Está en la página 1de 247
Rock Deformation from Field, Experiments and Theory MOM mm Macs Edited by D. R. Faulkner, E. Mariani and J. Mecklenburgh Geological Society Special Publication 409 THE GEOLOGICAL SOCIETY ‘The Geological Society of London (GSL) was founded in 1807. Its the oldest national geological society in the world and the largest in Europe. It was incorporated under Royal Charter in 1825 and is Registered Chatity 210161 “The Society is the UK national learned and professional society for geology with a worldwide Fellowship (FGS) of ‘over 10.000. The Society has the power to confer Chartered status on suitably qualified Fellows, and about 2000 of the Fellowship carry the title (CGeol). Chartered Geologists may also obtain the equivalent European title, European Geologist (EurGeol), One fifth of the Society's fellowship resides outside the UK. To find out more about the Society, log on to www.geolsoc.org.uk, ‘The Geologieal Society Publishing House (Bath, UK) produces the Society's international journals and books, and acts as European distributor for selected publications of the American Association of Petroleum Geologists (AAPG). the Indonesian Petroleum Association (IPA), the Geological Society of America (GSA), the Society for Sedimentary Geology (SEPM) and the Geologists’ Association (GA). Joint marketing agreements ensure that GSL Fellows may purchase these societies’ publications at a discount. The Society’s online bookshop (accessible from www.geolsoc ‘rg.uk) offers secure book purchasing with your credit or debit card. To find out about joining the Society and benefiting from substantial discounts on publications of GSL and other societies worldwide, consult www.geolsoc.org.uk, or contact the Fellowship Department at: The Geological Society, Burlington House, Piccadilly, London WIJ OBG: Tel. +44 (0)20 7434 9944; Fax +44 (0)20 7439 8975; E-mail: enquiries @geolsoe. org.uk, For information about the Society's meetings, consult Events on www.geolsoc.org.uk. To find out more about the Society's Corporate Affliates Scheme, write to enquiries@ geolsoe.org.uk. Published by The Geological Society from: The Geological Society Publishing House, Unit 7, Brassmill Enterprise Centre, Brassmill Lane, Bath BAI 33N, UK ‘The Lyell Collection: www lyellcollection.org Online bookshop: www.geolsoe.org.uk/ bookshop Orders: Tel. +44 (0)1225 445046, Fax +44 (0)1225 442836 ‘The publishers make no representation, express or implied, with regard to the accuraey of the information contained in this book and cannot accept any legal responsibility for any errors or omissions that may be made © The Geological Society of London 2015. No reproduction, copy or transmission of all or part of this publication may be made without the prior writen permission of the publisher. In the UK, users may clear copying permissions and ‘make payment to The Copyright Licensing Agency Ltd, Saffron House, 6~10 Kirby Street, London ECIN 8TS UK, and in the USA to the Copyright Clearance Center, 222 Rosewood Drive, Danvers, MA 01923, USA. Other countries ‘may have a local reproduction rights agency for such payments. Full information on the Society"s permissions policy ccan be found at: www.geolsoc.org.uk/ permissions British Library Cataloguing in Publication Data A catalogue record for this book is available trom the British Library ISBN 978-1-86239-688-3 ISSN 0305-8719 Distributors For details of international agents and distributors see: www.geolsoc.org.uk/agentsdistributors ‘Typeset by Techset Composition India (P) Ltd, Bangalore and Chennai, India. Printed and bound by CPI Group (UK) Ltd, Croydon CRO 4YY Contents Maniant, B., Mecktensunt, J. & Fautener, D. R. Towards an improved understanding of the ‘mechanical properties and rheology of the lithosphere: an introductory article to ‘Rock Deformation from Field, Experiments and Theory: A Volume in Honour of Emie Rutter’ Fautaner, D, R., Mariani, E., Meckiensuran, J. & Covey-Crump, S. E, H. Rutter: a biography Upper crustal deformation Applied rock mechanics , K, Percolation of the interacting elastic stress fields of aligned cracks as an alternative crack densities Ql, G. & Evans, B. Permeability and thermal cracking at pressure in Sioux Quartzite Sint, B. K. & Wenbr, A. S. Estimation of horizontal stress magnitudes using sonic data from vertical and deviated wellbores in a depleted reservoir axon, TR. & Hacan, J. T. The nature of stress state: numerical and laboratory experiments, and some field observations Pressure solution Piovsraxens, A. M.H, & Sprens, C, J. Compaction creep of simulated anhydrite fault gouge by pressure solution: theory v. experiments and implications for fault sealing Brown, R. H.C. & ALMgvist, B. S. G. The role of stress on chemical compaction of illite shale powder Lena, G., Barc, M. R., ALVAREZ, W., FELICI, F. & Mrveti, G. Mesostructural analysis of S fabrics in a shallow shear zone of the Umbria~Marche Apennines (Central Italy) Gapats, D, & Laouan Bre Bouol, A. Pegmatite mylonites: origin and significance Analy Vatexe, S. L. A., De Bresser, J. H. P., Pexwock, G.M. & Drury, M. R. Influence of deformation conditions on the development of heterogeneous recrystallization microstructures in experimentally deformed Carrara marble of microstructures and crystallographic preferred orientation Zocaus, M., Barserint, V., Vottount, M., OuLappiar, B., CuaTEIGNER, D., Mancint, L. & Lutrerormi, L. Quantitative 3D microstructural analysis of naturally deformed amphibolite from the Southern Alps (Italy): microstructures, CPO and seismic anisotropy from a fossil extensional margin, Mamnerice, D., BACHMANN, F., Hisiscitit, R., ScHasies, H, & Liovp, G. E. Calculating anisotropic piezoelecttic properties from texture data using the MTEX open source package Mainonice, D., BacuMann, F., Hitiscuee, R. & ScHazsex, H. Descriptive tools for texture projects with large datasets using MPEX: strength, symmetry and components ysis of Index 31 49 93 149 167 175, 201 251 23 image not available 2 E, MARIANI ET AL, understanding of the rheology of the Earth’s crust. This, together with Ernie’s intention to ‘retire’ (or perhaps ‘not retire”) and be able to dedicate his time to his main passions, the rock deformation lab- oratory and Earth’s natural laboratory, inspired the convening of « conference in his honour. This meet- ing was organized at Burlington House, Geological Society of London, and was held on 30-31 May 2012. ‘The conference included 50 participants attending from 10 different countries, and 55 contri- butions were presented, Conference contributions were varied and top- ics, all relevant to various aspects of the mechanical properties and theology of the lithosphere, ranged from applied studies of gas migration and perme- ability in shale and clays to the frictional properties (from slow to earthquake slip rates) and permeabil- ity of fault zones, to the microstructure of brittle faults and shear zones in relation to field studies in the Italian Alps and Apennines and Southern Spain, ‘Twelve of these contributions are published in this volume and will be introduced in the following sections, where we highlight the importance of combining field observations, experimental results and theoretical models to further our understand- ing of deformation processes in the Earth’s crust The concept of continental strength profiles, one of the key contributions of rock deformation studies to the Earth Sciences, is used here as a framework to present the contributions made to this volume in the context of Emnie’s work, as well as some of the challenges that still remain. Lithospherie strength profiles The idea of a continental crust divided into a shal- low schizosphere, or zone of fracture, and a deeper plastosphere, or zone of flow, developed amongst ts in the 1930s (Griggs 1936). e theological concepts developed along deconvolution methods in seismology that allowed geophysicists to determine the depth of earthquake hypocentres with more confidence (Macelwane 1936). Pioneering experimental studies on rock deformation were carried out that investigated fric- tional strength, through to brittle~plastic and fully plastic deformation behaviour (terminology after Rutter 1986) of a variety of rock types, at a range Of pressure (P) and temperature (T) conditions sim- ulating different depths in the continental crust (e.g Paterson 1969, 1978). Within this new and vibrant experimental rock-deformation community, Bye- lee (1978) first observed that, for most rock types, frictional strength is independent of lithology and may be represented by a simple linear relationship between shear stress and normal stress. (It is inter- esting to note that Leonardo Da Vinci, in his manu- seript Codex Arundel (early 1500, ff. 40v, 41r), explained how his experiments showed that friction is nominally independent of material and propor- tional to the weight of the objects in contact. Da Vinci proposed a universal coefficient of friction of | 0.25.] At the same time, the work of Goetze (1978) and the combined efforts of Heard & Carter (1968) and Christie et al. (1979) resulted in some of the first flow laws proposed in the geoscience commu- nity for olivine and quartz, respectively (Brace & Kohlstedt 1980). As a young geoscientist Eric Rutter explored the wet and dry rheology of calcite aggregates (Rutter 1974, 1976) and, extrapolating his results to geological strain rates, contributed one of the early strength profiles for carbonate rocks (Fig. 1) (Rutter 1972). In the decades that fol- lowed, owing to caleite’s natural occurrence in active and fossil fault zones in the shallow crust (Badertscher & Burkard 2000; Ebert et al. 2007; De Paola ef al. 2011; Smith er al. 2011) and its rele- vance as an analogue geo-material, deformable at the full range of conditions accessible in the laboratory, it became one of the most studied and better charac- terized minerals in the experimental rock deforma- tion community (e.g. Schmid et al. 1980; Rutter 1995; Covey-Crump 1998; Brodie & Rutter 2000; Barnhoorn et al. 2004). In 1980 Brace and Kohlstedt gathered existing experimental data to construct a model for the rhe- ology of the continental lithosphere. Around the same time, Goetze & Evans (1979) developed a similar approach for the oceanic lithosphere. hese and similar models have since been known as ‘strength profiles’. Brace & Kohistedt (1980) used Byerlee’s rule of friction to represent the strength of rocks in the upper 5—10km of the Earth’s crust. The lower crust was described with high-temperature experimental data for quartz. and olivine extrapolated to the (lower) temperatures nd slow strain rates predicted in nature. At depths where the highest density of earthquakes was recorded (c. 15 km, see also Fig. 2c), intersection of the brittle and plastic strength curves was inter- preted to represent the brittle—plastic transition (BPT), now often referred to as the frictional —vis- cous transition. Field observations of fault rock structures, indicative of the depth of the BPT, sug- gested that the model of Brace and Kohlstedt may be plausible (e.g. Sibson 1982). Thus, despite its many limitations, from the assumption of an Ander- sonian stress state (Anderson 1905), to that of steady-state deformation and geothermal gradient, to the effects of fluids and pore pressures at depth (eg, Carter & Tsenn 1987; Kohlstedt et al. 1995; Scholz 2002), the continental lithospheric strength profile proposed in 1980 has. inspired fervent research activities aimed at testing its validity and providing an improved understanding of the rheol- ogy of the lithosphere. image not available image not available image not available 6 E, MARIANI ET AL, methods in a North Sea reservoir. With a view on the «general state of stress in the crust, Harper & Hagan (2015) present a reappraisal of how we should think about stresses that are locked into rocks at depth, ‘They argue that terminology borrowed from engi- neering such as ‘residual stress’ is not very useful in geomechanics and propose the use of the term “inherent stress’ to describe the stresses that a rock can record in it from a previous stress state the rock was under. These authors show that inherent forces can have an impact on the way geological ‘materials behave, a phenomenon that currently is not widely considered. Pressure solution The shape and character of the BPT in crustal strength profiles are still not well constrained, It is shown in Figure 1 that the BPT, even in controlled experiments on monomineralic rocks, cannot be obtained simply by the intersection of Byerlee’s rule of friction with a flow law curve, but that instead it has a more complex character. During deforma- tion in the lithosphere the BPT may be dependent (on rock type, rate and temperature of deformation, pore pressure and rock—fluid interactions. Pressure solution is a mechanism of deformation that can accommodate large strains at a range of low tem- peratures in the upper crust, by the diffusion of matter in a thin, intergranular aqueous film, away from high stress interfaces (Rutter 1983), Numer- ous field studies have reported microstructural evidence indicating that pressure solution is impor- tant in regional scale active and fossil fault zones slog :o(0V3G) (e.g. Imber et af. 2001; Frost er al. 2011; Gratier etal. 2011; Lena et al. 2014; Richard et al. 2014) ‘This, combined with the low temperature sensitivity of pressure solution, suggests that slow fault creep at the BPT, under favourable conditions, may be largely accommodated by this mechanism (Rutter & Mainprice 1979). Two seminal papers by Ernie Rutter in the 1970s and 1980s (Rutter 1976; 1983) led to: (a) a flow law for pressure solution derived using a theoretical approach; and (b) one of the first deformation mech- anism maps in Geoscience to include pressure solu- tion (in the context of calcite deformation; Fig, 4), Further experimental and field studies by Chris Spiers, one of Emie’s past students, as well as other colleagues in the experimental rock deformation community (e.g. Spiers et al. 1990; Hickman & Evans 1995), later resulted in the development of a new microphysical model of frictional viscous Chrittle-plastic’) creep, where it is proposed that the combined action of frictional sliding and pres- sure solution may accommodate large displacements during the slow interseismic creep of faults (Bos et al. 2000; Bos & Spiers 2002b; Jefferies et al. 2006; Gratier et al. 2013). Additionally, pressure solution is proposed as one of the possible mecha- nisms for fluid-assisted fault healing, resulting in fault strengthening during periods of no or slow creep (e.g, Rutter & Mainprice 1978; Bos & Spiers 2002b; Gratier ef al. 2013; Richard et al. 2014). ‘This topic has inspired the contribution of Pluy- makers & Spiers (2014), who estimate the time- scale of sealing and healing processes in anhydrite fault gouges using detailed and rigorous kinetic logo (o/10° Pa) at 500°C 4, Deformation mechanism map for calcite plotted in temperature stress space from with contours forlog strain rate (Rutter 1976). One of the first applications to rocks of M. F. Ashby’s deformation mechanism maps (see review by Frost & Ashby 1982). image not available image not available image not available eee Fi E, MARIANI ET AL, log Differential Stress (MPa) 7. Log strain rate log differential stress geaph for serpentinite experimentally deformed both above and below the dehydration temperature. Dehydration weakening results in fine-grained reaction produets deforming by a linear viscous grain size-sensitive deformation mechanism (Rutter & Brodie 1988) underpining micro-physical parameters that may be used to construct predictive models of the inter- action between deformation and metamorphis Partial melting in the middle to lower crust lead- ing to large scale weakness, as evidenced by the lack of large topographic variations of the high pla- teau of Tibet, has been used to explain the tectonic evolution of the Himalaya with a channel flow ‘model (Beaumont et al. 2001; Harris 2007). The first experiments on partially molten granites were Fig.8. Back-scattered electron image of olivine forming fine-grained reaction product from the dehydration of serpentinite in micro shear zones deforming by grain Size-sensitive low (Rutter & Brodie 1988), performed by van der Molen & Paterson (1979), who defined a rheologically critical melt percentage for granitic rocks. This concept became powerful within the geological community although, in the experiments of van der Molen & Paterson (1979), the melt viscosity varied considerably over the melt fraction range. Rutter & Neumann (1995) used increasing temperature to increase the melt fraction, thus obtaining a more constant melt viscosity over the range of melt fractions and hence a more realis- tic melt fraction dependence of strength. ‘This was further improved by using synthetic granitic rocks where the melt fraction and viscosity were con- trolled independently (Fig. 9; Mecklenburgh & Rutter 2003; Rutter ef al. 2006). Many of the geo- dynamic models involving partially molten mate- rial use arbitrary weakening functions rather than experimental parameters (e.g. Beaumont et al. 2001; Shen etal. 2001). Ernie and his group showed that their empirical flow law for partially molten rock could be used effectively in geodynamic mod- celling (Rutter et al, 2011). Analysis of microstructures and crystallographic preferred orientation Jual 10 Ernie Rutter’s passion for fieldwork is his conviction that it is only through petrographic and microstructural analyses of both recovered image not available image not available image not available 4 E, MARIANI ET AL, (Cunistie, J. M., Koc, P. S. & GEORGE, R. P. 1979. Flow law of quartzite in the alpha-quartz field (abstract). Eos Trans. AGU, 60, 370. Cortex, A., Houunasworrn, J. & Berowan, E, 2012, Constraints on fault and lithosphere rheology from the coseismic slip and postseismic afteslip of the 2006 Mw 7.0 Mozambique earthquake. Journal of Geophysical Research, 117, BO3404, hutp://doi.org/ 10.1029/201 LJB008580 Covey-Crume, S. J, 1998, Evolution of mechanical state in Carrara marble during deformation at 400° to 700°C. Journal of Geophysical Research, 103, 781-729, 794. Css, R.J., RurTER, E. H. & Hottoway, R.F.2003. The application of critical state soil mechanies to the mechanical behaviour of porous sandstones. Inter national Journat of Rock Mechanics and Mining Sciences, 40, 847-862, _hitp://doi org/10.1016/, s1365-1609(03)00053-4 Davies, G. F. 2011. Manile Convection for Geologists. Cambridge University Press, Cambridge. Pao1.r, N., CHIODINI, G., HIROSE, T., CARDELLIN, C., CaLIRO, 8. & SHIMAMOTO, T. 2011. The geochemical Signature caused by earthquake propagation in carbonate-hosted faults. Earth and Planetary Science Letters, 340, 225-232. pe Ronpe, A. A., Sruntrz, H., Totus, J & Hetseon- NER, R. 2005. Reaction-induced weakening of plagio- clase—olivine composites. Tectonophysics, 408, 85-106, Dr Toro, G., HAN, R. 27 L, 2011. Fault lubrication during earthquakes. Narwre, 471, 494-498, Bnet, A., HeRWEGH, M. & PrirENER, O. A. 2007. Coole ing induced strain localization in carbonate mylonites withina large-scale shear zone (Glarus thrust, Switzer- land). Journal of Structural Geology, 29, 1164~1184. Farver, J. R.& Yonb, R. A. 1991. Oxygen diffusion in (quartz: dependence on temperature and water fugacity Geochimica Cosmochimica Acta, 4, 2053-2964, FAULKNER, D.R. & RUTTER, EB, H. 2000. Comparisons of water and argon permeability in natural clay-bearing gouge under high pressure at 20°C. Journat of Geophysical Research—Solid Earth, 108, 16415~ 16 426, hitp://doi.org,/'10.1029/2000jb900134 FAULKNER, D. R., Lewis, A. C. & Rurrex, E. H. 2003. On the internal structure and mechanics of large strike-slip, fault zones: field observations of the Carboneras fault in southeastem Spain, Tectonophysics, 367, 235-251, hutp://doi.org/10.1016/s0040-1951(03)00134-3, FAULKNER, D. Ry, JACKSON, C. A. Li, LUNN, Re Is Scutiscue, Ro W., Suton, ZK, WinsERLEY, . A.J. & WITHIACK, M. O, 2010. A review of recent developments concerning the structure, mechanics and fluid flow properties of fault zones. Journal of Struc tral Geology, 32, 1557-1575. Frost, E., Dotan, J., Ratscrmactre, L., Hacker, B. & Sewakb, G. 2011, Direct observation of fault zone structure at the brittle-ductile transition along the Salzach~Ennstal-Mariazell~Puchberg. fault system, Austrian Alps, Jowmal of Geophysicat Research, 116, BO2411, heip://doi.org/10.1029/2010N8007719 Frost, HJ. & Asupy, M. F. 1982. Deformation Mechanism Maps: The Plasticity and Creep of Metals and Ceramics. Pergamon Press, Oxford Di Fussets, F., SCHRANK, C., LIU, Ju, KARRECH, A., LLANA: FONkz, S., Xiao, X. & REGENAUER-Libs, K, 2012. Pore formation during dehydration of a polycrystalline gypsum sample observed and quantified in a time: series synehrotron X-ray micro-tomography experi ‘ment. Solid Earth, 3, 71-86. »als, D. & LAOUAN BREM BOUND, A. 2014. Pegmatite mylonites: origin and significance. Za: FAULKNER, D.R., Mariani, E. & MEcKLENBURGH, J. (eds) Rock Deformation from Field, Experiments and Theory: A Volume in Honour of Ernie Rutter. Geological Soci ety, London, Special Publications, 409, First pub- lished online August 4, 2014, hitp: //doi.org/10.1144/ SP409.7 Gortzr, C. 1978. The mechanisms of creep in olivine, Philosophical Transactions of the Royal Society, London, Series A, 288, 99-119. Gortze, C. & EVANS, B, 1979, Stress and temperature in the bending lithosphere as constrained by experimental rock mechanics. Geophysical Journal of The Royal Astronomical Society, $9, 463-478, Gotspsy, C. & Koutstept, D. L. 2001. Superplastic deformation of ice: experimental observations. Journal of Geophysical Research, 106, 11 017-11 030, Gratien, J.P,, GUIGUET, R., RENARD, F,, JENATION, L. & BeRNakD, D. 2009, Pressure solution ercep law for quartz from indentation experiments. Journal of Geo- Physical Research~Solid Earth, 114, 1-16, Gaamex, J-P., Dysrur, D.& Rea, F. 2013. The role ‘of pressure solution creep in the ductility of the earth's upper crust. Advances in Geophysics, 54, 47—179. Gane, J.P. RicHaRD, J. Er AL. 2011, Aseismic sliding of active faults by pressure solution ercep: evidence from the San Andreas Fault Observatory at depth. Geology, 39, 1131-1134. Garces, D. T. 1936. Deformation of rocks under high con: fining pressures. The Journal of Geology, 44, 541-577. Hanpy, M.R. & Bau, J. P. 2004. Seismicity, structure and strength of the continental lithosphere. Earn and Planetary Science Leuers, 223, 427-441, hutp://doi. org/ 10.1016 /j.cpsl.2004.04.021 Hagper, T. R. & HaGan, J. T. 2015. The nature of stress sale: numerical and laboratory experiments and some field observations. Jn: FAULKNER, D. R., MaR- ANI, E. & MECKLENBURGH, J. (eds) Rock Deformation from Field, Experiments and Theory: A Volume in Honour of Ernie Rurter. Geological Society, London, Special Publications, 409, First published online June 12, 2015, huup://doi.org/10. 1 144/SP409,12 Hanis, N. 2007, Channel flow and the Himalayan— ‘Tibetan orogen: a critical review. Journal of the Geo: logical Society. London, 164, $11~523, hp: //doi ‘org/10.1144/0016-76492006-133 Hawkeswormn, C.J. a. A. 1S. K. 2006, Evolution of the continental crust. Nature, 443, 811-817. Haep, H.C. & Carter, N. L. 1968. Experimentally induced ‘natural’ intragranular flow in quartz and quartzite. American Journal of Science, 266, \—2. Herrer, K. 2014, Percolation of the interacting clastic siress fields of aligned cracks as an alternative explana- tion of critical crack densities. Jn: FAULKNER, D. Rs. Manian, E, & MECKLENRURGH, J. (eds) Rock Defor mation from Field, Experiments and Theory: A Volume in Honow of Ernie Rutter. Geological Society, image not available image not available image not available 18 E, MARIANI ET AL, Field, Experiments and Theory: A Volume in Honour of Emie Ruter. Geological Society, London, Special Publications, 409. Fisst published online August 13, 2014, btip://doi.org/10.1144/SP409.4 ‘WAN DER MOLEN, I. & PATERSON, M. S. 1979. Experimen- tal deformation of partally-melted granite, Contribu- tions to Mineralogy and Petrology, 70, 299-318, Waker, ALN. Rorter, E. H. & Bkopir, K. H. 1990, Experimental study of grain-size sensitive flow of syn= thetic, hot-pressed calcite rocks. In: Knipe, R. J. & Rurrer, E. H. (eds) Deformation Mechanisms, Rheol- ‘ogy and Tectonics. Geological Society, London, Spe- cial Publications, $4, 259-284. Wartter, 1.2014, Dramatic effects of stress on metamor- phic reactions. Geology, 42, 647-650, http: //doi.org 10.1130/235718.1 Wusnexcey. C. A. J. & Suimamoro, T. 2003. Internal structure and permeability of major strike-slip fault zones: the median tectonic line in Mie Prefecture, southwest Japan. Jounal of Structural Geology. 23, 59-78. Winey, C. A. J, YIELDING, G. & Di ToRO, G. 2008. Recent advances in the understanding of fault zone intemal structure: a review. In: WinnERLEY, C. A.J Kunz, W., banex, J., Hovpswoxrn, RE. & Couer- tint, C. (eds) The Internal Structure of Fault Zones: Implications for Mechanical and Fluid-Flow Proper- ties. Geological Society, London, Special Publications, 299, 5-33. Woon, D. M. 1991, Soil Behaviour and Criticat State Soil ‘Mechani. Cambridge University Press, Cambridge. ZocaLt, M., BaRBERINI, V., VOLTOLINI, ML, OULADDIAF B., CuaTeiGNex, D., Mancini, L. & Lurrerornt, L 2014. Quantitative 3D microstructural analysis of nat- turally deformed amphibolite from the Southern Alps (ltaly): microstructures, CPO and seismic anisotropy from a fossil extensional margin. In: FAULKNER, D.R., Manian, E. & MECKLENSURGH, J. (eds) Rock Deformation from Field, Experiments and Theory: A Volume in Honour of Emie Ruter. Geological Society London, Special Publications, 409. First published online August 1, 2014 hitp://doi.org/10.1144/SP409.5 image not available image not available image not available 2 D. R. FAULKNER ET AL, Fig. 5. Emie with his family at the gypsum caves in the Sorbas Basin, SE Spain. Fig. 6. Emie at Deformation Mechanisms, Rheology and Tectonies Conference at ETH Zurich with Mervyn Paterson and John Ramsay image not available image not available image not available 26 35, 31, 40. 41, 2. a. 45. 46. 41. 48. 49. D. R. FAULKNER ET AL, Rurrer, B. H. & Covey-CRump, 8. J. 1985, Cal- cite graim-growth in marbles from Naxos, Greece. Journal of the Geological Society, London, 142, 1242-1242, Buenxinsor, T. G. & Rutter, E. H. 1986. Cata clastie deformation of quarvzie in the Moine thrust. zone, Journal of Structural Geology, 8, 669-68 Rurrer, B. H., Mapbock, R. H., Hatt, S. H. & Wanre, S. H. 1986. Comparative microstructures of natural and experimentally produced clay-bearing fault gouges. Pure and Applied Geophysics, 124, 3-30. RUTTER, EH. 1986, On the nomenclature of mode of failure transitions in rocks. Tectonophysics, 122, 387, £, 8. H.. BRETAN, P.G. & RUTTER, E. H. 1986, Fault-zone. reactivation—kinematies and me nisms. Philosophical Transactions of the Royal Soci- ety of London A, 317, 81-92. Rute, E. H. & HALL, S. H, 1986, The structure of Neogene fault ones and fault rocks of the Betic Zone, SE Spain. Journal of the Geological Society, London, 143, 326~326, Brooie, K, H. & Rutter, E. H. 1987. Deep crustal extensional faulting in the Ivrea Zone of northern aly. Tectonophysics, 140, 193-212. Bxopie, K. H. & Rutter, E, H. 1987. The role of, ‘wansiently fine-grained reaction products in syatec- tonic metamorphism: natural and experimental examples. Canadian Journal of Earth Sciences, 24, 556-564. Rurver, E. H. & Bropie, K. H. 1987, On the mechanical properties of oceanic transform faults Annales Tectonicae, 1, 81-96, Prox, D. J., Kntre, R. J, BATES, M. P., GRanr, N. T. Law, RD, Ltovp, G. E., WELON, A. ‘AGAR, S.M., Broott, K. H., Mppoex, R.H., Rut ree, E. H., Wurre, 8. H., Bett, T. H., FERGUSON, CC. & Warten, J. 1987, Orientation of speci- ‘mens~essential data forall fields of geology. Geol- ogy, 18, 829-831 Mappoce, R. H., Wits, SH. & Rurrex, EH. 1987, Electron-optical studies of experimentally deformed Tennessee Sandstone and quattz + kaolinite gouge. Mineralogical Magazine, Sl, 125-125, RUTTER, E. H, & Broie, K, H. 1988. Experimental approaches to the study of deformation metamor- phism relationships. Mineralogical Magazine, 52, 35-42, RUTTER, B. H. & BRopis, K, H. 1988, The role of tectonic grain size reduction inthe theological strati- fication of the lithosphere, Geologische Rundschaat, 77, 295-308. Rurrer, E. H. & Bronte, K. H. 1988. Experimental ‘syntectonic’ dehydration of serpentinite under con- ditions of controlled pore water pressure. Journal of Geophysical Research-Solid Earth and Planets, 93, 4907-4032, Covey-Ceump, S.J. & Rurte, BH. 1989, ‘Thermally-induced grain-geovth of calcite marbles ‘on Naxos Island, Greece. Contributions to Mineral- ogy and Petrology, 1, 69-86. 50, sl 2 sa. 56, 7 38. 39. 6 6 Bropir, K. H., Rex, D, & RUTTER, E. H, 1989, On the age of deep crustal extensional faulting in the Ivrea zone, northern Italy. fn: Cowarp, M. P., Dierwicu, D. & Park, R. G. (eds) Alpine Tectonics Geological Society, London, Special Publications, 45, 203-210, Rowe, K.J.& RUTEER, E. H, 1990, Palacostress e ‘mation using ealeite twinning experimental calibra tion and application to nature, Journal of Structural Geology, 12, 1=11. Rurrer, E. H. & BRopie, K. H. 1990. Some geo- physical implications of the deformation and meta- ‘morphism of the Ivrea zone, northern Maly. Tecionophysies, 182, 147~160. WaLker, A. N., RUTTER, E. H. & Broptr, K. H. 1990. Experimental study of grain-size sensitive flow of synthetic, hot-pressed calcite rocks. I Kuipe, RJ. & RurreR, E, H. (eds) Deformation Mechanisms, Rheology and Tectonics. Geological Society, London, Special Publications, $4, 259-284, Rurrer, EH. & Bropir, K. H. 1991. Lithosphere theology—a note of caution. Journal of Structural Geology, 13, 363-367 RUrver, E, H, & Habrzapen, J. 1991. On the influ ‘ence of porosity on the low-temperature brittle ductile transition in siliciclastic rocks. Journal of Siructural Geology: 13, 609-614. Rurrer, E. H. 1992. The mechanics of natural rock deformation, In: Hupsox, J. A., Brown, E. T., Farrwursr, C. & Hork, E. (eds) Compre- hensive Rock Engineering (Principles, Practice, and Projects), Vol. I: Fundamentals. Pergamon, Oxford, 63-92, Haiwwoon, E. A., MaDDock, R. H., FUNG, T. & Rurrex, E. H. 1992. Palacomagnetic analysis of fault gouge and dating fault movement, Anglesey, North Wales. Jounal of the Geological Society of London, 149, 273-284, Rurver, E. H. & Mappock, R. H. 1992. On the mechanical properties of synthetic kaolinite/quartz fault gouge. Terra Nova, 4, 489-500, RUrveR, E, H, & Broptt, K, H. 1992, The rheology of the Lower Continental Crust. Jn: Foustain, D., ARCULUS, R. & Kay, R. (eds) Continental Lower Crust. Elsevier, Oxford, Series on Developments in Geodynamics, 201-268. Brovir, K. H., RUTTER, E. H.& EVANS, P. 1992. On the structure of the Ivrea Verbano Zone (northern Ttaly) and its implications for present-day lower con- {inental crust geometry. Terra Nova, 4, 34-40, Rurrex, E. H., Bropie, K. H. & Evans, P. J. 1993. Structural geometry, lower erustal magmatic under plating and lithospheric stetching in the Ivrea Verbano zone, northern Italy. Journal of Structural Geology, 18, 647-662. Rurter, EH. 1993. Experimental rock deforma- tion: techniques, results and applications to tectonics, Geology Today, 9, 61-65. NEUMANN, D. & Ruttek, E, H. 1993. Experimental deformation of partially molten Westerly Granite and extraction mechanisms at low melt percentages. J Unzoa, W., WALLBRECHER, E, & BRANDMAYR, M. (eds) Structures and Tecionies at Different Litho- spheric Levels. Blackwell Scientific, Oxford, 24. image not available image not available image not available Percolation of the interacting elastic stress fields of aligned cracks as an alternative explanation of critical crack densities KES HEFFER Reservoir Dynamics Ltd., Kayness, Nightingale Avenue, West Horsley, Leatherhead, Surrey, KT24 6PB, UK (e-mail: kes@ reservoir-dynamics.co.uk) Abstract: The concept that the lithosphere s ina general critical, ornear-critical, mechanical state has previously assumed spatial intersection of fractures atthe critical point. This paper provides an initial basis for an alternative mechanism in which elastic interactions between aligned, open, spatially separated (miiero-Jeracks in rock can, by themselves, lead to a continuous phase change sal threshold of erack density lower than that required for crack coalescence, The existence ical density of aligned cracks is first demonstrated on a regular 2D hexagonal grid when subjected to a central stress perturbation; at this critical density the elastic tensile stress interactions have a long-range effect. From the results of those calculations an approximation of the tensile sitess field around a erack in 3D is provided. The percolation behaviour of such 3D bodies is dis ‘cussed and a critical crack density in 3D for elastic interaction of 0.035 is deduced. This falls within the range of crack densities (0.01: ‘many different rocks. The possibility that c ‘one explanation of the lon; large-scale seismicity. There is theory and evidence indicating that the Earth’s crust is in a near-critical state (e.g. Harper & Szymanski 1991; Main 1996; Bak 1997; Crampin 1999; Zoback et al. 2002; Al-Kindy & Main 2003; Rundle eral. 2003). This implies that there are perco- lating paths of faults, joints, incipient fractures and other discontinuities that are near mechanical failure in the prevailing stress states or that provide connectivity for conductivity (Madden 1983). In other words, the general structure of the subsurface is on the verge of mechanical failure or conductiv- ity on extended or even through-going paths. This paper is aimed at providing a slight modification to the detail of that concept, but one that is important for interpreting observations of production charac- teristics in hydrocarbon reservoirs, particularly in the low-stress environment of a waterflood. As particular indicators of near-criticality, common geomechanical characteristics have been observed in two completely independent techniques for investigating the condition of the subsurface as follows. (a) Production and injection flowrates at wells in oilfields generally show fluctuations over time. Most commonly, flowrates are reported. ‘on a monthly basis but fluctuations, often of high amplitude and frequency. occur over all timescales, partly due to intervention by the ‘operator in changing chokes and in well oper- ations but otherwise due to communication between wells, most importantly through the subsurface. Heffer et al. (1995) found that 0.045) interpreted from observations of sheas-wave splitting in cal rack interactions can be mostly clastie pr range nature of correlations in flowrate fluctuations in oilfield ides hhout flowrate correlations between injector—produ- cer wellpairs are significantly higher when the wellpairs’ subsurface locations are separated in a direction subparallel to the direction of the maximum horizontal principal stress axis, (Sima): In-a study of flowrate fluctuations in 6 North Sea fields Heffer (2012a, 6) found that the diffusivity tensor of flowrate fluctu- ations, derived from the time behaviour of cor- relations over a more local region around both injector and producer wells, showed peaks in the ‘shearing’ directions at c. 30 degrees 10 Symes (b) Shear-wave splitting (or birefringence), where the polarizations of the faster waves are sub- parallel to the direction Of Syma» has been observed to be an inherent characteristic of almost all rocks in the crust (and upper mantle) (e.g. Crampin 1987, 1994). This has been taken to indicate that most in situ rocks are pervaded by stress-aligned fluid-saturated microcracks. Stress-aligned fluid-saturated grain-boundary cracks, and possibly also low- aspect-ratio pores and pore-throats, are the most compliant elements of the reservoir or rock mass. An additional finding that is common to plentiful observations of both oilfield production data and of shear-wave splitting in all types of sedimentary and igneous rocks is that there is a long-range spatial influence. Heffer ef al. (1995), Main et al. (2007) and Heffer (20124, 6) found long-range correlations between temporal fluctuations in From: Fautkner, D. R., Maniant, E. & Mecktensunot, J. (eds) 2015. Rock Deformation from Field, Experiments and Theory: A Volume in Honour of Ernie Rutter. Geological Society. London, Special Publications, 409, 31—47. First published online September 3, 2014, hip: //dx.doi.org/10.1144/SP409.1 © The Geological Society of London 2015, Publishis disclaimer: www.geolsoe.org,uk/pub_ ethics image not available image not available image not available PERCOLATION OF STRESS Fl Fig. 1. Crack-normal stresses around a single central crack on the surface of which is imposed a unit tensile crack-normal stress. The stresses are calculated according to equation 8.4 of Pollard & Segall (1987), but only evaluated at the nodes of the hexagonal grid shown in Figure 2 and interpolated in between: the central region therefore shows the tensile stress on the central crack, rather than the compressive siress of the rock, and multiple crack array, respectively. Figure 5 (no boundary condition, or *no b.c.") plots the RMS difference between these two profiles (for points at DS OF ALIGNED CRACKS 35 10 different radii) against the value of ¢/H and confirms that the critical value of c/H is 0.5. For the calculation described above there was no outer boundary condition. This might be the most appropriate assumption for many subsurface situ- ations. In other cases a fixed outer stress state might be applicable. To examine this sensitivity, the tensile stress change was constrained to zero at new grid points added with equal spacing around a circle encompassing the grid: those extra constraints gave an even more over-determined system which was again solved with singular value decomposition to give a least-squares solution of stresses. at the cracks on the grid. The pattems of stress distribu- tion are shown in Figure 3b. They are qualitatively similar to those of Figure 3a, but differing in detail, In particular the critical value of ¢/H was again found to be 0.5, with the plot of RMS differ- ence between the multiple and single crack cases similar to that of the previous ‘no boundary con- dition’ case, although with slightly lower values (ee Fig. 5 ‘zero b.."). Percolation densities of randomly located, aligned cracks in a 3D continuum Having empirically determined a crack density on 2D lattice at which interferences rapidly increase, I crack density in a 3D. Fig. 2. (a) Hexagonal (or “diamond’) grid of aligned equal-length cracks used for 2D stress calculations. The array of cracks is limited to acircular region. A unit tensile erack-normal stress is applied to the central crack (thicker bar) and the consequent distribution of crack-normal stresses ucross the array is calculated. (b) Dimensions of a representative diamond around each crack in the array showing the crack half-length c, the erack spacing s and the distance between. erackenormal rows ofthe aay 17 (note tat ffs = 3/2) image not available image not available image not available PERCOLATION OF STRESS For parallel, regularly shaped, reasonably convex, equal-sized inter-penetrating objects randomly located in 3D space at the percolation threshold there isa universal value of V.v, where Nis the num- ber of objects per unit volume and v is the volume of each object. In 3D this universal value has been determined to be 0.35 + 0.02 (Balberg er al. 1984). With ras the distance along the 30° centre— centre line to the threshold contour in the 2D lattice simulations described above, r/s=0.5 is imposed for the limit of stress fields that just touch, Using the previous result that ¢/H = 0.5 for critical interference and the geometrical relation- ship H/s= ¥3/2, then (c/seriuea = V3/ 0.433 oF (5/Derica = 11S where L= 2c is the full crack length. The critical dimension of the volume of stress field associated with a crack is defined by (r/c)eritieat = (/9)/(C/Serivicat = LAS. Note also that the critical density of cracks fully occupying the 2D hexagonal lattice is pay =n C/A = Q/V3)(e/ serie = 0.22, where n is the number of eracks within area A. Shapes around cracks in 2D (or ‘dog-bones’) were converted to 3D bodies (or *yo-yos") by rotat- ing the 2D area of the critical area of stress field around the crack normal (see Fig. 7); this assumes a stress field that is axisymmetric around the crack normal. The percolation of these ‘yo-yos’ can then a Fl DS OF ALIGNED CRACKS 39 be estimated from established percolation theory, which requires the volume of each “yo-yo". Values of r/c corresponding to various contours of ten- sile stress were examined from the 2D simulations and the results of the simulations at the value of (r/oderiiea Were interpolated, from which the vol- ume of each object can be approximated. Corre- sponding to the critical value of (r/)ieat OF 1.15, the calculated value of normalized volume v/c* is 10.15. Each fracture and its immediate tensile stress field in 3D were viewed as ‘soft-core’ objects for the purposes of assessing percolation; that is, the objects are inter-penetrating and the stress inter- actions can be larger than when the bodies just touch, Cautiously appealing to the universality of such percolation outlined above, the threshold 3D crack density was determined as peap = Ne Ne v{v/e*) = 0.35/(v/e*) and hence, using the cal- culated normalized volume, psp = 0.35/10.15 = 0,035. This is a factor of 3~6 less than the range of densities required for percolation by intersection of just crack planes themselves (0.10.2 as noted previously), a consequence of the fact that the objects have 3D volume rather than being 2D dises, ‘Some discussion of 2D v. 3D crack densities is, pertinent. It is readily shown that the 2D density of crack traces on an areal cut (gn = (n/A) <(L/2)°>, where <(L/2)°> is the average of (b) 2¢ 45 00 08 ae as Fig. 7. (a) The crack-normal stress field in 2D around a single crack showing the contour at which the field is curtailed: and (b) illustration of how the 2D stress field is rotated around the crack normal (n-axis) to form the bounding surlace toa pseudo-3D stress field. image not available image not available image not available PERCOLATION OF STRI so! 110 110 110) 8 g = 3 2 = = 119 119 o ot . 1-10. Double logarithmic presentation of Figure 5, showing that the RMS difference between stress profiles for the single and multiple cases along transects at 30° to Sama ApPFOximates a power law with exponent in excess of +4, and implying that a unique critical ‘point’ depends upon the pertinent level of stresses. followed by rotation to approximate 3D stress fields, curtailed to a semi-arbitrary contour. In particular, the fact that interactions of the tensile stress fields were only considered at crack centres (rather than at continuously distributed points along each crack's surface) nplies that negative interactions between the overlapping tip portions of neighbouring en echelon cracks were neglected. This obviously becomes of greater significance where local crack densities exceed the threshold (i.e. for ¢/H > 0.5 in the 2D model- ling). On the other hand, linear (or planar) cracks have been assumed; it is however likely that, during their development, crack interactions will cause bending of their subsequent growth (e.g. fig. 8 of Fonseka et al. 1985). Such non-alignment will act t0 decrease the ‘shielding’ effect of overlapping tips, ameliorating inhibition of greater stress interactions between the central portions of the cracks. (@)_ The perturbation approach for effective elastic, constants (Crampin 1993, using results from Hudson 1980, 1981) may not be entirely ade~ quate as a means to estimate crack densities from percentage anisotropies of shear waves at the critical threshold for percolation of clastic tensile stress fields, The failure of per- turbative methods is an intrinsic difficulty in the study of continuous phase transitions at the critical point, as the correlation length diverges and many degrees of freedom are cor- related (e.g. Binney et al. 1992; Cardy 2000). OF ALIGNED CRACKS 43 (e) The fact that the stress field around a crack has both tensile and compressive domains, unlike the idealized ‘yo-yo’ comprising just regions of tensile stress. Configurations of cracks in which regions of compressive stress overlap are allowed in the idealized percolation theory will be inhibited in reality. (8) Shear stress interactions have been ignored in this analysis, (g)_ Stress interactions will affect the genesis and gfowth of microcracks, including sub-critical crack growth (Atkinson & Meredith 1987) and the development of wing and splay cracks (e.g. Schulson etal, 1999), such that their spa- tial locations are not uncorrelated as assumed in the simple percolation theory used here. (h)_ The peripheral stress field that exists in reality beyond the artificial finite spatial bound placed upon it for the purposes of using percolation theory. (i) The 3D percolation theory employed in this work assumes a threshold when there is a 1D continuous path through the system; this threshold might be appropriate for hydraulic connectivity, but may be too low when consid- ering connectivity of extended 2D surfaces for ‘mechanical failure. (Crack pairs much closer than the spatial limit placed in this work on their stress fields are likely to form a through-going shear rather than remain separate, contrary to the invoked percolation theory. (k)_ The open crack density can fall below the per- colation threshold when the local stress has been relieved; the density can rise above the percolation threshold along. through-going failure planes. () Heterogencities in the material properties of the host rock at various scales affecting, thy (micto-Jerack properties and the stress inter- actions between them (m) Variable crack sizes: the threshold will depend upon the particular frequency distribution of crack size. (n)_ Finite size scaling: the interpretation of crack density from shear-waves is over a scale that is connected to the wavelength of the shear- wave signal, which might not be effectively infinite compared with the size of maerocracks or their correlated clusters, Previous analyses of crack interactions have generally not found such a significant effect as seen in this work. Through 3D finite element model- ling, Grechka & Kachanov (2006) found that the nnon-interaction approximation (NIA) incurred only small errors in the estimation of moduli for a solid containing densities of randomly located and image not available image not available image not available PERCOLATION OF STRESS FIELDS OF ALIGI UESTEFELD, A., VERDON, J. P., KENDALL, J-M,, RUTLE- GE, J, CLARKE, H. & Wookey, J. 2011. Inferring rock fracture evolution during reservoir stimulation from seismic anisotropy. Geophysics. 76, WC137-WC166. ZATSEPIN, S. V. & CRAMPIN, S. 1997. Modelling the com= pliance of erustal rock: I, Response of shear-wave ED CRACKS "1 splitting to differential stress. Geophysical Journat International, 129, 477-494, Zowack, M. D., TowNEND, J. & GRoLLIMUND, B. 2002 Swady-state failure equilibrium and deformation of intraplate lithosphere. International Geology Review 44, 383-401 image not available image not available image not available image not available image not available image not available image not available 56 G, SIDDIQI & B. EVANS image not available image not available image not available 60 G, SIDDIQI & B. EVANS Fig. 8 Model geometry of square grain in an isotropic homogencous matrix after Fredrich & Wong (1986). Arrows in the grain denote the principal direction of thermal expansion, Cracks either protrude into the matrix at 37/4, which ‘maximizes strosses generated by thermal expansion anisotropy, or lie along quartz grain boundaries. : é 2 z Initial Flaw Size (um) £ £ o E Ze Bo i 2 z .* ae. 2 Initial Flaw Size (ym) Initial Flaw Size (jm) ig.9. Computed stress intensity factors plotted as a function of an initial law size at three temperatures. Irthe Ki. is 0.5 MPa m, crack propagation will occur for most short intergranular cracks. Note that all cracks grow into a decreasing slress field and will eventually arrest, Grain boundary crack growth arrests earlier than intergranular growth, Superimposing a confining pressure will lower stress intensity factor below critical, particularly at P. = 200 MPa. image not available image not available image not available 64 G, SIDDIQI& B. EVANS Ac’ deviatoric part of the thermal expansion tensor v Poisson's ratio K thermal conductivity 4 heat flow Fo principal values of stress field (F~To) maximum temperature difference 2L grain size x integration variable ranging from 0 to a 1 time References Aransox, B. & Mereprri, P. 1987. Experimental fracture mechanies data for rocks and minerals, fv ArKINESON, B. K. (ed.) Fracture Mechanics of Rock Academie Press, London, UK, 477-525. Bannon, T., Contns, J, Swimm, T. & Wutre, G. 1982. Hermal expansion, Gruneisen functions and. static lattice properties of quartz. Journal of Physies C: Solid Siate Physies, 18, 4311-4326, Bavsn, S. & JonNsox, B, 1981. Effects of slow uniform heating on the physical of westerly and charcoal gran- ites. Proceedings of the US Symposium on Rock Mech- anics, 20, 112. Bosier, N. M. & HICKMAN, S. H. 2004, Stress-induced, time-dependent fracture closure at hydrothermal condi tions, Journal of Geophysical Research: Solid Earth, 109, BO2211, hup://dx.doi.org/10.1029/20025B oor7s2 Beanané, Y. 1986, Pore volume and transport-properties changes during pressure cycling of several crystalline rocks. Mechanics of Materials, 8, 235-249, Brawané, Y. 19874. The effective pressure law for per- meability during pore pressure and confining pressure cycling of several crystalline rocks. Jounal af Geo- physical Research, 92, 649-657 BeRwanf, Y. 1987b, A wide range permeameter for use in rock physics. International Journal af Rock Mechanics and Mining Sciences & Geomechanics Abstracts, 24, 309-315. Breanne, Y. 1988. Comparison of the effective pressure law for permeability and resitivity formation factor in Chelmsford granite, Pure and Applied Geophysics, 127, 607-625. Bexxasé, Y. 1991, Pore geometry and pressure dependence of the transport-properties in sandstones. Geophysics, 56, 436-46, Brace, W.F. & Martin, R.J., IIT 1968. A test of the law of effective stress for erystalline rocks of low porosity International Journal of Rock Mechanics and Mining Sciences, 8, 415-426. Brace, W.F., WALSH, J.B. d& FRANGOS, W. T. 1968, Per- lity of granite under high pressure. Jounal of Geophysical Research, 73, 225-2236, Brace, W.E,, SILVER, E., Goetze, C. & HADLEY, K. 1972. Cracks and pores ~ a closer look. Seience, 178, 162. CARLSON, S. R., Wu, M. & WANG, HLF, 1990, Microme- chanical modeling of thermal cracking in granite. I: Dua, A. G., Dunnam, W. B, HANDIN, J. W. & WANG, H. F. (eds) Brittle-Ductile Transition in Rocks, AGU, Washington, DC, Geophysies Mono- graphs, Heard Volume, 37—48. CantwRiGT, D. & ROokE, D. 1979, Green’s functions in fracture mechanics. In Sure, R. (ed.) Fracuure Mechanics: Current Status, Future Prospects. Perga- mon Press, NY, 91-123. (Ciarke, D. R 1980. Microfracture in brittle solids result- ing from anisotropic shape change. Acta Metallurgica, 28, 913-924, Cov, RS. & PareRsoN, M. $. 1969. The a-B inversion in quartz: a coherent phase transition under nonhydro- static stress. Journal of Geophysical Research, 74, 4921-4048, Danor, M., Gusouen, Y. & Baratin, M. L. 1992. Per- meability of thermally cracked granite. Geophysical Research Letters, 19, 869-872. Davin, C., Won, T-F., Ziv, W. & ZHANG, J. X. 1994. Laboratory measurement of compaction-induced per- meability change in porous rocks ~ implications for the generation and maintenance of pore pressure excess in the crust. Pure and Applied Geophysics, 143, 425-456, DrMarriy, B., HintH, G. & EVANS, B. 2004. Experimen- tal constraints on thermal eracking of peridotite at ‘oceanic spreading centers. In: GERMAN, C. Ri, LIN, J & Parson, L. M. (els) Mid-Ocean Ridges: Hydrother- ‘mal Interactions Between the Lithosphere and Oceans. AGU Washington, DC, 167-186, Evans, A. G. 1978. Microfracture from thermal expan- sion anisotropy ~ i. Single phase system. Aeta Metal lurgica, 26, 1845—1853. Farpricn, J. T. & WoNG, TF. 1986, Micromechanics of thermally induced cracking in three crustal rocks. Journal of Geophysical Research: Solid Earth Planets, 91, 12743-12764, FUNG, L-K., BUCHANAN, L. & WAN, RG. 1994, Coupled geomechanical-thermal simulation for deforming heavy-oil reservoirs. Journal of Canadian Petroleum Technology, 433, 22-28, Gavel, A. F. 1978. Variation of whole and fractured orous rock permeability with confining. pressure International Journal of Rock Mechanics and Mining Sciences, 18, 249-257. Gana, A. F. & Carton, R. L, 1996, An asperity- deformation model for effective pressure. Tectono- hyies, 286, 241-251 Gavainenxo, P. & GUEGUEN, Y. 1989. Pressure depen- dence on permeability: A’ model for cracked rocks. Geophysical Journal International, 98, 158~172. GeRsvp, Y. 1994, Variations of connected porosity and inferred permeability in a thermally cracked granite Geophysical Research Letters, 21, 979-982. Genavp, Y., MAzEROLLE, F, & RavNavD, S, 1992, Com- parison between connected and overall porosity of thermally stressed granites. Journal of Structural Geology, 14, 981-990. Grnsvp, Y., MaztROLLE, F., RAYNAUD, S. & LEHON, P. 1998. Crack location in granitic samples submitted to beating, low confining pressure and axial loading. Geo- physical Journat International, 133, 353567, GrRAUD, Y., DIRAISON, M. & ORELLANA, N, 2006, Fault zone geometry of a mature active normal fault: a poten: tial high permeability channel (Pirgaki Fault, Corinth Rift, Greece). Tectonophysics, 426, 61~T6. GnianeZ100, S., SULEM, J, GUEDON, S. & MARTINEAU, F 2009. Effective stress law for the permeability of a image not available image not available image not available 68 B. K. SINHA & A. S. WENDT (Gorehote sits) ® > ‘ZZ ig 1. (a) Schematic diagram of a deviated wellbore with azimuth ¢ measured from the north and deviation @ from the Vertical. It is assumed that the maximum horizontal stress direction coincides with the north, Indicial notation for the axes are convenient for the orthogonal transformation of tensors in terms ofthe Euler angles ¢and 0. (b) Schematic of a image not available image not available image not available R B. K. SINHA & A. S, WENDT shear velocities from an acoustoelastic model of wave propagation in prestressed materials. We can obtain the following relationship between stress ratios dmsin/@v and Spex /'7v from equations (8) and (9): (Css ~ Codon (Cu = Coaov Sra = Oss oy Cu Cos a2) Note that this relationship between the three effective principal stresses, ov, Sirmax aNd Chinn at a given depth and the three shear moduli in equa- tion (12) are independent of the chosen isotropic reference state of rock. In particular, notice that only two of three equations (8)-(10) are independent, However, there are four unknowns: Ciys, Css, Gray AVA Fi To overcome these limitations, we generalize the applicability of shear moduli difference equa- tions in the presence of known stress distributions caused by the presence of a borehole, Near-wellbore stress distributions are known from the theory of elasticity that are valid for rock stresses that are less than the rock yield stress. Dipole shear moduli Cus and C35 change as we approach the near- wellbore region where the far-field stresses change to borehole cylindrical stresses. Radial and azi- muthal variations of these stresses are known from the linear elasticity (Jaeger & Cook 1979: Roegiers 1989; Kirsch 1898). Since these difference equa- tions contain two unknown nonlinear constants, Css and Css, we have to solve for four unknowns However, we can form two more difference equa- tions that relate changes in the dipole shear moduli Cuz and Css at two radial positions to the core- sponding changes in borehole cylindrical stresses ‘These borehole stresses can be expressed in ter of the three formation principal stresses. One of the equations relates the difference (Css(r/a= far)— Css(r/a = near)] to corresponding stresses at these two radial positions normalized by the borehole radius a. The second equation relates the difference [Cu(r/a = far)-Css(r/a = near)] to the stresses at these two radial positions. Radial variations of shear moduli Css and Cy are obtained from the dipole shear radial velocity profiling algorithm using the fast- and slow-dipole dispersion We can therefore solve these four equations to obtain the maximum and minimum horizontal stres- ses and the nonlinear constants Cis5 and Cras that are referred to a local reference state. Higher-order coefficients of nonlinear elasticity Cyay, Ciss and Case are also used to calculate stress coefficients of shear velocities from an acoustoelastic model of ‘wave propagation in prestressed materials (Thurston & Brugger 1964; Sinha 2002). Note that we do not use the shear modulus Cys in the near-wellbore region. This shear modulus is likely to be more affected by the plastic yielding of | rock in the high stress concentration region where breakouts are likely to occur. In addition, we do not use the shear modulus Css at radial positions too close to the borehole surface where this mod- ulus might also be affected by the near-wellbore stress concentrations and associated plastic defor- mations. The radial variation of Cys is generally uniform because this modulus is dependent on the sum of radial and hoop stresses ax we approach the borehole surface from the far field. This sum is largely independent of the radial position in the inter- mediate region of radial profile (3 < r/a <5). ‘The presence of a fluid-filled borehole in a formation subjected to an overburden effective stress ary and isotropic horizontal effective stresses (Griese = Fresin) CAUSES An inctease in the tangential hoop stress and a decrease in the radial stress approaching the wellbore fluid pressure as we approach the borehole surface from the far-field horizontal stresses. Generally, the wellbore pressure is less than the formation pore pressure during overbalanced drilling and is less than the forma- tion pressure in the case of underbalanced drilling, In contrast, differences in the horizontal stresses Ghioox ANA Cyyin CAUSE the tangential stress to vary azimuthally with position around the wellbore, Figure 2a shows typical near-wellbore stress distri- butions along the nay stress direction, whereas Figure 2b depicts stress distributions along. the Grin direction that coincides with the azimuth where potential breakouts might occur. The tange tial stress attains its maximum compressive value at the azimuth of the far-field minimum horizontal stress, and it has its minimum compressive value at the azimuth of the far-field maximum horizon- tal stress, as illustrated in Figures 2a, b. Note that these stress distributions are calculated based on linear elasticity and ignore any plastic yielding of rocks that is often present. Figure 3 displays a sche- matic of the borehole Stoneley wave dispersion and the two orthogonal flexural wave disper- sions aligned with the principal stress directions in the presence of stress distributions. shown in Figure 2a, b. Difference equations using radial profiles of shear moduli The two orthogonal flexural dispersions in a vertical well corresponding to radial polarizations parallel and perpendicular to the Grima stress direction exhibit crossovers in the presence of near-wellbore stress concentrations caused by the far-field maxi- mum and minimum horizontal stresses as. shown in Figure 2a, b (Sinha & Kosick 1996; Sinha er al. 2000; Liu & Sinha 2003). These two flexural image not available image not available image not available 16 B. K. SINHA & A. S. WENDT Notice that quantities in the square brackets in equations (19)-(21) denote linear and nonlinear constants for an effectively isotropic rock in the chosen reference state. Consequently, these material parameters are invariant under any orthogonal trans- formation of the reference axes. Two difference equations from the measured far~ field shear moduli referred (o the deviated wellbore axes take the form Che~ Cl = Ado - trax) (22) (23) (OV — Phen The other two difference equations from the radial profiles of the fast and stow dipole shear slow- nesses are given by CH rae) = CHP” eae) = Ker AY CI an) — CSS rad = Yer (25) where X, and ¥, contain the unknowns of..4.+ Tporins Cus and Ciss, together with the equivalent- isotropic shear modulus jz and Poisson's ratio v (in its first appearance, v was not defined as Poisson's ratio) of the chosen reference state as shown in equa- tions (19) and (20); ria, and Facae Tefer to the far and near radial positions where the shear moduli are estimated from the corresponding radial variations of the shear slownesses. At this point, we have four nonlinear algebraic difference equations (22)~(25) that can be solved for the four unknowns Cjfgays Finmigs Cra and Css We also have the following expression for the stress ratio from equations (22) and (23) Fitwin 4 _ C55 = Cs ae C66) Fray, of ty cy ey, cy) of Cis — Cos @ = 3) ita =o (Bos). eo Ca Coo” C= Ci) 0% At this point, we have estimates of all three stres- 8€5 Cinaxr Fin ANd Of referred to the deviated wellbore axes. Next we express deviated borehole stresses along the radial (jjg0)s HOOP (Gif) and axial (of) directions in terms of the formation principal sitesses and borehole azimuth g measured from the maximum horizontal stress direction, and borehole deviation @ from the vertical: owas o% cost pH +. sino ov ov sin? psin?@ TE + cos sin? 07 +. cos" Qn Where Oy, Cymax ANd Oyun denote the formation vertical, maximum horizontal and minimum, hori- zontal stresses, respectively. Similarly, we can express the ratio of borehole hoop (ofjqi,) and axial (of) stresses as Sin oe sin? feos? 97 + cos? deos? #7" + sin? ov a i ge Hi + cost fsin" 0+ co 28) Equations (27) and (28) can now be solved to obtain the formation effective principal stresses Grins ANE Syn iD terMS OF the effective over- burden stress ay and the wellbore azimuth g and deviation 0. Using equations (24)~(26), we can now express Cinin/ Fy i terms Of Grimax/ Fy. measured shear ‘moduli referred to the deviated borehole axes and deviated borehole orientation referred to the princi- pal stress directions: =G)G2)-@) © sin” 6 + Rosin” dcos? 8 + cos? 6. Rycos’ #-+ Rosin’ 4, By = Rycos’ dsin’@ + Racos" eos" 4, n= CBS Ciu= Cig 1 Gs= Ce" Gs~ Ci G0) ‘The ratio of minimum horizontal to overburden SUES Trmin/ Gy can also be expressed in another form: Fr -(2 Smut) Gs aw \O oy a’ image not available image not available image not available 80 B. K. SINHA & A. S. WENDT ig. 8. Estimated stress magnitudes in a nearly vertical wellbore: The first rack shows shale volume and the friction angle over a select depth interval. The second track depicts the pore pressure in pink with measured pressures shown by diamond markers. The overburden and maximum and minimum horizontal stresses are shown by the blue, red, and sgreen curves, respectively. The black diamond markers are horizontal stress magnitudes obtained from the inversion of radial profiles of shear slownesses in a vertical wellhore, The solid red cirele denotes the minimum horizontal stress ‘magnitude obtained from a minifrac test that is consistent with that determined from the borehole sonic data. image not available a You have either reached 2 page thts unevalale fer vowing or reached your ievina tit for his book. a You have either reached 2 page thts unevalale fer vowing or reached your ievina tit for his book. a You have either reached 2 page thts unevalale fer vowing or reached your ievina tit for his book. a You have either reached 2 page thts unevalale fer vowing or reached your ievina tit for his book. a You have either reached 2 page thts unevalale fer vowing or reached your ievina tit for his book. a You have either reached 2 page thts unevalale fer vowing or reached your ievina tit for his book. image not available image not available image not available image not available image not available a You have either reached 2 page thts unevalale fer vowing or reached your ievina tit for his book. ‘THE NATURE OF STRESS STATE 95 Maximum princi! stress “870x107 820x107 “60x10 “e10x10" 590x107 “5 70x10" S.50K107 “B20 107 10x10? 490x107 Contour interval= 1.0010 Xy-stress contours -2.50x10" 0.0010 2.50. 10° ‘Contour interval = 2.50x 10" Fig. 1. The central 2 km x 2 km (plan view) of a 5 km x 5 km elastoplastic (Moht-Coulomb) simulation of discontinuously fractured rock having uniform elastic properties in a normal faulting environment ata depth of 3000 m, (@) Maximum stress (filled contours at intervals of | MPa) and (b) shear stress (illed contours at intervals of 2.5 MPa), Stresses of 61.2 MPa and 40.8 MPa were applied at the boundaries of the model, with 33,9 MPa pore pressure and an. out-of-plane stress of 67.9 MPa. The mode! was allowed to approach equilibrium before imposing roller boundary conditions and than allowing 435 000 cycles to investigate the approach to equilibrium. The plots are derived from fracture distributions having equal numbers of extension fractures and shear fractures, the shear fractures forming two groups of equal numbers allocated at 45° clockwise and 45° anticlockwise of the maximum applied stress (aligned up ‘and down the page) and the extension fractures being aligned with the maximum applied stress. The total of 10% fractured zones comprise 5% extension fractures and two groups of 2.5% shear fractures. Fracture orientations have a standard distribution of 10°. Identical fracture strengths were allocated to al fractures: angle of frietion of 27.5° with a Standard deviation of I°, cohesion of 0.3 MPa with a standard deviation of 0.1 MPa, dilation angle of 5° and ensile strength of 0.1 MPa. ‘Thornton & Barnes (1986) and Cundall (1989) using ‘Thornton & Barnes (1986) simulated isotropic particle flow codes (a subset of the range of distinct loading of a circular region followed by shear defor- element codes which can also be used to model mation. These authors observed a force transmission blocky media), is evident. pattern which is random when both the structure and, 96 T.R. HARPER & J. T, HAGAN applied stress field are isotropic. The pattern is characterized by lines or ‘chains’ of force which are several times greater in magnitude than those in the surrounding clusters of particles which are relatively less loaded (Thornton & Bares 1986). During shearing, these authors observed a progress- ive, increasingly anisotropic force distribution. The approximately circular patterns enclosed by the large force chains in the isotropic stage changed to form the boundaries of more lens-shaped regions elongated in the direction of the principle shortening strain, Oda & Konishi (1974) observed similar pat- terns in assemblies of photoelastic discs. Thornton (1990) explained that during shear deformation, within the anisotropic distribution of forces which develops in a granular medium, the large forces are aligned with the larger compressive stress. From our simulations, we observe that a fabric of force transmission, which has long been recognized for ‘granular media, occurs in discontinuously fractured media such as we have simulated here, Damjanac et al. (2004) gave examples of the response of fractured rock to dynamic wave trans- mission, Their results also showed a strain chan- nelling effect. Couples (2013) noted the stress heterogeneity resulting from interactions between the fracture-bounded regions in a blocky medium (pervasively fractured in contrast to our discontinu- ously fractured models), and suggested that the behaviour of blocky systems is ‘reminiscent of the load chains that develop in granular systems’, Our mulations of discontinuously fractured rock demonstrate such a characteristic, Strain energy storage, irreversibility, time dependence and marginal stability All these observations suggest that heterogeneity and anisotropy of stress state are characteristics which span a very wide seale of observation from particulate media to discontinuously fractured and blocky media. These observations further imply that the energy storage related to the distribution of elastic strain energy around individual slipped frac- tures in our simulations also occurs in blocky and in granular media (corresponding to the pattern of high-force chains enclosing low-stress regions seen by the authors reporting simulations of granular and blocky media), Indeed, numerous observations and discussions of ‘residual stress’ or ‘inherent stress’, both the phenomenon and confusion with the termi- nology, date from approximately half a century ago in the rock mechanics literature (e.g. Denkhaus 1967). Friedman (1972) used X-ray diffraction to examine crystal distortions and determined differential stresses of 30-40 MPa in quartzites, sandstones and granites. X-ray diffraction detects ‘mean strain stored in specific crystallographic planes (Friedman 1972). All these observations, and many others related to the core scale but not reported here, substantiate the conclusion that a hetero- geneity and anisotropy of stress state exists for a range of scales extending upwards from a length sale much smaller than individual grains or crystals, Investigators of residual strain have long recog- nized that residual elastic strains represent a means of energy storage (e.g. Friedman & Logan 1970; Friedman 1972; Nichols & Abel 1975; Warpinski et al. 1993). The range of scales of stress heterogen- eity reported here directly implies that balanced forces exist at a very wide range of scales, possibly the extremes thereof. When deformed, both granular and fractured media experience rolling and/or sliding, implying energy dissipation and hence irreversibility. The complex interactions between fractures which occur as each fracture experiences an increment of slip in our simulations suggest a mechanism which is time-dependent. While such time-dependeney may be overridden or at least constrained in laboratory constant. strain-rate experiments, in field situa- tions a single event ~ say an engineering operation such as a single hydraulic fracturing operation — represents a perturbation after which the reservoir may be allowed to equilibrate without further artificial constraints. Under such conditions, an interaction-related time-dependency might become influential. The processes are non-linear and dissi- pative so are also expected to be path-dependent, Our numerical model showed a strong tendency for interaction between fractures, and between fractures and the model boundaries, which contin- ued to occur for one million cycles when modelled explicitly using FLAC. Given the rapid fall-off in strain with the distance from an individual sliding fracture, this interaction implies a sensitivity to very small strains. Harper (2014) analysed similar simulations that showed that a small change in the effective stress history (less than 0.5 MPa transient change in pore pressure) very substantially changed the response to hydraulic fracturing. This sensitivity may reflect a state of marginal stability (as defined by Jensen 1998) or self-organized criticality (e.g. Crampin 1999), but the numerical data only infer this without additional analysis. In the field, Heffer and co-authors (e.g. Heffer & Dowokpor 1990; Heffer et al. 2010) have reported long-range injector-producer flow-rate correlations which cannot be explained by Darcy flow, but could be explained by a state of marginal stability (as defined by Jensen 1998) in which state the response to minor perturbations can be highly ‘THE NATURE OF STRESS STATE 7 Experimental core programme History and objectives During the 1980s, a research programme investi- gating the potential to determine stress from rock core was conducted by BP. At the time, the tech- nique known as Anelastic Strain Relief (Teufel 1989), or ASR, was in use. This technique involves measurement of core strains immediately when an orientated core taken in a vertical well arrives at surface. Mainly because of the slightly unpredict- able nature of precise coring times (incurring per- sonnel waiting costs) and technical complications associated with cooling of the core and pore pres- sure reduction during and after core recovery, this can be a difficult and time-consuming process, especially offshore. It had been found that the orientation of the maximum strain observed during ASR Tests coin- cides with the orientation of the maximum horizon- {al stress as indicated by microseismic monitoring of hydraulic fracturing operations in conventional reservoirs (Lacy 1984). This has been attributed to microcracking (Lacy 1984; Teufel 1989; Holt & Kenter 1992) during the coring and core recovery processes, leading to anisotropy of elastic modulus in the horizontal plane, Recalling the conclusions drawn from early point load tests on core (Friedman & Logan 1970), it was recognized at BP that a method involving perturbation of the residual stress state of orientated core under normal labora- tory conditions might be a far more practicable, reproducible and efficient means of determining the orientation of maximum horizontal stress. Experiments involving a sawcut made across the diameter of unconfined, orientated core, previously strain gauged at each end and conducted in con- trolled temperature and humidity conditions, allowed a redundancy of results (Fig. 2). Firstly, strain gauge arrays at each end of the initial saweut core yielded at least two sets of results. (More results can be obtained by attaching strain gauge delta rosettes of more than one diameter on the ends of the core.) Secondly, additional diametral saweuts could be made on the two halves of th initial core, monitoring strains using the strain rosettes attached at the end of each core piece remaining from the initial sawcut. This technique, known as laboratory. inherent force evolution (LIFE), was combined with a demagnetization pro- cess for core orientation (Rolph ef al, 1995) and shown to be consistent with stress determinations made using the ASR procedure and interpreta- tions based on wellbore breakout (Fig. 3, Table 1). Table 1 shows the results of 11 field comparisons of maximum horizontal stress orientations inferred by three different methods. The results. include Binsin anenancalinennine ends of core Fig. 2. Schematic of the LIFE test procedure used 10 determine the orientation of the maximum and minimum horizontal stresses from orientated core. ‘comparisons from the northern North Sea, southern North Sea, offshore Holland, onshore UK, the USA and Colombia. The average difference between the LIFE determinations and one of the other two methods is 10.1 During the development of the LIFE technique, similar cores were subjected to other forms of dis- turbance including step cycles of minor temperature change. The cores used were mainly sandstones of low to moderate clay content, many from wells in the southern North Sea or onshore UK. ‘The cores were found to be sensitive to the various forms of perturbation and displayed a markedly time-dependent and irreversible strain response. We draw upon these experiments and some reported by others to evaluate in more detail the character- istics of the stress state at the micromechanical scale, or close to such, responsible for the observed strains The influence of the coring process ‘The anisotropy of core strains in the horizontal plane has been attributed to extensional microcracking (Lacy 1984; Teufel 1989; Holt & Kenter 1992) (normal to the axis of the maximum horizontal strain and therefore to the inferred maximum hori- zontal stress). Harper & Szymanski (1991) reported that acoustic emission also occurs following a dia- metral saweut (a Lower Leman sandstone core). Extensional cracking during and after the removal of a rock core from the in situ location is assumed 98 T.R. HARPER & J. T. HAGAN Fig. 3. Variation of the orientation of the maximum horizontal stress in a sector of the southern North St Red laboratory core (LIFE). Green: wellbore breakout, Yellow: field core (ASR). Natural fracture orientations denoted by black solid lines (open) and pecked lines (filled), North is aligned upwards parallel to the edge of the map. to be most likely to occur in the low-stress regions between the relatively high-stress-force chains which are more likely to remain in compression, ‘The microcracking, indicated by acoustic emission when extracted cores are perturbed in the laboratory (Harper & Szymanski 1991), may contribute to ani- sotropy of Young’s modulus in the horizontal plane (Holt & Kenter 1992). We suggest that the stress fabric (force chains) control the strain anisotropy when the sample is perturbed by some means and the microcracks are a less influential feature of the ‘micromechanical evolution, Strains induced by step changes of temperature ‘The microcrack hypothesis has been tested by thermal cycling tests similar to the test reported in Harper & Szymanski (1991). Typically, step changes of 5°C or 10°C above ambient were imposed. In almost all the many such tests we con- ducted, the thermoelastic anisotropy was substan- tially different to that of the time-dependent strain anisotropy. This demonstrated that the orientation of the crack fabric (indicated by the thermoelastic strain anisotropy) and of the time-dependent strain tensor are different for almost all the sandstones which we tested (mostly from southern North Sea gas reservoirs). Zang ef al. (1996) also found such differences in eight crystalline rock cores, observing that the ‘residual strain azimuth differed from the orientation of cracks and maximum wave speed’ Following a discussion with one of the authors, Yassir & Robertson (1994) tested the strain induced at successive temperature increments, rather than cycles of temperature change, in two carbonates and to much higher temperatures than employed in most of our experiments. They found that the orien- tation of the maximum horizontal strain induced by each increment was perpendicular to the orien- tation of hydraulic fractures in the region. This is ings that the thermoelastic strain anisotropy is quite different to the time- dependent strain anisotropy. This supports the above conclusion that microcracking, which mi nate thermoelastic anisotropy, is a feature of the time-dependent strain changes but not the dominant factor. The time-dependent, permanent strain aniso- tropy induced by minor temperature cycles can be used to indicate the orientation of the maximum in situ normal stress. This type of test is ‘non- destructive’, but the LIFE method was more usually adopted in practical applications at BP. ‘An alternative experimental examination of the response of unconfined core samples to temperature change involved application of a confining pressure in an autoclave. The temperature of a sandstone core from a reservoir in the northern North Sea, confined by a hydrostatic stress of 3.5 MPa, was raised to 30°C for 22.6 hours then raised to 60 °C for 41.2 hours before returning the temperature t0 30°C (Fig. 4). After 52.6 hours, the boundary condition was changed from constant stress (3.5 MPa) to con- stant strain and the core dimensions were restored “Table 1 Ac of maxi rina rest wer ‘aon Formation FE ASK ‘Welorebealcou Denk Arinwi) sepia ——epih Arm oo ‘ace ™ ‘ae o ice) ein UK SNS. Ie emimenguome—asasiesay 17s Mos SS 48/60 UK SNS, 1 Leman Sandstone dong 290808145 29385-2930 160 Bien UKSNS Totem Sinisowe 32 6E312619 148 Ugh fo Uicondow Ege anon hosesio27 oe . Renguonet UR ombore Danian Limsone 96845-8659 123G.I6-86R.35 150 2iiianG? Uk NonbScs Magn Sniuone 3307 7-39NSt 13H Kap DuchRorass heres Mewes SsoLe-ssoaan 140 sou swors2 14s 206/88 Wot Sheand Okt Red Sandstone 1y94.27-216906 thd ins Worsham! — OW Redsandame —f33ese-inioas 1S 500-1870 ° foyimwao—URNonsSes Fala Same Moras anes 5s 3650-800 a Linasiat Well Ohie, USA UcEwchirFormation "2IS72-as108t 8 «2EL12-SISRS BT ae) Chism 2s Calon ‘Gti Fomaton ios 1 100 T.R. HARPER & J. T. HAGAN 0c 400 go B20 100 og 5 og : ok & of 5 sme: (secon) Fig. 4, Change of applied stress required to return a confined example of North Sea sandstone sample t its initial dimensions after a 30°°C temperature eycle. to those pertaining prior to the temperature cycle. The stress required by the load frame to restore the strain condition was recorded, Figure 4 shows that both the axial and radial tractions decreased as a result of either the temperature cycle or the initial load application or both. Whatever the mechanism, involved, the essential point is that the process led to ‘permanent’ strains and was irreversible. Strains induced by a diametral saweut Each saweut induced a time-dependent extension of a vertical plug, typically anisotropic in the plane of bedding. An example was reported in Harper (1995). Typically, the duration of the strain devel- opment measurable by the strain gauges was approximately one day. Brujic et al. (2005) simu- lated stress relaxation in a uniaxially compressed «granular medium employing two damping mechan- isms: viscoelastic damping at grain contacts and a global damping simulated by immersion ina viscous fluid, Slow strain relaxation could only be observed above a critical damping, either contact or global. It can be speculated that cement around and/or between the grains in our sandstones, argil- laceous or otherwise, might provide a damping ‘mechanism. However, the presence of cement does not seem to be necessary for a time-dependence to be observed, as noted by Skinner (1974) who ‘monitored creep recovery of previously loaded clean sand. Sensitivity to perturbation Unander & Holt (1998) investigated the stress sur- face in space that encloses all stress states. that a reservoir sandstone can be changed to from the original state without damaging the original mate- rial, Damage was used by these authors to mean irreversible alterations to the rock. Unander & Holt (1998) created an artificial sandstone which was cemented under load in a triaxial cell. They detected damage by means of acoustic emission, In two tests, samples were subjected to a series of small stress cycles to check that no damage was incurred when the stress was cycled within the stress surface. The authors concluded that the micro- mechanical strength, with damage detected by the acoustic emission as noted, is negligible in the direc- tion of increased shear. Unander & Holt (1998) concluded: ‘This study does not support the assumption that a reser voir rock has a “safe” region within [which] the stress may be varied clastically without damage being induced in the rock. This means that itis not sufficient to use simple, Tinearised models when simulating eformation of a reservoir... For a normally consoli dated material, microscopic strength may be close 10 zero for shear loading even if itis cemented In the overwhelming majority of tests our sam- ple sandstones developed measurable, permanent strains in response to small temperature cycles (as low as 5 °C) or saweuts. For these types of tests which avoid imposing either predetermined strain rates or servo-controlled boundary conditions, sam- ples were found to be sensitive to processes which involve only minor quantities of energy transfer in relation to most engineering processes. (Some energy is transferred to the core by the work done by the diamond saw, including frictional heating which is limited by the flow of cooling fluid to the saw.) ‘An experiment designed to test the sensitivity of sandstone core to remote perturbation was re- ported by Harper (1995). A core of length 255 mm and 100mm diameter was strain-gauged at one end. A small, circular hole of depth 20 mm and diameter 8 mm was drilled at the opposite end. This amounts to a ratio of the distance of the gauges ‘THE NATURE OF STRESS STATE, 101 from the nearest point of the hole to the maximum dimension of the hole of more than 25:1, Small (c. 10 microstrains), permanent strains developed in response to the drilling of the hole, displaying a time-dependence. The hole was drilled in preserved, whole core (normally saturated) in two steps of 10 mm, allowing a lag of two days between steps. No strains were observed after the first step. Had each drilling increment created enough tempera- ture rise around the drillhole to detect strains at the remote gauges, a detectable effect should have been registered after the first 10mm, Approxi- mately 240 mL of brine was used to cool the drill bit during drilling, In the BP experimental pro- gramme, samples of sandstone were normally chosen which were sufficiently clean so that swel: ling strains would not complicate interpretation of the results. Given that the core was already satu- rated, transmission of fluid across the core by capil- lary action is not likely to have occurred. It is unlikely that any undetected swelling occurred in the vicinity of the drilled hole, but this possibility ‘was not eliminated. Even if this had been the case, it does not invalidate the conclusion that long-range transmission of the minor disturbance at the drilled end of the core was recorded. Interestingly, the magnitude of the strain induced by a sawcut in the LIFE experiments appeared to decrease with successive cuts. Although a depen- dence of the magnitude of strains upon the volume of the sample subject to a saweut has not been ruled out, this could alternatively indicate a decreas- ing sensitivity to perturbation. This would be con- sistent with the durability of sandstones in natural situations, such as the diurnal temperature changes to which the surface of an outcrop is repeatedly exposed without a complete loss of stored energy (eg. Engelder & Sbar 1976) or disintegration. An incidental test of the ability to retain strain energy ‘when subject to even quite severe diurnal fluctuation ‘was reported by Harper & Chambers (2004). These authors documented successful LIFE tests to deter- mine stress orientation on unpreserved core which had lain in an uninsulated core shed in the West Indies for 40 years. Changes of effective stress ‘Two experiments were conducted to test the effect of one or more cycles of applied effective stress on, respectively, the induced strain and the change of applied stress required to restore the sample to its initial dimensions. Figure 5 shows that exten- sional strains resulted from a hydrostatic stress eycle of 4.8MPa. In the same manner as unconfined saweut tests and thermal cycling tests, perma- nent extensional strains resulted. This demonstrates sample evolution when subjected to (a cycle of) Fig. 5. Strain response ofa North Sea reservoirsandstone toa hydrostatic stress cycle. (a) Hydrostatic pressure history. (b) Strain evolution (horizontal plane). (e) Strain evolution of dummy sample. (4) Temperature history surface forces, although a contribution resulting from minor ambient temperature changes (Fig. Sd) ‘cannot be ruled out ‘The second experiment involved changes of effective stress successively by cyclically incr ing the boundary stresses and then the pore pres- sure (Fig. 6). A sandstone from a southern North Sea reservoir was confined at 3 MPa and kerosene flowed through the sample. Axial and radial stresses of 10.6 MPa and 9.7 MPa, respectively, were then applied with a pore pressure of c. | MPa. After 44 hours, the pore pressure was increased to 6.7 MPa for a further period of 44 hours before being restored to 1 MPa, The constant stress boundary condition was maintained for 47 hours before using the servo system on the load frame to return the sample to the same dimensions (macroscopic strain state) as 102 T.R. HARPER & J. T. HAGAN 2000 1500 ‘STRESS 8 500 TIME (seconds) before the pore pressure cycle. The required axial sttess was 11% lower and the radial stress 4% lower. The sample also did not behave elastically and evolved following the application of effective stresses well below its compressive strength. Discussion ‘The results of the numerical and core experiments are entirely inconsistent with ‘the assumed notion of uniform and constant stress states’ disputed by Couples (2013). They characteristically demon strate or imply a heterogeneity of stress state at all les and corresponding energy storage, irrevers- ible processes and interaction at a wide range of scales. Interaction between slipping fractures gives rise to one mechanism of time-dependeney in deformation. A state of marginal stability has been implied for some situations. A memory of past stress states can be demonstrated. Memon y of past stress states Mechanical engineers have found it useful to divide residual stresses. into classes defined by length scales ranging from the scale of crystal defects to the macroscopic scale (IAEA 2014). It is also con- venient to consider the memory of past geological processes at a range of scales which relate to the scale of the experimentation, theoretical analysis, simulation or field observations employed, and here wwe cite examples for this purpose. Friedman (1972) used X-ray diffraction to inves- tigate the residual strains in sandstone grains and oncluded that he had mapped residual stresses relating to Mesozoic and possibly even to Precam- brian geological events. Friedman & Logan (1970) found that a ‘state of prestrain’ (determined by X-ray diffractometry) controlled the orientation of 6. Response of a southern North Sea gas reservoir sandstone to a cycle of reduced effective stress. experimental fractures induced by triaxial, point Toad and Brazilian testing in three quartzose sand- stones. These authors used core plugs of mainly 74 mm diameter for triaxial testing, plugs of diam- eter 111-268 mm for Brazilian testing and discs of 246-491 mm for point load testing. The only non-random fabric (other than the bedding) detected by the authors in these rocks was the trend of grain elongation. The direction of induced fractures was not consistently orientated with this trend. At this scale, greater than the scale of individual grains, the presence of a force chain fabric could now be inferred following the distinct element simulations which post-dated Friedman & Logan's (1970) work. Nichols & Abel (1975) overcored samples of granite, pegmatite and quartzite in the laboratory ing a 50 mm core bit. They concluded that large amounts of recoverable residual strain (potential) energy can be stored within most rock masses, suffi- cient to cause rock failures and some moderate earthquakes. Brereton & Muller (1991) used well- bore break-out analyses (representing a scale of a few metres) to reveal a major change in the orien- tation of the minimum horizontal stress in the English Midlands. This change in orientation coin- cided with an unconformity below which the sedi- ments had been subject to the Variscan Orogeny and above which they had been laid down after this tectonic event. Brady ef al. (1986) focused on the stresses induced in the vicinity of the termin- ation of a slipping fracture at the intersection with another fracture or bedding plane along boreholes of 50 m length and 1.81 m diameter, and interpreted the results to show an apparently non-systematic variation, These authors were able to dismiss lithol- ogy as the cause of the variation. Here we have discussed simulations of discontinuous fracture networks and given examples of field observations of normal stress at the kilometre scale, Because THE NATURE OF STRESS STATE, 103 the mechanism of ‘stress dipole” formation around a slipped fracture or fault is largely independent of the Iength of the slipping discontinuity, it is straightforward to extend the same reasoning to larger scales. In a letter advising on the interpretation of the stress state derived from a distinct element simu- lation, P. A. Cundall (pers. comm., 1986) succinctly emphasized the crucial role of the history of each location in the rock mass in determining the present equilibrium stress state: there is an infinite number of stress states that Satisfy equilibrium, The one that actually exists in the ground depends on the particular tectonic history at the location in question, with additional complications, thrown in from ercep, fracturing and so on. ‘The mechanism for persistence of previous stress states in rock requires further research. It is a sim- ple step to infer that a system of force chains can be stored by cementation. One could speculate that when a rock mass subsequently experiences th changes associated with typical geological pro- cesses, involving the formation of new systems of force or stress chains at a wide range of scales, some regions of the rock mass could retain the ‘pre-stressed’ state. Alternative perceptions of stress state Agreeing with the conclusion of Couples (2013) restated at the beginning of this paper, it follows that one or more alternative representations of the mechanical state should be identified. Couples (2013) proposed that appropriate simulation tech- niques should replace the simple assumptions about stress regimes. We have used a continuum mechanics repre- ie stress state in discon- tinuously fractured media and found this to be instructive. There are two main limitations to many continuum mechanics simulations: the impo- sition of artificial boundaries and the length scale assumed. Because of the interactions between the strain and stress changes around individual sliding fractures and the model boundaries, work ‘was done on our model at the boundaries to maintain the no-displacement condition we imposed. One typical solution used in practice is to use large models and assume that the region of interest (e.g. wellbore or tunnel complex) is sufficiently remote from the boundaries of the model that the state of the rock in the vicinity of the region of interest is insignificantly affected by the artificial boundaries (in accordance with St Venant’s principle). This is often adequate to give sufficient insight for engin- eering design but, given the considerations given above, is a simplification which should not be used without recognition of the simplifying assump- tions emphasized by Couples (2013) and extended here. Another solution can be to impose periodic boundaries. ‘This can be expected to better reflect the interactive processes, such as between slipping fractures as modelled here, but involves a choice of the area or volume of the model. When the stress state is heterogeneous, the results may depend upon the model dimensions. Finally, the boundary clement method as used by Brady et at. (1986) does not eliminate work done on the region of investigation associated with the enclosing elastic medium, A similar limitation arises at the small scale. The usual continuum mechanics description of rock stress assumes a scale A at which it is assumed that properties and forces can be averaged over a length scale which is intermediate between the grain or crystal scale d and the length scale of displacem- ents of interest L(L > A> d) (Nicolaevskiy 1996). ‘This ignores the grain or crystal scale “residual stre ses’, thereby implicitly assuming that they are insig nificant, and does not appropriately represent the force chains which are present in granular media, In geomechanics, it is impossible to include the wide range of heterogeneity which exists in the subsurface. Starfield & Cundall (1988) recog- nized that modelling in geomechanics should be different from that in mechanical engineering, ‘These authors recognized that it is futile to expect model to provide design data when there are massive uncertainties in the input data (Itasca 2008). Couples (2013) gave a clear example which illustrated the major influence of small changes in local flow pathways. Itasca (2008) therefore advo- cated that a numerical model is useful “for providing @ picture of the mechanisms that may occur in particular physical systems’. Itasca (2008) also pointed out that, for calculations involving discon tinuous materials, the results can be extremely sen- sitive to the initial conditions; they also note that the detailed evolution of chaotic systems is not predict- able, ‘even in principle’, and the best that can be expected is a finite spectrum of possibilities. tase (2008) point out that in most nonlinear, inelastic systems there are an infinite number of solutions that satisfy equilibrium and all possible solutions are correct if the path is not specified ‘We emphasize the limitations of the typical con- tinuum mechanics approach for two reasons. First, it is essential to be aware of these limitations and of the fundamental point that our simulations are no more than numerical experiments which give us insights into the mechanisms which may occur We may be able to derive simulation results which are consistent with our interpretations of field observations, but this does not imply that we have modelled the ‘correct’ solution, Second, the 104 T.R. HARPER & J. T, HAGAN limitations imply that if we can develop alternative or complementary perceptions of geomechanical processes such alternatives may be helpful, even if these perceptions also must involve limitations. Stresses at the grain scale which are normally ignored in continuum mechanics representations of rock stress (Nicolaevskiy 1996) are normally termed ‘residual stresses’, Amadei & Stephansson (1997) describe the term “residual stresses’ as a ‘much used and abused term.’ The terminology of residual stresses is derived primarily from mechan- ical engineering where some processes of fabri- cation, such as autofrettage of gun barrels and welding, resulted in a balanced internal system of forees in the absence of applied (surface) forces, Nadai (1950) defined residual stresses as follows: Ifa part or several parts of a body have been strained permanently beyond the limits of plasticity and the external forces and moments are removed, the mate rial in the overstrained region and around it will, in general, be subjected to inherent stresses which are then called residual stresses ... The term inherent stress is proposed as a general term for this kind of internal siress created by the body itself, residual stress being, reserved for the case where permanent strain is its Nadai noted that inherent stresses may be present in combination with tractions in certain circum- stances, providing that the resultant forces and resul- tant moments of the tractions vanish, Timoshenko & Goodier (1982) refer to initial stresses resulting from non-elastic deformations during the process of forming a body. In the geology and rock mech- nies we are normally concerned with subsurface conditions under which the ‘external forces and moments’ (Nadai 1950) have not been removed, yet numerous authors haye referred to both residual and locked-in stresses. Friedman & Bur (1974) stated In the general field or subsurface conditions, applied strains and loads are superposed on residual strains and stresses, Here, we have emphasized that the very process of applying strains or loads to at least certain rocks, such as the sandstones we have tested, causes irre versible change, even for very small magnitudes of strain and load, The process is not one of linear superposition as applicable for linearly elastic pro- cesses. This is an essential difference from the definition of residual stress derived for mechani- cal engineering purposes. Even the terminology adopted from mechanical engineering, usually so helpful, can in this instance be misleading. We have the choice of retaining the misleading termi- nology, seeking 4 new terminology or simply incor- porating ‘residual stress’ in the wide range of stress-state heterogeneity emphasized above. From the observations given above, the mechan- ical state of rock changes continually with time. The influence of mechanical thresholds associated with frictional and cohesive strength implies that changes are discrete rather than continuous. (These thresholds are one explanation why “residual” stres- ses have been perceived as ‘locked in’). Systems that change with time are well studied and known as dynamical systems, The state of such dynamical systems is the memory that the system possesses of its past (Mees 1981) such as, for example, the contents of a warehouse. If the system is determinis- tic its evolution is always completely determined by its current state and history (Thompson & Stewart 1986). Such systems can be expressed as follows (Mees 1981). If x(t) is the state of the system and summarizes the past history of the system, and the environment influences the system by means of a variable u, the state of the system is given by. dr GD =F. Wl, t} a ‘The output y(0) is described by: yO) = e[x, ul), 1) Q where f and ¢ are vector fields, 1 is the variable describing the influence of the environment on the system and 1 is time. These equations define an ordinary dynamical system (Mees 1981). To quan- tify these expressions in terms of a given body of rock is impracticable. This does not detract from the insight which the expressions convey. Equations (1) and (2) describe a dynamical system which is described by ordinary differential equations. Other kinds of systems can be described, and the inclusion of Equations (1) and (2) here is not meant to imply that any heterogeneous geomechanical system can be described as an ordinary dynamical system. Equations (1) and (2) serve to emphasize that the state of the system is a function of its history, chan- ging external influences and time. Much more research is required to better describe the state of the complex subsurface stress systems expressed as a dynamical system. Recognition that a geologi- cal state with heterogeneous stress existing at a wide range of scales can be represented as a dynami- cal system is only a first step in identifying alterna- tive perceptions of geomechanical state which may prove helpful in some situations. Some aspects of the applications of dynamics to Earth science have been explored (Hergarten 2002; Ribiero 2002). An alternative perception sug- gested by Harper & Szymanski (1991) is to treat geomechanical processes not solely as mechanical processes but as thermodynamic processes. Geo- logical systems are not isolated and exchange ‘THE NATURE OF STRESS STATE, 105 energy or energy and matter with their environment. Systems which are not isolated are capable of self- organization (Nicolis & Prigogine 1989), consis- tent with heterogeneous stress states organized into patterns such as we have simulated for discontinu- ously fractured rock and others have simulated for granular media. We recognize that there is likely to be little support for alternative means of perceiving stress state while the conceptual model of stress state as uniform and constant continues to be so universally favoured, Consequently, here we have concentrated on those observations which demonstrate or infer that stress state is heterogeneous at a wide range of scales and sensitive to change. At present, it is normal practice to assume that the state of stress throughout the body of rock enclosing any given volume of geological or engineering interest is uniform and can therefore be described by a sin- gle stress tensor representing some average of the heterogeneous state, We suggest that it would, ideally, be appropriate to justify this assumption but, ay a minimum, to state the assumption prior to its adoption for any given investigation of geologi- cal or engineering behaviour. “The laboratory experiments reported here were financed and supported by BP ple and the numerical experiments by Geosphere Lid. We are grateful for the helpful com- iments of the reviewers. References AMADEt, B. & STEPHANSSON, O. 1997, Rock Stress and its ‘Measurement. Chapman & Hall, London. Brapy, B. H. G., Lemos, J. V. & CUNDALL, P. A. 1986. Stress measurement schemes for jointed and frac~ tured rock. In: SrEPHANSSON, O. (ed.) Rock Stress and Rock Siress Measurements. Centek, Lule’, Sweden, 167-176. Bueaeron, R. & Mutek, B. 1991, European stress: tributions from borehole breakouts. In: WHITMARSH, R.B., Bort, M. H. P., Fairitean, J. D. & Kuszir, N. J. (eds) Tectonic Stress in she Lithosphere. The Royal Society, London, Buusic, 1, WANG. P., SONG, C., JOHNSON, D.L., SinDt, O. & Makst, H. A. 2005. Granular dynamics in compac- tion and stress relaxation. Piysfeal Review Letters, PTL 95, 128001-1-128001-4, CoupLes, G. D. 2013. Geomechanical impacts on flow in fractured reservoirs. In: SPENCE, G. H., REDFERN, J. AGUILERA, R., BEVAN, T. G., Coscrove, J. Wa. Coupes, G. D. & DaNtEt, JM. (eds) Advances in the Study of Fractured Reservoirs. Geological Society, London, Special Publications, 374. CCxasry, $. 1999. Caleulable fluid-rock interactions. Jour nal ofthe Geological Society, London, 136,501~314. Cunpatt, P. A. 1989, Numerical experiments on localiz- ation in frictional materials, Ingenteur-Archiv, 59, 148-159. (CUNDALL, P. A., DRESCHER, A. & STRACK, OD. L, 1982, ‘Numerical experiments on granular assemblies: mea surements and observations. fn: Proceedings of the IUTAM Symposium on Deformation and Failure of Granular Materials. Delfi, Balkema, 355~370. Dawanac, B., CUNDALL, P. A. & BraNDsitauG, T. 2004. Tasca presentations, Workshop on extreme. ground motions at Yucea Mountain, Menlo Park, Califor- nia. http://pubs.usgs.gov /of/2006/1277 appendixes appendix. articles/12_ Damjanac.pdt DeNKHAUS, H. 1967. Residual stresses in rock masses General Report Theme IV, Proceedings of the Ist Con: gress ofthe International Society for Rock Mechanics Lisbon, 3, 312-319. Exortprr, T. & Smar, M. L. 1976, Determination of the Regional Stress Patterns in New York State and ‘Adjacent Areas by in situ Strain Relief Measurements Report NYSERDA-15/182, Lamont-Doheny Geo- logical Observatory, Palisades, NY, November. Farepnax, M. 1972. Residual elastic strain in rocks. Tec tonophysies, 18, 297330. Fanpwax, M. & Bur, T. R. 1974. Investigations of the relations among residual strain, fabric, fracture and ultrasonic attenuation and velocity in rocks. Jnrer national Journal of Rock Mechanics, Mineral Sciences & Geomechanics Abstracts, 11, 221~234. Farepax, M. & Loaan, J. M. 1970, Influence of residual elastic strain on the orientation of experimental frac ‘ures in three quarizose sandstones. Journal of Geophy sical Research, 75, 387-405, HawerR, T. R. 1995, Dead or alive: two concepts of rock behaviour. In: Dav, P. & Cartow, CR. A. (eds) Bicycling to Utopia, Essays on Science and Technol- ‘ogy. Oxford University Press, The Royal Institution, Oxford, 4364. Haweer, T. R. 2011. Well Proese Hall, The mechanism ‘of induced seismicity, Geosphere Ltd. report prepared for Cuadrilla Resources Lid, 10 October. Available at_hiip:// www cuadrllaresources.com/wp-content/ uploads/2012/06/Geosphere-Final-Report pdt Hanees, T.R. 2014, Effective stress history and the poten- tial for seismicity associated with hydraulic fracturing. Jounal of the Geological Society, London, V71, 418-492, map://do: 10.1144/jgs2013-149 Hanerr, T.R. & Chrawers, J. L. 2004. Stress state and its influence on drilling performance in the Brighton Marine Field, Trinidad. Marine and Petroleum Geol ogy. 21, 947-963. HaweeR, T.R, & Last, N.C. 1990. Response of fractured rock subject to fluid injection, Part Il. Characteristic behaviour. Tectonophysics, 172, 33-51 Haweer, T. R. & SZYMANSKI, J. S. 1991. The nature and determination of stress in the accessible lithosphere. Philosophical Transactions of the Royal Society of London A, 337, 5-24. Herre, K.. Greexnovan, J.. MAIN, L G., Zi1aNa, X HUSSEIN, A. M, & Koursane.ouis, N. 2010. Low ‘cost monitoring of inte-well reservoir communication paths through correlations in_ well rate fluctuations: cease studies from mature fields in the North Sea. Society of Petroleum Engineers, Paper SPE 130734. er, K. J. & Dowoxror, A. B. 1990. Relationship between azimuths of flood anisotropy and local earth stresses in oil reservoirs. In: BULLER, A.T., BERG, E. 106 TR. HARPER & J. T. HAGAN HUELMELAND,O., KLEPPE, J, TORSETER, O. & AASEN, J. 0. (eds) North Sea Oil and Gas Reservoirs ~ 11. ‘The Norwegian Institute of Technology, Graham & ‘Trouman, London. Henaarten, S. 2002. Seif-Organized Criticality in Earth Systems, Springer, Heidelberg Hout, R. M, & KENTER, C. J. 1992. Laboratory simu- Iations of core damage induced by stress release. fr TILLERSON, J. R. & Waversix, W. R. (eds) Proceed- ings 33rd U.S. Rock Mechanics Symposium. Rotter dam, AA Balkema, 959-968, IAEA 2014, Development and applications of residual stress measurements using neutron beams. Technical Reports Series No. 477, International Atomic Energy Agency, Vienna, ITASCA 2008. FLAC Fast Lagrangian Analysis of Continua, Users Guide, Minneapolis, Minnesota. JENSEN, H. J. 1998. Self-organized crinicalty, Cambridge University Press, Cambridge Lecture Notes in Physics. Kerrex, A. A., Heinze, J. R., Danusts, J. L. & WATERS, G. 2008. 4 field study in optimizing completion strat- egies for fracture initiation in Barnett Shale horizontal wells. Society of Petroleum Engineers, Production & Operations, 373-378, Lacy, L. L. 1984, Comparison of hydraulic fracture orien tation techniques. Proceedings of $9%h Annual Techni cal Conference and Echibition, Houston, Texas, Paper SPE 13225, Mess, A. 1. 1981. Dynamies of Feedback Systems. John Wiley & Sons, Chichester. Napa, A. 1950, Theory of Flow and Fracture of Solids ‘MeGraws-Hill, New York. Nicuois, T. C. & Anet, J. F, 1975, Mobilized residual energy ~ a factor in rock deformation, Bulletin of the Association of Engineering Geologists, XML, 213~225. NIcOLAEVSKIY, V. N. 1996. Geomechanics and Fluido- dynamies, Kluwer Academic Publishers, Dordrecht Nicoiis, G. & Paicoaine, I 1989. Exploring Complexity Freeman, Oba, M. & Konisnt, J. 1974, Microscopic deformation ‘mechanism of granular material in simple shear. Soils and Foundations, 14, 25-38, Pouano, D. D. & SEGALL, P. 1987, Theoretical displace- ‘ments and stresses near fractures in rock: with appli cations to faults, joints, veins, dikes, and solution surfaces. In: ATKINSON, B. K. (ed.) Fracture Meche anics of Rock. Academie Press, London, 277-428, RimeRo, A. 2002. Soft Plate and Impact Tectonics. Springer, Berlin, Heidelberg \T. C., SHAW, J., HARPER, T. R. & HAGAN, J. T. 1995. Viscous remanent magnetization: a tool for Rot orientation of drill cores. fn: TURNER, P. & TURNER, A. (cds) Palacomagnetic Applications in Hydrocarbon Exploration and Production. Geological Society, London, Special Publications, 98, 230-243. Skinner, AE. 1974. The Effect of High Pore Water Pressures on the Mechanical Behaviour of Sediments PHD thesis, University of London. Srawriei, A’ M. & CUNDALL, P. A. 1988. Towards a methodology for rock mechanics modelling. Inter national Journal of Rock Mechanics, Mining Sciences & Geomechanics Abstracts, 25, 99~ 106. ‘Tevret, L. W. 1989. Acoustic emissions during anelastic Sirain recovery of cores from deep boreholes. Watian Kiara, A. & Wautan, A. (eds) Rock Mech- ‘anics as a Guide for Efficient Utilisation of Natural Resources: Proceedings of the 301k US Symposium, West Virginia, 19-22 June 1989. Balkema, Rotterdam, 269-276, ‘Tuowrson, J. M. T. & Srewanr, H. B. 1986, Non linear Dynamics and Chaos. John Wiley & Sons, Chichester ‘TuoRNTON, C. 1990. Induced anisotropy and energy dissi pation in particulate material~results from computer= Simulated experiments. Ir: Bower, J. P. EGFS (cd.) Yielding, Damage and Failure of Anisotropic Solids, Mechanical Engineering Publications, London, 113-130. ‘TuoxnToN, C.& BARNES, D. J. 1986, Computer simulated deformation of compact granular assemblies, Acta Mechanica, 64, 45-61, ‘TimosHENKo, 8. P, & Goopter, J. N. 1982. Theory of Blasticty. McGraw-Hill Book Co., Singapore, Unanver, T. E, & Hout, R. M. 1998, Virgin damage surface of a symthetic analogue to a reservoir sand- stone. In: SPE/ISRM Eurock ‘98, Trondheim, Paper SPE/ISRM 47201, 23-32. Wanrivsei, NR, True, LW. Lorenz, LC. & Horcous, D. J. 1993. Core based stress measure ments: a guide to their application. Sandia National Laboratories teport for Gas Research Institute, contract no, 5089-211-2059, WILLEMS, T. & DEGEARE, J, 2014. Quest for the most effi- cient feld-track network, Hart's E&P, July. Yassin, N. A. & ROBERISON, D. G. 1994. The use of thermal stain relaxation of core to determine stress directions. In: NELSON, P. P. & Laumactt, S. E. (eds) Rack Mechanies. Balkema, Rotterdam, 401—408, ZANG, Au. LIENERT, M., ZINKE, J. & BERCKHEIMER, H. 1996. Residual strain, wave speed and crack analysis of crystalline cores from the KTB-VB well. Tectono- physies, 263, 219-234. image not available image not available a You have either reached 2 page thts unevalale fer vowing or reached your ievina tit for his book. a You have either reached 2 page thts unevalale fer vowing or reached your ievina tit for his book. a You have either reached 2 page thts unevalale fer vowing or reached your ievina tit for his book. a You have either reached 2 page thts unevalale fer vowing or reached your ievina tit for his book. a You have either reached 2 page thts unevalale fer vowing or reached your ievina tit for his book. a You have either reached 2 page thts unevalale fer vowing or reached your ievina tit for his book. image not available image not available PRESS 1990; Spiers et al. 2004; De Meer er al. 2005; Zhang, and Spiers 2005; Zhang et al. 2010; Koelemeijer et.al. 2012). Our calculations on creep rate v. &/do, made using the full set of models, are compared in Figure 4. This shows that diffusion-controtled pressure solution is predicted to be 1-3 orders of magnitude slower than either dissolution-controlled (S) or precipitation-controlled (P) pressure solu- tion in anhydrite under the conditions of the experiments reported by Pluymakers er al. (2014) (Fig. 4), regardless of the specific diffusion model applied (DI—D4) and despite the range in the values taken for the grain boundary diffusion prod- uuct DS, Note here that the available data on anhy- drite dissolution kinetics determined by Blount & Dickson (1969) likely underestimate the true rates of dissolution and precipitation. This is because those reactions are so fast they are difficult to measure independently of diffusion effects. The implication is that pressure solution in anhydrite will generally be diffusion controlled for a wide range of grain sizes, effective stresses and tempera- tures as well as those shown in Figure 4. Comparison of the diffusion-controlled pressure solution models DI and D4 with each other in Figure 4a shows they are almost indistinguishable for each of the chosen values of the parameter DS. Regarding the linearized models for diffusion- controlled creep, that is, models D2 and D3 plotted in Figure 4b, these are equivalent to each other and so give identical results despite the differ- ence in derivation. They yield lower creep rates than either Di or D4 at normalized porosities 4/0 > 0.95, thats, for small grain-to-grain contact IRE SOLUTION OF ANHYDRITE GOUGE 7 areas and high contact stresses. At 4/Gy < 0.9 however, all four diffusion-controlled models yield strain rate predictions within a factor of 2 for any given value of DS and at }/dy < 0.6 the four models are indistinguishable for practical purposes. Figure 5 compares the experimental data for the wet fine-grained samples of Pluymakers et al. (2014) with the predictions of the diffusion- controlled rate models (D1—D4) applied for the same DS values used in Figure 4. This shows that the experimental strain rates match the diffusion- controlled rates within one order of magnitude, although the sensitivity of compaction strain rate to changing porosity (expressed as. &/dp) is much higher for the experimental samples than predicted by the diffusion-controlled models. At the same time, the experimentally determined strain rates are 1001000 times lower than the rates predicted for dissolution- and precipitation-controlled pres- sure solution (compare Figs 4 & 5). The similarity between the diffusion-controlled pressure solution models and the experimental strain rates strongly suggests that our diffusion-controlled pressure solution models offer a rough estimate and explana- tion of the compaction creep rates measured for the fine-grained anhydrite fault gouges, despite differ- ‘ences in assumed v. real grain shape, packing and grain size distribution. This supports the inference by Pluymakers ef al. (2014) that the behaviour of fine-grained anhydrite is dominated by pressure sol- ution and specifically by diffusion-controlled pressure solution. The order of magnitude agree- ment between the creep rates predicted by our diffusion-controlled models and the experimental data also imply that the models provide a basis for Table 2, Values of the parameters used in applying the present pressure solution models Symbol Definition Value /range Source and additional information (where applicable) @ Geometric factor 09 ‘Van Noort & Spiers (2009) A Geometric constant 4 Average value for simple cubic pack , of grains Cy Anhydtrite solubility (m* m~*) 4.74104 Blount & Dickson (1969) Ds Product equal to diffusion coefficient 10" to 107*” Range determined for diffasion- D times mean grain boundary fluid ‘controlled pressure solution in other thickness (ms) ionic compounds such as NaCl and calcite (e.g. De Meer et al. 2005; Koclemeijer et al. 2012; Zhang et al. 2010) F Grain shape factor = Value for simple cubic pack of grains ke Rate constant for dissolution/ 5.1 x 10-7 Minimum values determined by precipitation of anhydrite (m s-') pane Bildstein er af. (2001) R Gas constant (J mol-! K!) 8314 For example Chang (2000) Zz Coordination number 6 Value for simple cubic pack of grains 2 Molar volume of anhydrite CaSO, 4.6 x 10° Hummel e¢ ai, (2002); Thoenen & (n° mol") Kulik (2003) a You have either reached 2 page thts unevalale fer vowing or reached your ievina tit for his book. a You have either reached 2 page thts unevalale fer vowing or reached your ievina tit for his book. 120 rotation and translation, will most likely cause a (near-instantaneous) porosity decrease, whereas pressure solution will dominate the longer-term compaction behaviour and hence the time to sealing. For an effective stress of c. 10 MPa and a temperature of 80 °C (as used in our experiments), assuming pressure solution is the only process operating and taking d= 35 jm and DS 10°? m3 s- (in tine with the asymptotic approach of the experimental data in Figure Se towards the ‘model predictions for this value of DS) and using the data shown in Table 2 for diffusion-controlled pressure solution, the estimated sealing time for an anhydrite-bearing fault is around 10 years. However, for reservoir-bounding faults (cross- cutting the caprock) in a COp storage scenario, a 10 MPa effective normal stress is relatively low, assuming hydrostatic fluid pressure (e.g. Ramm 1992) and a normal stress equal to the overburden pressure. Using parameter values in Equation (26) (Model D4) listed in Table 2 plus a temperature of 80°C, but varying the effective normal. stress between 5 and 40 MPa (corresponding to depths of 1-3.5 km), shows that sealing times decrease A. M. H. PLUYMAKERS & C.J. ERS. rapidly with increasing normal stress and thus depth (Fig. 6). Additionally, in nature fault gouges consist of a broad range of (fine) grain sizes (e.g, Logan et al. 1979; Keulen et al. 2007) which speed up compaction creep (Niemeijer et al. 2009) and hence fault sealing. In practice sealing should occur over a few decades or less, which is short compared to the time scales relevant for COs storage (thousands of years; IPCC 2007). The impli- cation is that even if faults in an anhydrite caprock system are reactivated during the active CO> injec- tion phase, they should self-seal quickly enough to pose only a small long-term leakage risk. This will be especially so for a storage reservoir that is underpressured with CO; (i.e. the CO, pressure is less than the pore fluid pressure in the surrounding rocks), such that penetration of CO into the fault gouge and possible reaction with the gouge are avoided. Similar arguments apply for geological storage of methane and hydrogen. If faults in anhydrite formations can seal by compacting to approach the percolation threshold on time scales of a few decades and less, it is likely that fault healing or re-strengthening will depth (assuming P=0.850,) [km] 2 A 6 8 10 12 depth (assuming hydrostatic P,) [km] ¢ 3 o — 5 > i a: oo 4 Fig.6. Fault gouge sealing i as predicted by diffusion-controlled model D4 for gr indi 12 wv es effective stress [MPa] fora porosity decrease of 48.5 —> 3% (assumed to represent the percolation threshold), eters (d) between 5 and 80 jum as a function of effective normal stress (MPa) and depth (m), at constant temperature of $0 ‘C. Depth values are calculated assuming the normal stress on the fault equals the effe ve stress caused by the overburden pressure (4%). Depth values in black are representative of a storage reservoir setting, assuming a hydrostatic pore fluid pressure head, whereas depth values in fey are representative of the Italian Apennines where P; = 0.850, PRES occur even more rapidly following a fault reacti- vation event. This has implications not only for strength evolution following fault reactivation that may occur in geological storage systems, but also in relation to natural seismicity in anhydrite- bearing terrains. In the central Apennines in Italy, many of the destructive earthquakes and after- shocks that characterize the region are known to nucleate in the interbedded carbonate /anhydrite cover sequence known as the Burano Formation. In this region, high fluid pressures related to mantle degassing (up to 85% of the lithostatic pressure) are found at seismogenic depths (Collettini & Barchi 2002; Chiodini et al. 2004; Miller et al. 2004; Trippetta er af. 2013), leading to relatively Iow normal stresses. Assuming then that the pre sence of Ca-sulphates in exposed faults (De Paola et al. 2008) indicates that faults at depth contain anhydrite, application of our pressure solution model is reasonable for the expected normal stress range and for typical fault gouge grain sizes of 10-80 jum (cf. our experiments). Consider now the model predictions of fault sealing times for a temp- erature of 80 °C shown as a function of effective normal stress in Figure 6, Assuming the normal stress to be lithostatie with a pore fluid pressure of 85% of lithostatic, as appropriate for the Apennines, these predictions should roughly apply for depths between c 2 and 12.5 km (see upper scale, Fig. 6). ‘Taking the temperatures expected at depths beyond 3km, that is, higher than 80°C into account, the fault sealing times calculated in Figure 6 will likely be overestimates, since activation of other defor- mation processes such as crystal plasticity will tend to speed up healing/sealing. Seismic recurrence times for major earthquakes in the Apennines region (M > 5) are of the order 2000-5000 years (Pantosti et al, 1993; Cello et ai. 1997; Palumbo et al. 2004; Galli et al. 2008). This means that if anhydrite fault gouge is present within the fault cores it has the time to fully seal and heal between major ruptures, and that fault strength recovery in anhydrite-dominated faults is not the factor that controls the repeat frequency of events with M > 5, For such faults, the repeat frequency is more likely controlled by tectonic loading rate, pore fluid pressure build-up due to the natural CO, accumu- lation reported in the Apennines (Chiodini et al. 2004: Collettini et al. 2008; Walters et al. 2009; Ciotoli er al. 2013; Trippetta et al. 2013) or factors such as the re-strengthening behaviour of faults cut- ting the carbonate units present in the carbonate- anhydrite cover stratigraphy (ie. by the behaviour of calcite and especially dolomite-rich gouges). In contrast, since earthquake repeat frequency scales with magnitude according to the Gutenberg-Richter law, anhydrite healing and sealing may well play a role for events with M <5. IRE SOLUTION OF ANHYDRITE GOUGE 121 Conclusions We have derived kinetic models for compaction of granular aggregates by dissolution-controlled, precipitation-controlled and diffusion-controlled pressure solution, clarifying and/or avoiding many of the assumptions made in previous work, and explicitly including the effect of aggregate porosity on strain rate, We have compared our models with the experimental results on compaction creep of wet anhydrite fault gouge, previously reported by Pluymakers et al. (2014). Our conclusions are as follows. (1) Regardless of the detailed assumptions and simplifications made in deriving pressure solution models, diffusion-controlled pres- sure solution in anhydrite fault gouge will be rate controlling for grain sizes greater than 1-2 1m, at least for effective stresses of the order of 10-40 MPa and for upper crustal temperatures. This supports the conclusions ‘on the experimentally determined compaction creep rates for fine-grained anhydrite fault ‘gouge reported by Pluymakers er al. (2014), that is, compaction most likely occurred by diffusion-controlled pressure solution. (2) Kinetic models for diffusion-controlled pres- sure solution, derived by either classical approaches (e.g. Rutter 1976) or using a dissi- pation approach (Lehner 1990; Spiers & Schutjens 1990), produce essentially identical predictions for the compaction creep rate of fine anhydrite fault gouges for normalized porosities (b/d) below 0,8-0.9 (i.e. absolute porosities below 40-45%), regardless of approximations generally made in these models relating to grain contact stress magei tude (low stress approximation) and grain boundary solute concentration (3) The most rigorous model for diff controlled pressure solution derived here is, based on a dissipation balance approach, applied avoiding the approximations usually made for grain boundary solute concentra- tion as well as the assumption of low grain contact stress. This model, along with the other diffusion-controtled models considered, predicts sealing times (time to reach the per colation threshold) of up to several tens of years for faults filled with anhydrite gouge with a grain size of S~80 jum at in sitw upper crustal conditions. (4) Such time scales are short compared to the ime scales relevant for COs storage, implying that even if faults in an anhydrite caprock system are reactivated during the COs injection phase they should self-seal quickly 122 A. M. H. PLUYMAKERS & C. J. SPIERS enough to pose only a minor long-term leakage risk, particularly if the storage reser- voir is underpressured with COs. (5) Since recurrences times for earthquakes with M> 5 in the central Apennines are 2000- 5000 years, anhydrite-rich faults in this region will fully heal between these events. Fault strength recovery in. anhydrite-dominated faults therefore cannot be the factor that con- trols this repeat frequency, though it may play a role in controlling recurrence time for (much) smaller events. ‘This research was performed within Work Package 3.3 (Storage) of the Dutch national carbon capture and storage rescarch programme, CATO-2. We thank J. Pen- ninga (Nederlandse Aardolie Maatschappij B.V.) and S. Hangx (Shell Global Solutions) for supplying the sample material. We also thank two anonymous reviewers for their consiruetive and most useful comments on this paper References Axorvine, C. L., Toxcorte, D. L. & Furuistt, M.D. 1982. Pressure solution Iihification as a mechanism for the stick-stip behavior of faults. Tectonics, 1. 151~160, doi: 10,1029/TC0O1i002p00151 BILDSTEIN, O., WORDEN, R. H. & BROSSE, E. 2001, Assess- iment of anhydrite dissolution as the rate-limiting step during thermochemical sulfate reduction. Chemicat Geology, 176, 173-189, hitp://dx.doi org/10.1016/ 0009-254 (00}00398-3, Brounr, CW. & Dickson, F. W. 1969. The solubility of anhydrite (CaSO4) in NaCI-H20 from 100 t0 450 °C tnd 110 1000 bars. Geochimica et Casmachimioa Acta, 33, 227-245, hitp://dx.doi.org/ 10.1016/0016-7037 (69901409 Brant, S., Kumicky. J. & Ware, A. 2008, Kinetics of Water-Rock Interaction. Springer, New York, USA. Cantuccr, B., MonTraRosst, G., VASEtt, O., Tasst, F, Quarrrocemt, F. & Perkins, E. H. 2009, Geochem= ‘eal modeling of CO? storage in deep reservoirs: the Weyburn Project (Canada) case study. Chemical Geology, 268, 181-197, hitp://dx.doi ore/10.1016/ jichemgeo,2008.12.029 Cotto, G., Mxzzout, S., Tonpr, B. & Turco, B. 1997. ‘Active tectonics in the central Apennines and possible implications for seismic hazard analysis in peninsular Italy. Tectonophysics, 272, 43~68, http://dx.doi.org/ 10.1016/S0040-195 1(96)00275-2 CHANG, R. 2000, Physical Chemistry for the Chemical and Biological Sciences. University Science Books, Sausa- Tito, California Cute, J., Yano, X., Duan, Q, Sumsiaaoro, 7. & SrEKs, €.J.2013. importance of thermochemical pressuriz- ation in the dynamic weakening of the Longmenshan Fault during the 2008 Wenehuan earthquake: infer- ences from experiments and modeling. Jounal of Geo- Physical Research: Solid Earth, M18, 4145-4169, http://dx.doi.ons/10.1002,/jar.50260 CHoDINI, G., CaRDELLIN, C AMATO, A, BOSCH, E, CaLino, 8, Fkoxbis, F.& VenTuRa, G. 2004, Carbon dioxide Earth degassing and seismogenesis in central and southern lily. Geophysical Research Letters, 31, LLOTGIS, ttp://ax.doi.ory'10.1029/ 2001019480 Cioro1s, G., Enorr, G., FLominoo, F., Manns, F., RUG- ‘ito, L, & SuER,P. E, 2013, Sudden deep gas erup tion nearby Rome's airpor of Fiumicino. Geophysical Research Leners, 40, 2013GL058132, hup//ax.di ong/10.1002/2013g1088132 CoLtertist, C. & BaKent, M. R. 2002. A tow-angle normal faut in the Umbria region (Cental Italy). a ‘mechanical model forthe elated microseismiciy. Tee= tonophysies, 389, 97-115. Cotterti, C, CarDrtuist, Cy Ciopist, G., De Paota, N., Horpswortn, RE, & Swiri, S.A. F 2008. Fault weakening due to CO2 degassing in the Northern Apennines: short- and long-term processes In: Winwext ey, C.A. 1, Kure, W., une, 1, HOLD woum, RE, & Couterm, C. (eds) The Internal Siructure of Fault Zones: nplications for Mechanical ‘and. Fluid-Flow Properties. Geological Society, London, Special Publications, 299, 175~194, hup:// dx doi.org/10.1144/3p299, 11 CoLLeTtis, C, De PAoLa, N. & FaLxNeR, D. R, 2008. Insights on’ the geometry nd mechanics of the Umbria-Marche earthquakes (Central tay) from the integration of field and laboratory data. Tectonoph sies, 476, 99-109, lnp//dx.do.ong/ 10.1016] ecto, 2008.08.013 De Meer, 5. & SPrERS, C. J. 1997. Uniaxial compaction creep of wel gypsum aggregates. Journal of Geophysi- cal Research, 102, 875-891, hitp://dx.doiorg/10. 1029/96)602481 De Maen, ., Sruns,C.J. 8 Naxastusta,S, 2005. Structure and diffusive properties of fluid-filled grain boundaries an in-situ study using infrared (micro) spectroscopy Earth and Planetary Science Leners, 232, 403-414, tnup://axdoi.org/10.1016/j:eps 2004, 12.030 Dé PAOLA, Nu COLLETTINE, C., FAULKNER, D. R, & TRIP- verta, F. 2008, Fault zone architecture and defor- mation processes within evaporite rocks in the upper crust. Tectonics, 27, TCAO17, hp://dx.do.org/10. 1029/2007%<002230 Den Brox, S. W. J. 1998. Effect of microcracking on Pressure-solution stain rate: the Gratz grain-boundary model, Geology, 26, 915-918, hnip://dx.doi.org/10. 1130/0091-7613(1998)026-<09 S:comops>2.3.c0:2 Favtete, D.1.R. & RUrIER, EH. 2001, Can the main- \enanee of overpressured Mud in large strike-slip fault zones explain their apparent weakness? Geology, 29, ‘503-506, http://dx.doi.org/10.1130/0091-7613(2001) (129<0503:cuno0f> 2.0.00.2 FAULKNER, D. R., JACKSON, C. ALL, LUNN, R. J ScHLIscHE, Rl Wa, SHIPTON, Z. Ky WABWERLEY. C_A.1. & Wrriuack, M.O. 2010. A review of recent developments conceming the stucture, mechanies an fluid flow properties of fault zones. Journal of Structural Geology, 32, 1887-1575, hup://dx doi.org/10.1016/ {s8.2010.06,009 Gait P., GaLanint, F, & Paxrosri, D, 2008, Twenty years ‘of paleoscismology in. Waly. Earth-Science Reviews, 88, 89-117, hitp://ax.doi.org/10.1016/ jjearscire.2008,01.001 PRESSURE SOLUTION OF ANHYDRITE GOUGE, 123 GLeNN1E, K.W. 2001. Exploration activitiesin the Nether lands and North-West Europe since Groningen. Neth erlands Journal of Geosciences, 80, 33-52. Gunverse, E., Renanp, F., Dystie, D. K., BJORLYKKE, K. & JAmTveit, B. 2002. Coupling between pres- sure solution ereep and diffusive mass transport in porous rocks. Journal of Geophysical Research: Solid Earth, 107, 2317, huap://dx.doi.org,/10.1029/ 2001000287 Hawox, S.J. T., Sous, C.J. & Peactt, C.J. 2010, Creep of simulated reservoir sands and coupled chemical- mechanical effects of CO2 injection, Journal of Geo- physical Research, 11S, 1809205, bttp://éx.doi.org/ 10.1029 2008006939 Hickwax, S., SmsON, R, & BRUHN, R. 1995, Introduction to special seetion: mechanical involvement of fluids, in faulting. Journal of Geophysical Research: Solid Earth, 100, 1283112 840, hnp://dx.doiorg/10.1029/ 9jo01 121 Honotet, W., Beever, U., Corrt, E., PEARSON, F. J. & ‘THOFNEN, T. 2002. Chemical thermodynamic data- base 01/01. fn: uPublish (ed.) Nagra/ PSI Publishers, Wettingen. IPCC 2007, Climate change 2007: synthesis report, Sum- mary for policymakers. fi: CORE WRITING TEAM, PacHavit, R, K.& REISINGER, A. (eds) IPCC, Geneva. Keune, N., Heiesronner, R., Sriinrrz, H., Boutin, A. & [r0, H. 2007, Grain size distributions of fault rocks: a comparison between experimentally and natu- rally deformed granitoids. Jounal of Structural Geology, 29, 12821300, hitp://dx.doiore/10.1016/ {sisg.2007.04,003 KobLemever, P. J., Peacit, C. J. & Spiers, C. J. 2012. Surface diffusivity of cleaved NaCl crystals as a fune- tion of humidity: impedance spectroscopy measure- ments and implications for erack healing in rock sat. Journal of Geophysicat Research: Solid Earth, 117, 1801205, huup://dx.doi.org/ 10.1029/201 1jb008627 Lenner, F. K, 1990, Thermodynamics of rocks by pressure solution. /: BARBER, D. J. & MEREDITH, P. G. (eds) Deformation Processes in Minerals, Cer= camics and Rocks. Unwin Hyman Ltd, London. Lerner, F. K. 1995, A model for intergranular pres- sure’ solution in open. systems. Tectonophysics 248, 153-170, huip://dx.doi.org/10.1016/0040-1951 (04)00232-% Liteanv, E., NIeMEUER, A., Se1eRS, C.1.,PeacH, C1. & De Bresser, J. H. P, 2012, The effect of CO2 on creep of wet caleite aggregates. Jounal of Geophsi- cat Research, 117, 20, hitp://dxdoiorg/10.1029/ 2011JB008789 LOGAN, J. M., FRIEDMAN, M., HiGs, N. G., DENGO, C. & Sunwamoro, T. 1979, Experimental studies of simu- lated gouge and their application to studies of natural fault zones, In: SuRvEY, U. S. G. (ed.) Proceedings of Conference VIM, Analysis of Actual Fault zones in Bedrock, 79-1239, 305-343, MULLER, S. A., COLLETTINI, C., CHIARALUCE, L., COCCO, M., Baxcit, M. & Kaus, B.J. P. 2004. Aftershocks driven by a highpressure CO2 source at depth Nattre, 427, 724-727. MinapeLis, F., Barent, M.R. & Lurartents, A. 2008. Seismic reflection data in. the Umbria Marche Region: limits and capabilities to unravel the subsurface structure in a_ seismically active area, Annals of Geophysics, $1, hitp://dx.doi.ons/10. 4401 /an-3032 Nusneuer, A. R., Sours, C. J. & Bos, B. 2002. Compac- tion creep of quartz sand at 400~600 °C: experimental evidence for dissolution-controlled pressure solution, Earth and Planetary Science Letters, 198, 261-275, Iutp://dx.doi.org, 10.1016/30012-821x(01}0059 Niewruer, A. R., Eusworrt, D. & MARONE, €. 2009, Significant effect of grain size distribution on compac- tion rates in granular aggregates. Earth and Planetary Science Letters, 284, 386-391. On aN Gas JourNat. 2013. Worldwide look at reserves ‘and production, dn: Tiree, B. J. (ed) Oil and Gas Journal, 140, Series volume 12, 28-3 PaLuMo, L., BENEDETTI, L., BOURLES, D., CINQUE, A. & Finkel, R. 2004 Slip history of the Magnola fault (Apennines, Central Italy) from 36CI surface exposure dating: evidence for strong earthquakes over the Holocene. Earth and Planetary Science Letters, 225, 163176, bitp://dx.doi.org/10.1016/, epsl.2004.06.012 Paxtosti, D., SCHWARTZ, D. P. & VALENSISE, G. 1993, Paleoseismology along the 1980 surface rupture of the Irpinia Fault; implications for earthquake recur- rence in the southern Apennines, Italy. Jounal of Geo: physical Research: Solid Earth, 98, 6561-6577, http://dx.doi.org/ 10.1029/92j002277 Parreson, M.S. 1973, Thermodynamics and its geologi cal applications. Reviews of Geophysics, I, 355-389, PATERSON, M, S, 1995. A theory for granular Row aecom- modated by material transfer viaan intergranular Nid. Tectonophysics, 248, 135~151, hitp://éx.doi.org/ 10. 1016/0040-1951(94)00231-W Puoywanens, A.M. H,, Peact, J. & Spiers, C.J, 2014, Diagenetic compaction experiments on. simulated anhydrite fault gouge under static conditions. Journat of Geophysical Research: Solid Earth, M19, 4123— 4148, Ral, R. 1982. Creep in polycrystalline aggregates by ‘matter transport through a liguid phase. Jowmal of Geophysical Research, 87, A731-4739,_hitp://dx doi.org/10.1029/JBD87iB06p04731 Raw, M. 1992, Porosity-depth trends in reservoir sand- stones: theoretical models related Jurassic sandstones offshore Norway. Marine and Petrolewn Geology 9, $83—567, huip://dx.doi.org,/10.1016/0264-8172 (92390066-N RENARD, F,, ORTOLEVA, P. & Gxattex, J.P. 1997. Pressure solution in sandstones: influence of clays and depen: dence on temperature and stress. Tectonophysics, 280, 257-266, Rorovist, J., Revatnt, A. P., Capea, F, & Moris, G.J 2013. Modeling of fault weactivation and induced scis- micity during hydraulic fracturing of shale-gas reser- Voirs. Journal of Pesvolewm Science and Engineering, 107, 31-44, hup://Ax doi.org/10.1016/j,petrol.2013. 04.023 Rutter, E. H, 1976. The kinetics of rock deformation by pressure solution. Philosophical Transaetions of the Royal Society of London, Series A, Mathematical and Physical Sciences, 283, 203-219. Rurrer, E. H, 1983. Pressure solution in nature, theory and experiment. Journal of the Geological Society 124 A. M. H. PLUYMAKERS & C. J. SPIERS 140, 72: 5.0725 ScHUTIENS, P. T. M. 1991. Experimental compaction of quartz sand at low effective stress and temperature con- ditions. Journal of the Geological Society London, 148, 527-539, Inup://dx-doi.org/10.1144/gsjgs. 148. 3.0527 ScHUTIENS, P.T. M. & Spires, C. 1. 1999, Intergranular pressure solution in Nack: grain-to-grain contact exper- ments under the optical microscope. Oi! & Gas cience and Technology, $4, 729-750, hitp://4x.doi. ‘rg/10.2516/ogst:1999062 ‘Suimtzzu, 11995. Kinetics of pressure solution ereep in quartz: theoretical considerations. Teetonophysies, 245, 121-134, hitp: //dx.doi.org,/10.1016/0040-1951 (04)00230-7, ‘Susow, R. H, 1992. Implications of fault-valve behaviour for rupture nucleation and recurrence. Tectonophysics, 211, 283-293, hup://dx.doi.org/10.1016/0040-19 51(92)90065-E ‘Spies, C.J. & SCHUTIENS, P.'T, M. 1990. Densification of crystalline aggregates by fluid-phase ditfusional creep. i: Barer, D. J. & MeReDITH, P. G. (eds) Defor- mation Processes in Minerals, Ceramics and Rocks. ‘Unwin Hyman Lid, London. Spurs, C.J, De Mee, S., NIEMEMER, A. R. & ZHANG, X, 2004. Kineties of rock deformation by pressure Solution and the role of thin aqueous films. fr Naxastuna, S., Sotens, C.J., Mercury, L., FENTER, PLA. & Hocunita, M. F Jk. (eds) Phiysicockemisiry of Water in Geological ani! Biological Systems ~ Structures and Properties of Thin Aqueous Films. Uni versal Academy Press, Inc., Tokyo, 129-158, THOENEN, T. & KULIK, D. 2003, Chemical thermody ical database 01/01 for the GEM-Selektor (V.2- PSI). In: PSI (ed.) Geochemical Modeling Code. Villi= gen, LES PSI Thermodynamics Group. Tewrerta, F, Couternal, C., VINCIGUERRA, S. & Mek eprri, P. G. 2010, Laboratory measurements of the physical properties of Triassic Evaporites from Central aly and conelation with geophysical data, Tectonophysics, 492, 121-132, hutp://dx.doi.org/10. 1016/}.tecto.2010.06.001 740, hnup://dx oi.org/10.1 144 /esigs. 140. Trrperta, F., COLLETTINI, C., BARCH, M. R., LUPAP TELL, A. & MIRABELLA, F. 2013. A multidiscipli- nary study of a natural example of a CO2 geological reservoir in central Italy. International Journal of Greenhouse Gas Control, 12, 72-83, hitp://dx.doi org/10.1016/j.jgge.2012.11.010 Van Noort, R. & Spurs, C. J. 2009, Kinetic effects of microscale plasticity at grain boundaries during pressure solution. Journal of Geophysical Research: Solid Earth, M14, B03206, hitp://dx.doi.org/10. 1029/2008}b005634 Vanpeweuer, Vi. VAN DER Mee, B., Horstez, C., Mutpers, F., D'Hoore, D. & Graven, H. 2011 Monitoring the CO2 injection site: KIB. Energy Procedia, 4, 5471-478, bitp://dx.doi.org/10.1016/ Jiegypro.2011,02.532 Waurens, RJ, ELuiort, JR. er ar. 2009. The 2009 L’Aquila earthquake (central Italy): a source mechan ism and implications for seismic hazard. Geophysical Research Letters, 36, L17312, hitp://dxdoi.org/10. 1029/2009g1039337 Winery, C. A.J. & Suimanoro, 7.2003. Internal struc ture and permeability of major strike-slip fault zones: the Median Tectonic Line in Mie Prefecture, Southwest Japan, Jounal of Structural Geology, 2, 59-78, hutp://dx.doi.org/10.1016/S0191-8141(02)00014-7 Wissreiey, C. A. J, Yie one, G. & Dr Toxo, G. 2008. Recent advances in the understanding of fault zone internal structure: a review. In: Winnetey, C. A. J Kunz, W., Iaser, 1, HoLosworth, R. B. & Couer tint, C. (eds) The Internal Structure of Fault Zones: Implications for Mechanical and Fluid-Flow Proper ties. Geological Society, London, Special Publications, 299, 533, hup://dx.doiorg/ 10.1144 /sp299.2 ‘Zuaxe, X. & Spuexs, C. J, 2005. Compaction of granular calcite by pressure solution at room temperature and effects of pore fluid chemistry. International Journal of Rock Mechanics and Mining Sciences, 42, 950-960. ZHANG, X., Souews, C. J. & Pract, C. J. 2010. Compa tion creep of wet granular calcite by pressure solu- tion at 28°C to 150°C. Journal of Geophysical Research, 118, B09217, http://dx.doi.org/10.1029/ 20080005853 The role of stress on chemical compaction of illite shale powder ROLF H. C. BRUIJN!* & BJARNE S. G. ALMQVIST!S 'Geological Institute, ETH Zurich, Sonneggstrasse 5, CH-8092, Zurich, Switzerland ?Present address: Department of Earth & Planetary Sciences, Washington University in St. Louis, One-Brookings Drive, Saint Louis, MO 63130, USA ‘Present address: Department of Earth Sciences, Uppsala University, Villaviigen 16, 752 36 Uppsala, Sweden *Corresponding author (e-mail: bruijn@levee.wustledu) Abstract: Experimental compaction curves for mud are key factors in sedimentary basin models ‘but they are poor at predicting porosity evolution beyond 25 km depth, where chemical eompac- tion dominates shale diagenesis. This study presents the results of experimental simulation of dia- genesis and low-grade metamorphism of an illite shale powder. By applying amphibolite facies conditions using high-pressure-high-temperature deformation apparatus, the transformation of illite to phengite that takes place under anchizone and epizone metamorphic conditions could be accelerated and completed within laboratory timescales. In accordance with observation of natu rally compacted shales, our experimentally compacted metapelites range in porosity from 1% to 16% and contain authigenic phengite, biotite, and quartz. By compacting illite shale powder under both lithostatie and non-ithostatie conditions, the role of stress or tectonic forces during dia- ‘genesis could be evaluated, Based on microstructural and chemical similarities we infer that both Iypes of samples compact in the same way, indicating that chemical compaction is unaffected by differential stress. Basin evolution models can therefore extend the ID generalization ofthe stress field to depths where chemical compaction takes place. However, as chemical compaction ihances mineral alignment, anisotropy of physical properties such as permeability and sonie vel- ‘cities can change substantially Mudstones and shales are the most common sedi- mentary rocks (Tucker 2001), They form during the various stages of clay diagenesis that commences immediately after deposition in low-energy marine environments. Continued deposition of newly sup- plied material buries earlier deposits, which com- pact in response to the increasing vertical load, ‘Void ratio and porosity quantify the degree of com- paction, which is frequently coupled to hydraulic and seismic properties of clay-rich sediments (e.g. Neuzil 1994; Vasseur ef al. 1995; Dewhurst et al. 1998; Nygird et al. 2004; Mondol et al. 2007; Schneider et al. 2011). The compaction behaviour of argillaceous sediments is complex and control- led by both intrinsic parameters, such as mineral composition, clay fraction and fluid salinity, and extrinsic parameters such as temperature, effec- tive pressure, time and compaction rate (e.g. Rieke & Chilingarian 1974; Bjorlykke & Hoeg 1997), The complexities are visualized by the spread of compaction curves (porosity /depth relationships) obtained from mudstones and shales (Fig. 1a). In the top 2.5 km of the sedimentary column, mechani- cal processes that lead to improved grain align- ment and close packing, such as. grain rotation and fluid expulsion, are responsible for the compac- tion, In deeper domains thermally activated mineral reactions and mass transfer dominate the compac- tion (e.g. Athy 1930; Hedberg 1936; Weller 1959; Rieke & Chilingarian 1974; Velde 1996; Aplin et al. 2003; Charpentier et af. 2003; Worden et al. 2005; Peltonen et ai. 2009; Day-Stirrat et al. 2010) ‘To study the effect of mechanical compaction on the physical properties of rock, clay compaction has frequently been simulated in the laboratory by mimicking the kinetic and chemical conditions of the shallow portions of basins (ic. effective pressure <50 MPa or depth <2.5 km; eg. Chilingar & Knight 1960; Rieke & Chilingarian 1974; Vasseur et al. 1995; Dewhurst et al. 1998; Mondol et al. 2007; Voltolini et al, 2009; Schneider et al. 2011), The resulting compaction curves show good agree- ment with natural curves for low effective pressure. In experimentally compacted mud, porosity at 50 MPa ranges from 11% to 44% (Fig. 1b); at the equivatent depth of 2.5 km in nature, the porosity of mudstone and shales is however 10-15%. To reach porosity below 10-15% mudstones and shales require compaction by chemical processes such as dissolution-precipitation or clay trans formation. An important process that stimulates ‘chemical compaction is smectite illitization, which occurs between 60 °C and 110 °C (e.g. Towe 1962; Perry & Hower 1970; Hower ef al. 1976; Eber! From: Favtkwer, D. R., Maniant, E. & Mecktenaurctt, J. (eds) 2015. Rock Deformation from Field, Experiments and Theory: A Volume in Honour of Ernie Rutter. Geological Society. London, Special Publications, 409, 125— 147. First published online August 6, 2014, huup://dx.doi.org/10.1144/SP409.3 (The Geological Society of London 2015. Publishing disclaimer: www geolsoe-org.uk/pub_ethics a You have either reached 2 page thts unevalale fer vowing or reached your ievina tit for his book. CHEMICAL COMPACTION OF § 1993; Bjorlykke & Heg 1997). This process has a marked effect on porosity and compressional wave velocity evolution with depth in the North Sea Basin (Peltonen ef al. 2009). In the shallow domain of the basin where mechanical compac- tion dominates, porosity and compressional wave velocity in the well-log response are easily fit to a compositionally related experimental compaction curve, That fit is however lost below the depth where smectite illitization occurs and chemical compaction is initiated (Fig. Ic, d). Nevertheless, sedimentary basin evolution mod- els frequently resort to experimental mechanical compaction curves to couple rock properties to porosity, effective stress, depth and permeability (eg. Bethke 1985; England et al. 1987; Ungerer etal, 1990; Audet & Fowler 1992; Audet & MeCon- nell 1992; Hart et al. 1995; Luo et al, 1998; Pouya et al. 1998; Suetnova & Vasseur 2000). Athy’s law (Athy 1930), a natural compaction curve, was used in earlier models but, as more natural com- paction curves were collected, it became clear that each basin had a unique signature, rendering Athy’s Jaw inadequate. In using experimental compaction curves, modern models depend on a few assump- tions, simplifications and extrapolations: (1) a 1D approximation of deformation, i.e. burial (lithostatic pressure) in a passive basin setting; (2) mean effec- tive stress is the sole driving force (Terzaghi 1925; Rubey & Hubbert 1959; Smith 1971); G) model parameters are derived from low-stress oedometer tests, but extrapolated over two orders of magnitude to basin conditions; (4) the effect of differential stress, lateral strain or even time and temperature is often neglected; and both (5) the compressibility of individual phases and (6) chemical changes caused by mineral transformations are often neg- lected. These assumptions are justified for the part of the basin where mechanical compaction is domi- nant (i.e. top 2-3 km), following conclusions by Jones & Addis (1985) who analysed the critical state theory in the context of mud compaction and Luo et al, (1998) and Pouya et al. (1998) who incor- porated non-vertical stress in a modified Cam-Clay model. his paper presents the results of laboratory simulations of shale diagenesis and lower greens- chist facies pelite metamorphism in which chemical compaction was activated. By exposing illite shale powder to amphibolite and granulite facies con- ditions in the presence or absence of differential stress, the otherwise slow mineral transformations required for chemical compaction could be acceler- ated to finish within hours to days, To do so, a novel three-stage compaction procedure was developed. The first stage simulates cold burial in a fashion comparable to traditional mud compaction exper- iments and only activates mechanical compaction. Al # POWDER 127 Heat is introduced to the sample in the second stage, simulating deep burial under lithostatic pressure and activating mineral transformation, In the final stage, We simulated the transformation in pelites associ- ated with the transition from anchizone diagenesis to epizone metamorphism in tectonically active ns such as fold-and-thrust belts. This study therefore allows for the evaluation of the role of the stress field during chemical compaction of illite shale powder. The results are used to test whether the assumptions, justifications and simplifications of modem basin evolution models are appropriate for deep basins where chemical compaction domi- nates shale diagenesis. To achieve these evaluations we investigated the microstructural and chemi- al evolution with porosity of our experimentally re-compacted natural shale samples. This study builds on the work of Bruijn er al. (2013), who dis- cussed the development of magnetic suscepti and fabric in the same samples. Starting material Maplewood Shale from the Lower Clinton Group (Llandovery) in the Appalachian Basin of western New York, United States (e.g. Brett et al. 1990; Rogers ef al. 1990) was used as starting material Fragments of Maplewood Shale were acquired from Ward’s Natural Science® (Rochester, New York). This rock type has informally obtained the status. of laboratory standard for illite and ilite- rich shale during the past decade, having been used in friction experiments (Saffer & Marone 2003; Tkari er al, 2009; Den Hartog et af. 2012a, b; Den Hartog & Spiers 2013), a study on brine and super- critical CO, interaction with aquifer and aquitard rocks (Kaszuba et af. 2005) and a study on the devel- opment of magnetic susceptibility and fabric during chemical compaction (Bruijn er al. 2013) ‘The olive-green to greenish-grey fragments, range in length or width from 3 to 10 em with thick- ness of less than 3.cm (Fig, 2a). Colour variation within and between fragments is minimal and no veins or lenses of coarser particles are visible by the unaided eye, indicating homogeneous com- position and fabric. Scanning electron microscopy (SEM) imaging combined with energy-dispersive X-ray spectroscopy (EDX) analysis idemtfied illite, phengite and quartz as abundant phases and bio- tite, pyrite, apatite, rutile and K-feldspar as rare phases (Fig. 2b, c). The phyllosilicates and quartz represent ¢. 60-75% and 20-30% of the imaged area, respectively, ax determined by grid-point analysis, Based on SEM analysis of broken and polished surface imaging, porosity in Maplewood Shale is attributed to four generic pore types: (1) fractures; 128 R. H.C. BRUUN & B.S. G. ALMQVIST (2) matrix-clast-voids; (3) clay-micropores; and (4) matrix-pores, which are described and correlated with pore-types defined by Kwon ef al. (2004), Desbois et al, (2009) and Heath er al. (2011) in Table 1 Based on burial depth estimates of 5 km for nearby Late Ordovician Trenton Group carbonates (Friedman 1987), Maplewood Shale is inferred to have formed by chemical compaction under late diagenetic to anchizone conditions triggered by smectite illitization (e.g. Hower et ai. 1976; Merri- man & Peacor 1999). The absence of authigenic quartz (i.e. quartz cement), which would have been produced during the smectite to illite trans- formation, suggests an open system during burial of the Maplewood Shale source sediments (Van de Kamp 2008; Peltonen ef al. 2009; Day-Stirrat et al. 2010; Thyberg er al. 2010; Thyberg & Jahren 2011). X-ray powder diffraction analysis (XRD) verified the abundant presence of quartz, the diocta- hedral micas and some minor amounts of clino- chlore (Fig. 3a). Additionally, small peaks at 7.9 and 33.0° 20 suggest the presence of a smectite- group mineral and apatite, respectively. The narrow width of the illite basal plane (001) peak at 8.9° 26 indicates that sediment butial reached anchizone conditions. Under these conditions, the illite smec- tite ratio is typically >9 (Merriman & Frey 1999) With the exception of apatite, the concentration of rare phases identified by SEM is below the limit of| detection of XRD. ‘The major element composition of Maplewood Shale is consistent with illite shale devoid of carbon- ates (Table 2; Kaszuba er al. 2005). Thermogravi- metry analysis (TGA) revealed a minor amount of free water lost between 50 and 150°C, which was related either to smectite or absorbed from the air prior to analysis (Fig. 3b). The majority of water loss oceurs between 450 and 750°C, associated with subsequent dehydroxylation of illite, phengite and biotite (c.g. Guggenheim ef al, 1987; Earnest 1991). The TGA curve plots between curves for the smectite-tich Silver HilL-illite (Eamest 1991) and pure muscovite flakes (Mariani et al. 2006). For homogenization purposes, Maplewood Shale fragments were mechanically crushed into a fine-grained illite shale powder. After sieve separ- ation of the <125 um particle size fraction, the illite shale powder has a mean grain size of 21.1 pm with standard deviation of 2.4 um, and contains Fig. 2. (a) Photograph of Maplewood Shale fragments, (b,€) SEM micrographs of Maplewood Shale. Q. quartz I, illte; Py, pyrite; Ap, apatite; Bio, biotite; Ph, phengite. (b) Low-magnification BSE image showing abundant and rare mineral phases and lack of bedding in the illite matrix. (e) High-magnification BSE image of square in (b). Solid arrows depict matrix-clast voids: open arrows show examples of matrix-pores; and dotted squares highlight regions with cla CHEMICAL COMPACTION OF 129 Table 1. Pore types in Maplewood Shale Name Shape Size Heath et al. Kwon et a —_Desbois ott) 2004). eral (2009) Fractures Thin tabular Splitting along Pore-type VI planes cleavage plane Matrixlast- Tabula <2ym Geometrical misfit Porestype V__Pore-type Il voids pyramidal, between Spherical rmatiix-forming grains and coarser clasts Clay-micropores! <1001nm Misfit de to imegular Pore-type | Pore-types_Pore-type 1 clay surfaces and IL Matrixcpores! Tabular, 1-3 um Geometrical Pore-typesPoresypes—Pore-type Il pyramidal, mist among Wand Il Hand IV spherical rmatrix-forming lighted in Figure 2 by solid arrows ‘ighlghed in Figure 2e by dotted square ‘nighighted in Figure 2e by open arows. 15.1 vol% clay (<2 pum), 79.5 vol% silt (2. and 5.4 vol% sand (>63 jum) (Fig. 3c), Sample analysis, For microstructural imaging and mineral phase identification, a FEI Quanta 200 field emission gun SEM was used in combination with an EDAX Pegasus detector for EDX analysis on selected sample sites. High voltage of the electron beam for backscatter electron (BSE) imaging was 10.0KV, combined with a spot size of 4-6 units and beam aperture of 30 um. For EDX analysis, the high voltage was increased t0 20.0 kV and the beam aperture was changed to 40pm. The working distance was maintained at 10 mm under high vacuum (<10~* Pa) and at 7 mm under low vacuum (40 Pa) chamber conditions. EDX analysis was only performed under high vacuum conditions. SEM samples were prepared by either cutting or breaking a sample in half, The resulting fresh inter- face was imaged or used for EDX analysis. Cut samples were impregnated with cold mounting epoxy resin (Struers) to close pore space and support soft clay/mica matrix. Resulting pellets were mechanically polished using wet abrasive papers with inereasing grit and subsequently lapped with down to I um particle size diamond paste. The final preparation step involved the appli- cation of a 15-20 nm thick carbon coating. Broken surface samples required no impregnation or polish, and received an 8-10 nm thick gold coating. XRD analysis was performed using a Bruker AXS D8 Advance diffractometer, equipped with a Lynxeye superspeed detector and automatic 9-slot sample changer. Powder samples were measured between 4 and 89.9° 20 angle, with 0.025° step size per second. Scans were performed in locked coupled mode with a divergence and anti-scattering slit of 20 mm, while rotating samples. The Lynx-iris was positioned 12 mm from the samples. The pro- prietary software package EVA 2 (developed by SOCABIM, now Bruker AXS), version 11.0.03, was used to correct XRD diffractographs for back- ground noise and to apply curve smoothin A wavelength-dispersive X-ray fluorescence (ARF) spectrometer (Axios PANanalytical) equip- ped with five diffraction crystals was used for major element anal is was. per- formed on fused glass beads prepared from sample powder mixed with dilithium-tetroborate in a 1:5 ratio using a Claisse M4® fluxer. Glass bead prep- aration included loss on ignition (LOD) mea ments calculated from the weight loss of a c. powder during a 2 hr exposure to 1050 °C. ‘TGA was performed with a Netzsch STA 409 C/ CD thermobalance, with top-loaded samples and digital balance. A linear heating program with a rate of 10°C/min from room temperature to. 1500 °C was applied. Sample powders were placed in alu- mina crucibles of known mass. Simultaneously, a reference alumina crucible (1400 mg) with 50 mg alumina powder was measured to obtain sample holder correction values. Initial mass of sample powders varied between 50 and 230 mg. Resolu- tion of the mass balance system is 2 wg. To avoid reactions between sample and air, analyses were performed in an inert argon-gas atmosphere, main- tained with an input flow rate of 40 mL/min, Data were collected and processed using. proprietary Proteus® software 130 R. H.C. BRUUN & B, 8. G. ALMQVIS XRD 20 Spectrum ©) ze 3. q ua 1 w iw low Particle grain size (um) Fig. 3. (a) XRD spectrum for Maplewood Shale with selected interpreted mineral phases after 26 peaks. (b) ‘Thermogravimetry curves for Maplewood Shale powder and HIP synthetic metapelite after compaction stage 2, Light- grey and grey areas denote Maplewood Shale and HIPI dehydration, respectively. (€) Grain size distribution plot for ‘Maplewood Shale powder after fragment crushing and sieving. Particles smaller than 2 m ate considered clay: particles of size 2~63 jum are silt; and coarser material is considered sand, CHEMICAL COMPACTION OF § Table 2, Major element composition (1%) of Maplewood Shale and HIPI synthetics metapelites tas determined by XRF analysis Oxide Maplewood —-HIPL Kaszuba Shale et al. (2005) SiO. 63.617 63.737 60.66 TiO: 0.928 0.926 0.84 ALO, 19.474 19.359 17.78, FeO. 5.648. 5.547 482 MnO 0.045 0.045 0.03 MgO 2.636 2.651 232 C20 1.170 1.226 078 Na,O 0.236 0218, O19 K;0 6.271 6.227 5.88 P:0s 0.150 0.156 oz Total 100.175 100,092 93.42 Lor 3.439 2714 5.1 *LOI, lesson ignition (i.e H, + €0;) is excluded fom the tot Porosity and grain density were measured indirectly by a Micromerities AccuPye 1330 he- ium pyenometer. The pycnometer uses pressurized (1.44 MPa) helium gas to determine the grain volume of the solid material that occupies one of two chambers of calibrated space. The pycnometer chambers were calibrated with industry-certified standards. Detailed analysis of pycnometer accu- racy, precision and resolution yielded working values of 0.01%, 0.20% and 0.001 em’ respectively for the smallest and the most commonly used cali- brated chamber of 10 cm’, using an iron cylinder of known grain volume. The offset caused by accu- racy was corrected by applying a muhtiplication factor of 0.993 to grain volume determinations, Error analysis on an iron cylinder of known vol- ume and zero porosity revealed an appreciable temperature effect. Sample volume estimates were corrected linearly for temperature effects by apply- ing a multiplication factor of 0.0017 cm’ per degree deviation from 29 °C. The pycnometer was set to repeat volume measurements automatically six times to improve statistics and to rule out gas pres- sure effects on the pore structure of the sample. The ratio of corrected pore volume (or bulk volume minus grain volume) over bulk volume, expressed in percentage, defined porosity. Grain and bulk density are defined as the ratios of mass over cor- rected grain volume and bulk volume, respectively, expressed as g em”. Bulk volume was determined from 10 um resolution calliper measurements of cylinder diameter and length. Volume correction habitually decreased. porosity estimates. by 0.6~ 0.7%, while temperature corrections yielded a por- osity adjustment of up to 1%. Before pyenometer analysis, samples were oven dried at 110 °C for at Teast I hour to remove free water from the pores. Al # POWDER 131 Three-stage compaction procedure To achieve chemical compaction of illite shale powder in laboratory timescales and to investigate the role of stress in this process, a three-stage com- paction procedure was developed that makes use of three different types of rock deformation apparatus and three distinct stress fields. ‘The sequence of ‘experiments involves (1) load-sensitive mechanical compaction, followed by (2) temperature-sensitive chemical compaction with the absence or presence of directed constant strain-rate deformation, Essen- tially, this simulates diagenesis and very low-grade metamorphism of illite mud with and without tec- tonic forces. By applying and withholding differ- ential stress, this procedure allows the evaluation of the role of the stress field on the development of physico-chemical properties and fabric elements of illite mud during diagenesis and low-grade meta- morphism. The role of the stress field is evaluated by comparing microstructural and chemical devel- opment with porosity of lithostatically and non- lithostatically or dynamically compacted samples, Dynamic compaction refers to the experimental simulation of diagenesis in tectonically active basins. The simulation of diagenesis in tectonically passive basins is simulated experimentally during what is here referted to as lithostatic compaction tests. To highlight changes in mineral composition with respect to the illite shale starting material, the lithified and metamorphosed lower-porosity sam- ples are referred to as synthetic metapelites. For the purposes of interpretation, the synthetic metape- lites were grouped according to their experimental ‘compaction history rather than treated individually (Table 3). Compaction stage | In stage 1, illite shale powder was compressed inside stainless steel canisters (length = 220 mm and diameter = 51 mm) enclosed within a bolted stee! housing using a hydraulic press operated at room temperature. Sample material was added to the can- ster in a stepwise sequence, in amounts of 5~7 g. ‘The powder was compacted each time material was added using a 40 ton load, corresponding to 200 + 5 MPa. Potential water vapour released dur- ing subsequent compaction stages at elevated temp- erature was captured in a 2-4.em thick layer of alumina powder at the bottom and top of the canis- ters. The stress field of compaction stage I resem- bles that of confined compression, The bolted stee! housing ensured a non-zero horizontal stress. The excessive load (> 100 MPa) ensured exhaustion of mechanical compaction processes and allowed the inference that compaction in later stages could be solely ascribed to chemical processes (e.g. Rieke 132 R. H.C. BRUUN & B.S. G, ALMQVIST Table 3. Classification of synthetic metapelites according to compaction history and compaction Kinetics during the last experienced compaction stage ample group Compaction history Compaction kineties Notes last stage HPI 142 Confining pressure HIP? 142430 Confining pressure PATI (C) 142436 Confined compression PATI (.) 142436 Confining pressure PATI (Coo) 14243b Coniined compression Potential partial melt PATI (Cp) 142436 Confined compression With pore pressure PATI (T) 1424 3b Confined torsion & Chilingarian 1974; Vasseur er al. 1995; Nygard et al. 2004; Mondol et al. 2007), It was clear from visual inspection that illite shale powder remained unconsolidated after com- paction stage | and care was taken not to disturb the manufactured fabric, As also observed by Misra et al. (2009), no pronounced foliation developed: this was ascribed to the small grain size of the sheet silicates. Powder-filled canisters were stored in a 110°C drying oven until they were used for the next compaction stage. Stage 1 simulated cold sediment burial to a basin depth of 11—13 km and solely initiated mechanical compaction; no lithifica- tion or mineral alteration of the illite shale powder occurred. Compaction stages 2 and 3a In stage 2, the hermetically sealed canisters from age | were further compacted for 24 hours by the application of pressurized and heated argon gas (170 MPa and 590 °C) using a hot isostatic press (HIP) apparatus. The compressed gas in this appar- atus exerts an equal force to the canisters from all directions, producing a lithostatic pressure (i.e. con- fining pressure). Stage 2 lithified the illite shale powder and initiated the transformation of illite into phengite, thereby creating a moderately porous (porosity 11 15%) synthetic metapelite. There is no evidence that the buffering alumina powder con- taminated or intermixed with the newly formed metapelites After stage 2, the canisters were split into two groups. One group of canisters was placed back in the HIP apparatus for the third stage of compaction at 172MPa confining pressure and 480°C for 15 hours, referred to as stage 3a, Cylindrical sam- ples were prepared from the remaining canisters, The following section describes the preparation steps for these samples. Stages 2 and 3a simulated pelite transformation under lithostatie conditions during metamorphism up to lower greenschist facies by exposing illite shale powder for a short period to amphibolite facies conditions. Samples derived from the synthetic metapelites after com- paction stages 2 and 3a are referred to as HIP and HIP2, respectively. Compaction stage 3b In stage 3b, compaction of HIPI samples was con- tinued using a Paterson-type HPT gas-medium testing machine (Paterson apparatus) equipped with axial and torsion actuators and_pore-pressure system (Paterson 1990; Paterson & Olgaard 2000), Experiments were performed at 300 MPa argon pressure and a constant temperature of 500, 650, 700 or 750°C. Compaction in stage 3b tests was promoted by exposing the sample to heat and either (1) a confining pressure (PATI (P.)); (2) an axial or confined compression (PATI (C), (Caso) and (Cp); or (3) a confined torsion stress field (PATI (T)). These experiments led to predomi- nantly contraction or pure or simple shear deforma- tion of the sample, respectively. Confined torsion is an adaptation of confined compression in that instead of vertical force, a torque is applied on the sample by means of a servo-motor-controlled rota- tional twist (Paterson & Olgaard 2000). Activation of an externally mounted torsion ot axial actuator imposed. constant twist rates (torsion) or displace- i rales (confined compression) and resulted in eat and axial-strain rates, respectively, ranging from 10° to 10s”. To prevent sample bulging or buckling during confined compression, sample shortening was limited to 20%. Simple shear defor- mation (torsion) was limited to a shear strain of 0.3. Samples were placed inside the furnace hot zone to ensure a thermal gradient of <2 °C across the sample. To prevent argon gas penetration, the sample assembly was jacketed in either iron or copper sleeves. A list of experiments, including their conditions and sample characteristics, is pre- sented in Table 4. “able Condon: ant smple charter of eompacton tae So experiments Eapeinent Group Teva Dimer Tea TR) Daath Sia Thal ait ‘emigre cs) )— Ii sain) Pow PATI «cy na usr iis 53m a iow PATLG) a 186 3 si ato, 2 5 Prana Parc) Rs Hse 3 SS) a Pit Parle) 155 iy Ho 500 3001) 2” re) Prat Parc) a Ose ira} swat toa PRG PATI) 78 58 i SS) to Plates Parc) i 3 iis Sows Bs pay Pant) Soe Sie ug so ne bo Pian ParLc) ser Bx 1a es 3001) 1s Pia PATI) 8 oss 13 som at Pre PATI) Nea rt tis si sooty a Piso Parc) 199 si ia im ow a Peo Pare) re] Das ng io Gone to bisas Parte) os oe 102 5001 H Piss, Parc) 09 ass iba imo a Piss PATI) 7 35 im sou Ba Piso Parc) Ha 8 jm 300m) ia Piste PATLEG) 138 B89 3 sO as Pst PATLC) ne os jm doo iss Prats PATI) 2 as) 20 D1 2 Pa PATI) Do si) ama 2° " ison" Pare 5 55 im som = 5 Pisa Partie mst Ms im som ° 5 Plat parte) sa bs im som ww 6 ries Pari) a6 Os 7m 30) 5 Piss PATI) 337 Ske Mya a ist Pare ats 50 sow 18 Pi PATHCg) oor Se i "3002S Plas pares) 12s Do 130 9 oun Oe Sept, piss PAT Cp 738 Ds 8 sm 305 Sa pists PATI 219s 38) 1? 60 3082 IT Pon part) ei oe 3 im “Soom, so 30s pLin+Pi26 PATI sir oe Ho 700-0) 3x10 su 134 R. H.C. BRUUN & B.S. G, ALMQVIST The effect of pore pressure (P,) was investigated with confined compression tests at 500, 650 and 700°C and additional argon gas pressure inside the sample (P, = 20 or 50 MPa). The experiments were performed using the standard pore pres- sure system available for a Paterson apparatus that includes an upper and lower reservoir. Porous coarse-grained alumina spacers replaced a solid alumina spacer at the top and bottom of the sample to allow argon gas to penetrate the sample. Porous alumina spacers were also used in three expeti- ments where we intended to drain the samples. In those experiments that were conducted at 500 and 700 °C and with a confining pressure and confined compression stress field, the lower reservoir was intentionally left open. Samples used for compaction stage 3b were derived from synthetic metapelites that only experi- enced compaction stages I and 2, that is, HIPI ‘material. Canisters containing HIPL material were sawed open horizontally using a steel saw. Cores of 10 or 15 mm diameter were drilled from the can- ister halves using a hollow diamond drill-bit. The cylindrical cores were then trimmed down to a desired length ranging from 5 to 20mm. Sample ends were polished with wet abrasive paper to ensure parallelism within 20 ym, Sample length and diameter and grain volume were determined using a digital calliper and helium pycnometer, respectively. Before compaction stage 3b exper- iments, samples were initially stored in a 110°C drying oven for at least 12 hours, Stage 3b simulated pelite transformation under dynamic conditions during metamorphism up to lower-greenschist facies by re-heating and pres- surizing illite shale powder for a short period to amphibolite or granulite facies conditions while forcing deformation in a pure shear or simple shear deformation style. HIPI samples used for Paterson apparatus tests Were renamed according to the com- paction kinetics they experienced during stage 3 as described in Table 3 Forward phase stability model During compaction stages 2 and 3, illite shale powder was subjected to a higher temperature than the stability field of illite (80-350 °C). Exposure to the experimental conditions triggered mineral transformations and reactions at an accelerated rate (e.g. Merriman & Frey 1999). To understand which reactions and transformations would be acti- vated during experiments, a pseudosection showing stable mineral assemblages in pressure and tempera- ture space was calculated using Perple_X version 6.5.0 (Connolly 1990, 2009) combined with the 2002 update of the thermodynamic dataset of Holland & Powell (1998), The NCKFMASHO sys- tem (ie. Na,O, CaO, FeO + Fe,0s, KO, MgO, ‘AlsO3, SiO» weight percent composition data) was applied for the Perple_X calculations. Titanium and ‘manganese were found to have no visible effect on the location of the phase stability fields in the mod- elled pseudosection. ThermoCale solution model parameters were used to calculate phengite and staurolite solid solutions, Speciation models by Holland et al. (1998) and Powell & Holland (1999) were used to calculate chlorite and biotite solid solutions, respectively. The Waldbaum and Thomp- son mixing model (Waldbaum & Thompson 1968) ‘was applied to the solid solution calculation of K- feldspar. Garnet and chloritoid solid solution and melt were calculated using mixing models described by White er al, (2000), Holland & Powell (1998) and White er al. (2001) respectively. No phases were excluded from Gibb’s free energy minimization calculations. Except for H;0, Maplewood Shale major element concentrations (Table 2) were con- sidered as input values. To account for dehydration during compaction stage 2, we preferred to input the HO concentration determined for HIPI samples (Table 2) ‘A. simplified pseudosection based on Maple- wood Shale composition is shown in Figure 4, For the purposes of clarity, small phase stability fields were merged and stability limits of one of the reac- tion phases (ie. field boundaries) that do not involve mica or clay minerals were omitted. Ultimately, five principal equilibrium stability fields were ident- ified, based chiefly on phyllosilicate content, In fields with no Ca-bearing phases, apatite is con- sidered the sole Ca-bearing phase. The fields, label- led 1-V, are dominantly temperature controlled, especially above 50 MPa. Different confining pres- sure between compaction stages 2, 3a and 3b there- fore has a marginal effect on mineral reactions and stability. Quartz is stable in fields I-IV and water is in excess in all fields. lite, chlorite and phengite are stable in field I and bounded at about 350 °C by the disappearance of illite and appearance of bio- tite as a stable mineral phase. Compaction exper- iments performed with P—T conditions of field 1 are unlikely to activate chemical compaction, Chlorite, biotite and phengite are stable in field IT whose upper boundary (450-550 °C) is marked by the disappearance of chlorite. Compaction tests of stage 3a and 3b performed at 500 °C plot in this field. The equilibrium phyllosilicate assemblage in HIP2 and a selection of PATI (C-P,-Cp) samples is expected to be chlorite, biotite and phengite, Biotite and phengite are stable in field III, which includes the conditions of compaction stage 2 (HIPL_ samples) and 650 °C compaction stage 3b tests (some PATI C-P,-Cp samples). The phengite CHEMICAL COMPACTION OF $ 40: 351 30 251 20 P (MPa) 1s 101 s « wv aw suv au Ki indi 135 ow. ow v0 TeC) 4, Simplified Perple_X pscudosection for Maplewood Shale major clement composition. Labelled black dots Le operating conditions of respective compaction stages, The let es refer 10 Pe, Cp, T and C59 in parenth sample subgroup ID of compaction stage 3b ~ confined compression, confining pressure, confined compression with pore pressure, confined torsion and confined compression at 750 °C ~ respectively. The dashed horizontal line indicates the temperature range of the T/xH,0 diagram of Figure 5, dehydroxylation reaction (phengite + quartz + K-feldspar+ Al-silicate + H,0) marks the bound- ary between fields I and IV. Stable mineral phases in field IV are the reaction products of the phengite dehydroxylation reaction plus. biotite, The 700 °C experiments in compaction stage 3b plotin field IV. The equilibrium mineral assemb: in PATI (C-P.-Cp-T) samples consists of K- feldspar, an aluminosilicate (likely sillimanite) and biotite, The transition from field IV to V is the wet- solidus. Partial melt generates in field V at the expense of quartz, biotite and water. Sillimanite also contributes to melt at higher temperature; this is not indicated in Figure 4 A comparison of major element composition before and after compaction stage 2 revealed a negligible net loss or gain of elements, except for water (Table 2). Water content reduces by 50% during the first HIP apparatus compaction stage, attributed to high-temperature dehydration phyllosi- licates (Fig. 3b). The pseudosection of Figure 4 is calculated for a chemical environment with excess of water. The minimum water content required to maintain the equilibrium phase stability fields I1-V (Fig. 4) can be read off a T/xH,0 diagram, Such a diagram (Fig. 5) was calculated using the same input parameters as tor the pseudosection at a confining pressure of 300 MPa. Coupling between a pseudosection and T/xH;0 diagram is carried out via the HO content input value for the calculation of the pseudosection and the fixed value for pressure in the T/xHO diagram. The 136 R. H.C. BRUUN & B.S. G, ALMQVIST 1,0 we% ig. 8. Simplified Perple_X T/xH.O diagram for Maplewood Shale major element composition. Labelled black circles indicate operating conditions of respective ‘compaction stages. The dashed vertical line indicates the temperature range of the pseudosection where the HO was fixed at 5.18 wi% (Fig. 4). Bracketed letters C, Pe, Cp, T and Cyso, refer to the sample subgroup ID of compaction stage 3b, confined compression, conti pressure, confined compression with pore pressure, confined torsion and confined compression at 750°C, respectively. trace line (dashed grey line) in one diagram (Figs 4 & 5) displays the conditions for which the other diagram is constructed, Excess water exists for T< 640°C for HO > 2.2 wt% and for T<700°C for HO > 1.0 wt%. The boundary between field IV and V is identified by the wet solidus at 700 °C. Below 1.0 wt% HO, the solidus shifts to higher temperature (>750 °C). Besides fields II-V, two additional fields are identified at low H20 values (<2.2-2.3 wt%). Field VI exists below 0.9= 1.0 wt% HO and has a mineral assem- blage in equilibrium comprising biotite, an alumino- silicate (andalusite or sillimanite), K-feldspar and garnet. At water contents between 0.9 and 2.3 wt% HO, and temperature below 650 °C, field VIT has a stable mineral assemblage of biotite, phengite, andalusite or sillimanite, K-feldspar and quartz The accuracy of the phase stability model (Figs 4 & 5) was evaluated by comparing the tempera- ture range of mica dehydroxylation for Maplewood Shale, obtained from TGA with the boundary between fields III and IV in the pseudosection. ‘The phengite dehydroxylation boundary of field II shows a positive pressure dependency (Fig. 4). At 10 MPa, the maximum temperature for stable phen- gite coincides with the temperature range identified by TGA as the range for phengite dehydroxyla- tion (Fig. 3b). However, as the error of the molar volumes used in the calculations of phase stability fields yields greater uncertainties at lower pressure, the validity of this comparison is low Synthetic metapelites Porosity and density ‘The two- and three-stage compaction of illite shale powder resulted in lithified aggregates with con- nected porosity ranging from 1% to 17% and bulk density ranging from 2.35 gem * to 2.69 gem * (Fig. 6). Both compaction stages 3a (repetition of hot isostatic pressing, producing HIP2 samples) and 3b (Paterson apparatus tests, producing the PATI samples groups) reduced the porosity of most HIPL samples to less than 10%. PATI (C) and (P.) samples display a wide range of porosity and den- sity and exhibit comparable trends, in which other PATI samples also plot. The PAT! sample trend overlaps with that of the HIPI samples, but is shifted towards higher density relative to HIP2 samples. Maplewood Shale plots between the low- porosity HIP2 and PATI samples (Fig. 6). Chemistry Mineral composition of the synthetic metapelites is in disequilibrium during compaction stages 2 and 3. The progress of mineral reactions transforming illite assemblages to targeted equilibrium assem- blages was monitored by XRD analysis. The water content of samples after compaction stage 3b was measured to verify whether reactions towards stab- ility fields VI and VII (Fig. 5) were activated. XRD spectra. XRD spectra from all synthetic meta- pelite groups show a transformation of the original illite assemblage (NAT) into phengite-, biotite- and quartz-dominated assemblages (Fig. 7). ‘The synthetic metapelites show no characteristic peaks Connected porosity (%) Bulk density (g em™) Fig. 6 Bulk density of synthetic metapelites and Maplewood Shale as a function of porosity. Samples are labelled according to their last compaction stage. CHEMICAL COMPACTION OF $ t i 7 i i j 1 i | i i 137 Fig. 7. Selection of XRD spectra from each compaction (Table 3). Dashed and dotted grey lines indicate: mple group as d fined by their compaction history respectively, compared to the refere ‘natural Maplewood Shale (NAT) 26 pattern of mineral phases. Associated mineral erystal planes are identified. Grey circles highlight copper oxide peaks. for clinochlore basal (001) and (002) planes, smee- tite basal (QOL) and apatite (144) and (414) planes. Instead, we observed sharp peaks associated with the phengite (114) planes and biotite (020), (110), (131) and (402) planes. Additionally, copper- jacketed samples yield copper oxide peaks at 29 = 43.4", indicating minor jacket contamination (Fig. 7). Other peaks in the metapelite spectra are comparable with Maplewood Shale peaks and rep- resent quartz and detrital phyllosilicates. Water content. In natural Maplewood Shale (NAT), water content measured by Karl-Fischer titration is 3.7 £0.3 w%. During compaction stages 2 and 3a, mineral stability shifted towards stability fields Il and III (Fig. 5) because water content in HIPL and HIP2 samples is greater than 2.4 wt%. In the Paterson apparatus experiments, synthetic metape- lites experienced various water loss which resulted in 0.6-29 wt% HO remaining (Fig. 8). Most water loss is recorded for the 700 °C drained exper- iments. For a single temperature, undrained exper- iments display a decrease in water content with increasing experiment duration. The water content for half of the 500°C PATI metapelites decreased below the boundary of stability fields II and VIL at 2.4 wi%. For 650°C, 700°C and 750°C PATI samples, the boundary between mineral stability (CHEMICAL COMPACTION OF SHALE POWDER 139 Fig. 9. BSE and SE images of lithostatically compacted metapelites. Sample cylinder axis is vertical for all images. Examples of different pore types are highlighted (open white arrows: matrix-clast voids; dashed white boxes: clay micropores; solid white arrows: matrix pores; dashed white circles: intragranular fractures). Black ellipses exemplify structural features (dashed: complex authigenic biotite clusters and solid: mica-kinks or microfolds), Dashed white lines, denote traces oF the foliation. (a, b) High-porosity HIPI samples. (e, ) Low-porosity HIP2 samples. (e) Low-porosity PATI (P,) sample. (f) Intermediate porosity PAT! (P.) sample, CHEMICAL COMPACTION OF § A relatively higher content of clasts is observed in porous domains. In more crystalline domains, poros- ity is lower and alignment of clayey and silty mica gives rise to a weak foliation, consistent over at Teast 100 pm (Fig. 9f). The foliation is imegular due to mica wrapping around clasts, and is frequently interrupted or diverted by clusters of equigranular amoeboid authigenic biotite grains (black dashed ellipse in Fig. 9f). Quartz SPO is weak in silty par- ticles, but stronger in clay-sized particles. ‘A matrix of phengite-biotite dominates in meta- pelites with low porosity (g < 5%; Fig. 9e-e). The foliation is more pronounced and continuous than in higher-porosity samples. Mica wrapping around quartz clasts is common. Open and isoclinal micro- folds (1-3 1m) and mica kinks fold the foliation (Fig. 94). A qualitative analysis of the orientation and shape of detrital quartz clasts revealed that SPO is weak. In 2D cross-section, the long axis of the elliptical clasts aligns with the matrix foliation Gig. 9), Dynamic compaction, Detrital micas in high- porosity. metapelites that experienced confined compression possess a weak, imegular and folded foliation (Fig. 10a), Micro-folds and kinks are ocea- sionally observed in detrital and authigenic mi (Fig. 10a, c). Ilite-rich patches in the matrix are not foliated (Fig. 10d), and the silty quartz. grains in the matrix lack a SPO. Intermediate porosity metapelites have a more pronounced foliation, defined by detrital and authi- genic micas (Fig, 10d), The foliation is diverted by micas wrapping around clasts and is interrupted by patches of complexly organized authigenic biotite (Fig. 10d) or underdeveloped phengite. Both detrital and authigenic quartz grains. frequently align parallel to the foliation (Fig. 10d) In low-porosity synthetic metapelites, few illite-rich patches remain, The foliation is well developed (Fig. 10e, f), but occasionally interrupted by clusters of equigranular amoeboid authigenic biotite (Fig. 10e). Micas display I—3 jum open and isoclinal folds or kinks (Fig. 10f). Micro-folds and mica-wrapping fold the main foliation, Qualitative SPO analysis of detrital and authigenic quartz revealed a weak SPO for the former and strong SPO for the latter (Fig. 10e). In both cases, the SPO is parallel to the local foliation. Pores Some of the pore types found in Maplewood Shale, clay-micropores, matrix-clast-voids and matrix-pores are observed in the synthetic meta- pelites (Figs 9 & 10). Intragranular fractures (pore- type VI in Kwon ef al. 2004) form an additional generic pore type. Al # POWDER 141 ‘The submicron clay-micropores separate illite flakes and allow recognition of individual clay par- ticles (Figs 9 & 10). Clay-micropores are typically aligned with the illite flakes. As porosity decreases, illite becomes less abundant. Consequently, the amount of clay-micropores decreases with porosity Phengites in both fully and partially crystallized state are pore-free to at least 500m resolution (Figs 9F & 10f), Ultimately, phengite is distin- guishable from illite in broken surface SE images by tight stacking of ultra-thin mineral sheets. Clay- micropores also distinguish illite from phengite in polished surfaces. Matrix-clast-voids and matrix-pores represent ‘open spaces generated by geometric misfits between angular clasts and fine-grained matrix material and among matrix-forming minerals (Figs 9 & 10) Both pore types range in Iength from submicrons to 2-3 um. Matrix-clast-voids and_matrix-pores are frequently found in high-porosity metapelites (Figs 9a, b & 10a-c). Importantly however, com- paction does not eliminate them. The two pore types are frequently observed in both lithostatically compacted HIP2 and PATI (P_) metapelites (Fig, 9e-F) and high-strain PATI (C) and PATI (C350) samples (Fig. 10e, f) Intragranular fractures represent the openings created by fracturing of rigid (quartz) clasts. The white dashed circles in Figures 9 and 10 high- light several examples of these fractures. In. most ses, intragranular fractures chip off parts of the clasts or split them into two parts. Occasionally, the fractures die out within a clast and remain micro-fissures. Shapes range from slit-like to tab- ular to irregular pyramidal and semi-spherical Intragranular fractures are frequent in PATI (C) and PATI (Cp) samples of diverse porosities (Fig, 10a, b, d), but are rare to absent in PATI (Cyso) samples (Fig. 10e). In lithostatically compacted metapelites, intragranular fractures first appear in intermediate-porosity samples (Fig. 9f) and are more frequent in low-porosity samples (Fig. 9e) ‘There is no evidence for healing of intragranular fractures or filling in with authigenic minerals Discussion Chemical compaction Our three-stage compaction experiments on illite shale powder produced synthetic metapetites ranging in porosity from 1% to 17% (Fig. 6). In mechanical compaction experiments with clay-rich sediment, porosity decreases non-linearly with increasing effective normal stress to values that are still greater than 15% (e.g. Rieke & Chilingarian 1974; Vasseur et al. 1995; Mondol et al. 2007; Fawad etal. 2010). However, the minimum porosity image not available a You have either reached 2 page thts unevalale fer vowing or reached your ievina tit for his book. a You have either reached 2 page thts unevalale fer vowing or reached your ievina tit for his book. a You have either reached 2 page thts unevalale fer vowing or reached your ievina tit for his book. 150 G regime (Chester et af. 1985; Ramsay & Huber 1987) and are described in this regime using the terminol- ogy YPR (Tehalenko 1968; Rutter etal. 1986; Logan etal, 1992). Both tectonic pattems show sets of syn- genetic surfaces bounding relatively undeformed rock. We prefer to adopt the S-C terminology in the present study because this was adopted by previous papers published on the same rocks and in the same region (Alvarez.et al. 1978; Tesei et al. 2013) The S-C fabric is defined as a synthetic frame- work of two sets of surfaces formed at the same time and under the same stress field (Berthé et al. 1979). The main features of S-C tectonites within sedimentary rocks are (Bos & Spiers 2001; Lacroix et al. 2011): (1) S-planes, well-developed folia- tions (initially corresponding to the XY plane of the finite strain) bounding relatively undeformed rock volumes (lithons, e.g. Hippert 1999); and (2) C-planes, shear structures subparallel to the bulk shear plane (defined by the shear zone boundaties). Itis also possible to develop a third set of surfaces, referred to as C’-planes, formed as synthetic shear surfaces during progressively increasing deforma- tion. The C'-planes correspond geometrically to the “RI riedel shears’ in the terminology of Logan et al. (1979) and also to the ‘shear bands’ of White er al. (1980) or ‘extensional crenulation cleavages’ of Platt & Vissers (1980), both applied to higher-grade ‘metamorphic rocks (Rutter ef al. 1986). S-C struc- tures have been observed at very different scales of observation, from micrometres to several hun- dred kilometres (Davis 1984; Hippert 1999). The S-C fabric is commonly used as a reliable sense- of-shear indicator in foliated rocks (e.g. Platt 1984). Several examples of foliated fault zones charac- terized by a penetrative S-C fabric at a mesoscopic cale, dominated by pressure-solution and frictional sliding deformation mechanisms (Marshak et al. 1982), are exposed in the Umbria ‘These fault rocks formed at shallow depth (D <3 km) in sedimentary rocks (mostly consisting of pelagic limestones and marls) commonly associated with major thrust faults (Tesei et al. 2013), for example along the Sibillini Thrust Zone (Koopman 1983; Lavecchia 1985) and along the Monte Cos- cero Thrust (Barchi & Lemmi 1996). However, they have also been mapped as associated with normal faults (e.g. Gubbio normal fault, Collettini et al. 2003; Bussolotto et al. 2007; Bullock et al. 2014) and strike-slip faults (e.g. Valnerina fault, Barchi 1991). Foliated fault zones have also been mapped and described in adjacent regions of the Central Apennines, for example in Abruzzo (Gran Sasso Thrust Zone, Ghisetti 1987) and in western Umbria (Monte Rentella Shear Zone, Meneghini et al. 2012). ‘As commented in the seminal paper by Rutter et al, (1986), ‘geometrically similar structural ENA ETAL. features are frequently found in fault rocks devel- oped under radically different physical conditions and by different deformation mechanisms’ Although Berthé et af. (1979) introduced the S-C terminology for rocks deformed in the high- temperature plastic regime, in our case the geologi- cal and stratigraphic framework highlight that they formed at a depth of <3km (and possibly as small as 1.5 km) and that there was no crystalline plasticity. The main deformation mechanism for these tectonites is pressure-solution, as documented by microstructural analysis in the same region along shear zone developed in the same lithofacies (Tesei et al. 2013, figs 3 & 4). In this paper, we study in detail the mesostruc- tural pattern of the Scheggia Thrust Zone (STZ), a major shear zone affecting the Upper Cretaceou: Oligocene Scaglia Group, consisting of pelagic limestones and marls with few cherty horizons. The STZ is localized along the Scheggia—Foligno Line (Barchi er al, 2012), an important tectonic lineament ofthe northern Umbria~Marche Apennines (Fig. 1) ‘The formation of the STZ is related to a SW- dipping thrust that cuts down through the section of the back-limb of the Monte Nerone—Monte Cucco anticline (De Feyter & Menichetti 1986; Corsi & De Feyter 1991), emplacing younger-on- older rocks We carried out a detailed mesostructural analysis, along the STZ, aiming to: (1) provide a detailed qualitative and quantitative description of a signifi- cant example of S-C tectonites, developed in sedi- mentary rocks at shallow crustal levels; and (2) reconstruct both the general architecture and inter- nal structure of the shear zone. We also studied the factors controlling the genesis and development of the tectonic pattem, ‘The structural analysis of the STZ, considered in its regional framework, allows us to offer some hypotheses on the development of this shear zone and understand the evolution of a portion of the Scheggia—Foligno Line in the deformation history of the Umbria~Marche Apennines. Geological framework Representing the outer part of the Northern Apen- nines, the Umbria—Marche Apennines (UMA)arean are-shaped fold—thrust belt convex to the east and characterized by an eastwards vergence, formed in the framework of the convergence and collision between the African and European continental mar- gins, Located between the front of the external Tus can domain and the relatively undeformed Adriatic foreland, the UMA fold—thrust belt was formed during late Miocene~carly Pliocene time and was S-C FABRICS IN SHALLOW SHEAR ZONE 151 Quaterna Umbria-Marche Neogene + e ‘normal fault| Ligurides Upper Cretaceous-Palaeogene A A Sen] Tuscan Tertiary [EB rassictower-cretaceous thrust fault Fig. 1. Location of the study area and geological sketch of the Umbria~Marche Apennines, showing the major ‘morphostructural provinces and the Scheggi subsequently disrupted by late Pliocene—Quatern- ary extension (e.g. Elter et al, 1975; Barchi 2010). ‘The Scheggia~Foligno Line is an important tec~ tonic boundary between two major morphostruc- tural provinces of the UMA (Fig. 1): the Umbria pre-Apennines and the Umbria—Marche Ridge (Bally et al. 1986; Deiana & Pialli 1994), From the front of Tusean Nappe to the Monte Vicino syn- cline, the Umbria pre-Apennines are characterized by extensive outcrops of Miocene turbidites and by relatively low elevation. The Umbria~Marche Ridge isa belt of major anticlines with extensive exposures ‘oligno Line, The study area is indicated by the rectangle of Mesozoic—Palacogene carbonates (Umbria— Marche succession, e.g. Centamore et al. 1986), and forms the major mountain range of the region. The Scheggia—Foligno Line represents a major dis- continuity in both the deformation style and the tectono-sedimentary evolution of the region (Barchi et al, 2012), ‘Our study area is located close to the village of Scheggia at the northern portion of the Scheggia— Foligno Line, where it separates the Miocene cl tic succession of the Monte Vicino syncline from the Palaeogene carbonate rocks of the back-limb image not available 154 G.LENA ETAL & Rutter 2001), here extended to distributed defor- mation developed within the fault zone. The STZ is superimposed on previously deformed rocks that can be observed below the lower boundary of the STZ, within the middle— lower portion of the Scaglia Rossa Formation, at the base of the gorges cutting across the Monte Cerro Ridge. This background deformation consists of (Fig. 4): (1) steep_pressure-solution cleavages (Alvarez et al. 1976; Lavecchia etal. 1983) at a spacing, of 5~10¢m, mostly localized in the ealear= cous beds; these surfaces present an irregular wavy form and are filled by a thin film of red clay-rich material (most strike NW-SE. and dip SW, average 80/213); (2) subvertical (orthogonal to bedding) exten- sional veins, affecting the more competent (ie. careous and cherty) strata; the main set, trending SW—NE, traverses the fold axis; and (3) asymmetric NE-verging minor (‘parasitic’) open and tight folds, trending subparallel to the Monte Nerone~Monte Cucco anticline axis, associated with the folding of the major anticline (Lavecchia et al, 1983). All of these structures were generated in the early phases of layer-parallel compression just before and/or during the main folding phase. This event was referred to as “flattening pre-buckling” by Lavecchia (1985), This structural association is widespread over the entire Umbria~Marche region, where it has previously been recognized and 4. Structural pattern of the regional deformation. (a) Stereoplots (lower hemisphere projection) of the main structural features: bedding and fold hinges (left); pressure-solution cleavages and extensional veins (right). (b) High-angle pressure-solution cleavages in the calcareous strata of the Scaglia Rossa Formation (left); sub-vertical caleite veins (right). ‘S-C FABRICS IN SHALLOW S Fi exposed. Inset: stereagraphic projection of the fold hinges. described by several authors (e.g. Alvarez et al. 1976; Marshak et al. 1982; Lavecchia et al. 1983; Tavani er al. 2010), This study focuses on the characteristics of the tectonic assemblage developed through the fault core and the damage zone. During fieldwork within the damage zone of the STZ. (in the northern part of the study area), we also observed a regular set of asymmetric NE-verging well-developed open kink and chevron-like folds with a wavelength of 0.5— 5m (Fig, 5). These folds nucleated in the upper part of the Scaglia Rossa Formation, which is char- acterized by the rhythmic alternation of relatively thick calcareous beds and thin marly layers. We measured the attitudes of these folds along a valley (parallel to that shown in Fig. 5) where fold hinges trend NW-SE to NNW-SSE and plunge towards NW 5-10" (inset in Fig. 5). This fold assemblage is located between two detachments corresponding to sheared marly horizons, perpendicular to the orientation of the hinge lines, with bedding-parallel slickenlines. Structural analysi We collected our data at two main structural stat- ions (see location on Figs 2 & 3) within the study area, where we analysed: (1) the STZ fault core (structural station #1); and (2) the effects of 155 IEAR ZONE 5. Panoramic view of the upper patt of the Fiume Gorge, where the folded band of the STZ damage zone is lithology on the structural pattern of the STZ damage zone (structural station #2), The fault core ‘The first structural station is a recent quarry located along National Road 3 (the Via Flaminia) where the fault core of the STZ, is spectacularly exposed, showing a complete section through 30 m of thick- ness (Fig. 6a). The lithology involved in the deformation con- sists of (from bottom to top): (1) thin alternating limestones and marls (Scaglia Variegata Forma- tion); and (2) calcareous marls and argillites (Scaglia Cinerea Formation). The outcrop is characterized by a dense tectonic pattern, in which the bedding is highly transposed and obliterated (Fig. 6b, c). Only’a few remnants of the primary stratification are locally preserved, recognizable in the thyth- mie alternation of reddish marls and whitish lime- stones, and few lens-shaped reliets of cherty layers. ‘The fabric of the tectonites at the fault core (Figs 6 & 7) consists of the following planes. a S-planes: high-angle (dip >70°) pervasive foliation, mostly SW-dipping and closely spaced (0.3~3 cm). This foliation is mesosco- pically defined by anastomozing pressure~ solution seams and impingements (Fig. 6i) and has been reopened and filled with 156 G.LENA ETAL Fig. 6, STZ Fault Core (Structural Station# ~ 43°25.155'N, 12°39.209'E, WGS84): (a) panoramic view of the quarry front; (b) general view of a portion ofthe tectonites;(e) line-drawing ofthe S-C structural pattern (violet line, S-surfaces: black line, C-surfaces: black dot-lines, C’-surfaces: blue polylines, calcite veins; red lines, T2 thrusts); (de) Steeper reverse T? thrusts, displacing the previous formed S-C fabric, often witha staircase trajectories; (f) asymmetrically folded calcite veins, infilling the structural pattern; (gh) boudinage of the cherty layers (SL) highlights the down-section trajectory of C-surfaces; (i) detail of pressure-solution seams and impingements in limestones, 158 G millimette-thick striated calcite, indicating that S-planes were subsequently reactivated as minor reverse shear planes. planes: gently SW-dipping (spaced 5— 10cm) shear surfaces (subparailel to the main tectonic contact) with dip-slip slickenlines. These surfaces indicate a top-to-the-NE sense of shear, often cutting through the SW- dipping strata with a down-section trajectory. (3) planes: rare subhorizontal or gently SE- dipping shear surfaces, widely spaced (20- 50 em). These surfaces are minor shear-planes, synthetic to the previous C-planes, displacing, both C and $ surfaces. The attitudes of the C-planes with respect to the S-planes (Fig. 7) provides unmistakeable evidence of a gently SW-dipping shear zone with a top-to- the-east sense of shear. This is confirmed by: (1) the curved shape of the sigmoids in the S-foliation (Fig. 6c, d): (2) the dip-slip or slightly oblique slick- ensides on both C- and S-planes; and (3) the east- wards displacement of previously formed calcite veins and cherty layers (Fig. 6g, h), the latter repre- senting the remnants of the primary bedding The mesoscopic evidence indicates that S-planes, were generated by pressure-solution (the surfaces are characterized. by impingements and solution seams, Fig. 6i), subsequently opened and filled with striated calcite. The thrust event is therefore sub- sequent to the pressure-solution, even if they were formed during the same tectonic phase. Microstruc- tural analysis carried out in the same region and on the same formation shows that the dominant deformation mechanism for these planes is pressure- solution (Tesei ef al. 2013, figs 3 & 4). Steeper reverse shear planes, referred to here as “T2 thrusts’, can be observed at a spacing of 15— 25 cm and are characterized by dip-slip or slightly oblique slicks. T2 thrusts affect the earlier $-C fabric, displacing and/or rotating the previous S- type foliation (Fig. 6d, ). They are characterized by a staircase trajectory, often reactivating. por- tions of both C and $ surfaces. No new foliation ‘was generated in connection with the development of the T2 thrusts (Fig. 6d). Locally, both C-planes and T2 thrusts exploit the mechanical contrast between marl and carbon- ate horizons resulting in localized bedding-paralle! slip. Boudinage of more competent primary lay- ers (caleareous and/or cherty strata) occurs within the shear zone, due to the combined shear along C-planes, C’-planes and T2 thrusts (Fig. 6g, h). The whole fault zone is affected by a dense pattern of NW-SE-striking calcite veins which can be divi- ded into two sets according to attitude and origin ‘A set of high-angle (dip c. 70°) relatively unde- formed calcite veins (VO) is observed in the cherty ENA ETAL. layers or in some calcareous lenses. These veins are part of the background deformation (see the previous section, ‘The Scheggia Thrust Zone”), pre- dating the STZ shear zone. The other set of veins (V1). filling both foliation and shear surfaces (fabric-parallel veining, Fig. 6e), were probably formed during the deformation of the fault zone (e.g. Meneghini er al, 2012) Both VO and (particularly) V1. veins are asym- metrically folded and distorted (Fig. 6f) with sub- horizontal hinge lines normal to the slip vector (Fig. 7). These folds are NE-verging and coher- ently oriented with the sense of transport of the shear zone. The veins are formed contempora- neously to the S-C fabric, Folding of the veins occurs during the late stage of the evolution of this fabric. The T2 thrusts cut through all fabric elements, indicating that they were formed after the S-C fabric. This tectonic pattern is related to a single compressional phase, so they can also be viewed as the result of a continuum deformation: within this continuum, the sequence of events is however very well defined by the cross-cutting relationships. In the histograms of Figure 8, we analyse the cross-cutting relationship and the distributions of the dip angles of the structural assemblage within the fault core. The histograms show that: (1) S- planes are systematically steeper than both bedd- ing and C-planes, forming an angle of about 50 with the latter; (2) C-planes are generally at a shal- ow angle with respect to the bedding (down-section trajectory), forming an angle of $—10°; and (3) T2 thrusts (dissecting the previously formed S-C fabric) are parallel to or steeper than bedding, form- ing an angle of up to 25° with the latter. The STZ damage zone Structural station #2 is located 250m SE of the quarry of structural station #1, in a narrow canyon 100 m beneath the spectacular barrel-vaulted bridge called ‘Ponte a Botte’. Here the footwall block of the STZ is continuously exposed for a length of c. 300 m (sce location on Figs 2 & 3) ‘This outcrop provides an ideal section through the damage zone of the STZ, 60 m thick, affecting (from top to bottom) the Scaglia Variegata For- mation (thin alternations of marly and marly lime- stones, 10m thick) and the uppermost part of the Scaglia Rossa Formation (mainly micritic pelagic limestone with less-competent reddish and whitish ‘marls, 50 m thick). In this structural station (depicted in Fig. 9) we recorded a detailed lithologic and structural log of the damage zone (Fig. 10), with the aim of analys- ing the geometrical variations of the structural pat- tern with distance from the main tectonic contact 160 Fig. 9. STZ damage zone (structural station #2, 43°25.055'N, 12°39.334'E, WGS84): (a) S-C structural pattern in the calcareous lithosomes: the steep S-surfaces structural pattern in the marly lithosomes ate widely spaced and at high angles with respect to C-planes; (b) S-C surfaces are closely spaced and flattened with respect to the C-surfaces; (©) S-C fabric acts as.a small-scale duplex system, where the dip-angle of the $-surfaces increases toward the left; and (@) sharp contact between two different lithovomes (ic, calcareous v. marly) with contrasting mechanical properties, driving an alsrupt change in the structural style. and the control of the lithology on the observed structures Based on the lithology (calcareous v. marly) we divided the sections into 39 lithosomes, or lithos- tructural units, characterized by relatively homo- geneous lithology and structural pattern. Within these lithosomes we located 58 measurement points where we collected lithological and structural obser- vations, in particular: lithology (calcareous v. marly stratigraphic intervals); attitudes of primary and sec- ondary structures (ie. hedding, C-planes, S-planes); and spacing and angle between S- and C-planes, The spacing of C-planes was measured directly between two adjacent structures. S-planes are how- ever very closely spaced (commonly <1 em) and the spacing is rather variable; in order to obtain an average and representative measure, we therefore considered the length of a transect orthogonal to the structures as containing a fixed number (20) of S-planes. The following observations were made (Figs 9-11) (1) In the calcareous lithosomes the C-planes are generally widely spaced (5-20 em, mean value 10.2 cm, Fig. 9a), the C-planes dip gen- Uy (25~30°) towards WSW showing down- section trajectories and S-spacing is large, in the range of 1-5.5 em (average value 1.9 em). ‘The S-planes are steep (the average dip is 73° towards WSW), and the angle between S and C surfaces is large (average value 47°), (2) Marly lithosomes (generally less than 1m thick) show a penetrative S-C fabric (Fig. 9b), with a C-spacing of 5-12 em (average alue 7.8 cm) and an S-spacing in the range of 0.25-0.75 cm (average value 0.53 em). In these lithosomes the average dip of the C-planes is 35/240 while the S-planes dip S-C FABRICS IN SHALLOW SHEAR ZONE aly FS] creo limestones |_| mas Fig. 10. Lithostructural log of the STZ damage zone: the histograms represent the variation ofthe spacing of S-surfuces, with respect tothe distance from the main thrust, within the 60-m-thick STZ damage zone. The histograms in the centre column refer to the marly lithesomes and those in the right-hand column to the calcarcous lithosomes, Note that the horizontal scale is different for the two types; S-surfaces are much more closely spaced in the marly lithosomes than in, the calcareous lithosomes. It is possible to recognize three deformation bands with different characteristics. 161 162 Dip Angle. freq) (blue histogram bars: dip angle of the C-surfuees; red histogram bars: dip angle of the S-surfaces) 64/240 (Fig. 11). The angle between the S-planes and C-planes is about 29°, lower than that measured in the calcareous lithosomes, (3) The observed structural pattern changes abruptly at the interfaces between calcareous and marly lithosomes (Fig. 9d). S-planes show a markedly sigmoidal shape and often act as a minor duplex system, coupled with the C- planes (Fig. 9c), Along the whole section we observed that C- spacing is not strongly affected by lithology and/ or distance from the main tectonic contact. We argue that bedding thickness is the main factor controlling the spacing of the C-surfaces, whose development was initially driven by the primary ani- sotropy (due to the rhythmic alternations of lime- stones and marls). Dip Angle LL. Geometrical relationships between S- and C-surfaces in the STZ damage zone. The stereoplot and the 'y histogram on the eft referto the marly lithosomes; the graphies onthe right refer to the calcareous lithosomes On the other hand, S-spacing is affected by both lithology (Fig. 9d) and distance from the main tec- tonic contact, Considering the S-spacing value recognized along the whole section, we distin- guished three different deformation zones. (Fig, 10). Starting from the main tectonic contact, these are as follows. Zone 1: 0-21 m, close to the tectonic contact, S spacing is close in both mars (0.2—0.4 em) and limestones (0.9~2.2 em); ‘* Zone 2: 21-44 m, S spacing is still close in the limestones (2.02.1 em) but increases strongly in the marly layers (0.6-0.7 em); and Zone 3: 44~60 m; far from the tectonic contact, the S spacing also increases in the lime- stones (25.5 cm) while it remains constant in the marly layers (0.5—0.7 em), 164 G were already tilted (je. SW-dipping) when the hanging-wall block was transported by the thrust. In other words, we argue that at least part of the growth of the Monte Nerone-Monte Cucco anti- cline pre-dates the emplacement of the thrust sheet. Early folding of the Monte Nerone~Monte Cucco amticline can be suggested, considering that the Monte Vicino syncline (Fig. 2), located at the STZ hanging wall, derives from the folding of an carly Tortonian satellite basin superimposed on the eastern-most portion of the Marnoso—Arenacea foredeep succession, and representing the last epi- sode of turbiditic sedimentation in this area (Cen- tamore ef al. 1977). The deposition of the Monte Vicino sandstones occurred in a satellite basin bounded by the Gubbio anticline to the west and by the Monte Nerone-Monte Cucco anticline to the east (Fig. 1), The presence of this satellite basin therefore constrains the timing of the early folding phases in this area (late Tortonian—early Messinian time). Rarity of C’ surfaces C’ surfaces are synthetic shear planes, developed at a low angle with respect to the C-planes. They rep- resent the late, mature stage of the fault zone evol- ution and are thought to be related to processes of slip localization within the shear zone. With no strong mechanical discontinuity, the relatively homogeneous footwall rocks (Scaslia Group) could have favoured the development of a thick shear zone of more than 70 m thick where the slip is dis- tributed among many parallel C-surfaces, affecting the footwall rock. As slip amount progressively increases, the overall thickness of the shear zone also increases by developing new shear surfaces in the footwall rocks parallel to the main tectonic boundary. The thickening of the fault zone might be related to a strain hardening process. The lack (or scarcity) of C’ shear planes could be indicative of a relatively immature structure, related to a slip distribution between numerous thrusts, within a very thick fault zone. In fact, both regional geology considerations and available seismic data (Barchi et al. 1998) suggest that the STZ. is associated with an important thrust, accommodating a significant horizontal shortening along a regionally extended tectonic boundary (ie. Scheggia—Foligno Line, Barchi et al. 2012), Relationships between the S-C fabric and T2 thrusts ‘The S-C fabric is affected by subsequent hi T2 mesoscopic thrusts, whose staircase trajectory mainly follows the previously formed structures (both C- and S-planes); they are steeper than C ENA ETAL. surfaces and rotate the previously formed S-C pattern (Fig. 6). No new S-planes are generated during this event. T2 thrusts may represent either: (2) the late features of a progressive deformation episode affecting the fault zone; or (2) the evidence of a polyphase compressional deformation, consist- ing at the same position. The first hypothesis supports the idea that the T2 thrusts are the result of embrittle- ment of the fault rock during the late phases of slip, as suggested by the absence of new S-planes developed during this stage. Conclusions Foliated fault zones dominated by pressure-solution mechanisms (Tesei et al. 2013), where cataclastic deformation is minor to almost absent and located lose to the thrust surfaces, can develop at shallow depths (1—3 km) in sedimentary rocks. The Scheg- gia Thrust Zone (STZ) is a significant example of such a fault zone. The structural analysis of the STZ fault core shows a tectonic pattern strictly ana- ogous (from a geometrical and kinematic point of view) to the S-C fabric, typically developed at greater depth in the ductile regime (Berthé et al. 1979). The occurrence of similar shear zones along other thrust zones of the Umbria~Marche Apennines (Tesei ef al. 2013) suggests that the rhythmic alternation of calcareous and marly rocks can favour the development of such fault zones: in particular, the strong variation of clay content in these formations produces an increase of pressure- solution cleavage, enhanced by the presence of clay minerals (Bos et al. 2000) especially under shallow crustal conditions. Mesostructural data collected in the fault core of the STZ confirm the expected down-section trajectory of the thrust, as shown by the mean atti- tude of C-planes with respect to bedding. As expected, the magnitude of deformation decreases with the distance below the main tectonic contact Far below the fault core however, the magni- tude of deformation is lithologically controlled: almost undeformed calcareous beds alternate with intensely sheared intervals, localized along marly horizons. ‘The mechanical log measured through the thick damage zone of the STZ shows that the spacing of S-planes is mainly controlled by two factors: (1) the distance below the main tectonic contact (the pattern is more closely spaced near the main thrust); and (2) the lithology of the protolith: in the marly levels the S-surfaces are flattened and more closely spaced than in calcareous beds. ‘The distribution of the deformation facies within the STZ allows us to hypothesize a three-stage S-C FABRICS IN SHALLOW SHEAR ZONE 165 evolution of this thrust zone, where the shear defor- mation is superimposed on a earlier deformation stage (background deformation) which is respon- sible for the growth of the major anticline at the footwall of the thrust zone. (1) Shearing was initially localized along bed- ding. The pre-existing pressure-solution sur- faces, formed during the earlier folding phase, were progressively rotated and became sigmoidal, since the amount of rotation is stronger along the shear surfaces, This defor- mation facies was locally preserved in the more competent intervals of the damage zone. 2) New shear surfaces subsequently formed, cutting and displacing the primary bedding which was progressively disrupted, while new closely spaced S-planes developed as strain intensity increased. This deformation facies dominates in the STZ fault core close to the main tectonie contact, characterized by intense and pervasive S-C fabric where primary bedding was completely destroyed and can- not be recognized, (3) Asset of new, minor, brittle reverse faults (T2 thrusts) was generated during the last stages of thrusting, displacing and/or folding the pre- existing S-C tectonites and the calcite veins within the fault core, This paper was produced within the framework of the ERC Starting Grant GLASS (259256), Resp. (Supervisor) C.Colletiini. We are grateful to T. Tesei for useful diseus- sions. We also thank the editors D. Faulkner and J. Mecklenburgh, as well as the anonymous reviewers whose comments greatly improved the manuscript. References Atvarez, W., ENGEL per, T. & Lownie, W. 1976. For mation of spaced cleavage and folds in brittle lime- stone by dissolution. Geology, 4, 698-701 Atwaréz, W., ENGELDER, T. & Grist, P. A. 1978, Classi- fication of solution cleavage in pelagic limestones. Geotogy, 6, 263-266. Baty, A. W., Burst, L., Corer, C. & Guetaxpont, R, 1986. Balanced sections and seismic reflection pro- files neross the Central Apennines. Memorie Soctetd Geotogica Hatiana, 38, 257-310. Baacut, M. R, 1991. Una sezione geologica bilanciata tattraverso il settore meridional dell’ Appennino tumbro-marchigiano: I’ Acquasparts-Spoleto-Accumoli. In: PIALL G., Barc, M.R. & Menicuerti, M. (eds) Studi preliminari all’ acquisizione dati del profilo CROP 03 Punta Ala-Gabicce. Studi Geologici Camerti, Volume Speciale 1, 347-362. Baxci, M. R. 2010. The Neogene-Quaternary evolu tion’ of the Norther Apennines: cristal structure, style of deformation and seismicity, Jn: BELTRANDO, M., PECCERILLO, A., MATTEL, M., CONTICELLL S, & DOGLION!, C. (eds) The Geology of traly, Journal of the Virtual Explorer, 36, paper 10, hutp://dx.doi.org/ 10.3809/virtex.2010,00220. Barcut, M.R, & Lesa, M. 1996, Geologia dell'area di M. Coscerno-M.di Civitella (Umbria sud-orientale). Bollestino Societa Geologica Italiana, (18, 601-624. Barc, M,, Laveecuta, G, & MINELLI, G. 1989, Sezione ‘geologiche bilanciate attraverso il sistema a pieghe umbro-marchigiano: 2-La sezione Scheggia-Serra San Abbondio, Bolledino Societa Geologica Haliana 108, 69-81 Baxcut, M.R., De Fevren, A., MaGNaNt,M.B., MINELL, G., PIAL, G. & SoTER«, M. 1998. The structural style ‘of the Umbria-Marche fold and thrust-belt, Memorie Societit Geologica taliana, $2, 557-578. Barc, MR, ALVAREZ, W. & SHIMABUKURO, D. HL 2012. The Umbria-Marche Apennines as a double ‘orogen: observations and hypotheses. fralian Journal of Geosciences (Bollettino Socieid Geologica Ili ‘ana), 131, 258-271 Berruf, D., CHouxxoune, P. & Jecovzo, P. 1979. Onthogneiss, mylonite and non-coaxial deformation fof granites: the example of the South Armorican Shear Zone. Journal of Structural Geology, 1, 31-42. Boccacerni, M. & Cot, M, 1982, Carta strutturale del Appennino settentrionale 1:250,000. CNR Progetto Finalizzato Geodinamica, SELCA, Firenze Bos, B. & Spies, C. J. 2001. Experimental investigation into the microstructural and mechanical evolution ‘of phyllosilicate-bearing fault rock under conditions favouring pressure solution. Journal of Structural Geology, 23, 1187-1202, Bos, B., Peact, C. J, & Spiers, C. J, 2000, Frietional~ Viscous flow of simulated fault gouge caused by the combined effects of phyllosilicates and pressure sol- ution. Tectonophysics, 327, 173-194, Buttock, R. J., Dr Paota, N., Honsworrn, RB. & ‘TRARUCHO-ALEXANDRE, J. 2014, Lithological con: trols on the deformation mechanisms operating within carbonate-hosted faults during the seismic cycle Journal of Structural Geology, $8, 22-42. BUSSOLOTTO, M., BENEDICTO, A., INVERNIZZ1, C., MICAR- ELL, La PLAGNES, V. & DitAxa, G. 2007. Defor ‘mation features within an active normal fault zone in carbonate rocks: the Gubbio fault (Central Apen- nines, Italy). Journal of Structural Geology, 29, 2017-2037, CALAMITA,F., SATOLLL 8. & TuRro, A. 2012. Analysis of ‘thrust shear zones in curve-shaped belts: deformation ‘mode and timing of the Olevano-Antrodoco-Sibillin thrust (Central/Northemn Apennines of Italy). Journal of Structural Geology, 44, 179-187. (Centanone, E., Cimoccuint, U. & Micaxeta, A. 1977, Analisi “dell’evoluzione “tettonico-sedimentaria dei “bucini minori’ torbiditici del Miocene medio-supei iore nell’Appennino Umbro-Marchigiano ¢ Laziale- Abruzzese. 3) Le arenarie di M. Vicino, un modello i conoide sottomarina affogata (Marche Settentro- ale). Studi Geologici Camerti, 3, 7-56. Centawont, E., DEIANA, G., MieaRELLI, A. & Pork M. 1986. I Trias-Paleogene delle Marebe. fn: Centa- MonE, E. & Deiana, G. (eds) La Geologia delle Marche. Studi Geologici Camerti, Special Volume, 9-27. image not available Pegmatite mylonites: origin and significance DENIS GAPAIS* & AWA LAOQUAN BREM BOUNDI Géosciences Rennes, UMR CNRS 6118, Université de Rennes 1, 35042 Rennes cedex, France *Corresponding author (e-mail: denis.gapais @ univ-rennes!fr) Abstract: Pegmatites generally behave as resistant ebjects compared to their host rocks during solid-state deformation. They usually show folding or boudinage within their host racks, features ‘observed from low-grade to high-grade deformations, up to upper-amphibolite metamorphic con- ditions. Here we deseribe mylonitie pegmatite veins emplaced within S.C fabric-bearing two-mica ‘aranites, Pegmatites consist mainly of feldspar and quartz ribbons, with remnants of coarse-grained fs phenocrysts. Quartz ribbons are made of large grains elongated in the shearing direction, with evidence of dynamic recrystallization involving grain boundary migration. Ribbons show very strong lattice-preferred orientations of e-axes at low angles 10 the stretching li these, we propose that quartz ribbons are formed by oriented growth during sation, From. shearing crystal- lization of the pegmatite. Subsequent syn-shearing solid-state deformation involved grain size reduction accompanied by prism slip. The occurrence of pegmatite mylonites within grani- 1oid intrusions should be considered as a univocal argument for synkinematic emplacement. Geologists are used to observing pegmatite veins or dykes affected by folding or boudinage within their host rocks. This holds for all metamorphic environments, from low to high grade. Where emplaced within metasediments, such competency contrast may reflect a softer host rich in weak minerals such as phyllosilicates. However, the occurrence of strong coarse-grained pegmatitic or aplitic dykes within host rocks of comparable min- eralogy is a classical observation (e.g. Ramsay & Hubert 1987, fig. 19.3). Here we describe pegmatite veins that accom- panied the emplacement of peraluminous leuco- granites from the Hereynian South Armorican belt (Brittany, France). Some peamatites are affected by intense mylonitization within their syntectonic granitic host rock. We discuss their deformation mechanisms and their rheological and tectonic significance. Geological setting The South Armorican belt is marked by numerous intrusions of synkinematic peraluminous leucogra- nites (Fig. 1). These granites, basically composed of Kfs, Quz, Pl, Ms and Bt, result from partial melting of sediments (Bernard-Griffiths et at. 1985). Many of them are aligned along the South Armorican Shear Zone (SASZ, Fig. 1), a major dextral upper Carboniferous wrench one (Jégouzo 1980; Gapais & Le Corre 1980) where S-C fab- rics were initially defined (Berthé et al. 19794, 6). This WNW-NW-striking subvertical shear zone ‘crops out along ¢, 300 km (Fig. 1) in length and is marked by granite mylonites of 3-5 km in width (Jégouzo 1980). South of the SASZ, similar leuco- granites emplaced along extensional shear zones of upper Carboniferous age (Gapais et al. 1993) Granites are intruded by pegmatite veins and ‘dykes, some of them showing very intense myloni- tization, We describe structures of pegmatite mylonites observed within these granites, and also provide a detailed description of microstructures and quartz fabrics observed within the Quiberon granite (Fig. 1). Mineralogy and structures Most pegmatite veins and dykes associated with Carboniferous granites from the SASZ. (south Brit- tany, Fig. 1) are coarse-grained rocks composed of (Qtz-Kfs-Ms + Pl-Turm-Grt-Ber. Figure 2 shows examples of pegmatite mylon- ite veins within the Questembert granite deformed along the SASZ and affected by pervasive S-C fabrics attesting to dextral wrenching. Rocks show very elongate quartz and feldspar ribbons, align- ments of tourmaline grains and some centimetric- scale Kfs porphyroclasts. The tourmaline-bearing mineralogy and the occurrence of coarse Kfs por- phyroblasts underline the pegmatitic nature of the rock, Compared with their granitic host, pegmatites are strongly sheared. Figure 3 shows examples of pegmatites the Quiberon granite emplaced within an exten- sional detachment (Gapais et al. 1993; Fig. 1) From: Favtkwer, D. R., Maniant, E. & Mecktenaurctt, J. (eds) 2015. Rock Deformation from Field, Experiments and Theory: A Volume in Honour of Ernie Rutter. Geological Society, London, Special Publications, 409, 167—173. First published online August 4, 2014, up: //dx.doi.org/10.1144/SP409.7 (The Geological Society of London 2015. Publishing disclaimer: www geolsoe.org.uk/pub_ethics

También podría gustarte