Hostname: page-component-76fb5796d-2lccl Total loading time: 0 Render date: 2024-04-29T11:30:01.034Z Has data issue: false hasContentIssue false

Dissolution collapse of a growing diapir from radial, concentric, and salt-withdrawal faults overprinting in the Salinas de Oro salt diapir, northern Spain

Published online by Cambridge University Press:  24 March 2017

Jesús Guerrero*
Affiliation:
Earth Science Department, University of Zaragoza, 50009 Zaragoza, Spain
*
*Corresponding author at: Earth Science Department, University of Zaragoza, 50009 Zaragoza, Spain. E-mail address: jgiturbe@unizar.es.

Abstract

A geomorphic investigation of the Salinas de Oro salt diapir in the Pyrenees reveals that the ring fracture pattern related to the karstic collapse of the diapir crest may vary significantly depending on the rates of dissolution and salt flow, and the rheology of the overburden. The salt diapir has well-developed concentric faults related to salt dissolution subsidence throughout the Quaternary. Roof strata accommodate subsidence by a combination of downward sagging and brittle collapse leading to the development of a ring monocline that is broken by 5 to 20 m throw conjugated normal faults and a 40 m throw, 9.5-km-long and 200-m-wide keystone graben. The salt diapir top has >100-m-long sinkholes that coalesce to form hollows >70 m deep. Up to 3-km-long radial grabens with a 70 to 90 m vertical throw overprint concentric-ring faulting and displace Quaternary deposits demonstrating active salt flow and diapir rise. Radial faults are linked with salt-withdrawal faults of the Andia Fault Zone (AFZ). Salt flow from the AFZ into the Salinas de Oro salt diapir causes brittle gravitational extension of limestone strata leading to a sequence of grabens and Quaternary faults >10 km long and several hundred meters deep.

Type
Research Article
Copyright
Copyright © University of Washington. Published by Cambridge University Press, 2017 

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

REFERENCES

Alsop, G.I., 1996. Physical modeling of fold and fracture geometries associated with salt diapirism. In: Alsop, G.I., Blundell, D.J., Davison, I. (Eds.), Salt Tectonics. Geological Society Special Publication 100. Geological Society, London, pp. 227241.Google Scholar
Asensio, E., Khazaraddze, G., Echeverria, A., King, R.W., Vilajosana, I., 2012. GPS studies of active deformation in the Pyrenees. Geophysical Journal International 190, 913921.CrossRefGoogle Scholar
Barnolas, A., Pujalte, V., 2004. La Cordillera Pirenaica: Definición, límites y división. In: Vera, J.A. (Ed.), Geología de España. SGE-IGME, Madrid, pp. 233240.Google Scholar
Bertoni, C., Cartwright, J.A., 2005. 3D seismic analysis of circular evaporite dissolution structures, eastern Mediterranean. Journal of the Geological Society 162, 909926.Google Scholar
Biot, M., Odé, H., Roever, W., 1961. Experimental verification of the theory of folding of stratified viscoelastic media. Geological Society of America Bulletin 72, 16211631.Google Scholar
Burr, D.M., Tanaka, K.L., Yoshikawa, K., 2009. Pingos on Earth and Mars. Planetary and Space Science 57, 541555.Google Scholar
Carbonel, D., Gutiérrez, F., Linares, R., Roqué, C., Zarroca, M., McCalpin, J., Guerrero, J., Rodríguez, V., 2013. Differentiating between gravitational and tectonic faults by means of geomorphological mapping, trenching and geophysical surveys: the case of the Zenzano Fault (Iberian Chain, N Spain). Geomorphology 189, 93108.Google Scholar
Carbonel, D., Rosdríguez, V., Gutiérrez, F., McCalpin, J.P., Linares, R., Roqué, C., Zarroca, M., Guerrero, J., 2014a. Sinkhole characterization combining trenching, ground penetrating radar (GPR) and electrical resistivity tomography (ERT). Earth Surface Processes and Landforms 39, 214227.Google Scholar
Carbonel, D., Rodríguez-Tribaldos, V., Gutiérrez, F., Galve, J.P., Guerrero, J., Zaroca, M., Roqué, C., Linares, R., McCalpin, J.P., 2014b. Investigating a damaging buried sinkhole cluster in an urban area integrating multiple techniques: geomorphological surveys, DInSAR, GPR, ERT, and trenching. Geomorphology 229, 316.Google Scholar
Castiella, J., Sole, J., Segismundo, N., Otamendi, A., 1982. Las Aguas subterráneas en Navarra: Proyecto hidrogeológico. Diputación Foral de Navarra, Industrias Gráficas Castuera, Burlada, Spain.Google Scholar
Confederación Hidrográfica del Ebro (CHE), 2016. Inventario de Puntos de Agua Subterránea (IPA): Geoportal SitEbro. Gobierno de España, CHE, Zaragoza, Spain.Google Scholar
Chevrot, S., Sylvander, M., Delouis, B., 2011. A preliminary catalog of moment tensors for the Pyrenees. Tectonophysics 510, 239251.Google Scholar
Davison, I., Alsop, I., Birch, P., Elders, C., Evans, N., Nicholson, H., Rorison, P., Wade, D., Woodward, J., Young, M., 2000. Overburden deformation patterns and mechanism of salt diapir penetration in the Central Graben, North Sea. Marine and Petroleum Geology 17, 601618.Google Scholar
Davison, I., Insley, M., Harper, M., Weston, P., Blundell, D., Mcclay, K., Quallington, A., 1993. Physical modelling of overburden deformation around salt diapirs. Tectonophysics 228, 255274.Google Scholar
De Vicente, G., Cloetingh, S., Muñoz Martin, A., Olaiz, A., Stich, D., Vegas, R., Galindo- Zaldivar, J., Fernández-Lozano, J., 2008. Inversion of moment tensor focal mechanisms for active stresses around the microcontinental Iberia: tectonic implications. Tectonics 27, TC1009. http://dx.doi.org/10.1029/2006TC002093.Google Scholar
Dias, R.P., Cabral, J., 2002. Interpretation of recent structures in an area of cryptokarst evolution-neotectonics versus subsidence genesis. Geodinámica Acta 15, 233248.Google Scholar
Doelling, H.H., Ross, M.L., Mulvey, M.L., 2002. Geologic Map of the Moab 7.5’ Quadrangle, Grand County, Utah. Utah Geological Survey Map 181, scale 1:24.000. Utah Geological Survey, Utah Department of Natural Resources, Salt Lake City, UT.Google Scholar
Dooley, T., Jackson, M.P.A., Hudec, M.R., 2009. Inflation and deflation of deeply buried salt stocks during lateral shortening. Journal of Structural Geology 31, 582600.Google Scholar
Elliott, R., Johnson, J.A., 1980. Structural evolution in the northern part of the Moine thrust belt of NW Scotland. Transactions of the Real Society of Edinburgh Earth Science 71, 6996.Google Scholar
Faci, E., Galé, C., Gil, A., Lago, M., Larrasoaña, J.C., Piedrafita, J.L., Aretxabaleta, A., et al., 2013. El diapiro de Salinas de Oro. Geolodía 13, Jaitz-Salinas de Oro. Gobierno de Navarra y Sociedad Geológica de España. Instituto Geológico y Minero de España, Madrid.Google Scholar
Galve, J.P., Tonelli, C., Gutiérrez, F., Lugli, S., Vescogni, A., Soldati, M., 2015. New insights into the genesis of the Miocene collapse structures of the island of Gozo (Malta, central Mediterranean Sea). Journal of the Geological Society 172, 336348.CrossRefGoogle Scholar
Garcia-Castellanos, D., Vergés, J., Gaspar-Escribano, J., Cloetingh, S., 2003. Interplay between tectonics, climate and fluvial transport during the Cenozoic evolution of the Ebro Basin (NE Iberia). Journal of Geophysical Research: Solid Earth 108, 2347. http://dx.doi.org/10.1029/2002JB002073.Google Scholar
García-Mondéjar, J., 1996. Plate reconstruction of the Bay of Biscay. Geology 24, 635638.Google Scholar
Gaullier, V., Vendeville, B.C., 2005. Salt tectonics driven by sediment progradation: radial spreading of sedimentary lobes prograding above salt. AAPG Bulletin 89, 10811089.Google Scholar
Ge, H., Jackson, M.P.A., 1998. Physical modeling of structures formed by salt withdrawal: implications for deformation caused by salt dissolution. AAPG Bulletin 82, 228250.Google Scholar
Gil, I., Liesa, C., 1994. El campo de fallas de la Sierra de Andia: Modelo Genético. In: Muñoz, A., González, A., Pérez, A. (Eds.), II Congreso del Grupo Español del Terciario: Navarro and Navarro. Jaca, University of Zaragoza, Zaragoza, Spain, pp. 117120.Google Scholar
Guerrero, J., Brunh, R.L., McCalpin, J., Gutiérrez, F., Willis, G., 2015. Salt-dissolution faults versus tectonic faults from the case study of salt collapse in Spanish Valley, SE Utah (USA). Lithosphere 7, 4658.CrossRefGoogle Scholar
Guerrero, J., Gutiérrez, F., Galve, J.P., 2013. Large depressions, thickened terraces, and gravitational deformation in the Ebro River valley (Zaragoza area, NE Spain): evidence of glauberite and halite interstratal karstification. Geomorphology 196, 162176.Google Scholar
Gutiérrez, F., 2004. Origin of the salt valleys in the Canyonlands section of the Colorado Plateau: evaporite-dissolution collapse versus tectonic subsidence. Geomorphology 57, 423435.Google Scholar
Gutiérrez, F., Carbonel, D., Guerrero, J., McCalpin, J.P., Linares, R., Roque, C., Zarroca, C., 2012. Late Holocene episodic displacement on fault scarps related to interstratal dissolution of evaporites (Teruel Neogene Graben, NE Spain). Journal of Structural Geology 34, 219.Google Scholar
Gutiérrez, F., Carbonel, D., Kirkham, R.M., Guerrero, J., Lucha, P., Matthews, V., 2014. Can flexural-slip faults related to evaporite dissolution generate hazardous earthquakes? The case of the Grand Hogback monocline of west-central Colorado. Geological Society of America Bulletin 126, 14811494.Google Scholar
Gutiérrez, F., Galve, J.P., Lucha, P., Castañeda, C., Bonachea, J., Guerrero, J., 2011. Integrating geomorphological mapping, trenching, InSAR and GPR for the identification and characterization of sinkholes in the mantled evaporite karst of the Ebro Valley (NE Spain). Geomorphology 134, 144156.CrossRefGoogle Scholar
Gutiérrez, F., Guerrero, J., Lucha, P., 2008. A genetic classification of sinkholes based on the analysis of evaporite paleokarst exposures in Spain. Environmental Geology 53, 9931006.Google Scholar
Hill, C., 1996. Geology of the Delaware Basin, Guadalupe, Apache, and Glass Mountains, New Mexico and West Texas. Permian Basin Section, Society of Economic Paleontologists and Mineralogists (SEPM) Publication 96-39. Permian Basin Section–SEPM, Midland, TX.Google Scholar
Holohan, E.P., Troll, V.R., Walter, T.R., Mqnn, S., McDonnell, S., Shipton, Z.K., 2005. Elliptical calderas in active tectonic settings: an experimental approach. Journal of Volcanology and Geothermal Research 144, 119136.Google Scholar
Hudec, M.R., Jackson, M.P.A., 2007. Terra infirma: understanding salt tectonics. Earth Science Reviews 82, 128.CrossRefGoogle Scholar
Hudec, M.R., Jackson, M.P.A., 2011. The Salt Mine: A Digital Atlas of Salt Tectonics. American Association of Petroleum Geologists (AAPG) Memoir 99. Bureau of Economic Geology Udden Book Series 5. Bureau of Economic Geology, Jackson School of Geosciences, University of Texas at Austin, Austin, TX; AAPG, Tulsa, OK.Google Scholar
Hudec, M.R., Jackson, M.P.A., Schultz-Ela, D., 2009. The paradox of minibasin subsidence into salt: clues to the evolution of crustal basins. Geological Society of America Bulletin 121, 201221.Google Scholar
Instituto Geológico y Minero de España (IGME), 1990a. Documentos sobre la geología del subsuelo de España: Cantábrica. IGME, Madrid.Google Scholar
Instituto Geológico y Minero de España (IGME), 1990b. Documentos sobre la geología del subsuelo de España: Ebro-Pirineos. IGME, Madrid.Google Scholar
Jackson, M.P.A., Adams, J.B., Dooley, T.P., Gillespie, A.R., Montgomery, D.R., 2011. Modeling the collapse of Hebes Chasma, Valles Marineris, Mars. Geological Society of America Bulletin 123, 15961627.Google Scholar
Jackson, M.P.A., Vendeville, B.C., 1994. Regional extension as a geologic trigger for diapirism. Geological Society of America Bulletin 106, 5773.2.3.CO;2>CrossRefGoogle Scholar
Kennedy, B., Stix, J., Vallance, J., Lavallée, Y., Longpré, M.A., 2004. Controls on caldera structure: results from analogue sandbox modeling. Geological Society of America Bulletin 116, 515524.Google Scholar
Kirkham, R.M., Streufert, R.K., Kunk, M.J., Budahn, J.R., Hudson, W., Perry, W. Jr. 2002. Evaporite tectonism in the lower Roaring Fork River valley, west-central Colorado. In: Kirkham, R.M., Scott, R.B., Judkins, W. (Eds.), Late Cenozoic Evaporite Tectonism and Volcanism in West-Central Colorado. Geological Society of America, Special Papers 366, 7399.Google Scholar
Larrasoaña, J.C., Parés, J.M., Millán, H., del Valle, J., Pueyo, E.L., 2003. Paleomagnetic, structural, and stratigraphic constraints on transverse fault kinematics during basin inversion: the Pamplona Fault (Pyrenees, north Spain). Tectonics 22, 1071. http://dx.doi.org/10.1029/2002TC001446.Google Scholar
Lei, C., Ren, J., Clift, P.D., Wang, Z., Li, X., Tong, C., 2011. The structure and formation of diapirs in the Yinggehai–Song Hong Basin, South China Sea. Marine and Petroleum Geology 28, 980991.Google Scholar
Lenaroz, B., 1993. Geomorfología y Geología Ambiental de la Ribera de Navarra. PhD dissertation, University of Zaragoza, Zaragoza, Spain.Google Scholar
Lewis, C.J., McDonald, E.V., Sancho, C., Peña, J.L., Rhodes, E.J., 2009. Climatic implications of correlated Upper Pleistocene glacial and fluvial deposits on the Cinca and Gállego Rivers (NE Spain) based on OSL dating and soil stratigraphy. Global and Planetary Change 67, 141152.Google Scholar
Liesa, C., 1999. Fracturación y campos de esfuerzos compresivos alpinos en la Cordillera Ibérica y el NE peninsular. PhD dissertation, University of Zaragoza, Zaragoza, Spain.Google Scholar
Luzón, A., Pérez, A., Soriano, M.A., Pocoví, A., 2008. Sedimentary record of Pleistocene evolution in the Ebro Basin (NE Spain). Sedimentary Geology 205, 113.Google Scholar
Mackay, J.R., 1998. Pingo Growth and collapse, Tuktoyaktuk Peninsula Area, Western Arctic Coast, Canada: a long-term field study. Géographie physique et Quaternaire 52, 271323.Google Scholar
Marti, J., Ablay, G.J., Redshaw, L.T., Sparks, R.S.J., 1994. Experimental studies of collapse calderas. Journal of the Geological Society 151, 919929.Google Scholar
McCalpin, J.P., 2009. Paleoseismology in extensional tectonic environments. In: McCalpin, J.P. (Ed.), Paleoseismology. Academic Press, San Diego, CA, pp. 171277.Google Scholar
McDonnell, A., Loucks, R.G., Dooley, T., 2007. Quantifying the origin and geometry of circular sag structures in northern Fort Worth Basin, Texas: paleocave collapse, pull-apart fault systems, or hydrothermal alteration. AAPG Bulletin 91, 12951318.Google Scholar
Muñoz, J.A., 1992. Evolution of a continental collision belt: ECORS-Pyrenees crustal balanced cross-section. In: McKlay, K.R. (Ed.), Thrust Tectonics. Chapman and Hall, New York, pp. 235246.Google Scholar
Neal, J.T., Colpitts, R.M., Johnson, K.S., 1998. Evaporite karst in the Holbrook Basin, Arizona. In: Borchers, J.W. (Ed.), Land Subsidence: Case Studies and Current Research. Association of Engineering Geologists Special Publication 8. Star, Belmont, CA, pp. 373384.Google Scholar
Nocquet, J.M., 2012. Present-day kinematics of the Mediterranean: a comprehensive overview of GPS results. Tectonophysics 579, 220242.Google Scholar
Nocquet, J.M., Calais, E., 2004. Geodetic measurements of crustal deformation in the western Mediterranean and Europe. Pure Applied Geophysics 161, 661681.Google Scholar
Nutz, A., Franc, J., Ghienne, O., Torch, P., 2013. Circular, cryogenic structures from the Hirnantian deglaciation sequence (Anti-Atlas, Morocco). Journal of Sedimentary Research 83, 115131.Google Scholar
Olive, A., López-Horgue, M.A., Baceta, J., Niñero, S., Villanueva, E., 2010. Memoria y mapa geológico de Navarra de la hoja de Lezaun (140) a escala 1:25.000. Gobierno de Navarra, Departamento de obras públicas, transportes y comunicaciones, Pamplona, Spain.Google Scholar
Parker, T.J., McDowell, A.N., 1955. Model studies of saltdome tectonics. AAPG Bulletin 39, 23842470.Google Scholar
Payros, A., 1997. El Eoceno de la Cuenca de Pamplona: Estratigrafia secuencial y evolución paleogeográfica. PhD dissertation, Basque Country University, Bilbao, Spain.Google Scholar
Payros, A., Pujalte, V., Baceta, J.I., Orue-Etxebarria, X., Serrakiel, J., 1996. Las calizas eocenas del Oeste de Navarra: revisión, redefinición y nueva interpretación de sus unidades estratigráficas. In: Astibia, H., Valle, J., Murelaga, X., Serrakiel, J. (Eds.), Libro Homenaje a Máximo Ruiz de Gaona, Príncipe de Viana. Gobierno de Navarra, Pamplona, Spain, pp. 137153.Google Scholar
Pérez-Rivarés, F., Garcés, M., Arenas, C., Pardo, G., 2004. Magnetostratigraphy of the Miocene continental deposits of the Montes de Castejón (central Ebro Basin, Spain): geochronological and paleoenvironmental implications. Geológica Acta 2, 221234.Google Scholar
Pflug, R., 1973. El diapiro de Estella. Munibe Ciencias Naturales 2–4, 171202.Google Scholar
Pinto, V., Casas, A., Rivero, L., Torné, M., 2005. 3D gravity modeling of the Triassic salt diapirs of the Cubeta Alavesa (northern Spain). Tectonophysics 405, 6575.Google Scholar
Quintà, A., Tavani, S., Roca, E., 2012. Fracture pattern analysis as a tool for constraining the interaction between regional and diapir-related stress fields: Poza de la Sal Diapir (Basque Pyrenees, Spain). In: Alsop, G.I., Archer, S.G., Hartley, A.J., Grant, N.T., Hodgkinson, R. (Eds.), Salt Tectonics, Sediments and Prospectivity. Geological Society, London, Special Publications 363, 521532.Google Scholar
Ramírez, J.I., Olive, A., Alvaro, M., Ramírez del Pozo, J., Meléndez, A., Gutiérrez Elorza, M., Carbayo, A., Villalobos, L., León, L., Gabaldón, V., 1987. Mapa Geológico de España 1:50.000 Estella, Hoja 140. 1ª Edición. Instituto Geologico y Minero de España, Gráficas Topacio, Madrid.Google Scholar
Ramos, G., Azcón, A., Araguás, L., García, A., 2004. Estudio del impacto hidrogeológico de la inyección profunda de salmuera procedente de las operaciones mineras de potasas de Subiza (Navarra). Instituto Geológico y Minero de España, Madrid.Google Scholar
Ramsay, J., Huber, M., 1987. The Techniques of Modern Structural Geology: Folds and Fractures. Academic Press, London.Google Scholar
Ranalli, G., 1995. Rheology of the Earth. 2nd ed. Chapman and Hall, London.Google Scholar
Rigo, A., Vernant, P., Feigl, K.L., Goula, X., Khazaradze, G., Talaya, G., Morel, L., et al., 2015. Present-day deformation of the Pyrenees revealed by GPS surveying and earthquake focal mechanisms until 2011. Geophysical Journal International 201, 947964.Google Scholar
Rowan, M.G., Lawton, T.F., Giles, K.A., Ratliff, R.A., 2003. Near-salt deformation in La Popa basin, Mexico, and the northern Gulf of Mexico: a general model for passive diapirism. AAPG Bulletin 87, 733756.Google Scholar
Salvany, J.M., 1990. Introducción a las evaporitas triásicas de las cadenas periféricas de la cuenca del Ebro: Catalánides, Pirineo y Región Cantábrica. In: Orti, F., Salvany, J.M. (Eds.), Formaciones evaporíticas de la Cuenca del Ebro y cadenas periféricas y de la zona de Levante. ENRESA-GPPG, Universidad de Barcelona, Barcelona, Spain, pp. 920.Google Scholar
Schultz-Ela, D.D., Jackson, M.P.A., Vendeville, B.C., 1993. Mechanics of active salt diapirism. Tectonophysics 228, 275312.Google Scholar
Serrano, A., Martinez del Olmo, W., 1990. Tectónica salina en el Dominio Cántabro–Navarro: evolución, edad y origen de las estructuras salinas. In: Orti, F., Salvany, J.M. (Eds.), Formaciones evaporíticas de la Cuenca del Ebro y cadenas periféricas y de la zona de Levante: Nuevas Aportaciones y Guía de Superficie. Empresa Nacional De Residuos Radiactivos S.A. ENRESA-GPPG, Universidad de Barcelona, Barcelona, Spain, pp. 3953.Google Scholar
Serrano, A., Martínez del Olmo, W., 2004. Estructuras diapíricas de la zona meridional de la Cuenca Vasco-Cantábrica. In: Vera, J.A. (Ed.), Geología de España. Sociedad Geológica de España, Instituto Geológico y Minero de España, Madrid, pp. 334338.Google Scholar
Simón, J.L., Soriano, M.A., Arlegui, L.E., Gracia, J., Liesa, C., Pocoví, A., 2008. Space-time distribution of ancient and active alluvial karst subsidence. Examples from the central Ebro Basin, Spain. Environmental Geology 53, 10571065.Google Scholar
Smith, R.I., Hodgson, N., Fulton, M.M., 1993. Salt control on Triassic reservoir distribution, UKCS Central North Sea. In: Parker, J.R. (Ed.), Petroleum Geology of Northwest Europe: Proceedings of the 4th Conference. Geological Society, London, pp. 547557.Google Scholar
Sørensen, K., 1998. The salt pillow to diapir transition: evidence from unroofing unconformities in the Norwegian-Danish Basin. Petroleum Geoscience 4, 193202.Google Scholar
Squyres, S.W., Janes, D.M., Baer, G., Bindschadler, D.L., Schubert, G., Sharpton, L., Stofan, E.R., 1992. The morphology and evolution of coronae on Venus. Journal of Geophysical Research: Planets 97, 1361113634.Google Scholar
Stewart, S.A., 2006. Implications of passive salt diapir kinematics for reservoir segmentation by radial and concentric faults. Marine and Petroleum Geology 23, 843853.Google Scholar
Stewart, S.A., Clarke, J.A., 1999. Impact of salt on the structure of the central North Sea hydrocarbon fairways. Petroleum Geology Conference Series 5, 179200.Google Scholar
Stewart, S.A., Harvey, M.J., Otto, S.C., Weston, P.J., 1996. Influence of salt on fault geometry: examples from the UK salt basins. Geological Society, London, Special Publications 100, 175202.Google Scholar
Stüwe, K., 2007. Geodynamics of the Lithosphere: An Introduction. Springer-Verlag, New York.Google Scholar
Talbot, C.J., 1998. Extrusions of Hormuz salt in Iran. In: Blundell, D.J., Scott, A.C. (Eds.), Lyell: The Past Is the Key to the Present. Geological Society, London, Special Publications 143, 315334.Google Scholar
Talbot, C.J., Aftabi, P., 2004. Geology and models of salt extrusion at Qum Kuh, central Iran. Journal of the Geological Society 161, 114.Google Scholar
Troll, V.R., Walter, T.R., Schmincke, H.U., 2002. Cyclic caldera collapse: piston or piecemeal subsidence: field and experimental evidence. Geology 30, 135138.Google Scholar
Trudgill, B.D., 2011. Evolution of salt structures in the northern Paradox Basin: controls on evaporite deposition, salt wall growth and supra-salt stratigraphic architecture. Basin Research 23, 208238.Google Scholar
Trusheim, E., 1960. Mechanism of salt migration in northern Germany. AAPG Bulletin 44, 15191540.Google Scholar
Turner, J.P., 1996. Switches in subduction and lateral termination of mountain belts: Pyrenees-Cantabrian transition, Spain. Journal of the Geological Society 153, 563571.Google Scholar
Vendeville, B.C., Jackson, M.P.A., 1992a. The fall of diapirs during thin-skinned extension. Marine and Petroleum Geology 9, 354371.Google Scholar
Vendeville, B.C., Jackson, M.P.A., 1992b. The rise of diapirs during thin-skinned extension. Marine and Petroleum Geology 9, 331353.Google Scholar
Vergés, J., Fernàndez, M., Martínez, A., 2002. The Pyrenean orogen: pre-, syn-, and post-collisional evolution. In: Rosenbaum, G., Lister, G.S. (Ed.), Reconstruction of the Evolution of the Alpine-Himalayan Orogen. Journal of the Virtual Explorer 8, 5574.Google Scholar
Vernant, P., Hivert, F., Chéry, J., Steer, P., Cattin, R., Rigo, A., 2013. Erosion-induced isostatic rebound triggers extension in low convergent mountain ranges. Geology 41, 467470.Google Scholar
Walter, T.R., Troll, R., 2001. Formation of caldera periphery faults: an experimental study. Bulletin of Volcanology 63, 191203.Google Scholar
Withjack, M., Scheiner, C., 1982. Fault patterns associated with domes: an experimental and analytical study. AAPG Bulletin 66, 302316.Google Scholar
Yamada, Y., Okamura, H., Tamura, Y., Tsuneyama, F., 2005. Analog models of faults associated with salt doming and wrenching: application to offshore United Arab Emirates. In: Sorkhabi, R., Tsuji, Y. (Eds.), Faults, Fluid Flow, and Petroleum Traps. American Association of Petroleum Geologists (AAPG) Memoir 85. AAPG, Tulsa, OK, pp. 95106.Google Scholar