Hostname: page-component-76fb5796d-45l2p Total loading time: 0 Render date: 2024-04-26T12:01:42.371Z Has data issue: false hasContentIssue false

Palaeomagnetism and magnetic fabrics of the Late Palaeozoic volcanism in the Castejón-Laspaúles basin (Central Pyrenees). Implications for palaeoflow directions and basin configuration

Published online by Cambridge University Press:  07 November 2013

ESTHER IZQUIERDO-LLAVALL*
Affiliation:
Departamento de Ciencias de la Tierra, Universidad de Zaragoza, 50009 Zaragoza, Spain
ANTONIO CASAS-SAINZ
Affiliation:
Departamento de Ciencias de la Tierra, Universidad de Zaragoza, 50009 Zaragoza, Spain
BELÉN OLIVA-URCIA
Affiliation:
Geo-environmental Processes and Global Change, IPE-CSIC, Zaragoza, Spain
ROBERT SCHOLGER
Affiliation:
Chair of Applied Geophysics, University of Leoben, Leoben, Austria
*
Author for correspondence: estheriz@unizar.es

Abstract

The Castejón-Laspaúles basin is one of the South Pyrenean basins of Late Variscan age that were strongly inverted during the Alpine compression (Late Cretaceous–Tertiary). It is mainly composed by Stephanian pyroclastic and volcanic deposits that reach a maximum thickness of ~ 500 m, and are overlain by Permian and Triassic sedimentary units. A palaeomagnetic and magnetic fabrics (AMS) study was carried out in the Stephanian units, where the general absence of flow markers at the outcrop scale and the Alpine inversional structure prevent the straightforward reconstruction of the original volcanic and basinal configuration. Magnetic fabric data are not overprinted by Alpine internal deformation and can be interpreted in terms of primary volcanic and pyroclastic fabrics. The obtained directions coincide in the different sampled units, suggesting a constant source area during the development of the basin, and show the dominance of N–S-trending K1 axes that are interpreted to be parallel to flow directions. Palaeomagnetic data indicate the presence of a pre-folding palaeomagnetic component that is rotated clockwise by an average of +37° (±32°) with regards to the Stephanian reference. This rotation probably took place during Alpine thrusting since it is also registered by the overlying Triassic deposits. The whole dataset is interpreted in terms of basin development under sinistral transtension with two main fault sets: deep-rooted E–W-striking faults, probably responsible for magmatic emissions, and shallow-rooted, listric faults of N–S orientation.

Type
Original Articles
Copyright
Copyright © Cambridge University Press 2013 

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

Allerton, S. 1998. Geometry and kinematics of vertical-axis rotations in fold and thrust belts. Tectonophysics 299, 1530.Google Scholar
Arranz Yagüe, E., Galé-Bornao, C. & Lago-San Jose, M. 1998. Características geológicas y petrográficas del magmatismo del sector Surpirenaico de Denuy (Huesca). Lucas Mallada 10, 4565.Google Scholar
Arthaud, F. & Matte, P. 1977. Late Paleozoic strike-slip in southern Europe and northern Africa: result of a right-lateral shear zone between the Appalachians and the Urals. Geological Society of America Bulletin 88, 1305–20.Google Scholar
Bates, M. P. 1989. Paleomagnetic evidence for rotations and deformation in the Nogueras Zone, Central Southern Pyrenees, Spain. Journal of the Geological Society, London 146, 459–76.CrossRefGoogle Scholar
Bixel, F. & Lucas, Cl. 1983. Magmatisme, tectonique et sédimentation dans les fossés stéphano-permiens des Pyrénées occidentales. Revue de Géologie Dynamique et Géographie Physique 24, 329–42.Google Scholar
Bourquin, S., Durand, M., Diez, J. B., Broutin, J. & Fiuteau, E. 2007. The Permian-Triassic boundary and lower Triassic sedimentation in the Western European basins: an overview. Journal of Iberian Geology 33, 221–36.Google Scholar
Cagnoli, B. & Tarling, D. H. 1997. The reliability of anisotropy of magnetic susceptibility (AMS) data as flow direction indicators in friable base surge and ignimbrite deposits: Italian examples. Journal of Volcanology and Geothermal Research 75, 309–20.Google Scholar
Cañón-Tapia, E. 2004. Anisotropy of magnetic susceptibility of lava flows and dykes: a historical account. In Magnetic Fabric: Methods and Applications (eds Martin-Hernandez, F., Luneburg, C. M., Aubourg, C. & Jackson, M.), pp. 205–25. Geological Society of London, Special Publication no. 238.Google Scholar
Cañón-Tapia, E., Walker, G. P. L. & Herrero-Bervera, E. 1996. The internal structure of lava flows: insights from AMS measurements. Journal of Volcanology and Geothermal Research 70, 2136.CrossRefGoogle Scholar
Cañón-Tapia, E., Walker, G. P. L. & Herrero-Bervera, E. 1997. The internal structure of lava flows- insights from AMS measurements II: Hawaiian pahoehoe, toothpaste lava and ‘a’ā. Journal of Volcanology and Geothermal Research 76, 1946.CrossRefGoogle Scholar
Cantarelli, V., Aldega, L., Corrado, S., Invernizzi, C. & Casas-Sainz, A. In press. Thermal history of the Aragón-Béarn basin (Late Paleozoic, western Pyrenees, Spain): insights into basin tectonic evolution. Italian Journal of Geosciences. Doi 10.3301.UG.2012.41 Google Scholar
Cantarelli, V., Casas-Sainz, A., Corrado, S., Gisbert-Aguilar, J., Invernizzi, C. & Aldega, L. 2009. Late Paleozoic basin evolution in the Western Pyrenees. Geophysical Research Abstracts, EGU General Assembly 11, 8621.Google Scholar
Chadima, M. & Hrouda, F. 2006. Remasoft 3.0 a user-friendly paleomagnetic data browser and analyzer. Travaux Géophysiques 27, 20–1.Google Scholar
Deng, C., Zhu, R., Jackson, M. J., Verosub, K. L. & Singer, M. J. 2001. Variability of the temperature-dependent susceptibility of the Holocene eolian deposits in the Chinese loess plateau: a pedogenesis indicator. Physics and Chemistry of the Earth, Part A: Solid Earth and Geodesy 26, 873–8.Google Scholar
Dunlop, D. J. 1972. Magnetic mineralogy of unheated and heated red sediments by coercivity spectrum analysis. Geophysical Journal of the Royal Astronomical Society 27, 3755.Google Scholar
Ellwood, B. B. 1978. Flow and emplacement direction determined for selected basaltic bodies using magnetic susceptibility anisotropy measurements. Earth and Planetary Science Letters 41, 254–64.Google Scholar
Ellwood, B. B., Osipov, J. B., Kafafy, A. M., Henry, B. & Chlupakova, M. 1993. Primary fabrics in sedimentary and igneous rocks. In Magnetic Anisotropy of Rocks (eds Tarling, D. & Hrouda, F.), pp. 95168. London: Chapman and Hall.Google Scholar
Fisher, R. A. 1953. Dispersion on a sphere. Proceedings of the Royal Society of London A 217, 295305.Google Scholar
García Senz, J., Ramirez Merino, J. I., Navarro Juli, J. J., Rodríguez Santisteban, R., Castaño, R. M., Leyva, F., García Sansegundo, J. & Ramirez del Pozo, J. 2009 a. Memoria Explicativa de la Hoja 213 (Pont de Suert). Madrid: Instituto Geológico y Minero de España.Google Scholar
García Senz, J., Ramirez Merino, J. I., Navarro Juli, J. J., Rodríguez Santisteban, R., Castaño, R. M., Leyva, F., García Sansegundo, J. & Ramirez del Pozo, J. 2009 b. Hoja 213 (Pont de Suert) del Mapa Geológico de España a Escala 1:50000. Madrid: Instituto Geológico y Minero de España.Google Scholar
Gisbert, J. 1983. El Pérmico de los Pirineos españoles. In Carbonífero y Pérmico de España (ed. Martínez Díaz, C.). X Congreso Internacional de Estratigrafía y Geología del Carbonífero, Madrid.Google Scholar
Gisbert, J. 1984. Las molasas tardihércinícas del Pireneo. In Libro Jubilar de J.M. Rios (ed. Comba, J. A.), pp. 168–86. Madrid: Instituto Geológico y Minero de España, 2.Google Scholar
Gisbert, J., Martí, J. & Gascón, F. 1985. Guía de la excursión al Stephaniense, Pérmico y Triásico inferior del Pirineo catalán. II Congreso de Estratigrafía y Paleogeografía del Pérmico y Triásico de España, La Seu d'Urgell, 79 pp.Google Scholar
Halvorsen, E. 1974. The magnetic fabric of some dolerite intrusions, northeast Spitsbergen; implications for the emplacement. Earth and Planetary Science Letters 21, 127–33.CrossRefGoogle Scholar
Henry, B., Plenier, G. & Camps, P. 2003. Post-emplacement tilting of lava flows inferred from magnetic fabric study; the example of Oligocene lavas in the Jeanne d'Arc peninsula (Kerguelen Islands). Journal of Volcanology and Geothermal Research 127, 153–64.Google Scholar
Hrouda, F. 1994. A technique for the measurement of thermal changes of magnetic susceptibility of weakly magnetic rocks by the CS-2 apparatus and KLY-2 Kappabridge. Geophysical Journal International 118, 604–12.Google Scholar
Incoronato, A. F. T., Addison, D. H., Tarling, G. N. & Pescatore, T. 1983. Magnetic fabric investigation of some pyroclastic deposits from the Phlegrean Fields, southern Italy. Nature 306, 461–3.CrossRefGoogle Scholar
Izquierdo-Llavall, E., Casas, A. M., Oliva-Urcia, B. & Scholger, R. 2012 a. Are there vertical axis rotations associated with folded thrusts? Insights from a paleomagnetic study in the Nogueres Zone (Central Pyrenees). Contributions to Geophysics and Geodesy 42, Special Issue, 76–7.Google Scholar
Izquierdo-Llavall, E., Román-Berdiel, T., Casas, A. M., Oliva-Urcia, B., Gil-Peña, I., Soto, R. & Jabaloy, A. 2012 b. Magnetic fabric and structural study of the Eaux-Chaudes intrusion: understanding the Variscan deformation in the Western Axial Zone (Pyrenees). International Journal of Earth Sciences 101, 1817–34.Google Scholar
Jelinek, V. 1977. The statistical theory of measuring anisotropy of magnetic susceptibility of rocks and its application. Geofyzika, BrnoGoogle Scholar
Jelinek, V. 1981. Characterization of the magnetic fabric of rocks. Tectonophysics 79, 6370.Google Scholar
Khan, M. A. 1962. The anisotropy of magnetic susceptibility of some igneous and metamorphic rocks. Journal of Geophysical Research 67, 2873–85.Google Scholar
Knight, M. D., Walker, G. P. L., Ellwood, B. B. & Diehl, J. F. 1986. Stratigraphy, paleomagnetic, and magnetic fabric of the Toba Tuffs: constraints on the source and eruptive styles. Journal of Geophysical Research 91, 10.355382.Google Scholar
Kruiver, P. P., Dekkers, M. J. & Heslop, D. 2001. Quantification of magnetic coercivity by analysis of acquisitions curves of isothermal remanent magnetisation. Earth and Planetary Science Letters 189, 269–76.CrossRefGoogle Scholar
Lago, M., Arranz, E., Pocoví, A., Galé, C. & Gil-Imaz, A. 2004. Permian magmatism and basin dynamics in the southern Pyrenees: a record of the transition from late Variscan transtension to early Alpine extension. In Permo-Carboniferous Magmatism and Rifting in Europe (eds Wilson, M., Neumann, E.-R., Davies, G. R., Timmerman, M. J., Heeremans, M. & Larsen, B. T.), pp. 439–64. Geological Society of London, Special Publication no. 223.Google Scholar
Lowrie, W. 1990. Identification of ferromagnetic minerals in a rock by coercivity and unblocking temperature properties. Geophysical Research Letters 17, 159–62.Google Scholar
Lüneburg, C. M., Lampert, S. A., Hermann, I., Lebit, D., Hirt, A. M., Casey, M. & Lowrie, W. 1999. Magnetic anisotropy, rock fabrics and finite strain in deformed sediments of SW Sardinia (Italy). Tectonophysics 307, 5174.Google Scholar
MacDonald, W. D. & Palmer, H. C. 1990. Flow directions in ash-flow tuffs: a comparison of geological and magnetic susceptibility measurements, Tshirege member (upper Bandelier Tuff), Valles caldera, New Mexico, USA. Bulletin of Volcanology 53, 4559.Google Scholar
MacDonald, W. D., Palmer, H. C., Deino, A. L. & Shen, P.-Y. 2012. Insights into deposition and deformation of intra-caldera ignimbrites, central Nevada. Journal of Volcanology and Geothermal Research 245–246, 4054.Google Scholar
Martí, J. & Mitjavila, J. 1987. Calderas volcánicas pasivas: un ejemplo en el Estefaniense del Pirineo Catalán. Geogaceta 2, 1922.Google Scholar
Martí, J. & Mitjavila, J. 1988. El volcanismo tardihercínico del Pirineo Catalán, II: caracterización de la actividad explosiva. Acta Geológica Hispánica 23, 2131.Google Scholar
McCann, T., Pascal, C., Timmermann, M. J., Krzywiec, P., Lopez-Gomez, J., Wetzel, A., Krawczyk, C. M., Rieke, H. & Lamarche, J. 2006. Post-Variscan (end Carboniferous-Early Permian) basin evolution in Western and Central Europe. In European Lithosphere Dynamics (eds Gee, D. G. & Stephenson, R. A.), pp. 355–88. London: The Geological Society.Google Scholar
Molina-Garza, R. S., Geissman, J. W. & van der Voo, R. 1989. Paleomagnetism of the Dewey Lake Formation (Late Permian), northwest Texas. The end of the Kiaman superchron in North America. Journal of Geophysical Research 94, 17881–8.Google Scholar
Nagtegaal, P. J. C. 1969. Sedimentology, paleoclimatology, and diagenesis of Post-Hercynian continental deposits in the south-Central Pyrenees, Spain. Leidse Geologische Mededelingen 42, 143238.Google Scholar
Oliva-Urcia, B., Casas, A. M., Ramón, M. J., Leiss, B., Mariani, E. & Román-Berdiel, T. 2012 a. On the reliability of AMS in ilmenite-type granites: an insight from the Marimanha pluton, Central Pyrenees. Geophysical Journal International 189, 187203.Google Scholar
Oliva-Urcia, B. & Pueyo, E. L. 2007. Rotational basement kinematics deduced from remagnetized cover rocks (Internal Sierras, southwestern Pyrenees). Tectonics. 26, TC4014.Google Scholar
Osete, M. L. & Palencia-Ortas, A. 2006. Polos paleomagnéticos de Iberia de los últimos 300 millones de años. Física de la Tierra 18, 157–81.Google Scholar
Osete, M. L., Rey, D., Villalaín, J. J. & Juárez, M. T. 1997. The Late Carboniferous to Late Triassic segment of the apparent polar wander path of Iberia. Geologie en Mijnbouw 76, 105–19.Google Scholar
Paquereau-Lebti, P., Fornari, M., Roperch, P., Thouret, J.-C. & Macedo, O. 2008. Paleomagnetism, magnetic fabric, and 40Ar/39Ar dating of Pliocene and Quaternary ignimbrites in the Arequipa area, southern Peru. Bulletin of Volcanology 70, 977–97.Google Scholar
Parés, J. M., van der Voo, R. & Stamatakos, J. A. 1996. Paleomagnetism of Permian and Triassic red beds of NW Spain and implications for the tectonic evolution of the Asturian-Cantabrian arc. Geophysical Journal International 126, 893901.Google Scholar
Petrovský, E. & Kapička, A. 2006. On determination of the Curie point from thermomagnetic curves. Journal of Geophysical Research 111, B12S27. doi: 10.1029/2006JB004507.Google Scholar
Platt, J. P., Allerton, S., Kirker, A., Mandeville, C., Mayfield, A., Platzman, E. S. & Rimi, A. 2003. The ultimate arc: differential displacement, oroclinal bending and vertical axis rotation in the External Betic-Rif arc. Tectonics 22, 1017.Google Scholar
Potter, D. K. & Stephenson, A. 1988. Single-domain particles in rocks and magnetic fabric analysis. Geophysical Research Letters 15, 1097–100.Google Scholar
Pueyo, E. L., Millán, H. & Pocoví, A. 2002. Rotation velocity of a thrust: a paleomagnetic study in the External Sierras (Southern Pyrenees). Sedimentary Geology 146, 191208.Google Scholar
Pueyo, E. L., Pocoví, A., Millán, H. & Sussman, A. J. 2004. Map view models for correcting and calculating shortening estimates in rotated thrust fronts using paleomagnetic data. In Orogenic Curvature: Integrating Paleomagnetic and Structural Analyses (eds. Sussman, A. J., Weil, A. B.), pp. 5771. Geological Society of America Special Paper no. 383.Google Scholar
Rochette, P. 1988. Inverse magnetic fabric in carbonate-bearing rocks. Earth and Planetary Science Letters 90, 229–37.Google Scholar
Rodríguez, L., Cuevas, J. & Tubía, J. M. 2012. Structure of the Anayet Permian basin (Axial Zone, Central Pyrenees). Geophysical Research Abstracts, EGU General Assembly 14, 5422.Google Scholar
Román-Berdiel, T., Casas, A. M., Oliva-Urcia, B., Pueyo, E. L. & Rillo, C. 2004. The main Variscan deformation event in the Pyrenees: new data from the structural study of the Bielsa granite. Journal of Structural Geology 4, 659–77.Google Scholar
Saura, E. & Teixell, A. 2006. Inversion of small basins: effects on structural variations at the leading edge of the Axial Zone antiformal stack (Southern Pyrenees, Spain). Journal of Structural Geology 28, 1909–20.Google Scholar
Schott, J. J. & Perés, A. 1987. Paleomagnetism of Permo-Triassic red beds from the Asturias and Cantabrian Chain (northern Spain): evidence for strong lower Tertiary remagnetizations. Tectonophysics 140, 179–91.Google Scholar
Séguret, M. 1972. Etude tectonique des nappes et séries décollées de la partie centrale du versant sud des Pyrénées. Série géologie structurale no. 2. Montpellier: Publications de l'Université des Sciences et Techniques du Languedoc (Usetela), 155 pp.Google Scholar
Soula, J. C., Lucas, C. & Bessiere, G. 1979. Genesis and evolution of Permian and Triassic basins in the Pyrenees by regional simple shear acting on older Variscan structures: field evidence and experimental models. Tectonophysics 58, 34.Google Scholar
Speksnijder, A. 1985. Anatomy of a strike-slip fault controlled sedimentary basin, Permian of southern Pyrenees, Spain. Sedimentary Geology 44, 179223.Google Scholar
Stampfli, G. M., Mosar, J., Marquer, D., Marchant, R., Baudin, T. & Borel, G. 1998. Subduction and obduction processes in the western Alps. In Continents and their Mantle Roots (eds Vauchez, A. & Meissner, R.). Tectonophysics 296, 159204.Google Scholar
Sussman, A. J., Pueyo, E. L., Chase, C. G., Mitra, G. & Weil, A. B. 2012. The impact of vertical-axis rotations on shortening estimates. Lithosphere 4, 383–94.Google Scholar
Teixell, A. 1996. The Ansó transect of the southern Pyrenees: basement and cover thrust geometries. Journal of Geophysical Research 102, 20.325342.Google Scholar
Valero-Garcés, B. & Gisbert-Aguilar, J. 2004. La extensión post-varisca en la Cordillera Pirenaica. In Geología de España (ed. Vera, J. A.), pp. 231343. Madrid: SGE-IGME.Google Scholar
Van der Voo, R. 1969. Paleomagnetic evidence for the rotation of the Iberian Peninsula. Tectonophysics 7, 556.Google Scholar
Verwey, E. J. W. 1939. Electronic conduction of magnetite (Fe3O4) and its transition point at low temperatures. Nature 144, 327.Google Scholar
Watson, G. S. 1956. A test for randomness of directions. Geophysical Supplements to the Monthly Notices of the Royal Astronomical Society 7, 160–1.Google Scholar
Weil, A. B. & Sussman, A. 2004. Classification of curved orogens based on the timing relationships between structural development and vertical-axis rotations. In Paleomagnetic and Structural Analysis of Orogenic Curvature (eds Sussman, A. & Weil, A. B.), pp. 117. Geological Society of America, Special Paper no. 383.Google Scholar
Ziegler, P. A. 1988. Evolution of the Arctic-North Atlantic and the Western Tethys. American Association of Petroleum Geologists Memoir 43.CrossRefGoogle Scholar