitioning of the alkali and alkali earth metals Lithophile element systematics in gold deposits Magmatic-related ore deposits Massive base metal sulphide deposits Geothermal systems Lithophile element systematics and the origin of gold deposits S/Se fractionation Stable isotope systematics of oxygen and hydrogen Oxygen and hydrogen isotope systematics of silicates: Archean deposits Vein quartz-regional scale Intradeposit systematics Characteristics of the hydrothermal fluids Temperature and pressure H and O isotopic composition of fluids Isotopic relationships to host lithologies Other Archean gold deposits Au-Ag veins in high level felsic plutons Isotope systematics of carbonates Introduction C-isotope systematics of Archean deposits Governing parameters for C-isotope composition Carbon reservoirs Spilitized Archean volcanics Archean clastic and chemical sediments Source variables Variables during carbonate precipitation Interaction with graphitic rocks 129 129 130 133 133 133 136 136 133 137 138 139 139 139 140 142 144 145 145 146 146 149 149 149 150 152 152 153 155 156 159 160 160 161 161 162 162 164 166 168 169 5.92 5.9.3 5.94 5.10 5.10.1 5.10.1.1 5.10.1.2 5.10.1.3 5.10.2 5.10.21 5.10.22 5.11 Interpretation of C-isotope data 171 Intradeposit systematics 171 Carbonate dispersion halos 172 Interdeposit variations 173 Summary 174 Quartz-carbonate relationships 175 Sulphur isotopes 176 Proterozoic deposits 181 Homestake Mine 181 Churchill Province - Saskatchewan 181 Mesozoic and Cenozoic deposits 182 Foothills Metamorphic Belt and Mother Lode Districts, California 182 Cordilleran mesothermal deposits 183 Discussion of models for Cordilleran mesothermal deposits 184 Monte Rosa lodes, Western Alps 185 Discussion of alternative models 186 Mantle C0) - granulitization 186 LILE enrichment - depletion patterns 186 Carbon isotope data 189 Additional considerations 189 Magmatic models 191 Association with felsic stocks 191 Significance of ‘magmatic’ C-Isotope values 192 Concluding remarks 194 CHAPTER 6: RECOGNITION OF REGIONAL SHEAR ZONES IN SOUTH- CENTRAL AND NORTHWESTERN SUPERIOR PROVINCE OF ONTARIO AND THEIR ECONOMIC SIGNIFICANCE. agg. Introduction 199 South-central region 200 Regional shear zones 204 Big Duck - Killala Lake Deformation Zone (BD-KLSZ) 204 Quetico Fault Zone 204 Barton Bay Deformation Zone 205 Gravel River Fault 207 Pipestone-Cameron-Manitou Lakes Fault 207 Age and tectonic significance of shearing 207 Relationships between mineralization and shearing 208 Discussion 208 Northwestern Section 209 Regional shear-zones 209 Bear Head Fault Zone (BHEFZ) 210 Windigo Lake-Horseshoe Lake Shear Zone 212 63.14 Stull Lake-Wunnummin Lake Fault Zone (GSWFZ) 212 6.3.1.5 North and South Kenyon Fault Zones 213 6.3.1.6 Winisk River Fault (WRF) 14 6.3.1.7 Northeast-trending faults 215 6.3.1.8 Shear zones paralleling subprovince boundaries 215 64 Discussion 215 64.1 Greenstone belt - shear zone relationships 216 65 Conclusions 217 CHAPTER7 STRUCTURAL FRAMEWORK OF THE ABITIBI GREENSTONE BELT OF QUEBEC AND ITS IMPLICATIONS FOR MINERAL. EXPLORATION. 219 Part I: Structural framework within the belt (C. Hubert) 219 Introduction 219 The lozenge-shaped blocks 20 Deformation zones 223 Part IThe Dumagami Structural Zone: its evolution and features for the Doyon, Bousquet nos.1 and 2, and Donald J. LaRonde gold mines of southwestern Abitibi, Quebec. (P. Marquis and C. Hubert)226 Introduction 226 Structural features of the Dumagami Structural Zone 227 D1 structures (FI, S1, Ls) 227 D2 structures (F2, $2a, $2b, Li) 229 D3 structures (F3, $3) 231 Faults 231 Barren veins and fractures 232 Mineralized veins and fractures 232 Microscopic structures and mineral relationships 233 Structural features of the massive sulphide bodies 235 Microstructures 235 The brittle-ductile stage 237 . The brittle stage of the progressive deformation 238 24 Conclusions 238 CHAPTER 8: SHEAR ZONES AND GOLD PRACTICAL EXAMPLES FROM THE SOUTHERN CANADIAN SHIELD 29 81 Introduction 239 8.2 Regional context of shear zones and gold in the Southern Shield 239 8.2.1 District-scale patterns 243 8.22 Deposit-scale patterns 2413 83 Common characteristics of shear zones and veins 244 8.3.1 Shear zones 244 8.3.1.1 8.3.1.2 8.3.1.3 8.3.1.4 8.3.1.5 8.3.2 83.21 8.3.2.2 8.3.2.3 83.24 83.3 84 8.4.1 8.4.2 8.4.2.2 8.4.23 8.4.24 84.25 Nature of shear zones Geometric relations of foliation Secondary foliations Folding of foliation within shear zones Lineations Veins Shear veins Extensional veins Oblique-extension veins Breccia veins Plunge of ore shoots Complicating factors in practical shear zone analysis Rock anisotropy Shear zone reactivation and deformation Near-conjugate behaviour Non-conjugate behaviour Shear zone reactivation Deformation of shear zone Conclusions REFERENCES 244 247 249 249 249 250 251 253, 255 256 256 259 260 264 265 265 265 266 266 267 CHAPTER 1 INTRODUCTION: REVIEW OF MECHANICAL PRINCIPLES, DEFORMATION MECHANISMS AND SHEAR ZONE ROCKS J.T. Bursnall Department of Physics Geophysics Laboratory University of Toronto Toronto, Ontario MS5S 1A7 11 Introduction. This introductory chapter contains a number of components in its aim to ‘set the scene’ and to prime the reader for what follows. An introduction defining and describing the essential characteristics of shear zones precedes brief review sections on the way rocks behave mechanically and the nature of the strain patterns that occur within these zones. Subsequent sections describe shear zone rocks and the types of structures that are present within them. 12 Shear zones Shear zones are narrow zones of high deformation across which there has been significant zone parallel displacement. They are typically planar on the large scale, although localized variations in strike are characteristic and they commonly possess a length to width ratio greater than 5:1 (Ramsay and Huber, 1987). Some authors (e.g., Hobbs, Means, and Williams, 1976) distinguish shear zones from faults and describe shear zones simply as zones of "high ductile shear strain", that is, deformation zones within which there is no visible loss of cohesion across discrete structural breaks at the hand sample scale, whereas others describe shear zones as all instances of abrupt bulk displacement, ranging in character from purely, or dominantly, brittle (faults in the traditional sense) to purely ductile (Ramsay, 1980a). Unfortunately this has led to some confusion and the introduction of some seemingly meaningless (but nonetheless useful) expressions, in terms of traditional usage, such as ‘ductile fault’ (e.g. Mitra, 1978; Wise et al.,1984). Figure 1.1 illustrates the range of shear zone types, reflecting the dominant style of deformation that produced the zone: ductile shear zones demonstrate (a) (b) (c) N WN / N . \ \ . S \ Ny \ i “4 M Figure 1.1: Range in deformational style of shear zones (after Ramsay, 1980a): (a) brit (b) britile-ductile; and,(e) ductile. continuously changing deformation state from boundary to boundary, faults (brittle shear zones) exhibit separation across discrete discontinuities, and those shear zones that are intermediate in character are referred to as brittle-ductile shears and typically show a complex interplay between ductile and brittle elements (Ramsay,1980a); most vein-hosted deposits fall into the latter shear zone type. In detail, shear zones are typically very heterogeneous internally: structural discontinuities may intersect at low angles producing a ‘braided’ or anastomosing complex of shears that envelop lensoid domains of lower strain (Figure 1.2). Although pure shear (section 1.5) may be an important mechanism within a shear zone the predominant one is simple shear, that is, there is a dominant element of displacement parallel to the trend of the zone. Such displacement may range from microscopic to tens of kilometres. In general, predominantly brittle shears are relatively narrow when compared to their ductile counterparts, where the bulk strain is more widely distributed. Large-scale shear zones generated at moderate to high metamorphic grade, and deeper levels within the crust, are characteristically much broader than those at lower grade. (Hanmer and Connelly, 1986; Escher ét al., 1977; Sibson, 1977). Research trends within the past decade or so have perhaps leaned more towards the ductile aspects of shear zones and to ductile shear zones in general (Chapter 2) whereas previously the emphasis tended towards brittle aspecis. It is apparent, however, that most significant shear-zone hosted deposits are the result of rock failure through fracturing at all scales, even though there may have been a significant component of ductile strain that either preceded or was contemporaneous with localized brittle failure: Sibson (1988, 1989) and Robert and Brown (1986) have emphasised the important observation of repeated Figure 1.2: Examples of shear zones: (a) ductile, lines represent trend of shear zone foliation (modified from Coward, 1976); and (b), brittle failure of experimental test sample, lines represent microshears and extension fractures (after Borg and Hardin, 1966) (b) t t t alternating ductile and brittle response evidenced by the structures in many mineralized shear zones. By far the most common concentration of economically significant shear- zone hosted material occurs in veins. Although some vein-like structures may be replacive (Park and MacDiamid, 1964) many result from the filling of pre- existing voids in the rock by introduced fluids and gases or from the progressive filling by accretion of material as the opening in the rock develops (Ramsay, 1980b; Chapters 3, 4). Most vein-hosted ore bodies, therefore, result from brittle processes even though there may be significant ductile behaviour and permanent internal strain to the surrounding rock. Recent examples of work relating to vein geometry and origin are presented by Beach (1975), Pollard et al. (1982), Ramsay (1980a,b), and Yardley (1983). The fluids responsible for mineralization within actively deforming shears may be concentrated in a variety of ways, but most typically seem to be localized within broadly lensoid dilatant areas situated sporadically along the deformation zone. The resulting vein complexes are rarely structurally simple in detail, but may consist of a single array of en échelon gashes or even isolated veins; more commonly, the overall vein geometry is extremely complex being comprised of numerous intersecting veins. The sporadic nature of these areas, although annoying to the exploration geologist, is likely of critical importance to our understanding of how, why, and where mineralization occurs. (See later sections and Sibson (1989) for review). With increased ductility, and in the absence of a later period of vein- generating brittle behaviour, mineralization may be less concentrated and exhibit a dispersed granular habit. (¢,g., the Hemlo gold occurrence). Such disseminated occurrences of ore minerals are by no means restricted to the shear zone environment since they may occur as replacement of a non-dilatant host (Roberts, 1987). Pervasive microfracturing during shearing (fracturing invisible at hand sample scale) could also facilitate this mineralization habit. It is important to recognize that such a deformation mechanism can result in bulk ductile behaviour of the rock body and, therefore, the classification of shear zones into brittle and ductile types is, in part at least, scale dependent (see below). In many cases, therefore, it is difficult to recognise not only the mechanism of deformation but also the relative contribution of active shear strain to the development of concentrations of economic values. The spatial association of mineralization and shear zone, for example, may be fortuitous and may be due to some other cause - such as the superimposition of shearing on a pre-existing ore deposit. 13. Crustal forces, deformation, and stress 13.1 Introduction In order to understand the nature of rock deformation and why and how shear zones develop it is necessary to know something of the crustal forces resulting in strain, and the controlling factors to the mechanical response of rocks undergoing shear. What follows are brief reviews of basic concepts, including some definition of terms. A force may be described as that which can change the state of rest (relative position) of any material body. The important forces that exist in crustal rocks result from gravitational attraction and those generated by the dynamic inter- action between different rock volumes; the latter are driven fundamentally by thermal activity and may be on all scales from continental collisions, and other plate interactions, to the differential thermal expansion of mineral grains. Forces may be divided into body forces and surface forces: body forces are those like the gravitational force that relate to distance and the amount and nature of the material. The gravitational force of a column of rock will act with increasing force across surfaces at depth: the ratio of the surface forces per unit area is lithostatic pressure , being measured as the product of the density of overlying material, the distance to the surface, and the gravitational attraction: Pith = P-hg (1) where p is density, h is the height of the column, and g is the attraction due to gravity on a column of rock. The ability of an applied force to cause change depends largely on the area over which it is applied. A means of comparing the potential relative effectiveness of forces is therefore to relate them to area across which-they are effective. A force (F) related to the area across which it acts is known as stress o=F/a (1.2) where a is the surface area and o the stress. All rocks will deform if sufficiently high stresses are applied. Deformation is a general term used to describe the response of the rock to these and involves Kinematic activity such as translation (moving from one place of reference to another), bulk rotation (changing orientation of the body), and internal distortion of the rock body (or, strain). All are present during the development of shear zones. Rocks are said to behave in a brittle way if they readily fracture and lose internal cohesion, and in a ductile way if they are able to sustain relatively large amounts of internal permanent distortion (strain) without gross fracturing (see Figures 1.4 and 1.6). Ductility is therefore somewhat dependent on the scale of the observation: at thin-section scale the distortion may be seen as the product of slip across numerous discrete discontinuities whereas at hand-specimen scale these microfractures are invisible and no gross fracturing therefore is evident. Rock distortion may be elastic or result from some form of flow..Elastic strain is temporary and wholly recoverable whilst ductile flow is permanent: some portion of the elastic strain may be dissipated over time through intracrystalline modifications. Under most conditions and natural rates of strain there is always some significant component of recoverable elastic strain within the body during its deformation. The release of this elastic strain energy resulting from the rapid reduction of stresses during failure generates the seismic response; those shear zones, therefore, that result from brittle behavior to any degree may be termed ‘seismic’ and those that result from purely ductile processes and the dissipation of strain through internal crystalline modifications (see below) may be considered as ‘aseismic’. 132 Stress The following is a brief review of some fundamental concepts. A more complete treatment will be found in texts by, for example, Hobbs et al. (1976), Means (1976), Ramsay and Huber (1987), and Suppe (1986). Pressure is a condition where the stresses, acting at a point are equal in all directions, a condition that may result in a change of volume (volumetric strain) but will not result in a shape change in compositionally homogeneous and structurally isotropic material (i., rocks where grains are randomly oriented, if inequidimensional, and no dominant preferred material alignments are present - such as subparallel grain boundaries, grain aggregates, and fractures, or compositional layering). The effect of non-random orientations (structural anisotropies) can be very important in controlling mechanical behaviour and affecting the orientation of structures such as fractures and veins. See below and Chapters 3 and 4. Where stresses are unequal then they may be resolved into three orthogonal principal directions for three-dimensional, or two principal stresses for two- dimensional, analysis.They may be compressional (considered here positive) or tensional (considered here negative); under conditions of general compression, G1 > 62 > 63. Under conditions of hydrostatic stress no principal stresses can be defined but for some purposes it is useful to consider the condition as one where 61 = 02=03. The average, or mean stress (also referred to as the hydrostatic stress) is a useful concept since it is the base from which shape changing stresses deviate: mean stress (6) = 6, +6) +03/3 (13) deviatoric stress =0-3 (1.4) Differential stress is the difference between the maximum and minimum principal stresses: differential stress = 01 - 03 (15) In most structural geology texts compressional stresses are referred to as positive (+ve) and tensional stresses as negative (-ve). For mathematical tment employing tensors the opposite is required - a practice that is followed n engineering and some recent texts in structural geology (e.g., Ramsay and uber, 1987), a somewhat confusing situation to the uninitiated! In this volume follow the traditional practice in the earth sciences. 132.1 Stresses acting across a plane it is not the purpose of this chapter to review the derivation of principal stresses fing at a point or the resolution of stresses acting on an inclined plane: these are adequately covered in the texts mentioned above. It is sufficient to say that the resolution of plane (biaxial) stresses (two principal stresses lying in the same plane) is far simpler than the triaxial or three-dimensional stress state. Such plane stress analysis is quite adequate to explain the basic fundamentals of the relationship between shear zones and the principal stresses; it is, however, a gross oversimplification of the natural, three-dimensional state. Reches (1978, 1983) and Reches and Dieterich (1983) describe the effect of three-dimensional strain fields on faulting (Chapter 3). It is convenient to resolve stresses acting across a plane into a normal stress (Gn) and those acting along the surface of the plane, the sliding, or shear stress (os or, 2). From Figure 1.3 it can be visually determined that the angle of inclination of the principal stress (o)) to the plane is critical in controlling the relative values Of 6, and o>. Os-8-Oagt (b) Figure 13: (a) Stresses acting across a plane; (b) Mohr's construction for stress Where the maximum principal compressive stress is perpendicular or parallel to the plane then no shear stresses act across it and the maximum shear stress is attained when the angle between the principal stresses and the plane are 45°. Values for On and os may be found from the following equations provided that the angle between oj and the plane (6) is known: Gp = (61 + 03)/2 + [(G} - 63)/2].cos 26 (1.6) and, 05 = [(o1- 3)/2]-sin 26 (17) These are expressed in this form to satisfy Mohr's construction which is purely a graphical solution for plane stress (Figure.1.3b). So far, we have been concerned with forces and stresses acting across defined or specific surfaces. In previously unfractured material, however, the deviatoric stresses must be sufficient to overcome the inherent strength properties of the material. This and other criteria for brittle behaviour are discussed in a following section. 14 Responses to stress I: mechanical behaviour of rocks from experimental work 14.1 Introduction In laboratory stress-strain tests the specimen will typically show significant elastic strain (Figure 1.4) that is followed (above the elastic limit) by either a period of plastic flow followed by brittle failure, or by brittle failure directly. Although laboratory tests are at strain rates much higher than most natural conditions and there are problems of scaling their results can at least supply a model for rock deformation and lead to some very general approximations of natural behaviour: in particular the failure fracture pattern and distortional behaviour of specimens is instructive (see summary diagram in Figure 1.6). From experimental rock deformation of homogeneous and isotropic previously unfractured material it is apparent that the intermediate principal stress (62) lies within the plane of a developing shear fracture and consequently the maximum and minimum principal stresses lie in a plane perpendicular to the fracture: slip across the fracture (fault) occurs in this plane. The orientation of the fracture to 61 is related in a systematic way to the cohesive properties of the rock. This ideal relationship is well known and is reflected in Anderson's (1951) dynamic theory of faulting (Figure 1.8): see also Griggs and Handin’s (1960) summary diagram of experimental behaviour (Figure 1.6). It is important to note here that whereas brittle and semi-brittle faults such as those described by Anderson from the coalfields of Britain will develop at angles of less than 45° to the principal compression, ductile shear zones typically develop at higher angles (Ramsay 1980). Pre-existing structure can have a profound affect on the orientation of developing fractures (see below; Chapters 3, 4, and 8) (a) _ultimate strength 4-03 yield Strength at elastic limit I~ elastic behaviour strain a %3 o3 102MPa P= pore fluid pressure 2 6 10 2 6 10 % strain % strain Figure 14: Summary of high strain rate experimental rock behaviour: plots are strain (¢) against differential stress (¢) (a) General curve identifying characteristic behaviour; (b) Temperature variation on basalt in compression and 5kb confining pressure (Griggs ef al., 1960); (c) Effect of pressure variation on polycrystalline galena (Atkinson in Price, 1975) (A) Pore fluid pressure (Rutter, 1972). 10 MPa = 1 bar 14.2 Short duration triaxial tests From Figure 1.4 it will be noticed that increases in ambient temperature result in a lowering of strength and an increase in the amount of longitudinal shortening the specimen can sustain before failure occurs (increase in ductility). Increases in confining pressure result in increases in strength and an increase in ductility (Figure 1.4c). That is, in general, and for rocks of a narrow compositional 10 range and at a constant strain rate, the deeper the crustal level the more ductile the material will behave: well supported from field evidence in metamorphic terranes. Similar curves for strain rate indicate that as strain rate is decreased the resistance to strain is decreased and that failure occurs earlier at higher strain rates: in other words high strain rates tend to facilitate brittle behaviour whereas low strain rates favour greater internal distortion. Under any of these variable conditions differences in rock composition will produce distinctly different behaviour. Rocks that exhibit ductile shape changes compared to adjacent material that has undergone the same deformation conditions are referred to as being incompetent. Competent material tends to maintain its original dimensions, or fails brittly within a less competent envelope. 1.4.2.1 Structural anisotropy: Tests carried out on the strength and fracture behaviour of material possessing a strong planar anisotropy such as cleavage are instructive. Donath (1961,1970) demonstrated that such foliated rocks have a much lower strength (differential stress at failure) (Figure 1.5) and that the orientation of fractures was controlled to some extent by the inclination of the foliation to 01. Similarly, earlier work by Griggs et al. (1951) described the effects of planar anisotropies in the Yule marble under compression and under extension. Such mechanical variation according to orientation of compositional and structural anisotropies are obviously of great importance in controlling rock behaviour in shear zones and are discussed further in following chapters (3, 4. and 8) nt ts 143 Long durati Relatively long duration tests that measure strain against time under constant stress and which attempt to approach natural strain rates are known as creep tests. Creep is the slow distortional response to long applied stresses. From creep experiments it is observed that the total shape change is part permanent and part recoverable (Figure.1.7). Permanent distortion is accommodated primarily by intracrystalline deformation and exhibited by all rocks at low strain rates. Creep occurs at stress levels that are significantly lower than the instantaneous failure load. Note that complete, but in part delayed, elastic recovery occurs up to a certain point - within the period of primary or transient creep, where strain rate decreases with time; following this constant strain rate is attained and increasing levels of permanent strain occur with time (period of secondary or steady-state creep). The specimen exhibits ‘pseudo-viscous flow’ and is analogous to the typically small amount of natural rock creep (in addition to the typically more important elastic strain) against a shallow fault boundary prior to slip. If the stress levels are maintained then strain rate rapidly increases ultimately ending in failure (tertiary or accelerating creep, Figure 1.7). Hobbs et al. (1976) point out that tertiary creep is commonly associated with fracturing of the test sample but that in some instances rapidly accelerating strain rates may be initiated by 11 (a) 1 = A A 6000 a c o D Kg/em2 2000 Wir - extension V compression 19 % strain (b) , ° d \ O34 NY —o3 c) ‘ { t t mT 3000. 350 bars |) ‘1 fy 105 { al eet ie °3 t unornamented areas 0 indicate range in 30 0S~«S fracture orientation Inclination of cleavage Figure15: The effects of pre-existing planar anisotropy on rock strength and fracture orientation: (a) Deformation of Yule marble at 10,000 Kg/cm? (modified from Griggs ef al., 1951), Diagrams above indicate orientation of foliation to principal stresses. (b) Breaking strength against cleavage orientation (after Donath, 1961). (© Effect of cleavage orientation on shear fracture orientation (after Donath, 1961) v typical strain t belore a 1-5 2-8 5-10 >10 racturing Sel compressive 94199203 extension 03<0. typical stress — strain] curves Figure 1.6: Test deformation behaviour illustrating the transition from perfectly brittle to perfectly ductile behaviour. (After Griggs and Handin, 1960) recrystallisation of the deforming material - such strain softening may favour the localization of ductile to brittle-ductile shear zones (White ef al., 1980) Experimental creep behaviour is seen as providing a crude model for rock deformation under lower natural strain rates and may be modeled fairly simply using mechanical analogues (Price, 1966, 1975) (1.4). 15 Responses to stress II: deformation mechanisms 45. Brittle behaviour Brittle failure of test specimens under triaxial testing may occur through pure extension fractures that are parallel to the main compression axis (61) or varieties of shear fractures oriented at some angle to o1, which may occur as conjugate pairs with one orientation dominant, and with o2 lying parallel to their intersection; or failure may result as a combination of these (Figures 1.5, 1.6). 13 Secondary cep Primary creep STRAIN Terticry creep Constant stress J Permanent deformation TIME—> Figure 17: Typical form of time-dependent creep test. Note significant component of elastic strain and effects of strain softening towards end of experiment (after Price, 1975). (a) , (b) 2 ‘ = \ a WZ Ss = +03 S os = 4] e277 = et 6, O34 1 Figure 1.8; Anderson's (1951) dynamic theory of faulting: (a) normal, (b) reverse, and (¢) wrench faults A number of criteria that attempt to explain the stress conditions that control failure have been proposed. Of these, only two seem to adequately predict the relationship between principal stresses at failure and are known as the Navier- Coulomb, Coulomb-Mohr, or Coulomb criterion and the Griffith theory. The Navier-Coulomb criterion states that the shear stress (6s, or t) required for failure along a shear fracture in previously un sheared material must ‘overcome the cohesive strength of the material (S) and the frictional resistance 14 to movement (the product of the normal stress, On) and the coefficient of internal friction, 1j). 6s =S + Onli (1.8) This is the equation for a straight line and when plotted on the Mohr construction for stress (Figure 1.3) separates a field of stability, where os is insufficient to cause failure, from a field that cannot be attained - since the tequired shear stress for failure must have been reached (somewhat erroneously referred to as the unstable field by some authors). The angle that the shear fractures make with the principal stresses is also predicted (always provided that the material is homogeneous and isotropic, a somewhat rare condition). When the principal stress difference is sufficiently large so that the Navier-Coulomb failure line is tangent to Mohr's circle then the conditions for failure are satisfied, but only for a plane whose normal to oj is 26 (see Figure 1.3). Although the Navier-Coulomb relationship approximates the behaviour of many materials reasonably well in the compressive field, for most experimentally determined values of the ‘friction angle’, @ (< 45°: see Figure 1.3), it tends to exaggerate the tensile strength - to the extent that it commonly predicts a greater tensile strength than the cohesive strength (experimentally determined tensile strengths are approximately half the cohesive strength). The failure criterion in he tensile field may follow a curve as indicated in Figure 1.9. Similarly the criterion does not allow for changing values of p; resulting perhaps from greater grain boundary contact as On increases. Such behaviour is supported by experimental determinations of shear strength for a single material at varying differential stress levels where the constructed failure envelope is typically curved, and is better approximated by the Griffith criterion stated as follows: 0,2 + 4To, - 4T2=0 (1.9) where T is the tensile strength, or a modified Griffith criterion for the compressive field by Brace (1960) and McClintock and Walsh (1962). Os =2T + Only (1.10) where Hg is the coefficient of sliding friction 2T is equivalent to S of the Navier- Coulomb criterion. It is important to recognise that all these criteria, however, are over-simplifications, empirical or mathemetical approximations of some poorly understood but complex relationships. The Griffith criterion and its modifications are theoretically based on the behaviour of microflaws and voids in the rock and in particular the tensile stress concentrations developed around microcrack terminations preferentially oriented to the maximum shortening direction. It predicts that certain orientations will propogate rapidly and others will close (Griffith, 1920). There will be preference for crack growth as extension fractures (i.e. parallel to 6,) and the combination and intersection of these may lead to the development of sample-scale shear fractures. An important aspect of these fracture criteria is that predict that 61 will bisect the acute angle between conjugate shears. The criterion is also able to demonstrate how the differential stress (61- 63) is fo control fracture orientation (Figure 1.9) The role of pore fluid pressures: igure 14d is can be seen that higher pore fluid pressure (Prryia) can icantly lower rock strength. The pressure exerted by fluids within cracks and pore spaces may greatly influence failure levels by: (i) aiding Propogation through increasing the crack opening forces and, (ii) through ing the frictional resistance between grains and other surfaces in the rock by ively reducing the normal stress acting across them (Figure 1.3a). Terzaghi 3) suggested that the shearing resistance of water saturated soils may be ed by the modified Navier-Coulomb criterion, os=S+(On-p)ti aap ‘is relationship satisfies the behavior of rocks as well, given the modifications outlined above (Handin 1968). value (dy - p) is referred to as the effective normal stress (6,') and results in a m in the level of shear stress (Gs) required to overcome the resistance to ective principal stresses may be similarly stated. For example, oy = 01-p,and (1.12) 63'=03-p, (1.13) sh the relationship may be more complex than this (Garg and Nur, 1973). can therefore be seen how fracturing may be induced from a condition of by an increase in fluid pressure (Figure 1.9). This or related mechanisms likely to be extremely important in the generation of mineralised vein es. (See also discussion in Chapter 4.) ressure (Pflyig) May be increased in a number of ways. These are in Ramsay and Huber (1987) and include: rapid sediment burial, particularly where the sedimentary sequence includes meable shales and evaporites; onic reduction of pore space and resulting fluid transport to cause high sures elsewhere; ) rapid agitation of unconsolidated material resulting in pore volume ction as closer grain packing occurs; aquathermal pressuring, when a fluid containing dissolved material is d such as might occur in the aureole of igneous intrusions; mineral phase changes and prograde metamorphic reactions that generate and CO) (Chapter 2); and, 15 16 (vi) the release of fluids into the surrounding rock during magmatic crystallization. Os Figure 1.9: Mohr diagram illustrating the effects of Pfluid and differential stress in controlling the development of extension fractures at depth. Case I: High Pfluid will result in * shear fractures, demonstrated by a shift in the shift of the plot to the left (20 = ~120°); Case TI will result in extension fractures (20 = ~180°); and, Case III will result in no fracturing even though biaxial tension occurs. (Modified after Hobbs et dl, 1976). The concept of high Puig to explain the coherence of large, well-travelled thrust plates is well known from the work of Hubbert and Rubey (1959) but of possible great significance to the development of mineralised brittle and brittle- ductile shears is the relationship of fluid migration to episodic seismic activity (Sibson et al., 1975, 1987; Kerrich et al., 1987; Chapter 4) When shear stresses build up along active faults the wall rocks may dilate through the opening of microcracks, grain boundary sliding, and the growth of extension veins. This can significantly reduce pore fluid pressure locally, particularly if permeability is low in the surrounding material, and this in tum results in increased normal stresses and greater resistance to sliding. As fluids continue to migrate into the wall rocks fluid pressures rise and movement on the fault will occur when the effective normal stress is reduced to a level that induces failure (see Figure 1.9). Rapid upward migration of fluids away from the failure zone will occur from the abrupt relief of elastic strain at failure and associated partial collapse of pore spaces. This process, when repeated, is referred to as seismic pumping (Sibson et al., 1975) and provides a mechanism for the episodic movement of great volumes of fluids towards higher levels in fault zones. It may produce localized brittle failure throughhydraulic fracturing, initiated under conditions where Paid exceeds the minimum confining stress (03) and the tensile strength of the rock (T) Paid > 03 + T (1.14) ay (1980b), and Ramsay and Huber (1987) argue that seismic pumping ount for the progressive and episodic filling of veins, the process of ated fracturing and precipitation of a new increment of material being to as the crack-seal mechanism (Ramsay, 1980b).The orientation of sion veins (Chapters 3, 4, and 8) will be perpendicular to the minimum s direction provided sufficient differential stress levels exists and may be led at lower levels by rock anisotropies (Chapter 4). It is important to e that extension veins of this type occur in a different kinematic lent to shear veins (those that are generated parallel and within shear Ductile behaviour e two principal mechanisms for permanent major shape change of bodies without the development of through-going fracture systems or cale diffusion. These are the processes of cataclastic flow and crystal y. The former results from slip across microfractures within and at grain ries and can be induced through high pore fluid pressures (Figure 1.6), al plasticity results from intracrystalline distortional processes such as islocations and minor diffusion (translation gliding, twin gliding, etc.). These two processes are related since in order to deform a neralic aggregate without the separation of grain boundaries at least five ystalline slip directions are required (the Von Mises criterion): if t or inappropriately oriented slip directions exist then grain boundary producing voids and cracks may occur, leading to cataclasis on,1975). Higher temperatures than those requiring cataclastic flow or tic processes would favour significant grain boundary sliding, which erate voids that might be an important vehicle for the migration of hic reaction produced fluids. At higher temperatures still (‘ho!- ) diffusion is of major significance and recrystallisations will occur ithe continued nucleation and growth of new strain free grains and so nodate shape change of the rock body (syntectonic or dynamic sation). that are strained at lower temperatures during rock distortion may e that strain by the general processes of recovery which involves the ment of smalier strain free subgrains (polygonisation). Such lower e, ‘cold working’, strain may also be eradicated by subsequent elevated res that can produce post-deformation annealing recrystallisations. lay between rates of deformation and the strain reducing mechanisms ve (strain rate v 'recovery'; see Wise et al. (1984) ) are most impor- trolling the textural and internal structure of rocks affected by shears if 18 (see discussion below): they may be viewed as controlling whether or not brittle or ductile behaviour dominates the deformation. A process that has become increasingly recognised as an important deformation mechanism in the past few years is pressure solution. This produces shape change (bulk strain) by preferential dissolution and removal of material along characteristically curviplanar seams, and forms an important cleavage type in rocks of medium to low grade. Solution transfer of this type has been described by Beach (1977) where material removed by pressure solution supplied the filling for adjacent or closely situated veins within en échelon vein arrays. Strain softening, or deformation softening, occurs where a weakening of the material results from the grain reduction resulting from polygonisation and recrystallisation. Strain-induced weakening across grain boundaries can result in ‘runaway’ instability and may be one of the reasons why planar and localised zones of high ductile strain develop. The principal mechanisms may be identified as: (® grain size reduction during deformation (eg., mylonitization), (ii) mineral phase changes, (iii) development of planar anisotropies, and (iv) localization of fluids (Cobbold and Gapais,1987; Poirier,1980; White et al., 1980). gouge - breccia pseudotachylyte, if dry incohesives: cohesives cools silts. EE pseudotach. REGIME if dry 10-15 km 250 - 350°C ” eohedivetiitited! rod ap cohesive foliated rocks mmylonite series REGIME Fault zone broadens with depth Figure 1.10: Variation of behaviour with depth within major crustal shear systems (after Sibson, 1977, 1989) The brittle-ductile transition experimentally controlled observed transition between purely ductile and brittle behaviour is, in the crust, likely to be a complex zone within which rittle and ductile structures occur according to rock grain size, composition, fluid pressures, orientation of anisotropy, strain rate, etc. Solely spatial ition of ductile and brittle structures does not necessarily mark this : a temporal relationship has to be demonstrated also. on (1977, 1989) has suggested a three-fold division of behaviour within ar zones related to crustal depth (Figure 1.10). Within the upper 10 kms or so ttle features dominate in a seismogenic zone where unstable frictional g (FR) occurs. This is succeeded by a transitional zone, perhaps within a erature range of 250 to 300°C, separating it from deeper levels where quasi- (QP) deformation processes predominate. Resistance to shearing reaches a around this depth, generating the larger magnitude earthquakes. The on marks the onset of conditions favouring crystal plastic processes g the development of mylonites) and is punctuated by transient brittle e. The position of this zone will be strongly controlled by the geotherm and over limited distance according to the local crustal conditions. Review of strain Introduction Progressive strain history within semi-brittle and ductile shear zones is be extremely complex and much has been written of it during the past s. (For reviews see, for example: Ramsay and Graham, 1970; Ramsay, d Ramsay and Huber, 1983,1987). The terminology is also complex and lows is an attempt to simplify as much as possible without, hopefully, ing any misconceptions. may be viewed as the distortional response to stress and for this to ere must be particle movement and change within the deforming body. to state that a rock body is strained we must have some knowledge of its ype; and in order to say something about the strain process we need to ish between strain as measured by a change in dimension, or inal strain, and shear strain which results from deflections and changes agular relationship (Fig.1.11). There are relatively simple strain histories the orientation of all the progressive strain increments do not change me as the deformation proceeds (‘pure shear’ strain, or irrotational strain) fe are those where significant change in the orientation of strain axes oughout the deformation (rotational strain, and a special case known e shear'strain, described below). Rotational strains are most likely the case and certainly this is true in the case of shear zones. 19 1.62 The strain ellipsoid One of the simplest tools for viewing strain is to assume a spherical form for the unstrained state and then to shrink or expand the sphere to represent volumetric strain or to distort the sphere into an ellipsoid in order to represent a general strain (Figure 1.11). Such strain ellipsoids (Flinn, 1956; Ramsay, 1967) are not purely imaginary since many geological bodies contain essentially spherical elements that can be subsequently deformed: for example, reduction spots in mudstones, ooliths in limestone, and, maybe, some cobbles in a conglomerate or vesicles in lavas. The local strain for such rocks in the deformed state may be seen in the resulting material ellipsoids and the strain for these (but not necessarily the rock that contains them) is measured directly as the ratio of the principal axes of the ellipsoid. However, the quantification of strain in this manner does not tell us anything of the increments of the strain history, but only the state of the finite strain, the end product. If such originally spherical markers are absent then other strain indicators may allow us to approximate the orientation and even the shape of the strain ellipsoid (see Ramsay and Huber, 1983, for a comprehensive treatment and bibliography). ) Pure Shear x principal shortening direction Simple Shear —- (b) x : CA, , z XYZ Z Figure 1.11: (a) Pure and simple shear; (b) The strain ellipsoid.(In part after Ramsay, 1967) There are two kinematic systems for strain. These result from ‘pure shear’ and ‘simple shear’ (Figure 1.11) The resulting distortion, if it is known to have been derived by these mechanisms, is referred to as pure shear strain (irrotational strain or ‘flattening’) and simple shear strain (a type of rotational strain). The strain may be the same in orientation and amount throughout the domain in question and at the scale of the observation - in which case it is homogeneous, or, it may vary from place to place and is inhomogeous, or heterogeneous . Since shear zones by definition are localised then there must be a significant strain gradient from the unstrained boundary rocks to the highest levels of strain within the shear zone interior: shear zones are therefore necessarily inhomogeneous but there may be volumes within them that are essentially homogeneously strained. 163 The effects of simple shear ‘An important aspect of simple shear is that the resulting strain ellipsoid is one of plane strain (Figure 1.11) provided that there is no volume change or flattening and that, being rotational, the incremental strain axes differ in orientation to the finite strain axes, as illustrated in Figure 1.12. At low levels of shear strain under Zinc. X Xincrementat (a) learly! (b) \ Xtinite Y incremental (latest) Figure 1.12: Rotational strain in simple shear: (a) Incremental and finite strains illustrated by the strain cllipsoid; (b) The development of en echelon extension gashes (veins). (after Ramsay and Kligfied, 1983) wv v v brittle-ductile conditions this relationship may be illustrated by observing the behaviour of en échelon gashes during progressive simple shear. At each increment of infinitesimal strain the strain axes X and Z will be orientated at 45° to the shear direction as reflected in the orientation of extension veinlets (Figure 1.12). These will subsequently rotate as shear continues but increments of growth at the crack tips where tensile stresses will be extremely large (Griffith crack propogation theory) will continue towards the direction of principal compression within the zone, that is, at 45° to the shear zone boundary. The earliest parts of the veins may rotate into the extensional field of the finite strain ellipsoid at higher strains and thin or even suffer boudinage: complex vein arrays can develop in this way as successive veins are initiated within the developing shear zone. 1.64 Pure shear strains Localized high strain resulting from pure shear will result in boundary simple shear unless there is a significant volume loss to accommodate the shortening. Although many ductile and brittle-ductile shear zones are seemingly produced solely by simple shear many show evidence of concomitant or subsequent flattening. (See Figure 3.4) 17 Shear zone rocks 1.7.1 Introduction e products of high shear strain range in textural and structural character depending on the rock composition, strain rate, environmental conditions, etc. (for review see Groshong, 1988). A number of attempts have been made at classi- fication, but almost inevitably genetic connotations have crept in whether or not it was the author's intent (See for example: Higgins, 1971; Sibson, 1977; and, Wise et al., 1987). Table 1.1 is the classification scheme from Sibson (1977). 172 Brittle effects Brittle shear zone products result from the mechanical breakdown of material resulting from rock fracturing and will be best developed at shallow depths where confining pressures are low. Such comminution of material within the zone is referred to as cataclasis and the fine-grained products of this process as cataclasites (Table 1.1). Subdivisions, indicating progressive comminution as exhibited by the ratio of remnant protolith to crushed matrix ratio are: protocataclasite, cataclasite, and ultracataclasite with > 50%, 50-10%, and <10% parent rock respectively. Cataclasites are typically unlayered and unfoliated, a fact that has been used as a field means of separating them from mylonites (see INCOHE SIVE COHESIVE Proportion of matrix 010% }10-50% 50-90% 90100% _. £ —|— mylonite series — a eS ec 2 = aoe] £ 2 2 7 & oisl & 21. z 2 ale | 2 lz] § = 8. Ss = [5® 3 B.S) 2 3 Siwue| phyllonites. a | & | 8 S]—catactasite series — rae 3] 2 |828) 2] » 2 a] 3° e -| § jags! & a g Z13s/ 9 a 3 laze 3 2 3 ©) 2313 8 8 aoc) ag] 8 [38 z\/sa|S Ss Z 229/82] 2 [es al2°|s tT @ |\Pedla® | 3 [5° E. : Z\/ 3 al 8 a Teduction in grain size grain Z| = 8] © lslsss | dominates grain growth growth = = promint| Table1.1: Textural Classification of Fault Rocks (After Sibson. 1977) below); Chester et al. (1985), however, have suggested that banding may be a common feature. Coarse angular fragments representing perhaps an early stage of progressive cataclasis are referred to as fault breccia or micro-breccia (where the largest fragments are <1.0 mm). Fault gouge is extremely fine comminuted material in the range 0.1 to 100 um. All of these fault products are susceptible to subsequent cementation since in their original cohesionless state they provide a permeable conduit for circulating fluids. In general, normal faults (extensional faults) will result in coarser fault rocks than compressional ones (reverse, thrust, and strike-slip). Where confining pressures as effective normal stresses are sufficiently low separation of the two walls of the fault may occur at perturbations in trend. These typically platy to lensoid zones may be filled with mineral precipitate whose growth pattern reflects the displacement sense (Figure 3.19, 20) These slickensides, when viewed on the fault surface, typically possess a mineral lineation, or slickenline. In the brittle-ductile regime small-scale discontinuous and stepped brittle shears may develop parallel to and shear veins develop (Figure 1.14; Chapter 3, Figure 3). Frictionally induced heat during rapid displacement on faults may be sufficient to cause melting of ‘dry’ quartzo-feldspathic material at moderate confining pressures. The resulting glassy pseudotachylite typically occurs as ramifying or isolated veins within the fault zone and surrounding wall rocks; in 23 24 the field it is difficult, if not impossible, to distinguish between ultracataclasites and devitrified pseudotachylite and care should be taken in applying these terms. Pseudotachylite may be restricted to the upper 10-15 kms of the crust and at an intermediate depth within this range (around 5 km) a single rapid slip of around 5 m could produce a melt 1 cm thick (Sibson, 1975, 1977); McKenzie and Brun (1972) argue that melting could occur with very small displacements (1 mm) at shallow depths and below depths equivalent to 10 kbs frictionally induced melting is impossible. 1.73 Ductile effects An important product of the recent interest in shear zones is the recognition of the significance of non-brittle, or ductile, processes in shear zone development (e.g., Ramsay, 1980a; Ramsay and Graham, 1970). (These are discussed further in the following chapter). This is in contrast to much of the earlier work where the emphasis tended towards fracture analysis (and hence the brittle character of shear zones), and, indeed, the term shear zone was taken by many field geologists to imply a zone of high deformation manifested by a zonal increase in fracturing and alteration. Perhaps the most revolutionary example of the new perspective was the re-interpretation that that most, if not all, mylonites are products of ductile rather than brittle processes (Bell and Etheridge, 1973, White et al., 1980), in contrast to the original definition of Lapworth (1885). Mylonites are very fine-grained, and typically laminated even in nonlayered proioliths. They characteristically possess a strong stretching lineation defining the X axis of the finite strain ellipsoid (see Chapter 2). The foliation in mylonitic rocks may be a described as a schistosity if sufficient subparallel platy grain forms exist. The term pliyllonite has been used to describe mylonitic rocks of phyllitic appearance (low grade foliated rock intermediate between slate and schist) that are commonly associated with retrograde effects. Planar fabrics external to the zone of highest strain may be deflected into it in the style described by Ramsay and Graham (1970) and illustrated in Figure 1.13. The nature of planar and linear fabrics associated with shear zones are further described in the following chapters (particularly 2,3, and 8) and a brief review of structural and textural features that suggest the movement sense in these zones is given in Chapter 2 (Figure 2.2). A reliable indicator of shear sense in foliated and mylonitic shear zones is the relationship between 'C' and 'S' fabrics (Lister and Snoke, 1984; Chapters 2 and 3), provided that it can be demonstrated that they are both related to the same deformation phase and great care in the field must be taken to correctly identify them. Their progressive development is illustrated in Figure 1.13. 25 Low strain Intermediate strain levels High strain ¢ Figure 1.13: (a) Progressive development of structures in ductile shear zones (After Simpson, 1984) Figure 14: Identification of fractures and veins in brittle and brittle-ductile shears. (a) Shear and extensional veins (modified after Roberts, 1987): § and P shear veins, E, extension fractures (en echelon ‘gashes’). (b) Riedel shears, Ry and R2 (after Ramsay and Huber, 1987). Note that these may also be vein filled. 1.74 Discussion From the recognition of the likelihood of alternating conditions of ductile flow with transient brittle failure, and their likely strong relationship to fluid behaviour within the shear zone, it should be expected that the perhaps previously assumed trend of ductile through brittle conditions (e.g., from high grade to retrograde in regional metamorphic terranes or, during progressive uplift on major reverse fault systems) may be reversed and alternate over a protracted period - both deformation styles overprinting the other. This alternation may well be related to the competing influences of ‘recovery’ and strain rate (affected to a large extent by fluid mobility within the shear zone, as described in section 1.5.1.1). Wise et al. (1984) envisage that a typical path followed at depths which could account for this behaviour is that given in Figure 1.15. The nature and geometry of brittle and brittle ductile structures are discussed in later chapters but some are included in Figure 1.16. Their superimposition will dissect and partially destroy evidence of any earlier ductile shear displacement, and this will further be masked if pervasive veining occurs; the fea: Rie SHITE orghr.onTe erecils 4! / \ 2 1 a) iA te ‘ iA Lea BL uA PLAL AN eee 2) JALAL Sees Recovery Rate Figure 1.15: Possible pathway described by wallrocks close to a major fault zone (after Wise et al., 1984) resulting structural complex (some might refer to it as structural anarchy) may be further modified by a further period of ductile strain. In such situations it is very difficult to determine that the earliest (ductile) fabrics belong to the shear system or are part of a more extensive and unrelated event. It is important, therefore, to know the geology of the surrounding terrane as much as the intricacies of the shear controlled vein system Abrupt changes in the trend of shear zones, or ‘jogs’ and slight deflections in trend, or ‘bends’ are potentially important sites for mineralization (Sibson, 1988, 1989). They produce localized transpressional or transtensional zones and are illustrated in Figure 1.16. below. In the following chapters detailed descriptions are given of the patterns of mineralization observed at both regional and deposit scale, and appraisals are made of the likely tectonic environment that would facilitate the development of shear zones and the concentration within them of the appropriate mineralizing fluids jogs in strike slip systems: es al zone of dilation faults ‘step to right =. ‘antidilational bends: dilational or compressional depending on active segment transpression - transtension: transtension transpression 1.16: Nature of jogs and bends in fault traces and zones of transpression and transtension (in part after Sibson, 1983, 1989) 27 CHAPTER 2 TECTONIC ENVIRONMENT AND METAMORPHIC CHARACTERISTICS OF SHEAR ZONES J.B. Murphy Department of Geology, St. Francis Xavier University, Antigonish, Nova Scotia, B2G 1C0 21 Introduction Many studies indicate a spatial and genetic link between mineralized shear zones and metamorphism thereby suggesting that both are manifestations of the same regional tectonic processes. Of utmost importance in characterizing mineralized shear zones is (i) grade of metamorphism; (ii) the style of shear deformation since it is a function of pressure and temperature conditions and (iii) the thermal and pressure gradients since these are important with regards to controlling fluid transport and precipitation. In the latter situation, the geothermal gradient is an expression of heat flow and therefore the thermal engine that drives solution and precipitation of metals. In addition, an important genetic link between mineralization and metamorphism is suggested by isotopic studies of shear zones in Archean greenstone belts which indicate that mineralizing fluids may have a metamorphic source (e.g., Kerrich and Fyfe, 1988). Furthermore the type of ore deposit found in shear zones is temperature and pressure dependent. In general, low temperature and pressure (greenschist facies) during shearing favour vein deposits, whereas high temperatures and pressures (amphibolite facies) favour disseminated deposits. The approach followed here is an examination of the important geological features that characterize the style of metamorphism associated with shear zones in varied tectonic settings (rather than producing an inventory of associated metamorphic reactions). 29 Instantaneous Thickening Thinning by Erosion Thinning by Extension Figure 21: The effect of crustal thickening and thinning processes on the geothermal gradient (after England and Thompson 1986) Thickening, is assumed to have taken place instantaneously by a single thrust sheet 35 km thick and thinned by erosion over 120 Ma back to its original thickness. The initial geotherm is represented by the stippled line, the final geotherm by the solid line. ‘The lower diagram shows the effect of thinning, by extension on the geotherm. The stippled line shows the initial position, the solid line shows its final position. O€ 2.2 Tectonic environment: extensional and compressional regimes 221 Introduction The geothermal gradient is a function of the regional tectonic environment in which the shear zone occurs. Extensional shear zones predominate in regions of lithospheric extension, compressional shear zones in regions of lithospheric convergence. Strike-slip regimes are commonly characterized by complex local and regional distributions of both extensional and compressional shear zones. Model calculations show that marked differences between geothermal gradients in extensional and compressional tectonic environments occur on both regional and local scales (¢.g., Turcotte, 1983; England and Thompson, 1986, Figure 1). Extensional shear zones display high T low P metamorphism with diagnostic growth of andalusite-cordierite assemblages in rocks of suitable bulk composition. They occur in intra-continental rifts involving detachment faulting and the evolution of continental margins, in intra-are rifts and in transtensional zones within strike-slip regimes. Compressional shear zones display typical Barrovian assemblages. They occur in regions dominated by thrust and nappe tectonics and locally within transpressional environments of strike- slip regimes. 222 Inter acteristics of ducti] ir Zon Criteria to distinguish extensional and compressional shear zones in brittle Tegimes, such as the style of brecciation and fracture-vein distributions, are well _known. Within the ductile field, extensional and compressional shear zones can be distinguished on the basis of regional analysis of kinematic indicators including, for example, S-C fabrics, asymmetic augen and fold development (Figure 2.2). Until recently, the most commonly used features were the deflection ‘of pre-existing fabrics (i, Figure 2.2) as the shear zone is approached or the ‘observed the offset on at least two marker units that occur on either side of the shear zone. However, in many instances these markers do not occur or yield ambiguous results, hence the emphasis has been placed on the interpretation of fabrics that occur within the shear zone (e.g., Simpson and Schmid, 1983). Most ‘of these interpretations are based on the orientation of the inferred infinitesimal train ellipsoid for simple shear relative to the bounding walls of the shear zone. 2221 Mineral lineations: The plane of slip (or C plane) is generally parallel to the walls of the shear zone and is commonly defined by a mylonitic fabric. Mineral lineations parallel to the direction of slip consisting of a dimensionally preferred elongation of crystals, lie within the S plane, but rotate into parallelism with the C plane at high strains (9, 2). The orientation of these lineations may be used to distinguish between strike-slip, dip-slip and oblique-slip shear zones in the same manner as 31 a2 slickenlines are used in brittle regimes. However, extreme caution should be exercised in shear zones with evidence of multiple movements. Just as slickenlines only yield information of the last increment of strain in brittle regimes, the orientation of mineral lineations in complex ductile shear zones may vary locally. Figure 2.2: ‘A summary of the kinematic indicators used to determine the sense of shear. This diagram is compiled from many authors including Berthé et al. (1979) Simpson and Schmid (1983), Lister nd Snoke (1984), Huddleston (1986), and Cobbold and Quinguis (1980). (a) shows a variety of kinematic indicators within the shear zone and the finite strain ellipse associated with the shear. These features are 1: deflection of a pre-existing fabric; 2: S-C relationships (2a = C; 2b = S); 3: asymmetric augen; 4: folds (note that the axial plane is parallel to the S fabric of 2a and that if sheath folds are developed the fold axes (4a) are curvilinear); 5: intrusion of dykes and veins parallel to the plane of extension of the infintesimal strain ellipse (Sb). Note that intrusion during progressive shear may cause rotation of these features (5a) towards the plane of flattening; 6: mica “fish”; 7: synthetic Riedel shears with the same sense of shear as the shear zone; 8: antithetic shears (ie. the sense of shear is opposite to that of the shear zone); 9: stretching lineation assumed to be parallel to the movement direction of the shear zone; 10: S-C intersection lineation. On the finite strain ellipse; C is parallel to the plane of shear and the walls of the shear zone; S is, parallel to the long axis of the strain ellipse, known as the plane of flattening; E is the plane of extension; R is parallel to synthetic Riedel shears; R’ is parallel to antithetic Riedel shears. Related C and S fabrics: maximum extension direction of the finite strain ellipse lies within the ane of flattening (or S plane, Figure 2.2) which is commonly defined by a eferred orientation of phyllosilicates that define cleavage or schistosity surfaces. The relative orientation of C and S fabrics (when viewed in cross- section parallel to the direction of the mineral lineation) uniquely determine the of shear (2, 4, Figure 22, Berthé et al. 1979). Initially, the angle between C nd S fabrics is about 45°. However, with progressive deformation, the S fabric intensifies, stretches and rotates towards parallelism with the C plane. At very strains, the C and S planes may be parallel and determination of the shear se becomes problematic. Folds within ductile shear zones: During progressive simple shear, folds may develop in pre-existing or pene- ontemporaneous anisotropic surfaces such as $ planes, mylonitic foliations, yeins, or dykes. Folds commonly nucleate around regions of stiffer or softer terial or in the necks of boudins (Cobbold and Quinquis 1980) and grow BE R\ D i Plunge. (d) 22b-d: Sketch and stereoplot to illustrate the development of sheath folds. Note the ‘curvi-linear fold axes. Initially fold axes are perpendicular to the transport direction and the stretching lineation (2c). With progressive shear, these axes rotate towards parallelism with the stretching lineation. Note that most folds in the foreground of 2b are south plunging S folds (denoted by an anticlockwise arrow) whereas in the background of 2b folds, are north-plunging and predominantly Z folds (denoted by a clockwise arrow). A schematic stereoplot representation of the pattern is shown in 2d. 33 34 Figure 2.2e: Example of how analysis of kinematic indicators may lead to erroneous inferences as to the structural history of a shear zone. S-C fabrics are used in this example, however, many of the other indicators used in Figure 2a would also yield misleading kinematic data at X. asymmetrically such as their sense of asymmetry may be used to infer the sense of shear. However, Ramsay et al. (1983) caution the use of this method be because small scale folds with the opposite sense of symmetry may develop on the short limb of larger asymmetric folds (4, Figure 2.2). Fold axes initially develop perpendicular to the shear direction (parallel to the stretching lineation, Figure 2.2c). During progressive deformation, however, the fold axes rotate towards the shear direction (4a, Figure 2). As this occurs, the original cluster of fold axes on a stereographic projection are drawn out in a girdle (Figure 2.2d) as the fold hinges become curvilinear or "bow-shaped”, with the direction of the "arrow" indicating the sense of shear (4a, Figure 2.2). These folds are called sheath folds (Cobbold and Quinquis 1980). In cross-section, the culminations and depressions give rise to dome and basin interference patterns. If individual sheath folds are not identified, then parasitic folds developed on the limbs of these bow-shaped structures may also be used to determine the sense of shear. An example of this method is shown in the foreground of Figure 2.2b which looks up plunge at predominantly southerly plunging $ folds. Although several orders of parasitic folds may form in this area, S folds should significantly outnumber Z folds on a statistical basis. In the background, northerly plunging Z folds should predominate. 2224 Rotation of rigid inclusions: Porphyroclasts or augen commonly develop asymmetric tails that are elongate parallel to the stretching lineation and may also be used to uniquely determine sense of shear (3a,b Figure 2.2; Simpson and Schmid 1983). Mica fish which ent a special type of porphyroclast oriented with their 001 planes at a low le to the mylonitic fabric (Lister and Snoke 1984), can also be used to ‘ine sense of shear (6, Figure 2.2). 5 Extension fracture in brittle-ductile shear zones: plane of extension within the strain ellipsoid is perpendicular to the plane flattening. Hence, fracture systems that are parallel to this plane are often ilational and may be filled with veins (eg., tension gashes) or dykes ), Figure 2). As suites of veins or dykes may intrude at about 45° to the wails of shear zone at any time during deformation, but become progressively rotated deformation proceeds, the sense of rotation of these features may also be used an indicator of shear sense. 2226 Riedel shears: Sets of conjugate shears, known as Riedel shears, may also form within a shear zone at approximately 30 degrees to plane of extension (R, R’, Figure 2). One of these surfaces which lies at a small angle to the mylonitic fabric and has the same sense of shear as the shear zone is variously known as a C’ or R shear, or a shear band. The relationship of these surfaces to the S cleavage is commonly preserved and uniquely determines the sense of shear (7, Figure 2.2). Antithetic R’ shears lie at a high angle to the shear zone and have the opposite sense of offset (8, Figure 2.2) (Chapter 3). The above kinematic indicators have been used to determine the sense of shear with a considerable degree of success in many studies. However, extreme caution should be exercised in complexly deformed areas in which the locus of most intense deformation and displacement may shift within the shear zone with progressive deformation. In these situations (e.g., Figure 2.2e), kinematic indicators in some parts of the shear zone may lead to erroneous inferences as to the regional kinematic or polyphase history of the shear zone (see X, Figure 2e where, for example, C-S relationships are locally inconsistent with the dominant C-S relationships within the shear zone). This emphasizes the need for systematic kinematic analyses across the entire width of the shear zone. 22.3 Extensional shear zones Three models have been advanced to account for lithospheric extension, and hence, the development of extensional shear zones in intra-cratonic or intra-arc rifting environments (Figure 2.3). These include one involving essentially pure shear (Figure 2.3a), one involving simple shear in which low angle detachment faults traverse the entire lithosphere ("Wernicke" model, Figure 2.3b) and one involving a. combination of simple and pure shear in which intra-crustal and 35)

You might also like