Hostname: page-component-848d4c4894-jbqgn Total loading time: 0 Render date: 2024-06-20T06:27:25.205Z Has data issue: false hasContentIssue false

Neocrystallization of clay minerals in the Alhama de Murcia Fault (southeast Spain): implications for fault mechanics

Published online by Cambridge University Press:  05 April 2019

Isabel Abad*
Affiliation:
Departamento de Geología and CEACTierra, Unidad Asociada IACT (CSIC-UGR), Universidad de Jaén, Campus Las Lagunillas s/n, 23071 Jaén, Spain
Juan Jiménez-Millán
Affiliation:
Departamento de Geología and CEACTierra, Unidad Asociada IACT (CSIC-UGR), Universidad de Jaén, Campus Las Lagunillas s/n, 23071 Jaén, Spain
Catalina Sánchez-Roa
Affiliation:
Department of Earth Sciences, University College London, London, UK
Fernando Nieto
Affiliation:
Departamento de Mineralogía y Petrología and IACT (CSIC-UGR), Facultad de Ciencias, Universidad de Granada, Avda. Fuentenueva s/n, 18002 Granada, Spain
Nicolás Velilla
Affiliation:
Departamento de Mineralogía y Petrología, Facultad de Ciencias, Universidad de Granada, Avda. Fuentenueva s/n, 18002 Granada, Spain
*

Abstract

Two preferred textures were observed in the Alhama de Murcia Fault rocks: (a) foliated bands (>100 µm thick) rich in well-crystallized dioctahedral micas, quartz, hematite and dolomite; and (b) ultrafine-grained bands (<100 µm thick) made of patches composed of small mica crystals (<15 µm) and dispersed Fe-oxides. In both textures, kaolinite forms intergrowths or patches of randomly oriented crystals filling gaps or opening layers of presumably inherited detrital mica crystals, which is interpreted as an epitaxial growth from fluids. The Na/K ratio of mica crystals in the thin ultrafine-grained bands shows a wider range than the micas from the foliated bands including muscovitic, intermediate Na/K and paragonitic compositions. The absence of the 0.98 nm intermediate peak in the diffractograms indicates that the small micas are submicroscopically paragonite and phengite intergrowths. The d001 values of the K-dioctahedral micas in the <2 µm and whole fractions are clearly different from each other. The d001 values of micas of the <2 µm fraction are larger, indicating a higher K and lower Na content in the small micas. Their composition corresponds to lower temperatures, suggesting their growth during a genetic episode in the fault. The textural relationships indicate a late growth of kaolinite, probably due to the fluid–rock interaction along fault planes and fractures. The neoformed clay minerals might alter the stability of the fault plane. The absence of expandable clay minerals and the relatively high frictional strength of kaolinite under wet conditions might explain the observed velocity-neutral behaviour of this gouge and earthquake propagation towards the surface.

Type
Article
Copyright
Copyright © Mineralogical Society of Great Britain and Ireland 2019 

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.)

Footnotes

Guest Associate Editor: S. Potel

This paper was originally presented during the session: ‘GG01 – Clays in faults and fractures + MI-03 Clay mineral reaction progress in very low-grade temperature petrologic studies’ of the International Clay Conference 2017.

References

Abad, I., Jiménez-Millán, J., Schleicher, A.M. & van der Pluijm, B. (2017) Mineral characterization, clay quantification and Ar–Ar dating of faulted schists in the Carboneras and Palomares Faults (Betic Cordillera, SE Spain). European Journal of Mineralogy, 29, 1834.Google Scholar
Abad, I., Nieto, F., Peacor, D.R. & Velilla, N. (2003) Prograde and retrograde diagenetic and metamorphic evolution in metapelitic rocks of Sierra Espuña (Spain). Clay Minerals, 38, 123.Google Scholar
Alfaro, P., Delgado, J., de Galdeano C., Sanz, Galindo-Zaldívar, J., García-Tortosa, F.J., López-Garrido, A.C., López-Casado, C., Marín, C., Gil, A. & Borque, M.J. (2008) The Baza Fault: a major active extensional fault in the Central Betic Cordillera (south Spain). International Journal of Earth Sciences, 97, 13531365.Google Scholar
Alfaro, P., Delgado, J., García-Tortosa, F.J., Lenti, L., López, A., López-Casado, C. & Martino, S. (2012) Widespread landslides induced by the Mw 5.1 earthquake of 11 May 2011 in Lorca, SE Spain. Engineering Geology, 137–138, 4052.Google Scholar
Arostegui, J., Irabien, M.J., Nieto, F., Sangüesa, J. & Zuluaga, M.C. (2001) Microtextures and the origin of muscovite-kaolinite intergrowths in sandstones of the Utrillas Formation, Basque Cantabrian Basin, Spain. Clays and Clay Minerals, 49, 529539.Google Scholar
Azañón, J.M. & Crespo-Blanc, A. (2000) Exhumation during a continental collision inferred from the tectonometamorphic evolution of the Alpujárride Complex in the central Betics (Alborán Domain, SE Spain). Tectonics, 19, 549565.Google Scholar
Behnsen, J. & Faulkner, D. (2012) The effect of mineralogy and effective normal stress on frictional strength of sheet silicates. Journal of Structural Geology, 42, 4961.Google Scholar
Bell, J.W., Amelung, F. & King, G.C.P. (1997) Preliminary Late Quaternary slip history of the Carboneras fault, southeastern Spain. Journal of Geodynamics, 24, 5166.Google Scholar
Booth-Rea, G., Azañón, J.M., Azor, A. & García-Dueñas, V. (2004) Influence of strike-slip fault segmentation on drainage evolution and topography. A case study: the Palomares fault zone (southeastern Betics, Spain). Journal of Structural Geology, 26, 16151632.Google Scholar
Bourdelle, F. & Cathelineau, M. (2015) Low-temperature chlorite geothermometry: a graphical representation based on a T-R2+ diagram. European Journal of Mineralogy, 27, 617626.Google Scholar
Brantut, N., Schubnel, A., Rouzaud, J.N., Brunet, F. & Shimamoto, T. (2008) High-velocity frictional properties of a clay-bearing fault gouge and implications for earthquake mechanics. Journal of Geophysical Research, 113, B10401.Google Scholar
Byerlee, J. (1978) Friction of rocks. Pure and Applied Geophysics, 116, 615626.Google Scholar
Coggon, R. & Holland, J.B. (2002) Mixing properties of phengitic micas and revised garnet-phengite thermobarometers. Journal of Metamorphic Geology, 20, 683696.Google Scholar
Collettini, C., Niemejer, A., Viti, C. & Marone, C. (2009) Fault zone fabric and fault weakness. Nature, 462, 907910.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
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
Evans, J.P. & Chester, F.M. (1995) Fluid–rock interaction in faults of the San Andreas system; inferences from San Gabriel Fault rock geochemistry and microstructures. Journal of Geophysical Research, 100, 1300713020.Google Scholar
Faulkner, D.R., Jackson, C.A.L., Lunn, R.J., Schlische, R.W., Shipton, Z.K., Wibberley, C.A.J. & Withjack, M.O. (2010) A review of recent developments concerning the structure, mechanics and fluid flow properties of fault zones. Journal of Structural Geology, 32, 15571575.Google Scholar
Faulkner, D.R., Lewis, A.C. & Rutter, E.H. (2003) On the internal structure and mechanics of large strike-slip fault zones: field observations from the Carboneras fault, southeastern Spain. Tectonophysics, 367, 235251.Google Scholar
Ferrater, M., Booth-Rea, G., Pérez-Peña, J.V., Azañón, J.M., Giaconia, F., Masana, E. (2015) From extension to transpression: Quaternary reorganization of an extensional-related drainage network by the Alhama de Murcia strike-slip fault (eastern Betics). Tectonophysics, 663, 3347.Google Scholar
Frey, M. (1969) A mixed-layer paragonite/phengite of low-grade metamorphic origin. Contributions to Mineralogy and Petrology, 24, 6365.Google Scholar
Frey, M. (1978) Progressive low-grade metamorphism of a black shale formation, Central Swiss Alps, with special reference to pyrophyllite and margarite bearing assemblages. Journal of Petrology, 19, 95135.Google Scholar
Gracia, E., Pallas, R., Soto, J.I., et al. (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
Guidotti, C.V., Mazzoli, C., Sassi, F.P. & Blencoe, J.G. (1992) Compositional controls on the cell dimensions of 2M 1 muscovite and paragonite. European Journal of Mineralogy, 4, 283297.Google Scholar
Guidotti, C.V., Sassi, F.P., Blencoe, J.G., & Selverstone, J. (1994) The paragonite-muscovite solvus: I. P-T-X limits derived from the Na-K compositions of natural, quasi-binary paragonite-muscovite pairs. Geochimica et Cosmochimica Acta, 58, 22692275.Google Scholar
IGN (2001) Instituto Geográfico Nacional, Catálogo Sísmico Nacional hasta el 1900. IGN, Madrid, Spain.Google Scholar
Imber, J., Holdsworth, R.E., Butler, C.A. & Strachan, R.A. (2001) A reappraisal of the Sibson–Scholz fault zone model: the nature of the frictional to viscous (‘brittle–ductile’) transition along a long-lived, crustal-scale fault, Outer Hebrides, Scotland. Tectonics, 20, 601624.Google Scholar
Jiang, W.T. & Peacor, D.R. (1993) Formation and modification of metastable intermediate sodium potasium mica, paragonite, and muscovite in hydrothermally altered metabasites from north Wales. American Mineralogist, 78, 782793.Google Scholar
Jiménez-Millán, J., Abad, I., Hernández-Puentes, P. & Jiménez-Espinosa, R. (2015) 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). Clay Minerals, 50, 619638.Google Scholar
Livi, K.J.T., Christidis, G.E., Árkai, P. & Veblen, D.R. (2008) White mica domain formation: a model for paragonite, margarite, and muscovite formation during prograde metamorphism. American Mineralogist, 93, 520527.Google Scholar
Lockner, D.A., Morrow, C., Moore, D. & Hickman, S. (2011) Low strength of deep San Andreas Fault gouge from SAFOD core. Nature, 472, 8285.Google Scholar
López-Comino, J.A., Mancilla, F., Morales, J. & Stich, D. (2012) Rupture directivity of the 2011, Mw 5.2 Lorca earthquake (Spain). Geophysical Research Letters, 39, L03301.Google Scholar
Masana, E., Martínez-Díaz, 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, B01301.Google Scholar
Martínez-Díaz, J.J., Béjar-Pizarro, M., Álvarez-Gómez, J.A., Mancilla, F.L., Stich, D., Herrera, G. & Morales, J. (2012a) Tectonic and seismic implications of an intersegment rupture. The damaging May 11th 2011 Mw 5.2 Lorca, Spain, earthquake. Tectonophysics, 546–547, 2837.Google Scholar
Martínez-Díaz, J.J., Masana, E. & Ortuño, M. (2012b) Active tectonics of the Alhama de Murcia Fault, Betic Cordillera, Spain. Journal of Iberian Geology, 38, 269286.Google Scholar
Meijninger, B.M.L. & Vissers, R.L.M. (2006) Miocene extensional basin development in the Betic Cordillera, SE Spain revealed through analysis of the Alhama de Murcia and Crevillente faults. Basin Research, 18, 547571.Google Scholar
Montenat, C. & Ott D'Estevou, P. (1995) Late Neogene basins evolving in the Eastern Betic transcurrent fault zone: an illustrated review. Pp. 372386 in: Tertiary Basins of Spain (Friend, P.F. & Dabrio, C.J., editors). Cambridge University Press, Cambridge, UK.Google Scholar
Moore, D. & Lockner, D. (2004) Crystallographic controls on the frictional behavior of dry and water-saturated sheet structure minerals. Journal of Geophysical Research, 109, B03401.Google Scholar
Morrow, C., Moore, D. & Lockner, D. (2000) The effect of mineral bond strength and adsorbed water on fault gouge frictional strength. Geophysical Research Letters, 27, 815818.Google Scholar
Niemeijer, A.R. & Vissers, R.L.M. (2014) Earthquake rupture propagation inferred from the spatial distribution of fault rock frictional properties. Earth and Planetary Science Letters, 396, 154164.Google Scholar
Nieto, F., Velilla, N., Peacor, D.R. & Ortega-Huertas, M. (1994) Regional retrograde alteration of sub-greenschist facies chlorite to smectite. Contributions to Mineralogy and Petrology, 115, 243252.Google Scholar
Parra, T., Vidal, O. & Agard, P. (2002) A thermodynamic model for Fe–Mg dioctahedral K white micas using data from phase-equilibrium experiments and natural pelitic assemblages. Contributions to Mineralogy and Petrology, 143, 706732.Google Scholar
Pouchou, J.L. & Pichoir, F. (1985) ‘PAP’ (f) (r) (t) procedure for improved quantitative microanalysis. Pp. 104–106 in: Microbeam Analysis (Armstrong, J.T., editor). San Francisco Press, San Francisco, CA, USA.Google Scholar
Rutter, E.H., Faulkner, D.R. & Burgess, R. (2012) Structure and geological history of the Carboneras Fault Zone, SE Spain: part of a stretching transform fault system. Journal of Structural Geology, 42, 227245.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., Hofmann, H. & van der Pluijm, B.A. (2013) Constraining clay hydration state and its role in active fault systems. Geochemistry, Geophysics, Geosystems, 14, 10391052.Google Scholar
Schleicher, A.M., van der Pluijm, B.A. & Warr, L.N. (2010) Nanocoatings of clay and creep of the San Andreas Fault at Parkfield, California. Geology, 38, 667670.Google Scholar
Scholz, C.H. (1998) Earthquakes and friction laws. Nature, 391, 3742.Google Scholar
Shau, J.H., Feather, M.E., Essene, E.J. & Peacor, D.R. (1991) Genesis and solvus relations of submicroscopically intergrown paragonite and phengite in a blueschist from northern California. Contributions to Mineralogy and Petrology, 106, 367378.Google Scholar
Sibson, R.H. (1986) Earthquakes and rock deformation in crustal fault zones. Annual Review of Earth and Planetary Sciences, 14, 149175.Google Scholar
Solum, J., van der Pluijm, B., Peacor, D. & Warr, L. (2003) Influence of phyllosilicate mineral assemblages, fabrics, and fluids on the behavior of the Punchbowl fault, southern California. Journal of Geophysical Research, 108, 2233.Google Scholar
Torgersen, E. & Viola, G. (2014) Structural and temporal evolution of a reactivated brittle-ductile fault – part I: fault architecture, strain localization mechanisms and deformation history. Earth and Planetary Science Letters, 407, 205220.Google Scholar
Vidal, O., Lanari, P., Munoz, M., Bourdelle, F. & de Andrade, V. (2016) Deciphering temperature, pressure and oxygen-activity conditions of chlorite formation. Clay Minerals, 51, 615633.Google Scholar
Vrolijk, P. & van der Pluijm, B.A. (1999) Clay gouge. Journal of Structural Geology, 21, 10391048.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
Whitney, D.L. & Evans, B.W. (2010) Abbreviations for names of rock-forming minerals. American Mineralogist, 95, 185187.Google Scholar