Hostname: page-component-8448b6f56d-tj2md Total loading time: 0 Render date: 2024-04-17T15:21:38.496Z Has data issue: false hasContentIssue false

A lifetime of the Variscan orogenic plateau from uplift to collapse as recorded by the Prague Basin, Bohemian Massif

Published online by Cambridge University Press:  10 November 2017

FRANTIŠEK VACEK
Affiliation:
Department of Mineralogy and Petrology, National Museum, Václavské náměstí 68, Prague, 11579, Czech Republic Institute of Geology and Paleontology, Faculty of Science, Charles University, Albertov 6, Prague, 12843, Czech Republic
JIŘÍ ŽÁK*
Affiliation:
Institute of Geology and Paleontology, Faculty of Science, Charles University, Albertov 6, Prague, 12843, Czech Republic
*
Author for correspondence: jirizak@natur.cuni.cz

Abstract

The Ordovician to Middle Devonian Prague Basin, Bohemian Massif, represents the shallowest crust of the Variscan orogen corresponding to c. 1–4 km palaeodepth. The basin was inverted and multiply deformed during the Late Devonian to early Carboniferous Variscan orogeny, and its structural inventory provides an intriguing record of complex geodynamic processes that led to growth and collapse of a Tibetan-type orogenic plateau. The northeastern part of the Prague Basin is a simple syncline cross-cut by reverse/thrust faults and represents a doubly vergent compressional fan accommodating c. 10–19 % ~NW–SE shortening, only minor syncline axis-parallel extension and significant crustal thickening. The compressional structures were locally overprinted by vertical shortening, kinematically compatible with ductile normal shear zones that exhumed deep crust in the orogen's interior at c. 346–337 Ma. On a larger scale, the deformation history of the Prague Syncline is consistent with building significant palaeoelevation during Variscan plate convergence. Based on a synthesis of finite deformation parameters observed across the upper crust in the centre of the Bohemian Massif, we argue for a differentiated within-plateau palaeotopography consisting of domains of local thickening alternating with topographic depressions over lateral extrusion zones. The plateau growth, involving such complex three-dimensional internal deformations, was terminated by its collapse driven by multiple interlinked processes including gravity, voluminous magma emplacement and thermal softening in the hinterland, and far-field plate-boundary forces resulting from the relative dextral motion of Gondwana and Laurussia.

Type
Original Articles
Copyright
Copyright © 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

Aifa, T., Pruner, P., Chadima, M. & Štorch, P. 2007. Structural evolution of the Prague synform (Czech Republic) during Silurian times: an AMS, rock magnetism, and paleomagnetic study of the Svatý Jan pod Skalou dikes. Consequences for the nappes emplacement. Geological Society of America Special Paper 423, 249–65.Google Scholar
Aleksandrowski, P., Kryza, R., Mazur, S. & Zaba, J. 1997. Kinematic data on major Variscan strike-slip faults and shear zones in the Polish Sudetes, northeast Bohemian Massif. Geological Magazine 134, 727–39.Google Scholar
Andronicos, C. L., Velasco, A. A. & Hurtado, J. M. 2007. Large-scale deformation in the India–Asia collision constrained by earthquakes and topography. Terra Nova 19, 105–19.Google Scholar
Arthaud, F. & Matte, P. 1977. Late Paleozoic strike-slip faulting 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
Badham, J. P. N. 1982. Strike-slip orogens: an explanation for the Hercynides. Journal of the Geological Society, London 139, 493504.Google Scholar
Bajolet, F., Chardon, D., Martinod, J., Gapais, D. & Kermarrec, J. J. 2015. Synconvergence flow inside and at the margin of orogenic plateaus: lithospheric-scale experimental approach. Journal of Geophysical Research 120, 6634–57.Google Scholar
Barrande, J. 1852. Système Silurien du centre de la Bohême, 1ère Partie: Recherches Paléontologiques, Vol. 1, Planches. Crustacés: Trilobites. Prague, Paris, 935 pp.Google Scholar
Beaumont, C., Hamilton, J. & Fullsack, P. 1996. Mechanical model for subduction–collision tectonics of Alpine-type compressional orogens. Geology 24, 675–8.Google Scholar
Beaumont, C., Jamieson, R. A. & Nguyen, M. 2010. Models of large, hot orogens containing a collage of reworked and accreted terranes. Canadian Journal of Earth Sciences 47, 485515.Google Scholar
Beaumont, C., Jamieson, R. A., Nguyen, M. H. & Lee, B. 2001. Himalayan tectonics explained by extrusion of a low-viscosity crustal channel coupled to focused surface denudation. Nature 414, 738–42.Google Scholar
Becq-Giraudon, J. F., Montenat, C. & Van Den Driessche, J. 1996. Hercynian high-altitude phenomena in the French Massif Central: tectonic implications. Palaeogeography, Palaeoclimatology, Palaeoecology 122, 227–41.Google Scholar
Brenchley, P. J. & Štorch, P. 1989. Environmental changes in the Hirnantian (upper Ordovician) of the Prague Basin, Czechoslovakia. Geological Journal 24, 165–81.Google Scholar
Buggisch, W. & Mann, U. 2004. Carbon isotope stratigraphy of Lochkovian to Eifelian limestones from the Devonian of central and southern Europe. International Journal of Earth Sciences 93, 521–41.Google Scholar
Cao, W., Paterson, S. R., Memeti, V., Mundil, R., Anderson, J. L. & Schmidt, K. 2015. Tracking paleodeformation fields in the Mesozoic central Sierra Nevada arc: implications for intra-arc cyclic deformation and arc tempos. Lithosphere 7, 296320.Google Scholar
Chamberlin, R. T. 1910. The Appalachian folds of central Pennsylvania. Journal of Geology 18, 228–51.Google Scholar
Chlupáč, I. 1988. Possible global events and the stratigraphy of the Palaeozoic of the Barrandian (Cambrian–Middle Devonian, Czechoslovakia). Journal of Geological Sciences, Geology 43, 83146.Google Scholar
Chlupáč, I. 1993. Geology of the Barrandian: A Field Trip Guide. Frankfurt am Main: Waldemar Kramer, 163 pp.Google Scholar
Chlupáč, I. 1996. Neptunian dykes in the Koněprusy Devonian: geological and palaeontological observations. Bulletin of the Czech Geological Survey 71, 193208.Google Scholar
Chlupáč, I. 2003. Comments on facies development and stratigraphy of the Devonian, Barrandian area, Czech Republic. Bulletin of Geosciences 78, 299312.Google Scholar
Chlupáč, I., Havlíček, V., Kříž, J., Kukal, Z. & Štorch, P. 1998. Palaeozoic of the Barrandian (Cambrian to Devonian). Czech Geological Survey, 183 pp.Google Scholar
Copley, A. & Jackson, J. 2006. Active tectonics of the Turkish–Iranian Plateau. Tectonics 25, TC6006. doi: 10.1029/2005TC001906.Google Scholar
Crick, R. E., Ellwood, B. B., Hladil, J., El Hassani, A., Hrouda, F. & Chlupáč, I. 2001. Magnetostratigraphy susceptibility of the Přídolian–Lochkovian (Silurian–Devonian) GSSP (Klonk, Czech Republic) and a coeval sequence in Anti-Atlas Morocco. Palaeogeography, Palaeoclimatology, Palaeoecology 167, 73100.Google Scholar
Cruden, A. R., Nasseri, M. H. B. & Pysklywec, R. 2006. Surface topography and internal strain variation in wide hot orogens from three-dimensional analogue and two-dimensional numerical vice models. In Analogue and Numerical Modelling of Crustal-Scale Processes (eds Buiter, S. J. H. & Schreurs, G.), pp. 79104. Geological Society of London, Special Publication no. 253.Google Scholar
Culshaw, N. G., Beaumont, C. & Jamieson, R. A. 2006. The orogenic superstructure–infrastructure concept: revisited, quantified, and revived. Geology 34, 733–6.Google Scholar
Dahlstrom, C. D. A. 1969. Balanced cross sections. Canadian Journal of Earth Sciences 6, 743–57.Google Scholar
Dallmeyer, R. D. & Urban, M. 1994. Variscan vs. Cadomian tectonothermal evolution within the Teplá–Barrandian zone, Bohemian Massif, Czech Republic: evidence from 40Ar/39Ar mineral and whole-rock slate/phyllite ages. Journal of the Czech Geological Society 39, 21–2.Google Scholar
Dallmeyer, R. D. & Urban, M. 1998. Variscan vs Cadomian tectonothermal activity in northwestern sectors of the Teplá–Barrandian zone, Czech Republic: constraints from 40Ar/39Ar ages. Geologische Rundschau 87, 94106.Google Scholar
Da Silva, A. C., Hladil, J., Chadimová, L., Slavík, L., Hilgen, F. J., Bábek, O. & Dekkers, M. J. 2016. Refining the Early Devonian time scale using Milankovitch cyclicity in Lochkovian–Pragian sediments (Prague Synform, Czech Republic). Earth and Planetary Science Letters 455, 125–39.Google Scholar
DeCelles, P. G., Robinson, D. M. & Zandt, G. 2002. Implications of shortening in the Himalayan fold–thrust belt for uplift of the Tibetan Plateau. Tectonics 21, 1062. doi: 10.1029/2001TC001322.Google Scholar
de Sitter, L. U. & Zwart, H. J. 1960. Tectonic development in supra and infra-structures of a mountain chain. In Proceedings of the 21st International Geological Congress, Copenhagen, pp. 248–56.Google Scholar
Dörr, W. & Zulauf, G. 2010. Elevator tectonics and orogenic collapse of a Tibetan-style plateau in the European Variscides: the role of the Bohemian shear zone. International Journal of Earth Sciences 99, 299325.Google Scholar
Dörr, W. & Zulauf, G. 2012. Reply to W. Franke on W. Dörr and G. Zulauf elevator tectonics and orogenic collapse of a Tibetan-style plateau in the European Variscides: the role of the Bohemian shear zone. International Journal of Earth Sciences 101, 2035–41.Google Scholar
Dörr, W., Zulauf, G., Fiala, J., Franke, W. & Vejnar, Z. 2002. Neoproterozoic to Early Cambrian history of an active plate margin in the Teplá–Barrandian unit: a correlation of U–Pb isotopic-dilution-TIMS ages (Bohemia, Czech Republic). Tectonophysics 352, 6585.Google Scholar
Drost, K., Gerdes, A., Jeffries, T., Linnemann, U. & Storey, C. 2011. Provenance of Neoproterozoic and early Paleozoic siliciclastic rocks of the Teplá–Barrandian unit (Bohemian Massif): evidence from U–Pb detrital zircon ages. Gondwana Research 19, 213–31.Google Scholar
Drost, K., Linnemann, U., McNaughton, N., Fatka, O., Kraft, P., Gehmlich, M., Tonk, C. & Marek, J. 2004. New data on the Neoproterozoic–Cambrian geotectonic setting of the Teplá–Barrandian volcano-sedimentary successions: geochemistry, U–Pb zircon ages, and provenance (Bohemian Massif, Czech Republic). International Journal of Earth Sciences 93, 742–57.Google Scholar
Fatka, O., Kraft, J., Kraft, P., Mergl, M., Mikuláš, R. & Štorch, P. 1995. Ordovician of the Prague Basin: stratigraphy and development. In Ordovician Odyssey: Short Papers for the 7th International Symposium on the Ordovician System (eds Cooper, J. D., Droser, M. L. & Finney, S. C.), pp. 241–4. Las Vegas: The Pacific Section Society for Sedimentary Geology.Google Scholar
Fatka, O. & Mergl, M. 2009. The ‘microcontinent’ Perunica: status and story 15 years after conception. In Early Palaeozoic Peri-Gondwana Terranes: New Insights from Tectonics and Biogeography (ed. Bassett, M. G.), pp. 65101. Geological Society of London, Special Publication no. 325.Google Scholar
Ferrová, L., Frýda, J. & Lukeš, P. 2012. High-resolution tentaculite biostratigraphy and facies development across the Early Devonian Daleje Event in the Barrandian (Bohemia): implications for global Emsian stratigraphy. Bulletin of Geosciences 87, 587624.Google Scholar
Filip, J. & Suchý, V. 2004. Thermal and tectonic history of the Barrandian Lower Paleozoic, Czech Republic: is there a fission-track evidence for Carboniferous–Permian overburden and pre-Westphalian alpinotype thrusting? Bulletin of Geosciences 79, 107–12.Google Scholar
Franěk, J., Schulmann, K., Lexa, O., Tomek, Č. & Edel, J. B. 2011. Model of syn-convergent extrusion of orogenic lower crust in the core of the Variscan belt: implications for exhumation of high-pressure rocks in large hot orogens. Journal of Metamorphic Geology 29, 5378.Google Scholar
Franke, W. 2006. The Variscan orogen in Central Europe: construction and collapse. In European Lithosphere Dynamics (eds Gee, D. G. & Stephenson, R. A.), pp. 333–43. Geological Society of London, Memoir no. 32.Google Scholar
Franke, W. 2012. Comment on Dörr and Zulauf: elevator tectonics and orogenic collapse of a Tibetan-style plateau in the European Variscides: the role of the Bohemian shear zone. Int J Earth Sci (Geol Rundsch) (2010) 99: 299–325. International Journal of Earth Sciences 101, 2027–34.Google Scholar
Franke, W. 2014. Topography of the Variscan orogen in Europe: failed–not collapsed. International Journal of Earth Sciences 103, 1471–99.Google Scholar
Frýda, J. & Frýdová, B. 2014. First evidence for the Homerian (late Wenlock, Silurian) positive carbon isotope excursion from peri-Gondwana: new data from the Barrandian (Perunica). Bulletin of Geosciences 89, 617–34.Google Scholar
Frýda, J., Hladil, J. & Vokurka, K. 2002. Seawater strontium isotope curve at the Silurian/Devonian boundary: a study of the global Silurian/Devonian boundary stratotype. Geobios 35, 21–8.Google Scholar
Glasmacher, U. A., Mann, U. & Wagner, G. A. 2002. Thermotectonic evolution of the Barrandian, Czech Republic, as revealed by apatite fission-track analysis. Tectonophysics 359, 381402.Google Scholar
Glodny, J., Grauert, B., Fiala, J., Vejnar, Z. & Krohe, A. 1998. Metapegmatites in the western Bohemian massif: ages of crystallisation and metamorphic overprint, as constrained by U–Pb zircon, monazite, garnet, columbite and Rb–Sr muscovite data. Geologische Rundschau 87, 124–34.Google Scholar
Graveleau, F., Malavieille, J. & Dominguez, S. 2012. Experimental modelling of orogenic wedges: a review. Tectonophysics 538540, 166.Google Scholar
Groshong, R. H., Bond, C., Gibbs, A., Ratliff, R. & Wiltschko, D. V. 2012. Preface: Structural balancing at the start of the 21st century: 100 years since Chamberlin. Journal of Structural Geology 41, 15.Google Scholar
Hajná, J., Žák, J. & Kachlík, V. 2011. Structure and stratigraphy of the Teplá–Barrandian Neoproterozoic, Bohemian Massif: a new plate-tectonic reinterpretation. Gondwana Research 19, 495508.Google Scholar
Hajná, J., Žák, J. & Dörr, W. 2017. Time scales and mechanisms of growth of active margins of Gondwana: a model based on detrital zircon ages from the Neoproterozoic to Cambrian Blovice accretionary complex, Bohemian Massif. Gondwana Research 42, 6383.Google Scholar
Hajná, J., Žák, J., Kachlík, V. & Chadima, M. 2010. Subduction-driven shortening and differential exhumation in a Cadomian accretionary wedge: the Teplá–Barrandian unit, Bohemian Massif. Precambrian Research 176, 2745.Google Scholar
Hajná, J., Žák, J., Kachlík, V. & Chadima, M. 2012. Deciphering the Variscan tectonothermal overprint and deformation partitioning in the Cadomian basement of the Teplá–Barrandian unit, Bohemian Massif. International Journal of Earth Sciences 101, 1855–73.Google Scholar
Halavínová, M., Melichar, R. & Slobodník, M. 2008. Hydrothermal veins linked with the Variscan structure of the Prague Synform (Barrandien, Czech Republic): resolving fluid–wall rock interaction. Geological Quarterly 52, 309–20.Google Scholar
Havlíček, V. 1963. Tectogenetic disruption of the Barrandian Paleozoic. Journal of Geological Sciences, Geology 1, 77102.Google Scholar
Havlíček, V. 1980. Development of Paleozoic basins in the Bohemian Massif (Cambrian–Lower Carboniferous). Journal of Geological Sciences, Geology 34, 3165.Google Scholar
Havlíček, V. 1981. Development of a linear sedimentary depression exemplified by the Prague Basin (Ordovician–Middle Devonian; Barrandian area – central Bohemia). Journal of Geological Sciences, Geology 35, 748.Google Scholar
Havlíček, V. 1982. Ordovician of Bohemia: development of the Prague Basin and its benthic communities. Journal of Geological Sciences, Geology 37, 103–36.Google Scholar
Henk, A. 1999. Did the Variscides collapse or were they torn apart?: A quantitative evaluation of the driving forces for postconvergent extension in central Europe. Tectonics 18, 774–92.Google Scholar
Hladíková, J., Hladil, J. & Kříbek, B. 1997. Carbon and oxygen isotope record across Pridoli to Givetian stage boundaries in the Barrandian basin (Czech Republic). Palaeogeography, Palaeoclimatology, Palaeoecology 132, 225–41.Google Scholar
Hladil, J., Slavík, L., Vondra, M., Koptíková, L., Čejchan, P., Schnabl, P., Adamovič, J., Vacek, F., Vích, R., Lisá, L. & Lisý, P. 2011. Pragian–Emsian successions in Uzbekistan and Bohemia: magnetic susceptibility logs and their dynamic time warping alignment. Stratigraphy 8, 217–35.Google Scholar
Hofmann, M., Linnemann, U., Gerdes, A., Ullrich, B. & Schauer, M. 2009. Timing of dextral strike-slip processes and basement exhumation in the Elbe Zone (Saxo-Thuringian Zone): the final pulse of the Variscan Orogeny in the Bohemian Massif constrained by LA-SF-ICP-MS U–Pb zircon data. In Ancient Orogens and Modern Analogues (eds Murphy, J. B., Keppie, J. D. & Hynes, A. J.), pp. 197214. Geological Society of London, Special Publication no. 327.Google Scholar
Hollister, L. S. & Crawford, M. L. 1986. Melt-enhanced deformation: a major tectonic process. Geology 14, 558–61.Google Scholar
Holub, F. V., Cocherie, A. & Rossi, P. 1997. Radiometric dating of granitic rocks from the Central Bohemian Plutonic Complex: constraints on the chronology of thermal and tectonic events along the Barrandian–Moldanubian boundary. Comptes Rendus de L'Academie des Sciences, Series IIA, Earth and Planetary Science 325, 1926.Google Scholar
Horný, R. 1965. Tectonic structure and development of the Silurian between Beroun and Tachlovice. Journal for Mineralogy and Geology 10, 147–55.Google Scholar
Jamieson, R. A. & Beaumont, C. 2013. On the origin of orogens. Geological Society of America Bulletin 125, 1671–702.Google Scholar
Janoušek, V., Braithwaite, C. J. R., Bowes, D. R. & Gerdes, A. 2004. Magma-mixing in the genesis of Hercynian calc-alkaline granitoids: an integrated petrographic and geochemical study of the Sázava intrusion, Central Bohemian Pluton, Czech Republic. Lithos 78, 6799.Google Scholar
Janoušek, V. & Gerdes, A. 2003. Timing the magmatic activity within the Central Bohemian Pluton, Czech Republic: conventional U–Pb ages for the Sázava and Tábor intrusions and their geotectonic significance. Journal of the Czech Geological Society 48, 70–1.Google Scholar
Janoušek, V., Wiegand, B. A. & Žák, J. 2010. Dating the onset of Variscan crustal exhumation in the core of the Bohemian Massif: new U–Pb single zircon ages from the high-K calc-alkaline granodiorites of the Blatná suite, Central Bohemian Plutonic Complex. Journal of the Geological Society, London 167, 347–60.Google Scholar
Johnson, M. R. W. 2002. Shortening budgets and the role of continental subduction during the India–Asia collision. Earth-Science Reviews 59, 101–23.Google Scholar
Knížek, M., Melichar, R. & Janečka, J. 2010. Stratigraphic separation diagrams as a tool for determining fault geometry in a folded and thrusted region: an example from the Barrandian region, Czech Republic. Geological Journal 45, 536–43.Google Scholar
Konopásek, J. & Schulmann, K. 2005. Contrasting Early Carboniferous field geotherms: evidence for accretion of a thickened orogenic root and subducted Saxothuringian crust (Central European Variscides). Journal of the Geological Society, London 162, 463–70.Google Scholar
Koptíková, L., Bábek, O., Hladil, J., Kalvoda, J. & Slavík, L. 2010. Stratigraphic significance and resolution of spectral reflectance logs in Lower Devonian carbonates of the Barrandian area, Czech Republic; a correlation with magnetic susceptibility and gamma-ray logs. Sedimentary Geology 225, 8398.Google Scholar
Košler, J., Aftalion, M. & Bowers, D. R. 1993. Mid–late Devonian plutonic activity in the Bohemian Massif: U–Pb zircon isotopic evidence from the Staré Sedlo and Mirotice gneiss complexes, Czech Republic. Neues Jahrbuch für Mineralogie, Monatshefte 9, 417–31.Google Scholar
Košler, J., Bowes, D. R., Farrow, C. M., Hopgood, A. M., Rieder, M. & Rogers, G. 1997. Constraints on the timing of events in the multi-episodic history of the Teplá–Barrandian complex, western Bohemia, from integration of deformational sequence and Rb–Sr isotopic data. Neues Jahrbuch für Mineralogie, Monatshefte 5, 203–20.Google Scholar
Kroner, U. & Romer, R. L. 2013. Two plates – many subduction zones: the Variscan orogeny reconsidered. Gondwana Research 24, 298329.Google Scholar
Kříž, J. 1991. The Silurian of the Prague Basin (Bohemia) – tectonic, eustatic and volcanic controls on facies and faunal development. Special Papers in Palaeontology 44, 179203.Google Scholar
Kříž, J. 1992. Silurian field excursions: Prague Basin (Barrandian), Bohemia. National Museum of Wales, Geological Notes 13, 1111.Google Scholar
Krs, M., Krsová, M., Pruner, P., Chvojka, R. & Havlíček, V. 1987. Palaeomagnetism, palaeogeography and the multicomponent analysis of Middle and Upper Cambrian rocks of the Barrandian in the Bohemian Massif. Tectonophysics 139, 120.Google Scholar
Krs, M., Pruner, P. & Man, O. 2001. Tectonic and paleogeographic interpretation of the paleomagnetism of Variscan and pre-Variscan formations of the Bohemian Massif, with special reference to the Barrandian terrane. Tectonophysics 332, 93114.Google Scholar
Kubínová, Š., Faryad, S. W., Verner, K., Schmitz, M. D. & Holub, F. V. 2017. Ultrapotassic dykes in the Moldanubian Zone and their significance for understanding of the post-collisional mantle dynamics during Variscan orogeny in the Bohemian Massif. Lithos 272–273, 205–21.Google Scholar
Kukal, Z. & Jäger, O. 1988. Siliciclastic signal of the Variscan orogenesis: the Devonian Srbsko Formation of Central Bohemia. Bulletin of the Central Geological Survey 63, 6581.Google Scholar
Lease, R. O., Burbank, D. W., Zhang, H., Liu, J. & Yuan, D. 2012. Cenozoic shortening budget for the northeastern edge of the Tibetan Plateau: is lower crustal flow necessary? Tectonics 31, TC3011. doi: 10.1029/2011TC003066.Google Scholar
Lehnert, O., Frýda, J., Buggisch, W., Munnecke, A., Nützel, A., Kříž, J. & Manda, Š. 2007. δ13C records across the late Silurian Lau event: new data from middle palaeo-latitudes of northern peri-Gondwana (Prague Basin, Czech Republic). Palaeogeography, Palaeoclimatology, Palaeoecology 245, 227–44.Google Scholar
Lewandowski, M. 2003. Assembly of Pangea: combined paleomagnetic and paleoclimatic approach. Advances in Geophysics 46, 199235.Google Scholar
Li, Y., Wang, C., Dai, J., Xu, G., Hou, Y. & Li, X. 2015. Propagation of the deformation and growth of the Tibetan–Himalayan orogen: a review. Earth-Science Reviews 143, 3661.Google Scholar
Lister, G. & Foster, M. 2009. Tectonic mode switches and the nature of orogenesis. Lithos 113, 274–91.Google Scholar
Maierová, P., Schulmann, K., Lexa, O., Guillot, S., Štípská, Š., Janoušek, V. & Čadek, O. 2016. European Variscan orogenic evolution as an analogue of Tibetan–Himalayan orogen: insights from petrology and numerical modeling. Tectonics 35, 1760–80.Google Scholar
Manda, Š., Štorch, P., Slavík, L., Frýda, J., Kříž, J. & Tasáryová, Z. 2012. The graptolite, conodont and sedimentary record through the late Ludlow Kozlowskii Event (Silurian) in the shale-dominated succession of Bohemia. Geological Magazine 149, 507–31.Google Scholar
Martínez Catalán, J. R. 2011. Are the oroclines of the Variscan belt related to late Variscan strike-slip tectonics? Terra Nova 23, 241–7.Google Scholar
Martínez Catalán, J. R. 2012. The Central Iberian arc, an orocline centered in the Iberian Massif and some implications for the Variscan belt. International Journal of Earth Sciences 101, 1299–314.Google Scholar
Matte, P. 1986. Tectonics and plate tectonics model for the Variscan belt of Europe. Tectonophysics 126, 329–74.Google Scholar
Matte, P. 2001. The Variscan collage and orogeny (480–290 Ma) and the tectonic definition of the Armorica microplate: a review. Terra Nova 13, 122–8.Google Scholar
Mazur, S. & Aleksandrowski, P. 2001. The Teplá (?)/ Saxothuringian suture in the Karkonosze–Izera massif, western Sudetes, central European Variscides. International Journal of Earth Sciences 90, 341–60.Google Scholar
Mazur, S., Aleksandrowski, P., Kryza, R. & Oberc-Dziedzic, T. 2006. The Variscan Orogen in Poland. Geological Quarterly 50, 89118.Google Scholar
Melichar, R. 2004. Tectonics of the Prague Synform: a hundred years of scientific discussion. Krystalinikum 30, 167–87.Google Scholar
Mikuláš, R. 1998. Ordovician of the Barrandian area: reconstruction of the sedimentary basin, its benthic communities and ichnoassemblages. Journal of the Czech Geological Society 43, 143–59.Google Scholar
Mouthereau, F., Lacombe, O. & Vergés, J. 2012. Building the Zagros collisional orogen: timing, strain distribution and the dynamics of Arabia/Eurasia plate convergence. Tectonophysics 532–535, 2760.Google Scholar
Murphy, D. C. 1987. Suprastructure/infrastructure transition, east central Cariboo Mountains, British Columbia: geometry, kinematics and tectonic implications. Journal of Structural Geology 9, 1329.Google Scholar
Paris, F. & Robardet, M. 1990. Early Palaeozoic palaeobiogeography of the Variscan regions. Tectonophysics 177, 193213.Google Scholar
Patočka, F., Pruner, P. & Štorch, P. 2003. Palaeomagnetism and geochemistry of Early Palaeozoic rocks of the Barrandian (Teplá–Barrandian Unit, Bohemian Massif): palaeotectonic implications. Physics and Chemistry of the Earth 28, 735–49.Google Scholar
Patočka, F. & Štorch, P. 2004. Evolution of geochemistry and depositional settings of Early Palaeozoic siliciclastics of the Barrandian (Teplá–Barrandian Unit, Bohemian Massif, Czech Republic). International Journal of Earth Sciences 93, 728–41.Google Scholar
Pavlis, T. L. 1996. Fabric development in syn-tectonic intrusive sheets as a consequence of melt-dominated flow and thermal softening of the crust. Tectonophysics 253, 131.Google Scholar
Pertoldová, J., Verner, K., Vrána, S., Buriánek, D., Štědrá, V. & Vondrovic, L. 2010. Comparison of lithology and tectonometamorphic evolution of units at the northern margin of the Moldanubian Zone: implications for geodynamic evolution in the northeastern part of the Bohemian Massif. Journal of Geosciences 55, 299319.Google Scholar
Rajlich, P., Schulmann, K. & Synek, J. 1988. Strain analysis of conglomerates in the Central Bohemian shear zone. Krystalinikum 19, 119–34.Google Scholar
Ramsay, J. G. 2003. Folding and Fracturing of Rocks. Caldwell, NJ: Blackburn Press, 568 pp.Google Scholar
Ramsay, J. G. 1974. Development of chevron folds. Geological Society of America Bulletin 85, 1741–54.Google Scholar
Riller, U. & Oncken, O. 2003. Growth of the central Andean plateau by tectonic segmentation is controlled by the gradient in crustal shortening. Journal of Geology 111, 367–84.Google Scholar
Robardet, M. 2003. The Armorica ‘microplate’: fact or fiction? Critical review of the concept and contradictory palaeobiogeographical data. Palaeogeography, Palaeoclimatology, Palaeoecology 195, 125–48.Google Scholar
Röhlich, P. 2007. Structure of the Prague basin: the deformation diversity and its causes (Czech Republic). Bulletin of Geosciences 82, 175–82.Google Scholar
Royden, L. H., Burchfiel, B. C., King, R. W., Wang, E., Chen, Z., Shen, F. & Liu, Y. 1997. Surface deformation and lower crustal flow in eastern Tibet. Science 276, 788–90.Google Scholar
Royden, L. H., Burchfiel, B. C. & van der Hilst, R. D. 2008. The geological evolution of the Tibetan Plateau. Science 321, 1054–8.Google Scholar
Schulmann, K., Konopásek, J., Janoušek, V., Lexa, O., Lardeaux, J. M., Edel, J. B., Štípská, P. & Ulrich, S. 2009. An Andean type Palaeozoic convergence in the Bohemian Massif. Comptes Rendus Geoscience 341, 266–86.Google Scholar
Schulmann, K., Lexa, O., Janoušek, V., Lardeaux, J. M. & Edel, J. B. 2014. Anatomy of a diffuse cryptic suture zone: an example from the Bohemian Massif, European Variscides. Geology 42, 275–8.Google Scholar
Searle, M. P., Elliott, J. R., Phillips, R. J. & Chung, S. L. 2011. Crustal–lithospheric structure and continental extrusion of Tibet. Journal of the Geological Society, London 168, 633–72.Google Scholar
Servais, T. & Sintubin, M. 2009. Avalonia, Armorica, Perunica: terranes, microcontinents, microplates or palaeobiogeographical provinces? In Early Palaeozoic Peri-Gondwana Terranes: New Insights from Tectonics and Biogeography (ed. Bassett, M. G.), pp. 103–15. Geological Society of London, Special Publication no. 325.Google Scholar
Sláma, J., Dunkley, D. J., Kachlík, V. & Kusiak, M. A. 2008. Transition from island-arc to passive setting on the continental margin of Gondwana: U–Pb zircon dating of Neoproterozoic metaconglomerates from the SE margin of the Teplá–Barrandian Unit, Bohemian Massif. Tectonophysics 461, 4459.Google Scholar
Slavík, L. 2004. The Pragian–Emsian conodont successions of the Barrandian area: search of an alternative to the GSSP polygnathid-based correlation concept. Geobios 37, 454–70.Google Scholar
Slavík, L., Carls, P., Hladil, J. & Koptíková, L. 2012. Subdivision of the Lochkovian Stage based on conodont faunas from the stratotype area (Prague Synform, Czech Republic). Geological Journal 47, 616–31.Google Scholar
Slobodník, M., Melichar, R., Hurai, V. & Bakker, R. J. 2012. Litho-stratigraphic effect on Variscan fluid flow within the Prague synform, Barrandian: evidence based on C, O, Sr isotopes and fluid inclusions. Marine and Petroleum Geology 35, 128–38.Google Scholar
Stampfli, G. M., Hochard, C., Vérard, C., Wilhem, C. & von Raumer, J. F. 2013. The formation of Pangea. Tectonophysics 593, 119.Google Scholar
Štorch, P. 1986. Ordovician–Silurian boundary in the Prague Basin (Barrandian area, Bohemia). Journal of Geological Sciences, Geology 41, 69103.Google Scholar
Štorch, P. 1990. Upper Ordovician–lower Silurian sequences of the Bohemian Massif, central Europe. Geological Magazine 127, 225–39.Google Scholar
Štorch, P. 2006. Facies development, depositional settings and sequence stratigraphy across the Ordovician–Silurian boundary: a new perspective from the Barrandian area of the Czech Republic. Geological Journal 41, 163–92.Google Scholar
Štorch, P., Fatka, O. & Kraft, P. 1993. Lower Palaeozoic of the Barrandian area (Czech Republic) – a review. Coloquios de Paleontología 45, 163–91.Google Scholar
Štorch, P. & Frýda, J. 2012. The late Aeronian graptolite sedgwickii Event, associated positive carbon isotope excursion and facies changes in the Prague Synform (Barrandian area, Bohemia). Geological Magazine 149, 1089–106.Google Scholar
Štorch, P., Manda, Š., Slavík, L. & Tasáryová, Z. 2016. Wenlock–Ludlow boundary interval revisited: new insights from the offshore facies of the Prague Synform, Czech Republic. Canadian Journal of Earth Sciences 53, 666–73.Google Scholar
Strnad, L. & Mihaljevič, M. 2005. Sedimentary provenance of Mid-Devonian clastic sediments in the Teplá–Barrandian Unit (Bohemian Massif): U–Pb and Pb–Pb geochronology of detrital zircons by laser ablation ICP-MS. Mineralogy and Petrology 84, 4768.Google Scholar
Styron, R. H., Taylor, M. H. & Murphy, M. A. 2011. Oblique convergence, arc-parallel extension, and the role of strike-slip faulting in the High Himalaya. Geological Society of America Bulletin 7, 582–96.Google Scholar
Suchý, V., Dobeš, P., Filip, J., Stejskal, M. & Zeman, A. 2002 a. Conditions for veining in the Barrandian Basin (Lower Palaeozoic), Czech Republic: evidence from fluid inclusion and apatite fission track analysis. Tectonophysics 348, 2550.Google Scholar
Suchý, V., Dobeš, P., Sýkorová, I., Machovič, V., Stejskal, M., Kroufek, J., Chudoba, J., Matějovská, L., Havelcová, M. & Matysová, P. 2010. Oil-bearing inclusions in vein quartz and calcite and, bitumens in veins: testament to multiple phases of hydrocarbon migration in the Barrandian basin (lower Palaeozoic), Czech Republic. Marine and Petroleum Geology 27, 285–97.Google Scholar
Suchý, V., Rozkošný, I., Žák, K. & Franců, J. 1996. Epigenetic dolomitization of the Přídolí formation (Upper Silurian), the Barrandian basin, Czech Republic: implications for burial history of Lower Paleozoic strata. International Journal of Earth Sciences 85, 264–77.Google Scholar
Suchý, V., Sandler, A., Slobodník, M., Sýkorová, I., Filip, J., Melka, K. & Zeman, A. 2015. Diagenesis to very low-grade metamorphism in lower Palaeozoic sediments: a case study from deep borehole Tobolka 1, the Barrandian Basin, Czech Republic. International Journal of Coal Geology 140, 4162.Google Scholar
Suchý, V., Sýkorová, I., Dobeš, P., Machovič, V., Filip, J., Zeman, A. & Stejskal, M. 2012. Blackened bioclasts and bituminous impregnations in the Koněprusy Limestone (Lower Devonian), the Barrandian area, Czech Republic: implications for basin analysis. Facies 58, 759–77.Google Scholar
Suchý, V., Sýkorová, I., Melka, K., Filip, J. & Machovič, V. 2007. Illite ‘crystallinity’, maturation of organic matter and microstructural development associated with lowest-grade metamorphism of Neoproterozoic sediments in the Teplá–Barrandian unit, Czech Republic. Clay Minerals 42, 503–26.Google Scholar
Suchý, V., Sýkorová, I., Stejskal, M., Šafanda, J., Machovič, V. & Novotná, M. 2002 b. Dispersed organic matter from Silurian shales of the Barrandian Basin, Czech Republic: optical properties, chemical composition and thermal maturity. International Journal of Coal Geology 53, 125.Google Scholar
Tait, J., Bachtadse, V. & Soffel, H. 1994. New palaeomagnetic constraints on the position of central Bohemia during early Ordovician times. Geophysical Journal International 116, 131–40.Google Scholar
Tait, J., Bachtadse, V. & Soffel, H. 1995. Upper Ordovician paleogeography of the Bohemian Massif: implications for Armorica. Geophysical Journal International 122, 2112018.Google Scholar
Tasáryová, Z., Schnabl, P., Čížková, K., Pruner, P., Janoušek, V., Rapprich, V., Štorch, P., Manda, Š., Frýda, J. & Trubač, J. 2014. Gorstian palaeoposition and geotectonic setting of Suchomasty Volcanic Centre (Silurian, Prague Basin, Teplá–Barrandian Unit, Bohemian Massif). GFF 136, 262–5.Google Scholar
Timmermann, H., Dörr, W., Krenn, E., Finger, F. & Zulauf, G. 2006. Conventional and in situ geochronology of the Teplá Crystalline unit, Bohemian Massif: implications for the processes involving monazite formation. International Journal of Earth Sciences 95, 629–47.Google Scholar
Tomek, F., Žák, J. & Chadima, M. 2015. Granitic magma emplacement and deformation during early-orogenic syn-convergent transtension: the Staré Sedlo complex, Bohemian Massif. Journal of Geodynamics 87, 5066.Google Scholar
Vacek, F. 2011. Palaeoclimatic event at the Lochkovian–Pragian boundary recorded in magnetic susceptibility and gamma-ray spectrometry (Prague Synclinorium, Czech Republic). Bulletin of Geosciences 86, 259–68.Google Scholar
Vanderhaeghe, O. 2012. The thermal–mechanical evolution of crustal orogenic belts at convergent plate boundaries: a reappraisal of the orogenic cycle. Journal of Geodynamics 56–57, 124–45.Google Scholar
Venera, Z., Schulmann, K. & Kroner, A. 2000. Intrusion within a transtensional tectonic domain: the Čistá granodiorite (Bohemian Massif): structure and rheological modelling. Journal of Structural Geology 22, 1437–54.Google Scholar
Verner, K., Buriánek, D., Vrána, S., Vondrovic, L., Pertoldová, J., Hanžl, P. & Nahodilová, R. 2009. Tectonometamorphic features of geological units along the northern periphery of the Moldanubian Zone (Bohemian Massif). Journal of Geosciences 54, 87100.Google Scholar
Vodrážková, S., Frýda, J., Suttner, T. J., Koptíková, L. & Tonarová, P. 2013. Environmental changes close to the Lower–Middle Devonian boundary; the Basal Choteč Event in the Prague Basin (Czech Republic). Facies 59, 425–49.Google Scholar
Volk, H., Horsfield, B., Mann, U. & Suchý, V. 2002. Variability of petroleum inclusions in vein, fossil and vug cements: a geochemical study in the Barrandian Basin (Lower Palaeozoic, Czech Republic). Organic Geochemistry 33, 1319–41.Google Scholar
Weinerová, H., Hron, K., Bábek, O., Šimíček, D. & Hladil, J. 2017. Quantitative allochem compositional analysis of Lochkovian–Pragian boundary sections in the Prague Basin (Czech Republic). Sedimentary Geology 354, 4359.Google Scholar
Winchester, J. A. 2002. Palaeozoic amalgamation of Central Europe: new results from recent geological and geophysical investigations. Tectonophysics 360, 521.Google Scholar
Yin, A. & Harrison, T. M. 2000. Geologic evolution of the Himalayan–Tibetan orogen. Annual Review of Earth and Planetary Sciences 28, 211–80.Google Scholar
Žák, J., Dragoun, F., Verner, K., Chlupáčová, M., Holub, F. V. & Kachlík, V. 2009. Forearc deformation and strain partitioning during growth of a continental magmatic arc: the northwestern margin of the Central Bohemian Plutonic Complex, Bohemian Massif. Tectonophysics 469, 93111.Google Scholar
Žák, J., Holub, F. V. & Verner, K. 2005. Tectonic evolution of a continental magmatic arc from transpression in the upper crust to exhumation of mid-crustal orogenic root recorded by episodically emplaced plutons: the Central Bohemian Plutonic Complex (Bohemian Massif). International Journal of Earth Sciences 94, 385400.Google Scholar
Žák, J., Kraft, P. & Hajná, J. 2013. Timing, styles, and kinematics of Cambro–Ordovician extension in the Teplá–Barrandian Unit, Bohemian Massif, and its bearing on the opening of the Rheic Ocean. International Journal of Earth Sciences 102, 415–33.Google Scholar
Žák, J., Kratinová, Z., Trubač, J., Janoušek, V., Sláma, J. & Mrlina, J. 2011. Structure, emplacement, and tectonic setting of Late Devonian granitoid plutons in the Teplá–Barrandian unit, Bohemian Massif. International Journal of Earth Sciences 100, 1477–95.Google Scholar
Žák, J. & Sláma, J. 2017. How far did the Cadomian ʽterranesʼ travel from Gondwana during early Palaeozoic? A critical reappraisal based on detrital zircon geochronology. International Geology Review, published online 5 June 2017. doi: 10.1080/00206814.2017.1334599.Google Scholar
Žák, J., Sláma, J. & Burjak, M. 2017. Rapid extensional unroofing of a granite–migmatite dome with relics of high-pressure rocks, the Podolsko complex, Bohemian Massif. Geological Magazine 154, 354–80.Google Scholar
Žák, J., Verner, K., Holub, F. V., Kabele, P., Chlupáčová, M. & Halodová, P. 2012. Magmatic to solid state fabrics in syntectonic granitoids recording early Carboniferous orogenic collapse in the Bohemian Massif. Journal of Structural Geology 36, 2742.Google Scholar
Zulauf, G. 1994. Ductile normal faulting along the West Bohemian Shear Zone (Moldanubian/Teplá–Barrandian boundary): evidence for late Variscan extensional collapse in the Variscan Internides. Geologische Rundschau 83, 276–92.Google Scholar
Zulauf, G. 1997. From very low-grade to eclogite-facies metamorphism: tilted crustal sections as a consequence of Cadomian and Variscan orogeny in the Teplá–Barrandian unit (Bohemian Massif). Geotektonische Forschungen 89, 1302.Google Scholar
Zulauf, G. 2001. Structural style, deformational mechanisms and paleodifferential stress along an exposed crustal section: constraints on the rheology of quartzofeldspathic rocks at supra- and infrastructural levels (Bohemian Massif). Tectonophysics 332, 211–37.Google Scholar
Zulauf, G., Bues, C., Dörr, W. & Vejnar, Z. 2002. 10 km minimum throw along the West Bohemian shear zone: evidence for dramatic crustal thickening and high topography in the Bohemian Massif (European Variscides). International Journal of Earth Sciences 91, 850–64.Google Scholar
Zuza, A. V., Cheng, X. & Yin, A. 2016. Testing models of Tibetan Plateau formation with Cenozoic shortening estimates across the Qilian Shan–Nan Shan thrust belt. Geosphere 12, 501–32.Google Scholar
Zwart, H. J. 1967. The duality of orogenic belts. Geologie en Mijnbouw 46, 283309.Google Scholar