Hostname: page-component-8448b6f56d-42gr6 Total loading time: 0 Render date: 2024-04-24T10:53:48.297Z Has data issue: false hasContentIssue false

Coupled phengite 40Ar–39Ar geochronology and thermobarometry: P-T-t evolution of Andros Island (Cyclades, Greece)

Published online by Cambridge University Press:  21 November 2014

BENJAMIN HUET*
Affiliation:
Department of Geodynamics and Sedimentology, University of Vienna, Althanstrasse 14, 1090 Vienna, Austria
LOÏC LABROUSSE
Affiliation:
UPMC Université Paris 06, UMR 7193, ISTeP, 4 place Jussieu, 75252 Paris cedex 05, France CNRS, UMR 7193, ISTeP, 4 place Jussieu, 75252 Paris cedex 05, France
PATRICK MONIÉ
Affiliation:
Géosciences Montpellier, UMR 5573, Université Montpellier 2, France
BENJAMIN MALVOISIN
Affiliation:
Faculté des Géosciences et de l’Environnement, Université de Lausanne, Switzerland
LAURENT JOLIVET
Affiliation:
Université d’Orléans, ISTO, UMR 7327, 42071, Orléans, France CNRS/INSU, ISTO, UMR 7327, 42071, Orléans, France BRGM, ISTO, UMR 7327, BP 36009, 45060 Orléans, France
*
Author for correspondence: benjamin.huet@univie.ac.at

Abstract

Andros is a key island for understanding both the timing of high-pressure–low-temperature (HP-LT) metamorphism and the dynamics of crustal-scale detachment systems exhuming high-grade units in the Cyclades (Greece). Using phengite 40Ar–39Ar geochronology coupled with thermobarometry, as well as data from literature, we constrain the pressure–temperature–time (P-T-t) paths of the Makrotantalon and Attic–Cycladic Blueschist units on Andros. Peak conditions of the HP-LT episode in the Makrotantalon unit are 550°C and 18.5 kbar, dated at 116 Ma. We correlate this episode with Early Cretaceous blueschist facies metamorphism recognized in the Pelagonian zone of continental Greece. This is a new argument favouring a Pelagonian origin for the Makrotantalon unit. In the Attic–Cycladic Blueschist unit, the P-T-t path is characterized by: (1) exhumation after peak conditions in HP-LT conditions between 55 and 35 Ma; (2) isobaric heating at 7 kbar until 30 Ma; and (3) isothermal decompression until 21 Ma. This thermal evolution and timing are similar to those of the neighbouring Tinos Island, emphasizing major thermal re-equilibration at the transition between stable and retreating subduction. Modifications of the crustal thermal state played a major role in the evolution of the North Cycladic Detachment System, below which Andros HP-LT units were exhumed.

Type
Original Articles
Copyright
Copyright © Cambridge University Press 2014 

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

Agard, P., Monié, P., Jolivet, L. & Goffé, B. 2002. Exhumation of the Schistes Lustrés complex: in situ laser probe 40Ar/39Ar constraints, and implications for the Western Alps. Journal of Metamorphic Geology 20, 599618.Google Scholar
Agard, P., Yamato, P., Jolivet, L. & Burov, E. 2009. Exhumation of oceanic blueschists and eclogites in subduction zones: Timing and mechanisms. Earth-Science Reviews 92, 5379.CrossRefGoogle Scholar
Altherr, R., Kreuzer, H., Wendt, I., Lenz, H., Wagner, G. A., Keller, J., Harre, W. & Höhndorf, A. 1982. A late Oligocene/early Miocene high temperature belt in the Attic-Cycladic Crystalline Complex (SE Pelagonian, Greece). Geologisches Jahrbuch 23, 97164.Google Scholar
Altherr, R., Schliestedt, M., Okrusch, M., Seidel, E., Kreuzer, H., Harre, W., Lenz, H., Wendt, I. & Wagner, G. A. 1979. Geochronology of high-pressure rocks on Sifnos (Cyclades, Greece). Contributions to Mineralogy and Petrology 70, 245–55.Google Scholar
Arnaud, N., Tapponnier, P., Roger, F., Brunel, M., Scharer, U., Wen, C. & Zhiqin, X. 2003. Evidence for Mesozoic shear along the western Kunlun and Altyn-Tagh fault, northern Tibet (China). Journal of Geophysical Research: Solid Earth 108 (B1), doi:10.1029/2001JB000904.Google Scholar
Augier, R., Agard, P., Monié, P., Jolivet, L., Robin, C. & Booth-Rea, G. 2005. Exhumation, doming and slab retreat in the Betic Cordillera (SE Spain): in-situ 40Ar/39Ar ages and P-T-d-t paths for the Nevado-Filabride complex. Journal of Metamorphic Geology 23, 357–81.Google Scholar
Avigad, D. & Garfunkel, Z. 1989. Low-angle faults above and below a blueschist belt: Tinos Island, Cyclades, Greece. Terra Nova 1, 182–7.Google Scholar
Avigad, A., Garfunkel, Z., Jolivet, L. & Azañón, J. M. 1997. Back-arc extension and denudation of Mediterranean eclogites. Tectonics 16, 924–41.Google Scholar
Baldwin, S. L., Lister, G. S., Hill, E., Foster, D. A. & McDougall, I. 1993. Thermochronologic constraints on the tectonic evolution of active metamorphic core complexes, D’Entrecasteaux Islands, Papua New Guinea. Tectonics 12 (3), 611–28.Google Scholar
Baldwin, S. L., Monteleone, B. D., Webb, L. E., Fitzgerald, P. G., Grove, M. & Hill, E. J. 2004. Pliocene eclogite exhumation at plate tectonic rates in eastern Papua New Guinea. Nature 431 (7006), 263–7.Google Scholar
Baldwin, S. L., Webb, L. E. & Monteleone, B. 2008. Late Miocene coesite-eclogite exhumed in the Woodlark Rift. Geology 36 (9), 735–8.Google Scholar
Berman, R. G. 1991. Thermobarometry using multi-equilibrium calculations: a new technique, with petrological applications. Canadian Mineralogist 29, 833–55.Google Scholar
Bonneau, M. 1984. Correlation of the Hellenic nappes in the south-east Aegean and their tectonic reconstruction. In The Geological Evolution of the Eastern Mediterranean (eds Dixon, J. E. & Robertson, A. H. F.), pp. 517–27. Geological Society of London, Special Publication no. 17.Google Scholar
Brichau, S., Ring, U., Carter, A., Bolhar, R., Monie, P., Stockli, D. & Brunel, M. 2008. Timing, slip rate, displacement and cooling history of the Mykonos detachment footwall, Cyclades, Greece, and implications for the opening of the Aegean Sea basin. Journal of the Geological Society of London 165, 263–77.Google Scholar
Brichau, S., Ring, U., Carter, A., Monie, P., Bolhar, R., Stockli, D. & Brunel, M. 2007. Extensional faulting on Tinos Island, Aegean Sea, Greece; how many detachments? Tectonics 26, TC4009.Google Scholar
Bröcker, M. 1990. Blueschist-to-greenschist transition in metabasites from Tinos island, Cyclade, Greece: compositional control or fluid infiltration. Lithos 25, 2539.CrossRefGoogle Scholar
Bröcker, M. & Enders, M. 1999. U-Pb zircon geochronology of unusual eclogite-facies rocks from Syros and Tinos (Cyclades, Greece). Geological Magazine 136, 111–8.Google Scholar
Bröcker, M. & Enders, M. 2001. Unusual bulk-rock compositions in eclogite-facies rocks from Syros and Tinos (Cyclades, Greece): implications for U-Pb zircon geochronology. Chemical Geology 175, 581603.Google Scholar
Bröcker, M. & Franz, L. 1994. The contact aureole on Tinos (Cyclades, Greece). Part I: field relationships, petrography and P-T conditions. Chemie der Erde 54, 262–80.Google Scholar
Bröcker, M. & Franz, L. 1998. Rb-Sr isotope studies on Tinos Island (Cyclades, Greece): additional time constraints for metamorphism, extent of infiltration-controlled overprinting and deformational activity. Geological Magazine 135, 369–82.Google Scholar
Bröcker, M. & Franz, L. 2006. Dating metamorphism and tectonic juxtaposition on Andros Island (Cyclades, Greece): results of a Rb-Sr study. Geological Magazine 143, 609–20.Google Scholar
Bröcker, M., Kreuzer, H., Matthews, A. & Okrusch, M. 1993. 40Ar/39Ar and oxygen isotope studies of polymetamorphism from Tinos island, Cycladic blueschist belt, Greece. Journal of Metamorphic Geology 11, 223–40.Google Scholar
Bröcker, M. & Pidgeon, R. T. 2007. Protolith ages of metaigneous and meta-tuffaceous rocks from the Cycladic blueschist unit, Greece: results of a reconnaissance U-Pb zircon study. Journal of Geology 115, 8398.Google Scholar
Bulle, F., Bröcker, M., Gärtner, C. & Keasling, A. 2010. Geochemistry and geochronology of HP mélanges from Tinos and Andros, cycladic blueschist belt, Greece. Lithos 117, 6181.Google Scholar
Burg, J.-P. 2012. Rhodope: From Mesozoic convergence to Cenozoic extension. Review of petro-structural data in the geochronological frame. Journal of the Virtual Explorer 42, 144.Google Scholar
De Capitani, C. & Petrakakis, K. 2010. The computation of equilibrium assemblage diagrams with Theriak/Domino software. American Mineralogist 95, 1006–16.Google Scholar
Di Vincenzo, G., Ghiribelli, B., Giorgetti, G. & Palmeri, R. 2001. Evidence of a close link between petrology and isotop records: constraints from SEM, EMP, TEM and in situ 40Ar-39Ar laser analyses on multiple generations of white micas (Lanterman Range, Antartica). Earth and Planetary Science Letters 192, 389405.Google Scholar
Ernst, W. G. 1988. Tectonic history of subduction zones inferred from retrograde blueschist P-T paths. Geology 16, 1081–4.2.3.CO;2>CrossRefGoogle Scholar
Ernst, W. G. 2001. Subduction, ultrahigh-pressure metamorphism, and regurgitation of buoyant crustal slices: implications for arcs and continental growth. Physics of the Earth and Planetary Interiors 127, 253–75.Google Scholar
Evans, B. W. 1990. Phase relations in epidote-blueschists. Lithos 25, 323.Google Scholar
Evans, B. W. & Brown, E. H. (eds) 1986. Blueschists and Eclogites. Geological Society of America, Memoir no. 164.Google Scholar
Faupl, P., Pavlopoulos, A. & Migiros, G. 1999. The Palaeogene history of the Pelagonian Zone sl (Hellenides, Greece). Heavy mineral study from terrigenous flysch sediments. Geologica Carpathica 50, 449–58.Google Scholar
Faupl, P., Petrakakis, K., Migiros, G. & Pavlopoulos, A. 2002. Detrital blue amphiboles from the western Othrys Mountain and their relationship to the blueschist terrains of the Hellenides (Greece). International Journal of Earth Sciences 91, 433–44.Google Scholar
Gautier, P., Brun, J. P. & Jolivet, L. 1993. Structure and kinematics of Upper Cenozoic extensional detachment on Naxos and Paros (Cyclades Islands, Greece). Tectonics 12, 1180–94.Google Scholar
Grasemann, B., Schneider, D. A., Stöckli, D. & Iglseder, C. 2012. Miocene bivergent crustal extension in the Cyclades (Greece). Lithosphere 4, 2339.Google Scholar
Harrison, T. M., Célérier, J., Aikman, A. B., Hermann, J. & Heizler, M. T. 2009. Diffusion of 40Ar in muscovite. Geochimica Cosmochimica Acta 73, 1039–51.Google Scholar
Holland, T. J. B. & Powell, R. 1998. An internally consistent thermodynamic data set for phases of petrological interest. Journal of Metamorphic Geology 16, 309–43.Google Scholar
Huet, B., Labrousse, L. & Jolivet, L. 2009. Thrust or detachment? Exhumation processes in the Aegean: insight from a field study on Ios (Cyclades, Greece). Tectonics 28, doi: 10.1029/2008TC002397.Google Scholar
Huet, B., Le Pourhiet, L., Labrousse, L., Burov, E. & Jolivet, L. 2011 a. Formation of metamorphic core complex in inherited wedges: a thermomechanical modeling study. Earth and Planetary Science Letters 309, 249–57.Google Scholar
Huet, B., Le Pourhiet, L., Labrousse, L., Burov, E. & Jolivet, L. 2011 b. Post-orogenic extension and metamorphic core complexes in a heterogeneous crust; the role of crustal layering inherited from collision. Application to the Cyclades (Aegean domain). Geophysical Journal International 184 (2), 611–25, doi: 10.1111/j.1365-246X.2010.04849.x.Google Scholar
Huyskens, M. H. & Bröcker, M. 2014. The status of the Makrotantalon Unit (Andros, Greece) within the structural framework of the Attic-Cycladic Crystalline Belt. Geological Magazine 151 (3), 430–46, doi: 10.1017/S0016756813000307.CrossRefGoogle Scholar
Jacobshagen, V. 1986. Geologie von Griechenland. Berlin: Gebruder Borntraeger.Google Scholar
Jacobshagen, V., Dürr, S., Kockel, F., Kopp, K. O., Kowalczyk, G., Berckhemer, H. & Büttner, D. 1978. Structure and geodynamic evolution of the Aegean region. In Alps, Apennines, Hellenides (eds Cloos, H., Roeder, D. & Schmidt, K.), pp. 537–64. Stuttgart: IUGG.Google Scholar
Jolivet, L. & Brun, J. P. 2010. Cenozoic geodynamic evolution of the Aegean. International Journal of Earth Sciences 99, 109–38, doi: 10.1007/s00531-008-0366-4.Google Scholar
Jolivet, L. & Faccenna, C. 2000. Mediterranean extension and the Africa-Eurasia collision. Tectonics 19, 1095–106.Google Scholar
Jolivet, L., Faccenna, C., Huet, B., Labrousse, L., Le Pourhiet, L., Lacombe, O., Lecomte, E., Burov, E., Denèle, Y., Brun, J.-P., Philippon, M., Paul, A., Salaün, G., Karabulut, H., Piromallo, C., Monié, P., Gueydan, F., Okay, A., Oberhänsli, R., Pourteau, A., Augier, R., Gadenne, L. & Driussi, O. 2012. Aegean tectonics: Strain localisation, slab tearing and trench retreat. Tectonophysics 597–598, 133.Google Scholar
Jolivet, L., Lecomte, E., Huet, B., Denele, Y., Lacombe, O., Labrousse, L., Le Pourhiet, L. & Mehl, C. 2010. The North Cycladic detachment system. Earth and Planetary Science Letters 289, 87104.Google Scholar
Jolivet, L. & Patriat, M. 1999. Ductile extension and the formation of the Aegean Sea. In The Mediterranean Basins: Tertiary Extension within the Alpine Orogen (eds Durand, B., Jolivet, L., Horvàth, F. & Séranne, M.), pp. 427–56. Geological Society of London, Special Publication no. 156.Google Scholar
Jolivet, L., Rimmelé, G., Oberhänsli, R., Goffé, B. & Candan, O. 2004. Correlation of syn-orogenic tectonic and metamorphic events in the Cyclades, the Lycian Nappes and the Menderes massif, geodynamic implications. Bulletin de la Société Géologique de France 175 (3), 217–38.Google Scholar
Katzir, Y., Avigad, D., Matthews, A., Garfunkel, Z. & Evans, B. 2000. Origin, HP/LT metamorphism and cooling of ophiolitic mélanges in southern Evia (NW Cyclades), Greece. Journal of Metamorphic Geology 18, 699718.Google Scholar
Lacassin, R., Arnaud, N., Leloup, P. H., Armijo, R. & Meyer, B. 2007. Syn- and post-orogenic exhumation of metamorphic rocks in North Aegean. eEarth 2, 5163.Google Scholar
Lagos, M., Scherer, E. E., Tomaschek, F., Münker, C., Keiter, M., Berndt, J. & Ballhaus, C. 2007. High precision Lu/Hf geochronology of Eocene eclogite-facies rocks from Syros, Cyclades, Greece. Chemical Geology 243, 1635.Google Scholar
Leake, B. E., Woolley, A. R., Birch, W. D., Burke, E. A. J., Ferraris, G., Grice, J. D., Hawthorne, F. C., Kisch, H. J., Krivovichev, V. G., Schumacher, J. C., Stephenson, N. C. N. & Whittaker, E. J. W. 2007. Nomenclature of amphiboles: Additions and revisions to the International Mineralogical Association's amphibole nomenclature. American Mineralogist 89, 883–7.Google Scholar
Lecomte, E., Jolivet, L., Lacombe, O., Denèle, Y., Labrousse, L. & Le Pourhiet, L. 2010. Geometry and kinematics of Mykonos detachment, Cyclades, Greece: evidence for slip at shallow dip. Tectonics 29 (5), doi: 10.1029/2009TC002564.Google Scholar
Le Pichon, X. & Angelier, J. 1981. The Aegean Sea. Philosophical Transactions of the Royal Society of London 300, 357–72.Google Scholar
Lips, A. L. W., White, S. H. & Wijbrans, J. R. 1998. 40Ar/39Ar laserprobe direct dating of discrete deformational events: a continuous record of early Alpine tectonics in the Pelagonian Zone, NW Aegean area, Greece. Tectonophysics 298, 133–53.Google Scholar
Lips, A. L. W., Wijbrans, J. R. & White, S. H. 1999. New insights from 40Ar/39Ar laserprobe dating of white mica fabrics from the Pelion Massif, Pelagonian Zone, Internal Hellenides, Greece: implications for the timing of metamorphic episodes and tectonic events in the Aegean region. In The Mediterranean Basins: Tertiary Extension within the Alpine Orogen (eds Durand, B., Jolivet, L., Horvàth, F. & Séranne, M.), pp. 456–74. Geological Society of London, Special Publication no. 156.Google Scholar
Lister, G. S. & Baldwin, S. L. 1996. Modelling the effect of arbitrary P-T-t histories on argon diffusion in minerals using the MacArgon program for the Apple MacIntosh. Tectonophysics 253, 83110.Google Scholar
Maluski, H., Vergely, P., Bavay, D., Bavay, P. & Katsikatsos, G. 1981. 40Ar/39Ar dating of glaucophanes and phengites in southern Euboa (Greece), geodynamic implications. Bulletin de la Société Géologique de France 5, 469–76.Google Scholar
Massonne, H. J. & Schreyer, W. 1987. Phengite geobarometry based on the limiting assemblage with K-feldspar, phlogopite and quartz. Contributions to Mineralogy and Petrology 96, 212–24.Google Scholar
Maurel, O., Monié, P., Respaut, P., Leyreloup, A. F. & Maluski, H. 2003. Pre-metamorphic 40Ar/39Ar and U-Pb ages in HP metagranitoids from the Hercynian belt (France). Chemical Geology 193, 195214.Google Scholar
Mehl, C., Jolivet, L., Lacombe, O., Labrousse, L. & Rimmele, G. 2007. Structural evolution of Andros (Cyclades, Greece): a key to the behaviour of a (flat) detachment within an extending continental crust. In The geodynamics of the Aegean and Anatolia (eds Taymaz, T., Yilmaz, Y. & Dilek, Y.), pp. 4173. Geological Society of London, Special Publication no. 291.Google Scholar
Meyre, C., de Capitani, C., Zack, T. & Frey, M. 1999. Petrology of high-pressure metapelites from the Adula Nappe (Central Alps, Switzerland). Journal of Petrology 40, 199213.Google Scholar
Monié, P. & Agard, P. 2009. Coeval blueschist exhumation along thousands of kilometers: implications for subduction channel processes. Geochemistry Geophysics Geosystems 10 (7), doi: 10.1029/2009GC002428.CrossRefGoogle Scholar
Mulch, A. & Cosca, M. A. 2004. Recrystallization or cooling ages: in situ UV-laser 40Ar/39Ar geochronology of muscovite in mylonitic rocks. Journal of the Geological Society 161, 573–82.Google Scholar
Müller, W. 2003. Strengthening the link between geochronology, textures and petrology. Earth and Planetary Science Letters 206, 237–51.Google Scholar
Papanikolaou, D. J. 1978. Contribution to the geology of the Aegean Sea: the island of Andros. Annales Géologiques des Pays Helléniques 29, 477553.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, 706–32.Google Scholar
Parra, T., Vidal, O. & Jolivet, L. 2002. Relation between deformation and retrogression in blueschist metapelites of Tinos island (Greece) evidenced by chlorite-mica local equilibria. Lithos 63, 4166.Google Scholar
Patzak, M., Okrusch, M. & Kreuzer, H. 1994. The Akrotiri Unit of the island of Tinos, Cyclades, Greece: Witness to a lost terrane of Late Cretaceous age. Neues Jahrbuch für Geologie und Paläontologie Abhandlungen 194, 211–52.Google Scholar
Philippon, M., Brun, J.-P. & Gueydan, F. 2012. Deciphering subduction from exhumation in the segmented Cycladic Blueschist Unit (Central Aegean, Greece). Tectonophysics 524–525, 116–34.Google Scholar
Putlitz, B., Cosca, M. A. & Schumacher, J. C. 2005. Prograde mica 40Ar/39Ar growth ages recorded in high pressure rocks (Syros, Cyclades, Greece). Chemical Geology 214, 7998.Google Scholar
Reinecke, T. 1986. Phase relationships of sursassite and other Mn-silicates in highly oxidized low-grade, high-pressure metamorphic rocks from Evvia and Andros Islands, Greece. Contributions to Mineralogy and Petrology 84, 110–26.CrossRefGoogle Scholar
Ring, U., Glodny, J., Will, T. & Thomson, S. 2007 a. An Oligocene extrusion wedge of blueschist-facies nappes on Evia, Aegean Sea, Greece; implications for the early exhumation of high-pressure rocks. Journal of the Geological Society of London 164, 637–52.Google Scholar
Ring, U., Glodny, J., Will, T. & Thomson, S. 2010. The Hellenic subduction system: high-pressure metamorphism, exhumation, normal faulting, and large-scale extension. Annual Review of Earth and Planetary Sciences 38, 4576.Google Scholar
Ring, U., Will, T., Glodny, J., Kumerics, C., Gessner, K., Thomson, S., Gungor, T., Moie, P., Okrusch, M. & Drueppel, K. 2007 b. Early exhumation of high pressure rocks in extrusion wedges; Cycladic blueschist unit in the eastern Aegean, Greece, and Turkey. Tectonics 26 (2), doi: 10.1029/2005TC001872.Google Scholar
Schermer, E. R. 1990. Mechanism of blueschist creation and preservation in a A-type subduction zone, Mount Olympos region, Greece. Geology 18, 1130–3.Google Scholar
Schermer, E. R., Lux, D. R. & Burchfiel, B. C. 1990. Temperature-time history of subducted continental crust, Mount Olympos region, Greece. Tectonics 9, 1165–95.Google Scholar
Schliestedt, M. 1986. Eclogite-blueschist relationships as evidenced by mineral equilibria in the high-pressure metabasic rocks of Sifnos (Cycladic Islands). Journal of Petrology 27, 1437–59.CrossRefGoogle Scholar
Schneider, J., Bosch, D., Monié, P. & Bruguier, O. 2007. Micro-scale element migration during eclogitisation in the Bergen arcs (Norway): a case study on the role of fluids and deformation. Lithos 96 (3), 325–52.Google Scholar
Thompson, A. B. & England, P. C. 1984. Pressure-temperature-time paths of regional metamorphism II. Their inference and interpretation using mineral assemblages in metamorphic rocks. Journal of Petrology 25, 929–55.Google Scholar
Tomaschek, T., Kennedy, A. K., Villa, I. M., Lagos, M. & Ballhaus, C. 2003. Zircons from Syros, Cyclades, Greece: Recrystallization and mobilization of zircon during high-pressure metamorphism. Journal of Petrology 44, 19772002.Google Scholar
Trotet, F., Jolivet, L. & Vidal, O. 2001. Tectono-metamorphic evolution of Syros and Sifnos islands (Cyclades, Greece). Tectonophysics 338, 179206.Google Scholar
Trotet, F., Vidal, O. & Jolivet, L. 2001. Exhumation of Syros and Sifnos metamorphic rocks (Cyclades, Greece). New constraints on the P-T paths. European Journal of Mineralogy 13, 901–20.Google Scholar
Vidal, O., De Andrade, V., Lewin, E., Munoz, 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, 669–83.Google Scholar
Vidal, O. & Parra, T. 2000. Exhumation paths of high pressure metapelites obtained from equilibria for chlorite-phengite assemblages. Geological Journal 35, 139–61.Google Scholar
Vidal, O., Parra, T. & Trotet, F. 2001. A thermodynamic model for Fe-Mg aluminous chlorite using data from phase equilibrium experiments and natural pelitic assemblages in the 100–600°C, 1–25 kbar range. American Journal of Science 6, 557–92.CrossRefGoogle Scholar
Villa, I. M. 1998. Isotopic closure. Terra Nova 10, 42–7.Google Scholar
Villa, I. M. & Williams, M. L. 2012. Geochronology of metasomatic events. In Metasomatism and the Chemical Transformation of Rock (eds Harlov, D. E. & Austrheim, H.), pp. 171202. Berlin: Springer-Verlag.Google Scholar
Warren, C. J., Hanke, F. & Kelley, S. P. 2012. When can muscovite 40Ar/39Ar dating constrain the timing of metamorphic exhumation? Chemical Geology 291, 7986.Google Scholar
Whitney, D. L. & Evans, B. W. 2010. Abbreviations for names of rock-forming minerals. American Mineralogist 95, 185–7.Google Scholar
Wijbrans, J. R. & McDougall, I. 1986. 40Ar/39Ar dating of white micas from an alpine high-pressure metamorphic belt on Naxos (Greece); the resetting of the argon isotopic system. Contributions to Mineralogy and Petrology 93, 187–94.Google Scholar
Wijbrans, J. R. & McDougall, I. 1988. Metamorphic evolution of the Attic Cycladic Metamorphic Belt on Naxos (Cyclades, Greece) utilizing 40Ar/39Ar age spectrum measurements. Journal of Metamorphic Geology 6, 571–94.Google Scholar
Yamato, P., Agard, P., Goffé, B., De Andrade, V., Vidal, O. & Jolivet, L. 2007. New, high-precision P–T estimates for Oman blueschists: implications for obduction, nappe stacking and exhumation processes. Journal of Metamorphic Geology 25, 657–82.Google Scholar
Zeffren, S., Avigad, D., Heimann, A. & Gvirtzman, Z. 2005. Age resetting of hanging wall rocks above a Tertiary low-angle detachment, Tinos island, Aegean Sea. Tectonophysics 400, 125.Google Scholar
Ziv, A., Katzir, Y., Avigad, D. & Garfunkel, Z. 2010. Strain development and kinematic significance of the Alpine folding on Andros (western Cyclades, Greece). Tectonophysics 488, 248–55.Google Scholar