Hostname: page-component-8448b6f56d-qsmjn Total loading time: 0 Render date: 2024-04-25T03:33:40.679Z Has data issue: false hasContentIssue false

Influence of phyllosilicates and fluid–rock interaction on the deformation style and mechanical behaviour of quartz-rich rocks in the Carboneras and Palomares fault areas (SE Spain)

Published online by Cambridge University Press:  02 January 2018

J. Jiménez-Millán*
Affiliation:
Department of Geology, University of Jaén. 23071 Jaén, Spain
I. Abad
Affiliation:
Department of Geology, University of Jaén. 23071 Jaén, Spain
P. Hernández-Puentes
Affiliation:
Department of Geology, University of Jaén. 23071 Jaén, Spain
R. Jiménez-Espinosa
Affiliation:
Department of Geology, University of Jaén. 23071 Jaén, Spain
*

Abstract

Deformed quartzitic rocks from the Carboneras and Palomares fault areas (SE Spain) are enriched in phyllosilicates compared to their respective protoliths. Deformation is mainly localized in highly foliated chlorite-rich bands. Quartz-rich bands show brittle deformation developing dolomite-rich cross-cutting veins re-cementing microcataclasite areas. Undamaged lenses within the cataclastic rocks contain patches of phyllosilicates with randomly oriented chlorite and mica. Mg, Fe, water, As and Zn enrichment of the damaged rocks suggests a process of hydrothermal chloritization associated with the Cabo de Gata volcanism. Petrographic characteristics indicate that hydrothermal alteration that produced chlorite and mica-enrichment occurred before faulting. Phyllosilicates provided lubricating properties to the quartzitic rocks, favouring the predominance of creep over seismic stick-slip and reducing the possibility of large seismogenic events. Dolomite cementation as a consequence of fluid–rock interaction processes would have a limited effect, due to the presence of weak phyllosilicate surfaces.

Type
Research Article
Copyright
Copyright © The Mineralogical Society of Great Britain and Ireland 2015

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Bedrosian, P.A., Unsworth, M.J., Egbert, G.D. & Thuerber, C.H. (2004) Geophysical images of creeping segment of the San Andreas Fault: Implications for the role of crustal fluids in the earthquake process. Tectonophysics, 385, 137158.Google Scholar
Benito, R., Garcia-Guinea, J., Valle-Fuentes, F.J. & Recio, P. (1998) Mineralogy, geochemistry and uses of the mordenite-bentonite ash-tuff beds of Los Escullos, Almeria, Spain. Journal of Geochemical Exploration, 62, 229240.Google Scholar
Bolognesi, L. (1997) A tentative correlation between seismic activity and changes in the composition of thermal waters on Vulcano Island, Italy. Geothermics, 26, 379392.CrossRefGoogle Scholar
Bourdelle, F., Parra, T., Chopin, C. & Beyssac, O. (2013) A new chlorite geothermometer for diagenetic to low-grade metamorphic conditions. Contributions to Mineralogy and Petrology, 165, 723—735.Google Scholar
Brace, W.F., Paulding, B.W. Jr. & Scholz, C. (1966) Dilatancy in the fracture of crystalline rocks. Journal of Geophysical Research, 71, 39393953.Google Scholar
Brumm, M., Wang, C.Y. & Manga, M. (2009) Spring temperatures in the Sagehen Basin, Sierra Nevada, CA: Implications for heat flow and groundwater circulation. Geofluids, 9, 195207.Google Scholar
Byerlee, J. (1978) Friction of rocks. Pure and Applied Geophysics, 116, 615626.Google Scholar
Cerón, J.C., Martín-Vallejo, M. & García-Rossell, L. (2000) CO2-rich thermomineral groundwater in the Betic Cordilleras, southeastern Spain: Genesis and tectonic implications. Hydrogeology Journal, 8, 209217.Google Scholar
Chester, F., Evans, J.P. & Biegel, R.L. (1993) Internal structure and weakening mechanisms of the San Andreas Fault. Journal of Geophysical Research, 98, 771786.Google Scholar
Cliff, G. & Lorimer, G.W. (1975) The quantitative analyses of thin specimens. Journal of Microscopy, 103, 203207.Google Scholar
Clark, I. & Fritz, P. (2000) Environmental Isotopes in Hydrogeology. Lewis, New York. 328 pp.Google Scholar
Correns, C.W. (1978) Titanium: behavior in metamorphic reactions. pp. 22-H in: Handbook of Geochemistry, 1 (K.H. Wedepohl, editor). Springer, Berlin.Google Scholar
Dieterich, J.H. & Kilgore, B.D. (1994) Direct observation of frictional contacts: New Insights for state-dependent properties. Pure and Applied Geophysics, 143, 283302.Google Scholar
Dostal, J., Strong, D.F. & Jamieson, R.A. (1980) Trace element mobility in the mylonite zone within the ophiolite aureole, St. Anthony complex, Newfoundland. Earth and Planetary Science Letters, 49, 188192.Google Scholar
Duggen, S., Hoernle, K. & Van der Bogaard, H. (2004) Magmatic evolution of the Alboran region: The role of subduction in forming the western Mediterranean and causing the Messinian Salinity Crisis. Earth and Planetary Science Letters, 218, 91108.Google Scholar
Fagereng, A., Remitti, E. & Sibson, R.H. (2010) Shearveins observed within anisotropic fabric at high angles to the maximum compressive stress. Nature Geoscience, 3, 482485.Google Scholar
Faulkner, D.R. & Rutter, E.H. (2001) Can the maintenance of overpressured fluids in large strike-slip fault zones explain their apparent weakness. Geology, 29, 503506.Google Scholar
Faulkner, D.R., Mitchell, T.M., Hirose, T. & Shimamoto, T. (2009) The frictional properties of phyllosilicates at earthquake slip speeds. Geophysical Research Abstracts, 11, EGU2009-5751-3.Google Scholar
Floyd, P.A. & Winchester, I.A. (1983) Element mobility associated with meta-shear zones within the Ben Hope amphibolite suite. Scotland. Chemical Geology, 39, 115.Google Scholar
Glazner, A.F. & Bartley, J.M. (1991) Volume loss, fluid flow and state of strain in extensional mylonites from the central Mojave Desert, California. Journal of Structural Geology, 13, 587594.Google Scholar
Goette, T., Ramseyer, K. & Pettke, T. (2013) Implications of trace element composition of syntaxial quartz cements for the geochemical conditions during quartz precipitation in sandstones. Sedimentology, 60, 11111127.CrossRefGoogle Scholar
Gracia, E., Palla, R., Soto, J.I., Comas, M., Moreno, X., Masana, E., Santanach, P., Diez, S., García, M., Danobeitia, I. & HITS scientific party (2006) Active faulting offshore SE Spain (Alboran Sea): Implications for earthquake hazard assessment in the Southern Iberian Margin. Earth and Planetary Science Letters, 241, 734749.Google Scholar
Grant, I.A. (1986) The isocon diagram. A simple solution to Gresen's equations for metasomatic alteration. Economic Geology, 81, 19761982.Google Scholar
Gresens, R.L. (1967) Composition-volume relations of metasomatism. Chemical Geology, 2, 47—65.Google Scholar
Hernández-Puentes, P., Jiménez-Espinosa, R. & Jiménez-Millan, I. (2015) Geochemical patterns of groundwater from the Palomares-Carboneras active fault area aquifers (SE Spain): determination of a network of sensitive sites indicators of seismic events. Environmental Earth Sciences, 73, 6341—6354.Google Scholar
Humphris, S.E., Alt, J.C., Teagle, D.A.H. & Honnorez, J.J. (1998) Geochemical changes during hydrothermal alteration of basement in the stockwork beneath the active TAG hydrothermal mound (P.M. Herzig, S.E. Humphris, D.J. Miller & R.A. Zierenberg, editors). Pp. 255-276 in: Proceedings of the Ocean Drilling Programme Scientific Results, 158.Google Scholar
Ikari, M., Niemeijer, A. & Marone, C. (2011) The role of fault zone fabric and lithification state on frictional strength, constitutive behavior, and deformation microstructure. Journal of Geophysical Research, 116, B08404.Google Scholar
Inoue, A., Meunier, A., Patrier-Mas, P., Rigault, C., Beaufort, D. & Vieillard, P. (2009) Application of chemical geothermometry to low temperature trioctahedral chlorites. Clays and Clay Minerals, 57, 371382.Google Scholar
IGN (2001) Instituto Geográfico Nacional, Catálogo Sísmico Nacional hasta el 1900. Madrid, Spain, http:// www.ign.es/ign/resources/sismologia/publicaciones/ Catalogohasta1 900.pdfGoogle Scholar
Italiano, F., Bonfanti, P., Pizzino, L. & Quattrocchi, F. (2010) Geochemistry of fluids discharged over the seismic area of the Southern Apennines (Calabria region, southern Italy): Implications for fluid-fault relationships. Applied Geochemistry, 25, 540554.Google Scholar
Karner, S.L., Marone, C. & Evans, B. (1997) Laboratory study of fault healing and lithification in simulated fault gouge under hydrothermal conditions. Tectonophysics, 277, 41—55.Google Scholar
Kretz, R. (1983) Symbols for rock-forming minerals. American Mineralogist, 68, 277—279.Google Scholar
Lachenbruch, A.H. & Sass, J.H. (1988) The stress heat-flow paradox and thermal results from Cajon Pass. Geophysics Research Letter, 15, 981984.Google Scholar
Lackschewitz, K.S., Singer, A., Botz, R., Garbe-Schönberg, D. & Stoffers, P. (2000) Mineralogical and geochemical characteristics of clay minerals in the region of a major hydrothermal site in the Escanaba Trough, Gorda Ridge, northeast Pacific Ocean, Leg 169. Pp. 1-24 in: Proceedings of the ODP Scientific Results, 169 (R.A. Zierenberg, Y Fouquet, D.J. Miller & W.R. Normark, editors).Google Scholar
Lackschewitz, K.S., Devey, C.W., Stoffers, P., Botz, R., Eisenhauer, A., Kummetz, M., Schmidt, M. & Singer, A. (2004) Mineralogical, geochemical and isotopic characteristics of hydrothermal alteration processes in the active, submarine, felsic-hosted PACMANUS field, Manus Basin, Papua New Guinea. Geochimica et Cosmochimica Acta, 68, 44054427.Google Scholar
Lanari, P., Wagner, T. & Vidal, O. (2014) A thermodynamic model for di-trioctahedral chlorite from experimental and natural data in the system MgO-FeO-Al2O3-SiO2-H2O: applications to P-T sections and geother-mometry. Contributions to Mineralogy and Petrology, 167, 968.Google Scholar
Lloyd, G., Geoffrey, E. & Knipe, R.J. (1992) Deformation mechanisms accommodating faulting of quartzite under upper crustal conditions. Journal of Structural Geology, 14, 127143.Google Scholar
Lubben, J.D., Cline, J.S. & Barker, S.L. (2012) Ore fluid properties and sources from quartz-associated gold at the Betze-Post Carlin-Type Gold Deposit, Nevada, United States. Economic Geology, 107, 13511385.CrossRefGoogle Scholar
Manga, M. & Rowland, J.C. (2009) Response of Alum Rock Springs to the October 30, 2007 Alum Rock earthquake and implications for the origin of increased discharge after earthquakes. Geofluids, 9, 237250.Google Scholar
Masana, E., Martinez-Diaz, J.J., Hernández-Enrile, J.L. & Santanach, P. (2004) The Alhama de Murcia fault (SE Spain), a seismogenic fault in a diffuse plate boundary: Seismotectonic implications for the Ibero-Magrebian region. Journal of Geophysical Research, 109, 117.Google Scholar
Montenat, C. & Ott D'Estevou, P. (1995) Late Neogene basins evolving in the Eastern Betic transcurrent fault zone: An illustrated review. Pp. 372—386 in: Tertiary Basins of Spain (P.F. Friend & C.J. Dabrio, editors). Cambridge University Press, Cambridge, UK.Google Scholar
Montgomery, D.R. & Manga, M. (2003) Streamflow and water well responses to earthquakes. Science, 300, 20472049.Google Scholar
Mook, W.G., Gat, J.R., Roznski, K., Froehlich, K., Geyh, M., Séller, K.L. & Konikow, L.F. (2002) Isótopos Ambientales en el Ciclo Hidrológico. Principios y Aplicaciones. W.G. Publicaciones IGME, Madrid, Spain 596 pp.Google Scholar
Morales-Ruano, S.M., Rosua, F.J.C. & Hach-Ali, P.F. (2000) Epithermal Cu-Au mineralization in the Palai-Islica deposit, Almeria, southeastern Spain: Fluid-inclusion evidence for mixing of fluids as a guide to gold mineralization. The Canadian Mineralogist, 38, 553565.Google Scholar
Morrow, C.A., Moore, D.E. & Lockner, D.A. (2000) The effect of mineral bond strength and adsorbed water on fault gouge frictional strength. Geophysical Research Letters, 27, 815818.Google Scholar
Morrow, C.A., Shi, L.Q. & Byerlee, J.D. (1984) Permeability of fault gouge under confining pressure and shear stress. Journal of Geophysical Research, 89, 31933200.Google Scholar
Muhuri, S.K., Dewers, T.A., Scott, T.E. & Reches, Z. (2003) Interseismic fault strengthening and earthquake-slip instability: Friction or cohesion. Geology, 31, 881884.Google Scholar
Nieto, F., Ortega-Huertas, M., Peacor, D.R. & Arostegui, I. (1996) Evolution of illite/smectite from early diagen-esis through incipient metamorphism in sediments of the Basque-Cantabrian Basin. Clays Clay Minerals, 44, 304323.Google Scholar
O'Hara, K. (1988) Fluid flow and volume loss during mylonitization: an origin for phyllonite in an over-thrust setting, North Carolina, U.S.A. Tectonophysics, 156, 2136.Google Scholar
O'Hara, K. & Blackburn, W.H. (1989) Volume loss model for trace element enrichment in mylonites. Geology, 17, 524527.2.3.CO;2>CrossRefGoogle Scholar
Olsen, M., Scholz, C.H. & Leger, A. (1998) Healing and sealing of a simulated fault gouge under hydrothermal conditions: Implications for fault healing. Journal of Geophysical Research, 103(B4), 74217430.Google Scholar
Rusk, B.G., Lowers, H.A. & Reed, M.H. (2008) Trace elements in hydrothermal quartz: relationships to cathodoluminescent textures and insights into vein formation. Geology, 36, 547—550.Google Scholar
Rutter, E.H., Maddock, R.H., Hall, S.H. & White, S.H. (1986) Comparative microstructures of natural and experimentally produced clay-bearing fault gouges. Pure and Applied Geophysics, 124, 330.Google Scholar
Rutter, E.H., Burgess, R. & Faulkner, D.R. (2014) Constraints on the movement history of the Carboneras Fault Zone (SE Spain) from stratigraphy and 40Ar—39Ar dating of Neogene volcanic rocks. Pp. 79—99 in: Deformation Structures and Processes within the Continental Crust (S. Llana —Fúnez, A. Marcos & F. Bastida, editors). Special Publication, 394, Geological Society, London.Google Scholar
Sanchez-Espana, I., Velasco, F. & Yusta, I. (2000) Hydrothermal alteration of felsic volcanic rocks associated with massive sulphide deposition in the northern Iberian Pyrite Belt (SW Spain). Applied Geochemistry, 15, 12651290.Google Scholar
Sanz de Galdeano, C. (1990) Geologic evolution of the Betic Cordilleras in the western Mediterranean, Miocene to present. Tectonophysics, 172, 107119.Google Scholar
Schleicher, A.M., Vander Pluijm, B.A., Solum, J.G. & Warr, L.N. (2006) The origin and significance of clay minerals on surfaces, in fractures and in veins from SAFOD borehole samples (Parkfield, California). Geophysical Research Letters, 33, L16313.CrossRefGoogle Scholar
Sibson, R.H. (1990) Conditions for fault-valve behaviour. Pp. 15—28 in: Deformation Mechanisms, Rheology and Tectonics (R.J. Knipe & E.H. Rutter, editors). Special Publications, 54, Geological Society, London.Google Scholar
Singh, C.K., Rina, K., Singh, R.P. & Mukherjee, S. (2014) Geochemical characterization and heavy metal contamination of groundwater in Satluj River Basin. Environmental Earth Sciences, 71, 201—216.Google Scholar
Sinha, A.K., Hewitt, D.A. & Rimstidt, J.D. (1986) Fluid interaction and element mobility in the development of ultramylonites. Geology, 14, 833886.Google Scholar
Solum, J.G. & Van der Pluijm, B.A. (2004) Phyllosilicate mineral assemblages of the SAFOD Pilot Hole and comparison with an exhumed segment of the San Andreas Fault System. Geophysical Research Letters, 31, L15S19.Google Scholar
Tesei, T., Collettini, C., Carpenter, B.M., Viti, C. & Marone, C. (2012) Frictional strength and healing behavior of phyllosilicate-rich faults. Journal of Geophysical Research, 117, B09402.Google Scholar
Uehara, S.I. & Shimamoto, T. (2004) Gas permeability evolution of cataclasite and fault gouge in triaxial compression and implications for changes in fault-zone permeability structure through the earthquake cycle. Tectonophysics, 378, 183195.Google Scholar
Vannucchi, P., Remitti, F. & Bettelli, G. (2008) Geological record of fluid flow and seismogenesis along an erosive subducting plate boundary. Nature, 451, 699703.Google Scholar
Vidal, O., Parra, T. & Vieillard, P. (2005) Thermodynamic properties of the Tschermak solid solution in Fe-chlorite: application to natural examples and possible role of oxidation. American Mineralogist, 90, 347358.Google Scholar
Vidal, O., de Andrade, V., Lewin, E., Muñoz, M., Parra, T. & Pascarelli, S. (2006) P-T-deformation-Fe3+/Fe2+ mapping at the thin section scale and comparison with XANES mapping: application to a garnet-bearing metapelite from the Sambagawa metamorphic belt (Japan). Journal of Metamorphic Geology, 24, 669683.Google Scholar
Walker, R.J., Holdsworth, R.E., Imber, J., Faulkner, D.R. & Armitage, P.J. (2013) Fault zone architecture and fluid flow in interlayered basaltic volcaniclastic crystalline sequences. Journal of Structural Geology, 51, 92104.Google Scholar
Wang, C.Y. (1984) On the constitution of the San Andreas fault zone in central California. Journal of Geophysical Research, 89, 58585866.Google Scholar
Warr, L.N. & Cox, S. (2001) Clay mineral transformations and weakening mechanisms along the Alpine Fault, New Zealand. Pp. 85-101 in: The Nature and Tectonic Significance of Fault Zone Weakening (R.E. Holdsworth, R.A. Strachan, J.F. Magloughlin & R.J. Knipe, editors). Special Publications, 186, Geological Society, London.Google Scholar
Winchester, J.A. & Max, M.D. (1984) Element mobility associated with syn-metamorphic shear zones near Scotchport, NW Mayo, Ireland. Journal of Metamorphic Geology, 2, 1—11.Google Scholar
Wu, F.T., Blatter, L. & Robertson, H. (1975) Clay gouges in the San Andreas Fault system and their possible implications. Pure and Applied Geophysics, 113, 8796.Google Scholar
Yuce, G. (2007) A geochemical study of the groundwater in the Misli basin and environmental implications. Environmental Geology, 51, 857—868.Google Scholar
Zhang, S. & Cox, S.F. (2000) Enhancement of fluid permeability during shear deformation of a synthetic mud. Journal of Structural Geology, 22, 13851393.Google Scholar
Zhang, S., Tullis, T.E. & Scruggs, V.J. (1999) Permeability anisotropy and pressure dependency of permeability in experimentally sheared gouge materials. Journal of Structural Geology, 21, 795806.Google Scholar
Zoback, M.D. (2000) Strength of the San Andreas. Nature, 405, 3132.Google Scholar