Hostname: page-component-76fb5796d-dfsvx Total loading time: 0 Render date: 2024-04-26T03:00:14.049Z Has data issue: false hasContentIssue false

Middle Pleistocene seismically induced clay diapirism in an intraplate zone, western Brittany, France

Published online by Cambridge University Press:  31 August 2018

Brigitte van Vliet-Lanoë*
Affiliation:
UMR 6538 Géosciences Ocean, UBO- CNRS, IUEM, pl. N. Copernic, 29280 Plouzané, France
Christine Authemayou
Affiliation:
UMR 6538 Géosciences Ocean, UBO- CNRS, IUEM, pl. N. Copernic, 29280 Plouzané, France
Stéphane Molliex
Affiliation:
UMR 6538 Géosciences Ocean, UBO- CNRS, IUEM, pl. N. Copernic, 29280 Plouzané, France
Michael Hugh Field
Affiliation:
Archaeo- and Palaeo-botany Laboratory, Faculty of Archaeology, Leiden University, P.O. Box 9515, 2300 RA Leiden, the Netherlands
Manfred Frechen
Affiliation:
Leibniz-Institut für Angewandte Geophysik, Stilleweg 2, 30655 Hannover, Germany
Pascal Le Roy
Affiliation:
Archaeo- and Palaeo-botany Laboratory, Faculty of Archaeology, Leiden University, P.O. Box 9515, 2300 RA Leiden, the Netherlands
Julie Perrot
Affiliation:
Archaeo- and Palaeo-botany Laboratory, Faculty of Archaeology, Leiden University, P.O. Box 9515, 2300 RA Leiden, the Netherlands
Valérie Andrieu-Ponel
Affiliation:
Aix-Marseille Université, Avignon Université, CNRS, IRD, IMBE, Technopôle Arbois-Méditerranée, Bât. Villemin, BP 80, F-13545 Aix-en-Provence Cedex 04, France
Gwendoline Grégoire
Affiliation:
Archaeo- and Palaeo-botany Laboratory, Faculty of Archaeology, Leiden University, P.O. Box 9515, 2300 RA Leiden, the Netherlands
Bernard Hallégouët
Affiliation:
Retired Geography depart. UBO, 40 rue Commandant Boennec, 29490 Guipavas, France
*
*Corresponding author at: UMR 6538 Géosciences Ocean, UBO- CNRS, IUEM, pl. N. Copernic, 29280 Brest, France. E-mail address: brigitte.vanvlietlanoe@univ-brest.fr (B. van Vliet-Lanoë).

Abstract

The Brittany region of France is located in a low seismicity intraplate zone. Most of the instrumented earthquakes are limited to a shallow crustal depth without surface rupture. A paleoseismological analysis was performed on deposits on the Crozon Peninsula and in the Elorn estuary. We highlight hydroplastic deformations induced by liquefaction leading to clay diapirism, which were likely triggered by past earthquakes. This diapirism seems to be frequent in continental nonconsolidated sediments and to develop on the inherited tectonic structures, when a shallow water table and confining layers exist. Timing of deformation is dated using paleoenvironmental data, and electron spin resonance and infrared-stimulated luminescence dating methods. Two seismic periods were identified in western Europe during early Marine Oxygen Isotope Stage (MIS) 10 (~380 ka) and early MIS 8 (~280–265 ka). The lack of similar deformations affecting the Holocene tidal deposits in the Bay of Brest suggests that the magnitude of the triggering paleoearthquakes is probably higher (Mw ~6) than the recent events (Mw 5.4). These unusual intraplate major paleoearthquakes need specific factors affecting the far-field crustal stress loading to be triggered, such as a brief acceleration of the Africa-Eurasia lithospheric plate convergence, glacio-isostatic stress perturbations associated with the onset of major glaciations in northern Europe, or other processes induced by orbital forcing.

Type
Research Article
Copyright
Copyright © University of Washington. Published by Cambridge University Press, 2018 

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

Andreieff, P., Bouysse, P., Horn, R., Monciardini, C., 1972. Contribution à l’étude géologiques des approches de la Manche. Mémoires du Bureau de recherches géologiques et minières 79, 3148.Google Scholar
Andrieu-Ponel, V., Demory, F., Perrin, M., Cihat Alcicek, M., Lebatard, A.-E., Nomade, S., Djamali, M., Rochette, P., Helvacı, C., 2016. The long climatic sequence of Acigol Lake, NW Turkey: first chronological and palynological results. 25eme Réunion des sciences de la Terre, Caen, France.Google Scholar
Anell, I., Thybo, H., Artemieva, I.M., 2009. Cenozoic uplift and subsidence in the North Atlantic region: geological evidence revisited. Tectonophysics 474, 78105.Google Scholar
Arroucau, P., 2006. Sismicité du Massif armoricain: relocalisation et interprétation tectonique. PhD dissertation, Earth Sciences, University of Nantes, Nantes, France.Google Scholar
Bahain, J.J., Falguères, C., Laurent, M., Voinchet, P., Dolo, J.M., Antoine, P., Tuffreau, A., 2007. ESR chronology of the Somme River Terrace system and first human settlements in Northern France. Quaternary Geochronology 2, 356362.Google Scholar
Ballèvre, M., Bosse, V., Ducassou, C., Pitra, P., 2009. Palaeozoic history of the Armorican Massif: models for the tectonic evolution of the suture zones. Comptes Rendus Geoscience 341, 174201.Google Scholar
Beck, C., 2009. Late Quaternary lacustrine paleo-seismic archives in north-western Alps: examples of earthquake-origin assessment of sedimentary disturbances. Earth-Science Reviews 96, 327344.Google Scholar
Bendick, R., Bilham, R., 2017. Do weak global stresses synchronize earthquakes? Geophysical Research Letters 44, 83208327.Google Scholar
Bergerat, F., 1987. Stress fields in the European platform at the time of Africa-Eurasia collision. Tectonics 6, 99132.Google Scholar
Bessin, P., Guillocheau, F., Robin, C., Schroëtter, J.-M., Bauer, H., 2015. The Armorican Massif (western France): a two times exhumed relief shaped by planation surfaces in response to Iberia-Eurasia relative movements. Geomorphology 233, 7591.Google Scholar
Bolton, A., Maltman, A., 1998. Fluid-flow pathways in actively deforming sediments: the role of pore fluid pressure and volume change. Marine and Petroleum Geology 15, 281297.Google Scholar
Bonnet, S., Guillocheau, F., Brun, J.-P., Van Den Driessche, J., 2000. Large-scale relief development related to Quaternary tectonic uplift of a Proterozoic-Paleozoic basement: the Armorican Massif, NW France. Journal of Geophysical Research 105, 1927319288.Google Scholar
Cagnard, , 2008. Carte géologique harmonisée du département du Finistère. BRGM/RP-56273-FR, BRGM ed., 435 pp.Google Scholar
Calais, E., Camelbeeck, T., Stein, S., Liu, M., Craig, T.J., 2016. A new paradigm for large earthquakes in stable continental plate interiors. Geophysical Research Letters 43, 1062110637.Google Scholar
Caputo, R., 2005. Ground effects of large morphogenic earthquakes. Journal of Geodynamics 40, 113118.Google Scholar
Cara, M., Cansi, Y., Schlupp, A., Arroucau, P., Béthoux, N., Beucler, E., Bruno, S., et al., 2015. SI-Hex: a new catalogue of instrumental seismicity for metropolitan France. Bulletin de la Société Géolologique de France 186, 319.Google Scholar
Carbon, D., Combes, P., Cushing, M., Granier, T., Grellet, B., 1995. Rupture de surface post-Pléistocène moyen dans le Bassin aquitain. Comptes Rendus de l’Academie des Sciences, Serie IIa 320, 311317.Google Scholar
Chardon, D., Hermite, D., Ngyen, F., Bellier, O., 2005. First paleoseismological constraints on the strongest earthquake in France (Provence) in the twentieth century. Geology 33, 904911.Google Scholar
Chauris, L., Plusquellec, Y., Hallegouët, B., Darboux, J.-R., Melou, M., Chauvel, J.-J., Le Corre, C., Babin, C., Morzadec, P., Thonon, P., 1980. Notice de la carte géologique 1/50 000, feuille Brest (274). Bureau de recherches géologiques et minières, Orléans, France.Google Scholar
Cordier, S., Harmand, D., Frechen, M., Beiner, M., 2006. Fluvial system response to Middle and Upper Pleistocene climate change in the Meurthe and Moselle valleys (Eastern Paris Basin and Rhenish Massif). Quaternary Science Reviews 25, 14601474.Google Scholar
Darboux, J.R., Marcoux, E., Hallégouët, B., Lebret, P., Thomas, E., Margerel, J.-P., Blanchet, S., Carn, A., 2010. Notice explicative, Carte géol. France (1/50 000), feuille Landerneau (239). Bureau de recherches géologiques et minières, Orléans, France.Google Scholar
Derakhshandi, M., Rathje, E.E., Hazirbaba, E.E., Mirhosseini, S.M., 2007. The effect of plastic fines on the pore pressure generation characteristics of saturated sands. Soil Dynamics and Earthquake Engineering 28, 376386.Google Scholar
Diot, M.F., 1999. Le Pléistocène de la façade atlantique du Nord-Médoc (France): synthèse sur la palynologie des «Argiles du Gurp» s.l. et comparaison avec les données de l’Aquitaine. Quaternaire 10, 213225.Google Scholar
Dubreuilh, J., 1976. Contribution à l’étude sédimentologique du système fluviatile Dordogne-Garonne dans la région bordelaise. Les ressources en matériaux alluvionnaires du département de la Gironde. PhD dissertation, Université de Bordeaux, Bordeaux, France.Google Scholar
Dubreuilh, J., Marionnaud, J.M., 1973. Notice explicative, carte géologique de la France (1/50000), feuille St-Vivien de Médoc - Soulac-sur-Mer (730). Bureau de recherches géologiques et minières, Orléans, France.Google Scholar
Eissmann, L., 1987. Lagerungsstörungen im Lockergebirge: Endogene und exogene Tektonik im Lockergebirge des nördlichen Mitteleuropas. Geophysik und Geologie 3, 778.Google Scholar
Encrenaz, T., Bibring, J.-P., Blanc, M., 2003. The Solar System. Springer, Berlin.Google Scholar
Field, M.H., Velichkevich, F.Y., Andrieu-Ponel, V., Woltz, P., 2000. Significance of two new Pleistocene plant records from western Europe. Quaternary Research 54, 253263.Google Scholar
Gély, J.P., Sztràkos, K., 2000. L‘évolution paléogéographique et géodynamique du Bassin aquitain au Paléogène: enregistrement et datation de la tectonique pyrénéenne. Géologie de la France 2, 3157.Google Scholar
Giardini, D., Woessner, J., Danciu, L., Cotton, F., Crowley, H., Grünthal, G., Pinho, R., et al., 2013. Seismic Hazard Harmonization in Europe (SHARE): Online Data Resource (accessed). http://dx.doi.org/10.12686/SED-00000001-SHARE. Last accessed May 2017.Google Scholar
Golke, M., Coblentz, D., 1996. Origin of the European regional stress field. Tectonophysics 266, 1124.Google Scholar
Grégoire, G., Le Roy, P., Ehrhold, A., Jouet, G., Garlan, T., 2017. Control factors of Holocene sedimentary infilling in a semi-closed tidal estuarine-like system: the Bay of Brest (France). Marine Geology 385, 84100.Google Scholar
Hallégouët, B., 1989. La presqu’île de Crozon: évolution géomorphologique. Historien-Géographe 318, 141148.Google Scholar
Hallégouët, B., 1994. Formation de la rade de Brest. In: Corlaix, J.-P. (Ed.), Atlas permanent de la mer et du littoral. Editmar No. 1. CNRS, Nantes, France, pp. 22.Google Scholar
Hallégouët, B., Ollivier-Pierre, M.F., Esteoule-Choux, J., 1976. Découverte d’un dépôt oligocène inférieur dans la haute vallée de l’Aber Ildut, au nord-ouest de Brest (Finistère). Comptes Rendus de l’Academie des Sciences 283D, 17111714.Google Scholar
Haugmard, M., 2017. Détermination non-linéaire des paramètres hypocentraux et structuraux: application à la sismicité intracontinentale du Massif armoricain. PhD dissertation, Earth Sciences, University of Nantes, Nantes, France.Google Scholar
Hide, R., Dickey, J.O., 1991. Earth’s variable rotation. Science 253, 629637.Google Scholar
Iaffaldano, G., Bunge, H.-P., 2015. Rapid plate motion variations through geological time: observations serving geodynamic interpretation. Annual Review of Earth and Planetary Sciences 43, 571592.Google Scholar
Irfan, T.Y., 1996. Mineralogy, fabric properties and classification of weathered granites in Hong Kong. Quarterly Journal of Engineering Geology 29, 536.Google Scholar
Ito, K., Tamura, T., Tsukamoto, S., 2017. Post-IR IRSL dating of K-feldspar from last interglacial marine terrace deposits on the Kamikita coastal plain, northeastern Japan. Geochronometria 44, 352365.Google Scholar
Koutsodendris, A., Müller, U.C., Pross, J., Brauer, A., Kotthoff, U., Lotter, A.F., 2010. Vegetation dynamics and climate variability during the Holsteinian interglacial based on a pollen record from Dethlingen (northern Germany). Quaternary Science Reviews 29, 32983307.Google Scholar
Lambeck, K., Johnston, P., 2000. Reply to comment by W. Fjeldskaar ‘What about asthenosphere viscosity? Sea-level change, glacial rebound and mantle viscosity for northern Europe. Geophysical Journal International 142, 279281.Google Scholar
Landgraf, A., Kuebler, S., Hintersberger, E., Stein, S. (Eds.), 2016. Seismicity, Fault Rupture and Earthquake Hazards in Slowly Deforming Regions. Special Publication, No. 432. Geological Society, London.Google Scholar
Larson, R.L., Olson, P., 1991. Mantle plumes control magnetic reversal frequency. Earth and Planetary Science Letters 107, 437447.Google Scholar
Laut, P., 2003. Solar activity and terrestrial climate: an analysis of some purported correlations. Journal of Atmospheric and Solar-Terrestrial Physics 65, 801812.Google Scholar
Le Gall, B., Authemayou, C., Ehrhold, A., Paquette, J.-L., Bussien, D., Chazot, G., Aouizerat, A., Pastol, Y., 2014. LiDAR offshore structural mapping and U/Pb zircon/monazite dating of Variscan strain in the Leon metamorphic domain, NW Brittany. Tectonophysics 630, 236250.Google Scholar
Le Roy, P., Gracia-Garay, C., Guennoc, P., Bourillet, J.-F., Reynaud, J.-Y., Thinon, I., Kervevan, P., Paquet, F., Menier, D., Bulois, C., 2011. Cenozoic tectonics of the Western Approaches Channel basins and its control of local drainage systems. Bulletin de la Société Géologique de France 182, 451463.Google Scholar
Levi, T., Weinberger, R., Aïfa, T., Eyal, Y., Marco, S., 2006. Earthquake-induced clastic dikes detected by anisotropy of magnetic susceptibility. Geology 34, 6972.Google Scholar
Lolli, B., Gasperini, P., Vannucci, G., 2014. Empirical conversion between teleseismic magnitudes (mb and Ms) and moment magnitude (Mw) at the Global, Euro-Mediterranean and Italian scale. Geophysical Journal International 199, 805828.Google Scholar
Manchuel, K., Traversa, P., Baumont, D., Cara, M., Nayman, E., Durouchoux, C., 2018. The French seismic CATalogue (FCAT-17). Bulletin of Earthquake Engineering 16, 22272251.Google Scholar
Mansy, J.L., Manby, G.M., Averbuch, O., Everaerts, M., Bergerat, F., van Vliet-Lanoë, B., Lamarche, J., 2003. Role of basement reactivation in the formation and inversion of the Weald-Boulonnais basin. Tectonophysics 373, 161179.Google Scholar
Marotta, A.M., Mitrovica, J.X., Sabadini, R., Milne, G., 2004. Combined effects of tectonics and glacial isostatic adjustment on intraplate deformation in central and northern Europe: applications to geodetic baseline analyses. Journal of Geophysical Research 109, B01413.Google Scholar
Mazabraud, Y., Bethoux, N., Delouis, B., 2013. Is earthquake activity along the French Atlantic margin favoured by local rheological contrasts? Comptes Rendus Geoscience 345, 373382.Google Scholar
Mazabraud, Y., Béthoux, N., Guilbert, J., Bellier, O., 2005. Short scale stress determination in central and western France, an intraplate slow deforming region. Geophysical Journal International 160, 161178.Google Scholar
McCalpin, J.P. (Ed.), 1996. Paleoseismology. Academic Press, San Diego, CA.Google Scholar
Melou, M., 1968. Contribution à l’étude sédimentologique du Quaternaire Sud-Finistérien. L’anse de Trez Rouz et la ria de l’Odet. PhD dissertation, Université de Paris, centre d’Orsay, Paris.Google Scholar
Ménabréaz, L., Thouveny, N., Bourlès, D.L., Vidal, L., 2014. The geomagnetic dipole moment variation between 250 and 800 ka BP reconstructed from the authigenic 10Be/9Be signature in West Equatorial Pacific sediments. Earth and Planetary Science Letters 385, 190205.Google Scholar
Montgomery, D.R. and Manga, M. 2003. Streamflow and Water Well Responses to Earthquakes. Science 300, 20472049.Google Scholar
Morzadec-Kerfourn, M.-T, 1974. Variations de la ligne de rivage armoricaine au Quaternaire. Analyses polliniques de dépôts organiques littoraux. Mémoiresde la Société géologique et minéralogique de. Bretagne, 17. Institut de géologie, Université de Rennes, Rennes, France.Google Scholar
Morzadec-Kerfourn, M.-T., 1999. Littoraux pléistocènes de l’ouest du Massif armoricain: de la rade de Brest à la Baie d’Audierne. Quaternaire 10, 171179.Google Scholar
Muir-Wood, R., King, G.C.P., 1993. Hydrologic signatures of earthquake strain. Journal of Geophysical Research 98, 2203522068.Google Scholar
Mulargia, F., Bizzarri, A., 2014. Anthropogenic triggering of large earthquakes. Scientific Reports 4, 6100.Google Scholar
Nocquet, J.M., 2012. Present-day kinematics of the Mediterranean: a comprehensive overview of GPS results. Tectonophysics 579, 220242.Google Scholar
Obermeier, S.F., 1996. Using liquefaction induced features for palaeoseismic analysis. In: McCalpin, J. (Ed.), Paleoseismology. Academic Press, San Diego, CA, pp. 331396.Google Scholar
Obermeier, S.F., Martin, J.R., Frankel, A.D., Youd, T.L., Munson, P.J., Munson, C.A., Pond, E.C., 1993. Liquefaction Evidence for One or More Strong Holocene Earthquakes in Wabash Valley of Southern Indiana and Illinois, with a Preliminary Estimate of Magnitude. U.S. Geological Survey Professional Paper 1536. U.S. Government Printing Office, Washington, DC.Google Scholar
O’Brien, C.E., Jones, R.L., 2003. Early and Middle Pleistocene vegetation history of the Medoc region, southwest France. Journal of Quaternary Science 18, 557579.Google Scholar
Olaiz, A.J., Muñoz-Martín, A., De Vicente, G., Vegas, R., Cloetingh, S., 2009. European continuous active tectonic strain–stress map. Tectonophysics 474, 3340.Google Scholar
Olson, S., Green, R., Obermeier, S.F., 2005. Revised magnitude-bound relation for the Wabash Valley seismic zone of the central United States. Seismological Research Letters 76, 756771.Google Scholar
Perrot, J., Arroucau, P., Guilbert, J., Déverchère, J., Mazabraud, Y., Rolet, J., Mocquet, A., Mousseau, M., Matias, L., 2005. Analysis of the Mw 4.3 Lorient earthquake sequence: a multidisciplinary approach to the geodynamics of the Armorican Massif, westernmost France. Geophysical Journal International 162, 935950.Google Scholar
Plusquellec, Y., Rolet, J., Darboux, J.-R., Bosold, A., Chantraine, J., Chauris, L., Chauvel, J.-J., et al., 1999. Notice explicative de la feuille Châteaulin, 1/50 000. Bureau de recherches géologiques et minières, Orléans, France.Google Scholar
Powley, D.E., 1999. Shale Domes. Search and Discovery Article #60001. Revised from Amoco Production Company report. http://www.searchanddiscovery.com/documents/Shale/shale.htm.Google Scholar
Railsback, L.B., Gibbard, P.L., Head, M.J., Voarintsoa, N.R.G., Toucanne, S., 2015. An optimized scheme of lettered marine isotope substages for the last 1.0 million years, and the climatostratigraphic nature of isotope stages and substages. Quaternary Science Reviews 111, 94106.Google Scholar
Rohling, E.J., Grant, K., Bolshaw, M., Roberts, A.P., Siddall, M., Hemleben, C., Kucera, M., 2009. Antarctic temperature and global sea level closely coupled over the past five glacial cycles. Nature Geosciences 2, 500504.Google Scholar
Rolet, J., 1997. The concealed basement of Aquitaine. Mémoires de la Société géolologique de France 171, 115141.Google Scholar
Schneider, J.L., Van Vliet-Lanoë, B., Sitzia, L., 2012. Déformations co-sismiques à Cestas-Pot-aux-Pins et Larrusey. In: Bertran, P., Lenoble, A. (Eds.), Le Quaternaire continental d’Aquitaine: Un point sur les travaux récents. Field Guide. Association Française pour l’Etude du Quaternaire, Talence, France, pp. 140152.Google Scholar
Scourse, J.D., Haapaniemi, A.I., Colmenero-Hidalgo, E., 2009. Growth, dynamics and deglaciation of the last British–Irish ice sheet: the deep-sea ice-rafted detritus record. Quaternary Science Reviews 28, 30663084.Google Scholar
Segall, P., Rice, J.R., 1995. Dilatancy, compaction, and slip instability of a fluid-infiltrated fault. Journal of Geophysical Research 100, 2215522171.Google Scholar
Semblat, J.F., Kham, M., Parara, E., Bard, P.Y., Pitilakis, K., Makra, K., Raptakis, D., 2005. Site effects: basin geometry vs soil layering. Soil Dynamics and Earthquake Engineering 25, 529538.Google Scholar
Sibson, R.H., 1994. Crustal stress, faulting and fluid flow. In: Parnell, J. (Ed.), Geofluids: Origin, Migration and Evolution of Fluids in Sedimentary Basins. Special Publication, No. 78. Geological Society, London, pp. 6984.Google Scholar
Sitzia, L., Bertran, P., Bahain, J., Bateman, M.D., Hernandez, M., Garon, H., de Lafontaine, G., et al., 2015. The Quaternary coversands of southwest France. Quaternary Science Reviews 124, 84105.Google Scholar
Tastet, J.-P., 1999. Le Pléistocène de la façade atlantique du Nord-Médoc (France): état des connaissances sur la lithologie et la chronostratigraphie des «Argiles du Gurp» s.l. Quaternaire 10, 199212.Google Scholar
Thiel, C., Buylaert, J.P., Murray, A.S., Terhorst, B., Hofer, I., Tsukamoto, S., Frechen, M., 2011. Luminescence dating of the Stratzing loess profile (Austria) – testing the potential of an elevated temperature post-IR IRSL protocol. Quaternary International 234, 2331.Google Scholar
Van Vliet-Lanoë, B., Bonnet, S., Hallegouët, B., Laurent, M., 1997. Neotectonic and seismic activity in the Armorican and Cornubian Massifs: regional stress field with glacio-isostatic influence? Journal of Geodynamics 24, 219239.Google Scholar
Van Vliet-Lanoë, B., Brulhet, J., Combes, P., Duvail, C., Ego, F., Baize, S., Cojan, I., 2017. Quaternary thermokarst and thermal erosion features in northern France: origin and palaeoenvironments. Boreas 46, 442461.Google Scholar
Van Vliet-Lanoë, B., Hibsch, C., Csontos, L., Jegouzo, S., Hallégouët, B., Laurent, M., Maygari, A., Mercier, D., Voinchet, P., 2009. Seismically induced shale diapirism: the Mine d’Or section, Vilaine estuary, southern Brittany. International Journal of Earth Sciences 98, 969984.Google Scholar
Van Vliet-Lanoë, B., Laurent, M., Bahain, J.-J., Balescu, S., Falguères, C., Field, M., Hallégouët, B., Keen, D.H., 2000. Middle Pleistocene raised beach anomalies in the English Channel: regional and global stratigraphic implications. Journal of Geodynamics 29, 1541.Google Scholar
Van Vliet-Lanoë, B., Maygari, A., Meilliez, F., 2004. Distinguishing between tectonic and periglacial deformations of quaternary continental deposits in Europe. Global Planetary Change 43, 103127.Google Scholar
Van Vliet-Lanoë, B., Vandenberghe, N., Laignel, B., Laurent, M., Lauriat-Rage, A., Louwye, S., Mansy, J.L., et al., 2002. Paleogeographic evolution of northwestern Europe during the Upper Cenozoic. Geodiversitas 24, 511541.Google Scholar
Voinchet, P., Despriée, J, Tissoux, H., Falguères, C., Bahain, J.-J., Gageonnet, R., Dépont, J., Dolo, J.-M., 2010. ESR chronology of alluvial deposits and first human settlements of the Middle Loire Basin (Region Centre, France). Quaternary Geochronology 5, 381384.Google Scholar
Wells, D.J., Coppersmith, K.J., 1994. New empirical relationships among magnitude, rupture length, rupture width, rupture area, and surface displacement. Bulletin of Seismological Society of America 84, 9741002.Google Scholar
Woith, H., Wang, R., Milkereit, C., Zschau, J., Maiwald, U., Pekdeger, A., 2003. Heterogenous response of hydrogeological systems to the Izmit and Düzce (Turkey) earthquakes of 1999. Hydrogeology Journal 11, 113121.Google Scholar
Worm, H.U., 1997. A link between geomagnetic reversals and events and glaciations. Earth and Planetary Science Letters 147, 5567.Google Scholar
Wyns, R., 1991. Evolution tectonique du bâti armoricain oriental au Cénozoïque d’après l’analyse des paléosurfaces continentales et des formations géologiques associées. Géologie de la France 3, 1142.Google Scholar
Youd, T.L., Steidl, J.H., Nigbor, R.L., 2004. Lessons learned and need for instrumented liquefaction sites. Soil Dynamics and Earthquake Engineering 24, 639646.Google Scholar
Zagwijn, W.H., 1996. The Cromerian Complex Stage of the Netherlands and correlation with other areas in Europe. In: Turner, C. (Ed.), The Early Middle Pleistocene in Europe. Balkema, Rotterdam, the Netherlands, pp. 145172.Google Scholar
Ziegler, P., Dèzes, P., 2006. Crustal evolution of western and central Europe. Geological Society, London, Memoirs 32, 4356.Google Scholar