Hostname: page-component-76fb5796d-r6qrq Total loading time: 0 Render date: 2024-04-25T17:22:03.275Z Has data issue: false hasContentIssue false

The pre-orogenic detrital zircon record of the Peri-Gondwanan crust

Published online by Cambridge University Press:  08 February 2018

TOBIAS STEPHAN*
Affiliation:
Institut für Geologie, TU Bergakademie Freiberg, B. v. Cotta Str. 2, 09599 Freiberg, Germany
UWE KRONER
Affiliation:
Institut für Geologie, TU Bergakademie Freiberg, B. v. Cotta Str. 2, 09599 Freiberg, Germany
ROLF L. ROMER
Affiliation:
Deutsches GeoForschungsZentrum GFZ, Telegrafenberg, 14473 Potsdam, Germany
*
Author for correspondence: tobias.stephan@geo.tu-freiberg.de

Abstract

We present a statistical approach to data mining and quantitatively evaluating detrital age spectra for sedimentary provenance analyses and palaeogeographic reconstructions. Multidimensional scaling coupled with density-based clustering allows the objective identification of provenance end-member populations and sedimentary mixing processes for a composite crust. We compiled 58 601 detrital zircon U–Pb ages from 770 Precambrian to Lower Palaeozoic shelf sedimentary rocks from 160 publications and applied statistical provenance analysis for the Peri-Gondwanan crust north of Africa and the adjacent areas. We have filtered the dataset to reduce the age spectra to the provenance signal, and compared the signal with age patterns of potential source regions. In terms of provenance, our results reveal three distinct areas, namely the Avalonian, West African and East African–Arabian zircon provinces. Except for the Rheic Ocean separating the Avalonian Zircon Province from Gondwana, the statistical analysis provides no evidence for the existence of additional oceanic lithosphere. This implies a vast and contiguous Peri-Gondwanan shelf south of the Rheic Ocean that is supplied by two contrasting super-fan systems, reflected in the zircon provinces of West Africa and East Africa–Arabia.

Type
Original Articles
Copyright
Copyright © 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

Abati, J., Aghzer, A. M., Gerdes, A. & Ennih, N., 2010. Detrital zircon ages of Neoproterozoic sequences of the Moroccan Anti-Atlas belt. Precambrian Research 181, 115–28.10.1016/j.precamres.2010.05.018Google Scholar
Altumi, M. M., Elicki, O., Linnemann, U., Hofmann, M., Sagawe, A. & Gärtner, A. 2013. U–Pb LA-ICP-MS detrital zircon ages from the Cambrian of Al Qarqaf Arch, central-western Libya: provenance of the West Gondwanan sand sea at the dawn of the early Palaeozoic. Journal of African Earth Sciences 79, 7497.10.1016/j.jafrearsci.2012.11.007Google Scholar
Andersen, T. 2005. Detrital zircons as tracers of sedimentary provenance: limiting conditions from statistics and numerical simulation. Chemical Geology 216, 249–70.10.1016/j.chemgeo.2004.11.013Google Scholar
Andersen, T., Griffin, W. L. & Pearson, N. J. 2002. Crustal evolution in the SW part of the Baltic Shield: the Hf isotope evidence. Journal of Petrology 43, 1725–47.10.1093/petrology/43.9.1725Google Scholar
Andersen, T., Kristoffersen, M. & Elburg, M. A. 2016. How far can we trust provenance and crustal evolution information from detrital zircons? A South African case study. Gondwana Research 34, 129–48.10.1016/j.gr.2016.03.003Google Scholar
Andresen, A., Abu El-Enen, M. M., Stern, R. J., Wilde, S. A. & Ali, K. A. 2014. The Wadi Zaghra metasediments of Sinai, Egypt: new constraints on the late Cryogenian–Ediacaran tectonic evolution of the northernmost Arabian–Nubian Shield. International Geology Review 56, 1020–38.10.1080/00206814.2014.907755Google Scholar
Arboit, F., Collins, A. S., Morley, C. K., King, R. & Amrouch, K. 2016. Detrital zircon analysis of the southwest Indochina terrane, central Thailand: unravelling the Indosinian orogeny. Geological Society of America Bulletin 128, 1024–43.10.1130/B31411.1Google Scholar
Arndt, N. T. & Goldstein, S. L. 1987. Use and abuse of crust-formation ages. Geology 15, 893–95.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, 1305–20.10.1130/0016-7606(1977)88<1305:LPSFIS>2.0.CO;22.0.CO;2>Google Scholar
Avigad, D., Gerdes, A., Morag, N. & Bechstädt, T. 2012. Coupled U–Pb–Hf of detrital zircons of Cambrian sandstones from Morocco and Sardinia: implications for provenance and Precambrian crustal evolution of North Africa. Gondwana Research 21, 690703.10.1016/j.gr.2011.06.005Google Scholar
Avigad, D., Kolodner, K., McWilliams, M., Persing, H. & Weissbrod, T. 2003. Origin of northern Gondwana Cambrian sandstone revealed by detrital zircon SHRIMP dating. Geology 31, 227–30.10.1130/0091-7613(2003)031<0227:OONGCS>2.0.CO;22.0.CO;2>Google Scholar
Avigad, D., Sandler, A., Kolodner, K., Stern, R. J., McWilliams, M., Miller, N. & Beyth, M. 2005. Mass-production of Cambro–Ordovician quartz-rich sandstone as a consequence of chemical weathering of Pan-African terranes: environmental implications. Earth and Planetary Science Letters 240, 818–26.Google Scholar
Avigad, D., Stern, R. J., Beyth, M., Miller, N. & McWilliams, M. O. 2007. Detrital zircon U–Pb geochronology of Cryogenian diamictites and Lower Paleozoic sandstone in Ethiopia (Tigrai): age constraints on Neoproterozoic glaciation and crustal evolution of the southern Arabian–Nubian Shield. Precambrian Research 154, 88106.10.1016/j.precamres.2006.12.004Google Scholar
Avigad, D., Weissbrod, T., Gerdes, A., Zlatkin, O., Ireland, T. R. & Morag, N. 2015. The detrital zircon U–Pb–Hf fingerprint of the northern Arabian-Nubian Shield as reflected by a Late Ediacaran arkosic wedge (Zenifim Formation; subsurface Israel). Precambrian Research 266, 111.10.1016/j.precamres.2015.04.011Google Scholar
Balintoni, I., Balica, C., Ducea, M. N., Chen, F., Hann, H. P. & Şabliovschi, V. 2009. Late Cambrian–Early Ordovician Gondwanan terranes in the Romanian Carpathians: a zircon U–Pb provenance study. Gondwana Research 16, 119–33.10.1016/j.gr.2009.01.007Google Scholar
Ballèvre, M., Le Goff, E. & Hébert, R. 2001. The tectonothermal evolution of the Cadomian belt of northern Brittany, France: a Neoproterozoic volcanic arc. Tectonophysics 331, 1943.Google Scholar
Ballèvre, M., Marchand, J., Godard, G., Goujou, J.-C., Christian, J. & Wyns, R. 1994. Eo-Hercynian Events in the Armorican Massif. In Pre-Mesozoic Geology in France and Related Areas (eds. Chantraine, J., Rolet, J., Santallier, D. S., Piqué, A. & Keppie, J. D.), pp. 183–94. Berlin, Heidelberg: Springer.10.1007/978-3-642-84915-2_19Google Scholar
Barr, S. M., Hamilton, M. A., Samson, S. D., Satkoski, A. M. & White, C. E. 2012. Provenance variations in northern Appalachian Avalonia based on detrital zircon age patterns in Ediacaran and Cambrian sedimentary rocks, New Brunswick and Nova Scotia, Canada. Canadian Journal of Earth Sciences 49, 533–46.10.1139/e11-070Google Scholar
Bea, F., Montero, P., Talavera, C., Abu Anbar, M., Scarrow, J. H., Molina, J. F. & Moreno, J. A. 2010. The palaeogeographic position of Central Iberia in Gondwana during the Ordovician: evidence from zircon chronology and Nd isotopes. Terra Nova 22, 341–6.Google Scholar
Bearden, W. O., Sharma, S. & Teel, J. E. 1982. Sample size affects on chi square and other statistics used in evaluating causal models. Journal of Marketing Research 19, 425–30.10.1177/002224378201900404Google Scholar
Be'eri-Shlevin, Y., Gee, D., Claesson, S., Ladenberger, A., Majka, J., Kirkland, C., Robinson, P. & Frei, D. 2011. Provenance of Neoproterozoic sediments in the Särv nappes (Middle Allochthon) of the Scandinavian Caledonides: LA-ICP-MS and SIMS U–Pb dating of detrital zircons. Precambrian Research 187, 181200.10.1016/j.precamres.2011.03.007Google Scholar
Belka, Z., Ahrendt, H., Franke, W. & Wemmer, K. 2000. The Baltica-Gondwana suture in central Europe: evidence from K–Ar ages of detrital muscovites and biogeographical data. In Orogenic Processes: Quantification and Modelling in the Variscan Belt (eds Franke, W., Altherr, R., Haak, V., Ocncken, O. & Tanner, D.), pp. 87102. Geological Society of London, Special Publication no. 179.Google Scholar
Belka, Z. & Narkiewicz, M. 2008. Devonian. In The Geology of Central Europe. Volume 1: Precambrian and Palaeozoic (ed. McCann, T.), pp. 383410. London: The Geological Society of London.10.1144/CEV1P.8Google Scholar
Berger, J., Féménias, O., Ohnenstetter, D., Bruguier, O., Plissart, G., Mercier, J.-C. C. & Demaiffe, D. 2010. New occurrence of UHP eclogites in Limousin (French Massif Central): age, tectonic setting and fluid–rock interactions. Lithos 118, 365–82.Google Scholar
Blatt, H. 1967. Provenance determinations and recycling of sediments. Journal of Sedimentary Petrology 37, 1031–44.Google Scholar
Boger, S. D., Carson, C. J., Wilson, C. J. L. & Fanning, C. M. 2000. Neoproterozoic deformation in the Radok Lake region of the northern Prince Charles Mountains, east Antarctica; evidence for a single protracted orogenic event. Precambrian Research 104, 124.10.1016/S0301-9268(00)00079-6Google Scholar
Borg, I. & Groenen, P. J. F. 2005. Modern Multidimensional Scaling: Theory and Applications, 2nd edition. New York: Springer.Google Scholar
Boyce, W. D., Ash, J. S. & O'Brien, B. H. 1991. A new fossil locality in the Bay of Exploits, central Newfoundland. Current Research, Newfoundland Department of Mines and Energy, Geological Survey Branch Report 91-1, 7982.Google Scholar
Braid, J. A., Murphy, J. B., Quesada, C. & Mortensen, J. 2011. Tectonic escape of a crustal fragment during the closure of the Rheic Ocean: U–Pb detrital zircon data from the Late Palaeozoic Pulo do Lobo and South Portuguese zones, southern Iberia. Journal of the Geological Society, London 168, 383–92.10.1144/0016-76492010-104Google Scholar
Bream, B. R., Hatcher, R. D., Miller, C. F. & Fullagar, P. D. 2004. Detrital zircon ages and Nd isotopic data from the southern Appalachian crystalline core, Georgia, South Carolina, North Carolina, and Tennessee: new provenance constraints for part of the Laurentian margin. In Proterozoic Tectonic Evolution of the Grenville Orogen in North America (eds Tollo, R. P., McLelland, J., Corriveau, L. & Bartholomew, M. J.), pp. 459–75. Geological Society of America, Memoirs no. 197.Google Scholar
Breitkreuz, C. & Kennedy, A. 1999. Magmatic flare-up at the Carboniferous/Permian boundary in the NE German Basin revealed by SHRIMP zircon ages. Tectonophysics 302, 307–26.10.1016/S0040-1951(98)00293-5Google Scholar
Bröcker, M., Klemd, R., Cosca, M., Brock, W., Larionov, A. N. & Rodionov, N. 2009. The timing of eclogite facies metamorphism and migmatization in the Orlica-Śnieżnik complex, Bohemian Massif: constraints from a multimethod geochronological study. Journal of Metamorphic Geology 27, 385403.10.1111/j.1525-1314.2009.00823.xGoogle Scholar
Campello, R. J. G. B., Moulavi, D. & Sander, J. 2013. Density-based clustering based on hierarchical density estimates. In Advances in Knowledge Discovery and Data Mining. PAKDD 2013. Lecture Notes in Computer Science (eds Pei, J., Tseng, V. S., Cao, L., Motoda, H. & Xu, G.), pp. 160–72. Berlin, Heidelberg: Springer.Google Scholar
Carmignani, L., Franceschelli, M., Pertusati, P. C., Memmi, I. & Ricci, C. A. 1982. An example of compositional control of the celadonitic content of muscovite and the incoming of biotite in metapelites (Nurra, NW Sardinia). Neues Jahrbuch fur Mineralogie, Monatshefte 7, 289311.Google Scholar
Carson, C. J., Boger, S. D., Fanning, C. M., Wilson, C. J. L. & Thost, D. E. 2000. SHRIMP U–Pb geochronology from Mount Kirkby, northern Prince Charles Mountains, East Antarctica. Antarctic Science 12, 429–42.10.1017/S0954102000000523Google Scholar
Cawood, P. A., Hawkesworth, C. J. & Dhuime, B. 2012. Detrital zircon record and tectonic setting. Geology 40, 875–8.10.1130/G32945.1Google Scholar
Cawood, P. A., Hawkesworth, C. J. & Dhuime, B. 2013. The continental record and the generation of continental crust. Geological Society of America Bulletin 125, 1432.Google Scholar
Chantraine, J., Egal, E., Thiéblemont, D., Le Goff, E., Guerrot, C., Ballèvre, M. & Guennoc, P. 2001. The Cadomian active margin (North Armorican Massif, France): a segment of the North Atlantic Panafrican belt. Tectonophysics 331, 118.Google Scholar
Chen, F., Siebel, W., Satir, M., Terzioğlu, M. & Saka, K. 2002. Geochronology of the Karadere basement (NW Turkey) and implications for the geological evolution of the Istanbul zone. International Journal of Earth Sciences 91, 469–81.10.1007/s00531-001-0239-6Google Scholar
Cocco, F. & Funedda, A. 2017. The Sardic Phase: field evidence of Ordovician tectonics in SE Sardinia, Italy. Geological Magazine, published online 14 September 2017. doi: 10.1017/s0016756817000723. 14 pp.Google Scholar
Cocks, L. R. M. & Fortey, R. A. 1982. Faunal evidence for oceanic separations in the Palaeozoic of Britain. Journal of the Geological Society, London 139, 465–78.10.1144/gsjgs.139.4.0465Google Scholar
Cocks, L. R. M., McKerrow, W. S. & Van Staal, C. R. 1997. The margins of Avalonia. Geological Magazine 134, 627–36.Google Scholar
Cocks, L. R. M. & Torsvik, T. H. 2002. Earth geography from 500 to 400 million years ago: a faunal and palaeomagnetic review. Journal of the Geological Society, London 159, 631–44.10.1144/0016-764901-118Google Scholar
Cocks, L. R. M. & Torsvik, T. H. 2006. European geography in a global context from the Vendian to the end of the Palaeozoic. In European Lithosphere Dynamics (eds Gee, D. G. & Stephenson, R. A.), pp. 8395. The Geological Society of London, Memoirs no. 32.Google Scholar
Collins, A. S., Kinny, P. D. & Razakamanana, T. 2012. Depositional age, provenance and metamorphic age of metasedimentary rocks from southern Madagascar. Gondwana Research 21, 353–61.Google Scholar
Corvino, A. F., Boger, S. D., Henjes-Kunst, F., Wilson, C. J. L. & Fitzsimons, I. C. W. 2008. Superimposed tectonic events at 2450 Ma, 2100 Ma, 900 Ma and 500 Ma in the North Mawson Escarpment, Antarctic Prince Charles Mountains. Precambrian Research 167, 281302.10.1016/j.precamres.2008.09.001Google Scholar
Corvino, A. F., Boger, S. D., Wilson, C. J. L. & Fitzsimons, I. C. W. 2005. Geology and SHRIMP U–Pb zircon chronology of the Clemence Massif, central Prince Charles Mountains, East Antarctica. Terra Antarctica 12, 5568.Google Scholar
Corvino, A. & Henjes-Kunst, F. 2007. A record of 2.5 and 1.1 billion year old crust in the Lawrence Hills, Antarctic Southern Prince Charles Mountains. Terra Antarctica 14, 13.Google Scholar
Crowley, Q. G., Floyd, P. A., Winchester, J. A., Franke, W. & Holland, J. G. 2000. Early Palaeozoic rift-related magmatism in Variscan Europe: fragmentation of the Armorican Terrane Assemblage. Terra Nova 12, 171–80.Google Scholar
Dallmeyer, R. D., Catalán, J. R. M., Arenas, R., Gil Ibarguchi, J. I., Gutiérrez, Alonso, G., Farias, P., Bastida, F. & Aller, J. 1997. Diachronous Variscan tectonothermal activity in the NW Iberian Massif: evidence from 40Ar/39Ar dating of regional fabrics. Tectonophysics 277, 307–37.10.1016/S0040-1951(97)00035-8Google Scholar
Dean, W. T., Monod, O., Rickards, R. B., Demir, O. & Bultynck, P. 2000. Lower Palaeozoic stratigraphy and palaeontology, Karadere–Zirze area, Pontus Mountains, northern Turkey. Geological Magazine 137, 555–82.Google Scholar
Dickinson, W. R. & Gehrels, G. E. 2009a. U–Pb ages of detrital zircons in Jurassic eolian and associated sandstones of the Colorado Plateau: evidence for transcontinental dispersal and intraregional recycling of sediment U–Pb ages of detrital zircons in Colorado Plateau eolianites. Geological Society of America Bulletin 121, 408–33.Google Scholar
Dickinson, W. R. & Gehrels, G. E. 2009b. Use of U–Pb ages of detrital zircons to infer maximum depositional ages of strata: a test against a Colorado Plateau Mesozoic database. Earth and Planetary Science Letters 288, 115–25.10.1016/j.epsl.2009.09.013Google Scholar
Díez Fernández, R. Catalán, J. R. M., Gerdes, A., Abati, J., Arenas, R. & Fernández-Suárez, J. 2010. U–Pb ages of detrital zircons from the Basal allochthonous units of NW Iberia: provenance and paleoposition on the northern margin of Gondwana during the Neoproterozoic and Paleozoic. Gondwana Research 18, 385–99.Google Scholar
Dill, H. G., Sachsenhofer, R. F., Grecula, P., Sasvári, T., Palinkaš, L. A., Borojevic-Soštaric, S., Strmic-Palinkaš, S., Prochaska, W., Garuti, G., Zaccarini, F., Arbouille, D. & Schulz, H.-M.. 2008. Fossil fuels, ore and industrial minerals. In The Geology of Central Europe. Volume 1: Precambrian and Palaeozoic (ed. McCann, T.), pp. 1341–449. London: The Geological Society of London.10.1144/CEV2P.9Google Scholar
Dörr, W., Zulauf, G., Gerdes, A., Lahaye, Y. & Kowalczyk, G. 2015. A hidden Tonian basement in the eastern Mediterranean: age constraints from U–Pb data of magmatic and detrital zircons of the External Hellenides (Crete and Peloponnesus). Precambrian Research 258, 83108.Google Scholar
Dudek, A. 1995. Moravo-Silesian Zone – metamorphic evolution. In Pre-Permian Geology of Central and Eastern Europe (eds Dallmeyer, R. D., Franke, W. & Weber, K.), pp. 508–11. Berlin: Springer.10.1007/978-3-642-77518-5_50Google Scholar
Dunning, G. R., Swinden, H. S., Kean, B. F., Evans, D. T. W. & Jenner, G. A. 2009. A Cambrian island arc in Iapetus: geochronology and geochemistry of the Lake Ambrose volcanic belt, Newfoundland Appalachians. Geological Magazine 128, 117.Google Scholar
Ester, M., Kriegel, H.-P., Sander, J. & Xu, X. 1996. A density-based algorithm for discovering clusters in large spatial databases with noise. Proceedings of the Second International Conference on Knowledge Discovery and Data Mining (KDD-96) 96, 226–31.Google Scholar
Ezzouhairi, H., Ribeiro, M. L., Ait Ayad, N., Moreira, M. E., Charif, A., Ramos, J. M. F., de Oliveira, D. P. S. & Coke, C. 2008. The magmatic evolution at the Moroccan outboard of the West African craton between the Late Neoproterozoic and the Early Palaeozoic. In The Boundaries of the West African Craton (eds Ennih, N. & Liégeois, J.-P.), pp. 329–43. Geological Society of London, Special Publication no. 297.Google Scholar
Fedo, C. M., Sircombe, K. N. & Rainbird, R. H. 2003. Detrital zircon analysis of the sedimentary record. Reviews in Mineralogy and Geochemistry 53, 277303.10.2113/0530277Google Scholar
Fernández Suárez, J., Gutiérrez Alonso, G., Jenner, G. A. & Tubrett, M. N. 1998. Edades del basamento pre-varisco en Iberia: Herencia Icartiense, Grenville y Cadomiense en rocas del Complejo Olio de Sapo (NW de España). Estudio geocronológico mediante ablación láser. Studia Geologica Salmanticensia 34, 103–21.Google Scholar
Fernández-Suárez, J., Gutiérrez-Alonso, G., Pastor-Galán, D., Hofmann, M., Murphy, J. B. & Linnemann, U. 2014. The Ediacaran–Early Cambrian detrital zircon record of NW Iberia: possible sources and paleogeographic constraints. International Journal of Earth Sciences 103, 1335–57.10.1007/s00531-013-0923-3Google Scholar
Franke, W. 1989. Tectonostratigraphic units in the Variscan belt of central Europe. In Terranes in the Circum-Atlantic Paleozoic Orogens (ed. Dallmeyer, R. D.), pp. 6790. Geological Society of America, Special Papers no. 230.10.1130/SPE230-p67Google Scholar
Franke, W. 2000. The mid-European segment of the Variscides: tectonostratigraphic units, terrane boundaries and plate tectonic evolution. In Orogenic Processes: Quantification and Modelling in the Variscan Belt (eds Franke, W., Haak, V., Oncken, O. & Tanner, D.), pp. 3561. Geological Society of London: Special Publication no. 179.Google Scholar
Franke, W., Cocks, L. R. M. & Torsvik, T. H. 2017. The Palaeozoic Variscan oceans revisited. Gondwana Research 48, 257–84.Google Scholar
Franz, C., Linnemann, U., Hofmann, M., Winkler, R. & Ullrich, B. 2013. U–Pb ages of detrital zircons, fossils, and facies of the Cambro-Ordovician overstep sequence of the eastern Lausitz Block (Dubrau and Ober-Prauske formations, Saxo-Thuringian Zone). Geologica Saxonica 59, 4563.Google Scholar
Franz, L. & Romer, R. L. 2007. Caledonian high-pressure metamorphism in the Strona-Ceneri Zone (Southern Alps of southern Switzerland and northern Italy). Swiss Journal of Geosciences 100, 457–67.10.1007/s00015-007-1232-2Google Scholar
Friedl, G., Finger, F., McNaughton, N. J. & Fletcher, I. R. 2000. Deducing the ancestry of terranes: SHRIMP evidence for South America-derived Gondwana fragments in central Europe. Geology 28, 1035–8.10.1130/0091-7613(2000)28<1035:DTAOTS>2.0.CO;22.0.CO;2>Google Scholar
Garcia-Alcalde, J. L., Carls, P., Alonso, M. V. P., López, J. S., Soto, F. T., Truyols-Massoni, M. & Valenzuela-Rios, J. I. 2002. Devonian. In The Geology of Spain (eds Gibbons, W. & Moreno, T.), pp. 6791. London: Geological Society of London.Google Scholar
Garfunkel, Z. 2015. The relations between Gondwana and the adjacent peripheral Cadomian domain—constraints on the origin, history, and paleogeography of the peripheral domain. Gondwana Research 28, 1257–81.Google Scholar
Gee, D. G., Ladenberger, A., Dahlqvist, P., Majka, J., Be'eri-Shlevin, Y., Frei, D. & Thomsen, T. 2014. The Baltoscandian margin detrital zircon signatures of the central Scandes. In New Perspectives on the Caledonides of Scandinavia and Related Areas (eds Corfu, F., Gasser, D. & Chew, D.), pp. 131–55. Geological Society of London, Special Publication no. 390.Google Scholar
Geweke, J. F. & Singleton, K. J. 1980. Interpreting the likelihood ratio statistic in factor models when sample size is small. Journal of the American Statistical Association 75, 133–7.Google Scholar
Geyer, G., Elicki, O., Fatka, O. & Zylinska, A. 2008. Cambrian. In The Geology of Central Europe. Volume 1: Precambrian and Palaeozoic (ed. McCann, T.), pp. 155202. London: The Geological Society.10.1144/CEV1P.4Google Scholar
Ghienne, J.-F. 2003. Late Ordovician sedimentary environments, glacial cycles, and post-glacial transgression in the Taoudeni Basin, West Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 189, 117–45.10.1016/S0031-0182(02)00635-1Google Scholar
Ghienne, J. F., Monod, O., Kozlu, H. & Dean, W. T. 2010. Cambrian–Ordovician depositional sequences in the Middle East: a perspective from Turkey. Earth-Science Reviews 101, 101–46.Google Scholar
Giacomini, F., Bomparola, R. M. & Ghezzo, C. 2005. Petrology and geochronology of metabasites with eclogite facies relics from NE Sardinia: constraints for the Palaeozoic evolution of Southern Europe. Lithos 82, 221–48.10.1016/j.lithos.2004.12.013Google Scholar
Gil Ibarguchi, J. I., Navidad, M. & Ortega, L. A. 1990. Ordovician and Silurian igneous rocks and orthogeneisses in the Catalonian Coastal Ranges. Acta Geológica Hispánica 25, 23–9.Google Scholar
Golonka, J. & Ford, D. 2000. Pangean (Late Carboniferous-Middle Jurassic) paleoenvironment and lithofacies. Palaeogeography, Palaeoclimatology, Palaeoecology 161, 134.10.1016/S0031-0182(00)00115-2Google Scholar
Gutiérrez-Marco, J. C., De San José, M. A. & Pieren, A. P. L. 1990. Post-Cambrian Palaeozoic stratigraphy, Central Iberian Zone. In Pre-Mesozoic Geology of Iberia (ed. Dallmeyer, R. D.), pp. 3149. Berlin: Springer.Google Scholar
Gutiérrez-Marco, J. C., Robardet, M., Rábano, I., Sarmiento, G. N., San José Lancha, M. A., Herranz, P. & Pieren Pidal, A. P. 2002. Ordovician. In The Geology of Spain (eds Gibbons, W. W. & Moreno, T.), pp. 3149. London: Geological Society of London.10.1144/GOSPP.4Google Scholar
Halpin, J. A., Daczko, N. R., Milan, L. A. & Clarke, G. L. 2012. Decoding near-concordant U–Pb zircon ages spanning several hundred million years: recrystallisation, metamictisation or diffusion? Contributions to Mineralogy and Petrology 163, 6785.10.1007/s00410-011-0659-7Google Scholar
Hammann, W. 1992. The Ordovician trilobites from the Iberian chains in the province of Aragon, NE-Spain. 1. The trilobites of the Cystoid Limestone (Ashgill series). Beringeria 6, 1219.Google Scholar
Handy, M. R., Ustaszewski, K. & Kissling, E. 2015. Reconstructing the Alps–Carpathians–Dinarides as a key to understanding switches in subduction polarity, slab gaps and surface motion. International Journal of Earth Sciences 104, 126.Google Scholar
Hawkesworth, C., Cawood, P. & Dhuime, B. 2013. Continental growth and the crustal record. Tectonophysics 609, 651–60.10.1016/j.tecto.2013.08.013Google Scholar
Heinrichs, T., Siegesmund, S., Frei, D., Drobe, M. & Schulz, B. 2012. Provenance signatures from whole-rock geochemistry and detrital zircon ages of metasediments from the Austroalpine basement south of the Tauern window (eastern Tyrol, Austria). GeoAlp 9, 156–85.Google Scholar
Henderson, B. J., Collins, W. J., Murphy, J. B., Gutierrez-Alonso, G. & Hand, M. 2016. Gondwanan basement terranes of the Variscan–Appalachian orogen: Baltican, Saharan and West African hafnium isotopic fingerprints in Avalonia, Iberia and the Armorican Terranes. Tectonophysics 681, 278304.Google Scholar
Hietpas, J., Samson, S., Moecher, D. & Chakraborty, S. 2011. Enhancing tectonic and provenance information from detrital zircon studies: assessing terrane-scale sampling and grain-scale characterization. Journal of the Geological Society, London 168, 309–18.10.1144/0016-76492009-163Google Scholar
Hofmann, M., Linnemann, U., Rai, V., Becker, S., Gärtner, A. & Sagawe, A. 2011. The India and South China cratons at the margin of Rodinia — synchronous Neoproterozoic magmatism revealed by LA-ICP-MS zircon analyses. Lithos 123, 176–87.10.1016/j.lithos.2011.01.012Google Scholar
Horbe, A. M. C., Motta, M. B., de Almeida, C. M., Dantas, E. L. & Vieira, L. C. 2013. Provenance of Pliocene and recent sedimentary deposits in western Amazônia, Brazil: consequences for the paleodrainage of the Solimões-Amazonas River. Sedimentary Geology 296, 920.Google Scholar
Javier Álvaro, J., Colmenar, J., Monceret, E., Pouclet, A. & Vizcaïno, D. 2016. Late Ordovician (post-Sardic) rifting branches in the North Gondwanan Montagne Noire and Mouthoumet massifs of southern France. Tectonophysics 681, 111–23.10.1016/j.tecto.2015.11.031Google Scholar
Jimenez Millan, J. & Velilla, N. 1998. Mn-Fe spinels and silicates in manganese-rich rocks from the Ossa-Morena Zone, southern Iberian Massif, southwestern Spain. The Canadian Mineralogist 36, 701–11.Google Scholar
Johnson, S. C. & McLeod, M. J., 1996. The New River Belt: a unique segment along the western margin of the Avalon composite terrane, southern New Brunswick, Canada. In Avalonian and Related Peri-Gondwana Terranes of the Circum-North Atlantic (eds Nance, R. D. & Thompson, M. D.), pp. 149–64. Geological Society of America Special Papers no. 304.10.1130/0-8137-2304-3.149Google Scholar
Julivert, M., Fontbote, J. M., Ribeiro, A. & Conde, L. N. 1972. Mapa Tectónico de la Península Ibérica y Baleares. Madrid: Instituto Geológico y Minero de España.Google Scholar
Kelly, N. M., Clarke, G. L. & Fanning, C. M. 2002. A two-stage evolution of the Neoproterozoic Rayner Structural Episode: new U–Pb sensitive high resolution ion microprobe constraints from the Oygarden Group, Kemp Land, East Antarctica. Precambrian Research 116, 307–30.10.1016/S0301-9268(02)00028-1Google Scholar
Kelly, N. M., Clarke, G. L. & Fanning, C. M. 2004. Archaean crust in the Rayner Complex of east Antarctica: Oygarden Group of islands, Kemp Land. Transactions of the Royal Society of Edinburgh: Earth Sciences 95, 491510.Google Scholar
Kennan, P. S. & Morris, J. H. 1999. Manganiferous ironstones in the early Ordovician Manx Group, Isle of Man: a protolith of coticule? In In Sight of the Suture: The Palaeozoic Geology of the Isle of Man in its Iapetus Ocean Context (eds Woodcock, N. H., Quirk, D. G., Fitches, W. R. & Barnes, R. P.), pp. 109–19. Geological Society of London, Special Publication no. 160.Google Scholar
Keppie, J. D. & Krogh, T. E. 2000. 440 Ma igneous activity in the Meguma Terrane, Nova Scotia, Canada; part of the Appalachian overstep sequence? American Journal of Science 300, 528–38.Google Scholar
Kolodner, K., Avigad, D., McWilliams, M., Wooden, J. L., Weissbrod, T. & Feinstein, S. 2006. Provenance of north Gondwana Cambrian–Ordovician sandstone: U–Pb SHRIMP dating of detrital zircons from Israel and Jordan. Geological Magazine 143, 367–91.Google Scholar
Košler, J., Konopásek, J., Sláma, J. & Vrána, S. 2014. U–Pb zircon provenance of Moldanubian metasediments in the Bohemian Massif. Journal of the Geological Society, London 171, 8395.Google Scholar
Kossmat, F. 1927. Gliederung des varistischen Gebirgsbaues. Abhandlungen des Sächsischen Geologischen Landesamts 1, 139.Google Scholar
Kramm, U. 1976. The coticule rocks (spessartine quartzites) of the Venn-Stavelot Massif, Ardennes, a volcanoclastic metasediment? Contributions to Mineralogy and Petrology 56, 135–55.10.1007/BF00399600Google Scholar
Krawczyk, C. M., McCann, T., Cocks, L. R. M., England, R. W., McBride, J. H. & Wybraniez, S. 2008. Caledonian tectonics. In The Geology of Central Europe. Volume 1: Precambrian and Palaeozoic (ed. McCann, T.), pp. 3034381. London: The Geological Society.Google Scholar
Kreuzer, H., Seidel, E., Schüssler, U., Okrusch, M., Lenz, K.-L. & Raschka, H. 1989. K–Ar geochronology of different tectonic units at the northwestern margin of the Bohemian Massif. Tectonophysics 157, 149–78.Google Scholar
Kroner, U., Hahn, T., Romer, R. L. & Linnemann, U. 2007. The Variscan orogeny in the Saxo-Thuringian zone—heterogenous overprint of Cadomian/Paleozoic Peri-Gondwana crust. In The Evolution of the Rheic Ocean: From Avalonian–Cadomian Active Margin to Alleghenian–Variscan Collision (eds Linnemann, U., Nance, R. D., Kraft, P. & Zulauf, G.), pp. 153–72. Geological Society of America, Special Papers no. 423.Google Scholar
Kroner, U. & Romer, R. L. 2010. The Saxo-Thuringian Zone–tip of the Armorican spur and part of the Gondwana plate. In Pre-Mesozoic Geology of Saxo-Thuringia–From the Cadomian Active Margin to the Variscan Orogen (eds Linnemann, U. & Romer, R. L.), pp. 371–94. Stuttgart: Schweizerbart.Google Scholar
Kroner, U. & Romer, R. L. 2013. Two plates – many subduction zones: the Variscan orogeny reconsidered. Gondwana Research 24, 298329.10.1016/j.gr.2013.03.001Google Scholar
Kroner, U., Roscher, M. & Romer, R. L. 2016. Ancient plate kinematics derived from the deformation pattern of continental crust: Paleo- and Neo-Tethys opening coeval with prolonged Gondwana–Laurussia convergence. Tectonophysics 681, 220–33.Google Scholar
Kruskal, J. B. 1964. Multidimensional scaling by optimizing goodness of fit to a nonmetric hypothesis. Psychometrika 29, 127.Google Scholar
Kruskal, J. B. & Wish, M. 1978. Multidimensional Scaling. Newbury Park: Sage Publications.10.4135/9781412985130Google Scholar
Kryza, R. & Fanning, C. M. 2007. Devonian deep-crustal metamorphism and exhumation in the Variscan Orogen: evidence from SHRIMP zircon ages from the HT-HP granulites and migmatites of the Góry Sowie (Polish Sudetes). Geodinamica Acta 20, 159–75.Google Scholar
Küster, D., Romer, R. L., Tolessa, D., Zerihun, D., Bheemalingeswara, K., Melcher, F. & Oberthür, T. 2009. The Kenticha rare-element pegmatite, Ethiopia: internal differentiation, U–Pb age and Ta mineralization. Mineralium Deposita 44, 723–50.10.1007/s00126-009-0240-8Google Scholar
Langone, A., Braga, R., Massonne, H.-J. & Tiepolo, M. 2011. Preservation of old (prograde metamorphic) U–Th–Pb ages in unshielded monazite from the high-pressure paragneisses of the Variscan Ulten Zone (Italy). Lithos 127, 6885.Google Scholar
Lardeaux, J. M., Ledru, P., Daniel, I. & Duchene, S. 2001. The Variscan French Massif Central—a new addition to the ultra-high pressure metamorphic ‘club’: exhumation processes and geodynamic consequences. Tectonophysics 332, 143–67.Google Scholar
Ledent, D., Patterson, C. & Tilton, G. R. 1964. Ages of zircon and feldspar concentrates from North American beach and river sands. The Journal of Geology 72, 112–22.10.1086/626967Google Scholar
Le Heron, D. P., Ghienne, J.-F., El Houicha, M., Khoukhi, Y. & Rubino, J.-L. 2007. Maximum extent of ice sheets in Morocco during the Late Ordovician glaciation. Palaeogeography, Palaeoclimatology, Palaeoecology 245, 200–26.Google Scholar
Linan, E., Gozalo, R., Palacios, T., Vintaned, J. A. G., Ugidos, J. M. & Mayoral, E. 2002. Cambrian. In The Geology of Spain (eds Gibbons, W. & Moreno, T.), pp. 1730. London: Geological Society of London.Google Scholar
Linnemann, U., D'Lemos, R., Drost, K., Jeffries, T., Gerdes, A., Romer, R. L., Samson, S. D. & Strachan, R. A. 2008a. Cadomian tectonics. In The Geology of Central Europe. Volume 1: Precambrian and Palaeozoic (ed. McCann, T.), pp. 103–54. London: The Geological Society.Google Scholar
Linnemann, U., Gehmlich, M., Tichomirowa, M., Buschmann, B., Nasdala, L., Jonas, P., Lützner, H. & Bombach, K. 2000. From Cadomian subduction to Early Palaeozoic rifting: the evolution of Saxo-Thuringia at the margin of Gondwana in the light of single zircon geochronology and basin development (Central European Variscides, Germany). In Orogenic Processes: Quantification and Modelling in the Variscan Belt (eds. Franke, W., Altherr, R., Haak, V., Ocncken, O. & Tanner, D.), pp. 131–53. Geological Society of London: Special Publication no. 179.Google Scholar
Linnemann, U., Gerdes, A., Hofmann, M. & Marko, L. 2014. The Cadomian Orogen: Neoproterozoic to Early Cambrian crustal growth and orogenic zoning along the periphery of the West African Craton—constraints from U–Pb zircon ages and Hf isotopes (Schwarzburg Antiform, Germany). Precambrian Research 244, 236–78.Google Scholar
Linnemann, U., Herbosch, A., Liégeois, J.-P., Pin, C., Gärtner, A. & Hofmann, M. 2012. The Cambrian to Devonian odyssey of the Brabant Massif within Avalonia: a review with new zircon ages, geochemistry, Sm–Nd isotopes, stratigraphy and palaeogeography. Earth-Science Reviews 112, 126–54.Google Scholar
Linnemann, U. & Heuse, T. 2000. The Ordovician of the Schwarzburg Anticline: geotectonic setting, biostratigraphy and sequence stratigraphy (Saxo-Thuringian Terrane, Germany). Zeitschrift der deutschen Geologischen Gesellschaft 151, 471–91.Google Scholar
Linnemann, U., Hofmann, M., Romer, R. L. & Gerdes, A. 2010a. Transitional stages between the Cadomian and Variscan orogenies: basin development and tectonomagmatic evolution of the southern margin of the Rheic Ocean in the Saxo-Thuringian Zone (North Gondwana shelf). In Pre-Mesozoic Geology of Saxo-Thuringia — from the Cadomian Active Margin to the Variscan Orogen (ed. Linnemann, U. & Romer, R. L.), pp. 5998. Stuttgart: Schweizerbart.Google Scholar
Linnemann, U., McNaughton, N. J., Romer, R. L., Gehmlich, M., Drost, K. & Tonk, C. 2004. West African provenance for Saxo-Thuringia (Bohemian Massif): did Armorica ever leave pre-Pangean Gondwana? – U/Pb-SHRIMP zircon evidence and the Nd-isotopic record. International Journal of Earth Sciences 93, 683705.Google Scholar
Linnemann, U., Ouzegane, K., Drareni, A., Hofmann, M., Becker, S., Gärtner, A. & Sagawe, A. 2011. Sands of West Gondwana: an archive of secular magmatism and plate interactions—a case study from the Cambro-Ordovician section of the Tassili Ouan Ahaggar (Algerian Sahara) using U–Pb–LA-ICP-MS detrital zircon ages. Lithos 123, 188203.Google Scholar
Linnemann, U., Pereira, F., Jeffries, T. E., Drost, K. & Gerdes, A. 2008b. The Cadomian Orogeny and the opening of the Rheic Ocean: the diacrony of geotectonic processes constrained by LA-ICP-MS U–Pb zircon dating (Ossa-Morena and Saxo-Thuringian Zones, Iberian and Bohemian Massifs). Tectonophysics 461, 2143.10.1016/j.tecto.2008.05.002Google Scholar
Linnemann, U., Romer, R. L., Gerdes, A., Jeffries, T. E., Drost, K. & Ulrich, J. 2010b. The Cadomian orogeny in the Saxo-Thuringian zone. In Pre-Mesozoic Geology of Saxo-Thuringia–From the Cadomian Active Margin to the Variscan Orogen (eds Linneman, U. & Romer, R. L.), pp. 3758. Stuttgart: Schweizerbart.Google Scholar
Lotze, F. 1945. Zur Gliederung der Varisziden in der Iberischen Meseta. Geotektonische Forschungen 6.Google Scholar
Lotze, F. 1956. Das Präkambrium Spaniens. Neues Jahrbuch für Geologie und Paläontologie, Monatschefte 8, 373–80.Google Scholar
Lucks, H., Schulz, B., Audren, C. & Triboulet, C. 2002. Variscan pressure-temperature evolution of garnet pyroxenites and amphibolites in the Baie d'Audierne metamorphic series, Brittany (France). In Variscan–Appalachian Dynamics: The Building of the Late Paleozoic Basement (eds Catalán, J. R. M., Hatcher, J. R. D., Arenas, R. & García, F. D.), pp. 89103. Geological Society of America, Special Papers no. 364.Google Scholar
MacDonald, L. A., Barr, S. M., White, C. E. & Ketchum, J. W. F. 2002. Petrology, age, and tectonic setting of the White Rock Formation, Meguma terrane, Nova Scotia: evidence for Silurian continental rifting. Canadian Journal of Earth Sciences 39, 259–77.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
Martínez Catalán, J. R., Rubio Pascual, F. J., Montes, A. D., Fernández, R. D., Barreiro, J. G., Dias Da Silva, Í., Clavijo, E. G., Ayarza, P. & Alcock, J. E. 2014. The late Variscan HT/LP metamorphic event in NW and Central Iberia: relationships to crustal thickening, extension, orocline development and crustal evolution. In Geomechanics and Geology (eds. Turner, J. P., Healy, D., Hillis, R. R. & Welch, M. J.), pp. 225–47. Geological Society of London, Special Publication no. 405.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
Matteini, M., Dantas, E. L., Pimentel, M. M., de Alvarenga, C. J. S. & Dardenne, M. A. 2012. U–Pb and Hf isotope study on detrital zircons from the Paranoá Group, Brasília Belt Brazil: constraints on depositional age at Mesoproterozoic – Neoproterozoic transition and tectono-magmatic events in the São Francisco craton. Precambrian Research 206–207, 168–81.Google Scholar
Mattinson, J. M. 1987. U–Pb ages of zircons: a basic examination of error propagation. Chemical Geology: Isotope Geoscience Section 66, 151–62.Google Scholar
Mazur, S., Szczepański, J., Turniak, K. & McNaughton, N. J. 2012. Location of the Rheic suture in the eastern Bohemian Massif: evidence from detrital zircon data. Terra Nova 24, 199206.Google Scholar
McLennan, S. M., Hemming, S., McDaniel, D. K. & Hanson, G. N. 1993. Geochemical approaches to sedimentation, provenance, and tectonics. In Processes Controlling the Composition of Clastic Sediments (eds Johnsson, M. J. & Basu, A.), pp. 2140. Geological Society of America, Special Papers no. 284.10.1130/SPE284-p21Google Scholar
Meinhold, G., Kostopoulos, D., Frei, D., Himmerkus, F. & Reischmann, T. 2010. U–Pb LA-SF-ICP-MS zircon geochronology of the Serbo-Macedonian Massif, Greece: palaeotectonic constraints for Gondwana-derived terranes in the Eastern Mediterranean. International Journal of Earth Sciences 99, 813–32.10.1007/s00531-009-0425-5Google Scholar
Meinhold, G., Morton, A. C. & Avigad, D. 2013. New insights into peri-Gondwana paleogeography and the Gondwana super-fan system from detrital zircon U–Pb ages. Gondwana Research 23, 661–5.Google Scholar
Meinhold, G., Morton, A. C., Fanning, C. M., Frei, D., Howard, J. P., Phillips, R. J., Strogen, D. & Whitham, A. G. 2011. Evidence from detrital zircons for recycling of Mesoproterozoic and Neoproterozoic crust recorded in Paleozoic and Mesozoic sandstones of southern Libya. Earth and Planetary Science Letters 312, 164–75.10.1016/j.epsl.2011.09.056Google Scholar
Mélou, M., Oulebsir, L. & Paris, F. 1999. Brachiopodes et chitinozoaires ordoviciens dans le NE du Sahara algérien: implications stratigraphiques et paléogéographiques. Geobios 32, 822–39.10.1016/S0016-6995(99)80865-1Google Scholar
Mikhalsky, E. V., Beliatsky, B. V., Sheraton, J. W. & Roland, N. W. 2006. Two distinct Precambrian terranes in the Southern Prince Charles Mountains, East Antarctica: SHRIMP dating and geochemical constraints. Gondwana Research 9, 291309.Google Scholar
Moecher, D. P. & Samson, S. D. 2006. Differential zircon fertility of source terranes and natural bias in the detrital zircon record: implications for sedimentary provenance analysis. Earth and Planetary Science Letters 247, 252–66.Google Scholar
Moghadam, H. S., Li, X.-H., Griffin, W. L., Stern, R. J., Thomsen, T. B., Meinhold, G., Aharipour, R. & O'Reilly, S. Y. 2017. Early Paleozoic tectonic reconstruction of Iran: tales from detrital zircon geochronology. Lithos 268–271, 87101.Google Scholar
Molzahn, M., Anthes, G. & Reischmann, T. 1998. Single zircon Pb/Pb age geochronology and isotope systematics of the Rhenohercynian basement. Terra Nostra 98, 67–8.Google Scholar
Monod, O., Kozlu, H., Ghienne, J. F., Dean, W. T., Günay, Y., Hérissé, A. L., Paris, F. & Robardet, M. 2003. Late Ordovician glaciation in southern Turkey. Terra Nova 15, 249–57.10.1046/j.1365-3121.2003.00495.xGoogle Scholar
Morton, A. C. & Hallsworth, C. 1994. Identifying provenance-specific features of detrital heavy mineral assemblages in sandstones. Sedimentary Geology 90, 241–56.10.1016/0037-0738(94)90041-8Google Scholar
Moryc, W. & Łydka, K. 2012. Sedimentation and tectonics of the Upper Proterozoic-Lower Cambrian deposits of the southern Małopolska Massif (SE Poland). Geological Quarterly 44, 1247.Google Scholar
Murphy, J. B., Fernández-Suárez, J., Keppie, J. D. & Jeffries, T. E. 2004a. Contiguous rather than discrete Paleozoic histories for the Avalon and Meguma terranes based on detrital zircon data. Geology 32, 585–8.10.1130/G20351.1Google Scholar
Murphy, J. B., Pisarevsky, S. A., Nance, R. D. & Keppie, J. D. 2004b. Neoproterozoic—Early Paleozoic evolution of peri-Gondwanan terranes: implications for Laurentia-Gondwana connections. International Journal of Earth Sciences 93, 659–82.Google Scholar
Murphy, J. B., Strachan, R. A., Nance, R. D., Parker, K. D. & Fowler, M. B. 2000. Proto-Avalonia: a 1.2–1.0 Ga tectonothermal event and constraints for the evolution of Rodinia. Geology 28, 1071–4.Google Scholar
Nägler, T. F., Schafer, H.-J. & Gebauer, D. 1995. Evolution of the Western European continental crust: implications from Nd and Pb isotopes in Iberian sediments. Chemical Geology 121, 345–57.Google Scholar
Nance, R. D., Gutiérrez-Alonso, G., Keppie, J. D., Linnemann, U., Murphy, J. B., Quesada, C., Strachan, R. A. & Woodcock, N. H. 2010. Evolution of the Rheic Ocean. Gondwana Research 17, 194222.Google Scholar
Nance, R. D. & Linnemann, U. 2008. The Rheic Ocean: origin, evolution, and significance. GSA Today 18, 412.10.1130/GSATG24A.1Google Scholar
Nie, J., Peng, W., Möller, A., Song, Y., Stockli, D. F., Stevens, T., Horton, B. K., Liu, S., Bird, A., Oalmann, J., Gong, H. & Fang, X. 2014. Provenance of the upper Miocene–Pliocene Red Clay deposits of the Chinese loess plateau. Earth and Planetary Science Letters 407, 3547.Google Scholar
Noblet, C. & Lefort, J. P. 1990. Sedimentological evidence for a limited separation between Armorica and Gondwana during the Early Ordovician. Geology 18, 303–6.Google Scholar
Nutman, A. P., Green, D. H., Cook, C. A., Styles, M. T. & Holdsworth, R. E. 2001. SHRIMP U–Pb zircon dating of the exhumation of the Lizard Peridotite and its emplacement over crustal rocks: constraints for tectonic models. Journal of the Geological Society, London 158, 809–20.Google Scholar
O'Brien, S. J., Wardle, R. J. & King, A. F. 1983. The Avalon Zone: a Pan-African terrane in the Appalachian Orogen of Canada. Geological Journal 18, 195222.Google Scholar
Oczlon, M. S., Seghedi, A. & Carrigan, C. W. 2007. Avalonian and Baltican terranes in the Moesian Platform (southern Europe, Romania, and Bulgaria) in the context of Caledonian terranes along the southwestern margin of the East European craton. In The Evolution of the Rheic Ocean: From Avalonian-Cadomian Active Margin to Alleghenian-Variscan Collision (eds Linneman, U., Nance, R. D., Kraft, P. & Zulauf, G.), pp. 375400. Geological Society of America, Special Papers no. 423.Google Scholar
Ordóñez Casado, B., Gebauer, D., Schäfer, H. J., Gil Ibarguchi, J. I. & Peucat, J. J. 2001. A single Devonian subduction event for the HP/HT metamorphism of the Cabo Ortegal complex within the Iberian Massif. Tectonophysics 332, 359–85.Google Scholar
Orejana, D., Merino Martínez, E., Villaseca, C. & Andersen, T. 2015. Ediacaran–Cambrian paleogeography and geodynamic setting of the Central Iberian Zone: constraints from coupled U–Pb–Hf isotopes of detrital zircons. Precambrian Research 261, 234–51.10.1016/j.precamres.2015.02.009Google Scholar
Paris, F. & Robardet, M. 1990. Early Palaeozoic palaeobiogeography of the Variscan regions. Tectonophysics 177, 193213.Google Scholar
Pereira, M. F. 2015. Potential sources of Ediacaran strata of Iberia: a review. Geodinamica Acta 27, 114.Google Scholar
Pereira, M. F., Chichorro, M., Linnemann, U., Eguiluz, L. & Silva, J. B. 2006. Inherited arc signature in Ediacaran and Early Cambrian basins of the Ossa-Morena Zone (Iberian Massif, Portugal): paleogeographic link with European and North African Cadomian correlatives. Precambrian Research 144, 297315.Google Scholar
Pérez-Cáceres, I., Martínez Poyatos, D., Simancas, J. F. & Azor, A. 2017. Testing the Avalonian affinity of the South Portuguese Zone and the Neoproterozoic evolution of SW Iberia through detrital zircon populations. Gondwana Research 42, 177–92.10.1016/j.gr.2016.10.010Google Scholar
Perroud, H. & Bonhommet, N. 1984. A Devonian palaeomagnetic pole for Armorica. Geophysical Journal of the Royal Astronomical Society 77, 839–45.Google Scholar
Perroud, H., Calza, F. & Khattach, D. 1991. Paleomagnetism of the Silurian volcanism at Almaden, southern Spain. Journal of Geophysical Research: Solid Earth 96, 1949–62.10.1029/90JB02226Google Scholar
Pettersson, C. H., Pease, V. & Frei, D. 2010. Detrital zircon U–Pb ages of Silurian–Devonian sediments from NW Svalbard: a fragment of Avalonia and Laurentia? Journal of the Geological Society, London 167, 1019–32.10.1144/0016-76492010-062Google Scholar
Pharaoh, T. C. 1999. Palaeozoic terranes and their lithospheric boundaries within the Trans-European Suture Zone (TESZ): a review. Tectonophysics 314, 1741.Google Scholar
Pharaoh, T. C. & Carney, J. N. 2000. Introduction. In Precambrian Rocks of England and Wales (eds Carney, J., Horak, J. M., Pharao, T. C., Gibbons, W., Wilson, D., Barclay, W. J., Bevins, R. E., Cope, J. C. W. & Ford, T. D.), pp. 1–18. Geological Conservation Review Series no. 20.Google Scholar
Pique, A. & Michard, A. 1989. Moroccan Hercynides; a synopsis; the Paleozoic sedimentary and tectonic evolution at the northern margin of West Africa. American Journal of Science 289, 286330.Google Scholar
Poli, M. E. & Zanferrari, A. 1992. The Agordo basement (NE Italy): a 500 Ma-long geological record in the Southalpine crust. In Contribution to the Geology of Italy with Special Regard to the Paleozoic Basements: A Volume Dedicated to Tommaso Cocozza (eds Carmignani, L. & Sassi, F. P.), pp. 283–96. IGCP Project 276, Newsletter no. 5.Google Scholar
Pollock, J. C., Hibbard, J. P. & van Staal, C. R. 2011. A paleogeographical review of the peri-Gondwanan realm of the Appalachian orogen. Canadian Journal of Earth Sciences 49, 259–88.Google Scholar
Pullen, A., Ibáñez-Mejía, M., Gehrels, G. E., Ibáñez-Mejía, J. C. & Pecha, M. 2014. What happens when n = 1000? Creating large-n geochronological datasets with LA-ICP-MS for geologic investigations. Journal of Analytical Atomic Spectrometry 29, 971–80.10.1039/C4JA00024BGoogle Scholar
Razali, N. M. & Wah, Y. B. 2011. Power comparisons of Shapiro-Wilk, Kolmogorov-Smirnov, Lilliefors and Anderson-Darling tests Journal of Statistical Modeling and Analytics 2, 2133.Google Scholar
Rittner, M., Vermeesch, P., Carter, A., Bird, A., Stevens, T., Garzanti, E., Andò, S., Vezzoli, G., Dutt, R., Xu, Z. & Lu, H. 2016. The provenance of Taklamakan desert sand. Earth and Planetary Science Letters 437, 127–37.Google Scholar
Robardet, M. 2002. Alternative approach to the Variscan Belt in southwestern Europe: preorogenic paleobiogeographical constraints. In Variscan–Appalachian Dynamics: The Building of the Late Paleozoic Basement (eds Catalán, J. R. M., Hatcher, J. R. D., Arenas, R. & García, F. D.), pp. 115. Geological Society of America, Special Papers no. 364.Google Scholar
Robardet, M. & Doré, F. 1988. The late Ordovician diamictic formations from southwestern Europe: North-Gondwana glaciomarine deposits. Palaeogeography, Palaeoclimatology, Palaeoecology 66, 1931.10.1016/0031-0182(88)90078-8Google Scholar
Robardet, M. & Gutiérrez-Marco, J. C. 2002. Silurian. In The Geology of Spain (eds Gibbons, W. W. & Moreno, T.), pp. 5166. London: Geological Society of London.Google Scholar
Robardet, M., Verniers, J., Feist, R. & Paris, F. 1994. The pre-Variscan Palaeozoic successions in France, palaeogeographic and geodynamic setting. Geologie de la France 3, 331.Google Scholar
Rogers, N., van Staal, C. R., McNicoll, V., Pollock, J., Zagorevski, A. & Whalen, J. 2006. Neoproterozoic and Cambrian arc magmatism along the eastern margin of the Victoria Lake Supergroup: a remnant of Ganderian basement in central Newfoundland? Precambrian Research 147, 320–41.Google Scholar
Rolet, J., Gresselin, F., Jegouzo, P., Ledru, P. & Wyns, R. 1994. Intracontinental Hercynian Events in the Armorican Massif. In Pre-Mesozoic Geology in France and Related Areas (eds Chantraine, J., Rolet, J., Santallier, D. S., Piqué, A. & Keppie, J. D.), pp. 195219. Berlin, Heidelberg: Springer.Google Scholar
Romer, R. L. & Hahne, K. 2010. Life of the Rheic Ocean: scrolling through the shale record. Gondwana Research 17, 236–53.Google Scholar
Romer, R. L., Kirsch, M. & Kroner, U. 2011. Geochemical signature of Ordovician Mn-rich sedimentary rocks on the Avalonian shelf. Canadian Journal of Earth Sciences 48, 703–18.Google Scholar
Romer, R. L. & Kroner, U. 2015. Sediment and weathering control on the distribution of Paleozoic magmatic tin–tungsten mineralization. Mineralium Deposita 50, 327–38.Google Scholar
Romer, R. L. & Kroner, U. 2016. Phanerozoic tin and tungsten mineralization – tectonic controls on the distribution of enriched protoliths and heat sources for crustal melting. Gondwana Research 31, 6095.Google Scholar
Rösel, D., Boger, S. D., Möller, A., Gaitzsch, B., Barth, M., Oalmann, J. & Zack, T. 2014. Indo-Antarctic derived detritus on the northern margin of Gondwana: evidence for continental-scale sediment transport. Terra Nova 26, 6471.Google Scholar
Rossi, P., Oggiano, G. & Cocherie, A. 2009. A restored section of the “southern Variscan realm” across the Corsica–Sardinia microcontinent. Comptes Rendus Geoscience 341, 224–38.10.1016/j.crte.2008.12.005Google Scholar
Rousseeuw, P. J. 1987. Silhouettes: a graphical aid to the interpretation and validation of cluster analysis. Journal of Computational and Applied Mathematics 20, 5365.Google Scholar
Rubatto, D., Ferrando, S., Compagnoni, R. & Lombardo, B. 2010. Carboniferous high-pressure metamorphism of Ordovician protoliths in the Argentera Massif (Italy), southern European Variscan belt. Lithos 116, 6576.10.1016/j.lithos.2009.12.013Google Scholar
Satkoski, A. M., Barr, S. M. & Samson, S. D. 2010. Provenance of Late Neoproterozoic and Cambrian sediments in Avalonia: constraints from detrital zircon ages and Sm–Nd isotopic compositions in Southern New Brunswick, Canada. The Journal of Geology 118, 187200.Google Scholar
Schaltegger, U., Schneider, J.-L., Maurin, J.-C. & Corfu, F. 1996. Precise UPb chronometry of 345–340 Ma old magmatism related to syn-convergence extension in the Southern Vosges (Central Variscan Belt). Earth and Planetary Science Letters 144, 403–19.Google Scholar
Schmid, S. M., Fügenschuh, B., Kissling, E. & Schuster, R. 2004. Tectonic map and overall architecture of the Alpine orogen. Eclogae Geologicae Helvetiae 97, 93117.Google Scholar
Schulz, K. J., Stewart, D. B., Tucker, R. D., Pollock, J. C. & Ayuso, R. A. 2008. The Ellsworth terrane, coastal Maine: geochronology, geochemistry, and Nd–Pb isotopic composition—implications for the rifting of Ganderia. Geological Society of America Bulletin 120, 1134–58.Google Scholar
Scotese, C. R., Boucot, A. J. & McKerrow, W. S. 1999. Gondwanan palaeogeography and palaeoclimatology. Journal of African Earth Sciences 28, 99114.Google Scholar
Servais, T., Dzik, J., Fatka, O., Heuse, T., Vecoli, M. & Verniers, J. 2008. Ordovician. In The Geology of Central Europe. Volume 1: Precambrian and Palaeozoic (ed. McCann, T.), pp. 203–48. London: The Geological Society.Google Scholar
Sharland, P. R., Archer, R., Casey, D. M., Davies, R. B., Hall, S. H., Heward, A. P., Horbury, A. D. & Simmons, M. D. 2001. The Chrono-sequence Stratigraphy of the Arabian Plate. GeoArabia Special Publication no. 2.Google Scholar
Sharman, G. R. & Johnstone, S. A. 2017. Sediment unmixing using detrital geochronology. Earth and Planetary Science Letters 477, 183–94.Google Scholar
Shaw, J., Gutiérrez-Alonso, G., Johnston, S. T. & Pastor Galán, D. 2014. Provenance variability along the Early Ordovician north Gondwana margin: paleogeographic and tectonic implications of U–Pb detrital zircon ages from the Armorican Quartzite of the Iberian Variscan belt. Geological Society of America Bulletin 126, 702–19.Google Scholar
Shu, L. S., Deng, X. L., Zhu, W. B., Ma, D. S. & Xiao, W. J. 2011. Precambrian tectonic evolution of the Tarim Block, NW China: new geochronological insights from the Quruqtagh domain. Journal of Asian Earth Sciences 42, 774–90.Google Scholar
Sircombe, K. N. 2000. Quantitative comparison of large sets of geochronological data using multivariate analysis: a provenance study example from Australia. Geochimica et Cosmochimica Acta 64, 1593–616.Google Scholar
Slama, J. 2016. Rare late Neoproterozoic detritus in SW Scandinavia as a response to distant tectonic processes. Terra Nova 28, 394401.Google Scholar
Sláma, J. & Košler, J. 2012. Effects of sampling and mineral separation on accuracy of detrital zircon studies. Geochemistry, Geophysics, Geosystems 13.Google Scholar
Squire, R. J., Campbell, I. H., Allen, C. M. & Wilson, C. J. L. 2006. Did the Transgondwanan Supermountain trigger the explosive radiation of animals on Earth? Earth and Planetary Science Letters 250, 116–33.10.1016/j.epsl.2006.07.032Google Scholar
Stampfli, G. M. & Borel, G. D. 2002. A plate tectonic model for the Paleozoic and Mesozoic constrained by dynamic plate boundaries and restored synthetic oceanic isochrons. Earth and Planetary Science Letters 196, 1733.10.1016/S0012-821X(01)00588-XGoogle Scholar
Stampfli, G. M., Hochard, C., Vérard, C., Wilhem, C. & von Raumer, J. 2013. The formation of Pangea. Tectonophysics 593, 119.Google Scholar
Stampfli, G. M., von Raumer, J. F. & Borel, G. D. 2002. Paleozoic evolution of pre-Variscan terranes: from Gondwana to the Variscan collision. In Variscan–Appalachian Dynamics: The Building of the Late Paleozoic Basement (eds Catalán, J. R. M., Hatcher, J. R. D., Arenas, R. & García, F. D.), pp. 263–80. Geological Society of America, Special Papers no. 364.Google Scholar
Stephan, T., Kroner, U., Hahn, T., Hallas, P. & Heuse, T. 2016. Fold/cleavage relationships as indicator for late Variscan sinistral transpression at the Rheno-Hercynian–Saxo-Thuringian boundary zone, Central European Variscides. Tectonophysics 681, 250–62.Google Scholar
Stevens, T., Carter, A., Watson, T. P., Vermeesch, P., Andò, S., Bird, A. F., Lu, H., Garzanti, E., Cottam, M. A. & Sevastjanova, I. 2013. Genetic linkage between the Yellow River, the Mu Us desert and the Chinese Loess Plateau. Quaternary Science Reviews 78, 355–68.Google Scholar
Strachan, R. A. 2012. Late Neoproterozoic to Cambrian accretionary history of eastern Avalonia and Armorica on the Active Margin of Gondwana. In Geological History of Britain and Ireland (eds Woodcock, N. H. & Strachan, R. A.), pp. 133–49. Chichester: John Wiley & Sons, Ltd.Google Scholar
Strachan, R. A., Collins, A. S., Buchan, C., Nance, R. D., Murphy, J. B. & D'Lemos, R. S. 2007. Terrane analysis along a Neoproterozoic active margin of Gondwana: insights from U–Pb zircon geochronology. Journal of the Geological Society, London 164, 5760.Google Scholar
Strachan, R. A., Linnemann, U., Jeffries, T., Drost, K. & Ulrich, J. 2014. Armorican provenance for the mélange deposits below the Lizard ophiolite (Cornwall, UK): evidence for Devonian obduction of Cadomian and Lower Palaeozoic crust onto the southern margin of Avalonia. International Journal of Earth Sciences 103, 1359–83.Google Scholar
Tait, J. A., Bachtadse, V., Franke, W. & Soffel, H. C. 1997. Geodynamic evolution of the European Variscan fold belt: palaeomagnetic and geological constraints. Geologische Rundschau 86, 585–98.10.1007/s005310050165Google Scholar
Talavera, C., Montero, P., Martínez Poyatos, D. & Williams, I. S. 2012. Ediacaran to Lower Ordovician age for rocks ascribed to the Schist–Graywacke Complex (Iberian Massif, Spain): evidence from detrital zircon SHRIMP U–Pb geochronology. Gondwana Research 22, 928–42.Google Scholar
Tatsumoto, M. & Patterson, C. 1964. Age studies of zircon and feldspar concentrates from the Franconia Sandstone. The Journal of Geology 72, 232–42.Google Scholar
Tibshirani, R., Walther, G. & Hastie, T. 2001. Estimating the number of clusters in a data set via the gap statistic. Journal of the Royal Statistical Society: Series B (Statistical Methodology) 63, 411–23.Google Scholar
Torsvik, T. H. & Cocks, L. R. M. 2013. Gondwana from top to base in space and time. Gondwana Research 24, 9991030.Google Scholar
Torsvik, T. H. & Rehnström, E. F. 2003. The Tornquist Sea and Baltica–Avalonia docking. Tectonophysics 362, 6782.Google Scholar
Torsvik, T. H., Smethurst, M. A., Meert, J. G., Van Der Voo, R., McKerrow, W. S., Brasier, M. D., Sturt, B. A. & Walderhaug, H. J. 1996. Continental break-up and collision in the Neoproterozoic and Palaeozoic – a tale of Baltica and Laurentia. Earth-Science Reviews 40, 229–58.Google Scholar
Trombetta, A., Cirrincione, R., Corfu, F., Mazzoleni, P. & Pezzino, A. 2004. Mid-Ordovician U–Pb ages of porphyroids in the Peloritan Mountains (NE Sicily): palaeogeographical implications for the evolution of the Alboran microplate. Journal of the Geological Society, London 161, 265–76.Google Scholar
Ugidos, J. M., Sánchez-Santos, J. M., Barba, P. & Valladares, M. I. 2010. Upper Neoproterozoic series in the Central Iberian, Cantabrian and West Asturian Leonese Zones (Spain): geochemical data and statistical results as evidence for a shared homogenised source area. Precambrian Research 178, 51–8.Google Scholar
Ustaömer, P. A., Ustaömer, T., Gerdes, A. & Zulauf, G. 2011. Detrital zircon ages from a Lower Ordovician quartzite of the İstanbul exotic terrane (NW Turkey): evidence for Amazonian affinity. International Journal of Earth Sciences 100, 2341.Google Scholar
Valverde-Vaquero, P. & Dunning, G. R. 2000. New U–Pb ages for Early Ordovician magmatism in Central Spain. Journal of the Geological Society, London 157, 1526.Google Scholar
Van der Voo, R. 1988. Paleozoic paleogeography of North America, Gondwana, and intervening displaced terranes: comparisons of paleomagnetism with paleoclimatology and biogeographical patterns. Geological Society of America Bulletin 100, 311–24.Google Scholar
Van Staal, C. R., Dewey, J. F., Niocaill, C. M. & McKerrow, W. S. 1998. The Cambrian–Silurian tectonic evolution of the northern Appalachians and British Caledonides: history of a complex, west and southwest Pacific-type segment of Iapetus. In Lyell: The Past is the Key to the Present (eds Blundell, D. J. & Scott, A. C.), pp. 197242. Geological Society of London, Special Publication no. 143.Google Scholar
Veevers, J. J. 2004. Gondwanaland from 650–500 Ma assembly through 320 Ma merger in Pangea to 185–100 Ma breakup: supercontinental tectonics via stratigraphy and radiometric dating. Earth-Science Reviews 68, 1132.Google Scholar
Vermeesch, P. 2004. How many grains are needed for a provenance study? Earth and Planetary Science Letters 224, 441–51.Google Scholar
Vermeesch, P. 2013. Multi-sample comparison of detrital age distributions. Chemical Geology 341, 140–6.Google Scholar
Vermeesch, P., Resentini, A. & Garzanti, E. 2016. An R package for statistical provenance analysis. Sedimentary Geology 336, 1425.Google Scholar
Verniers, J., Maletz, J., Křiž, J., Žigaitė, Z., Paris, F., Schönlaub, H. P. & Wrona, R. 2008. Silurian. In The Geology of Central Europe. Volume 1: Precambrian and Palaeozoic (ed. McCann, T.), pp. 249302. London: The Geological Society.Google Scholar
Verniers, J., Pharaoh, T., André, L., Debacker, T. N., De Vos, W., Everaerts, M., Herbosch, A., Samuelsson, J., Sintubin, M. & Vecoli, M. 2002. The Cambrian to mid Devonian basin development and deformation history of Eastern Avalonia, east of the Midlands Microcraton: new data and a review. In Palaeozoic Amalgamation of Central Europe (eds Winchester, J. A., Pharao, T. C. & Verniers, J.), pp. 4793. Geological Society of London, Special Publication no. 201.Google Scholar
Villas, E., Vennin, E., Álvaro, J. J., Hammann, W., Herrera, Z. A. & Piovano, E. L. 2002. The late Ordovician carbonate sedimentation as a major triggering factor of the Hirnantian glaciation. Bulletin de la Societe geologique de France 173, 569–78.Google Scholar
von Raumer, J. F., Stampfli, G. M. & Bussy, F. 2003. Gondwana-derived microcontinents — the constituents of the Variscan and Alpine collisional orogens. Tectonophysics 365, 722.Google Scholar
Vorster, C., Kramers, J. A. N., Beukes, N. I. C. & Van Niekerk, H. 2015. Detrital zircon U–Pb ages of the Palaeozoic Natal Group and Msikaba Formation, Kwazulu-Natal, South Africa: provenance areas in context of Gondwana. Geological Magazine 153, 460–86.10.1017/S0016756815000370Google Scholar
Waldron, J. W. F., Schofield, D. I., White, C. E. & Barr, S. M. 2011. Cambrian successions of the Meguma Terrane, Nova Scotia, and Harlech Dome, North Wales: dispersed fragments of a peri-Gondwanan basin? Journal of the Geological Society, London 168, 8398.Google Scholar
Waldron, J. W. F., White, C. E., Barr, S. M., Simonetti, A. & Heaman, L. M. 2009. Provenance of the Meguma terrane, Nova Scotia: rifted margin of early Paleozoic Gondwana. Canadian Journal of Earth Sciences 46, 18.10.1139/E09-004Google Scholar
White, C. E., Barr, S. M., Bevier, M. L. & Kamo, S. 1994. A revised interpretation of Cambrian and Ordovician rocks in the Bourinot belt of central Cape Breton Island, Nova Scotia. Atlantic Geology 30, 123–42.Google Scholar
Williams, S. H. 1993. More Ordovician and Silurian Graptolites from the Exploits Subzone. Newfoundland Department of Mines and Energy, Geological Survey Branch, Report 93-1, pp. 311–15.Google Scholar
Willner, A. P., Barr, S. M., Gerdes, A., Massonne, H.-J. & White, C. E. 2013. Origin and evolution of Avalonia: evidence from U–Pb and Lu–Hf isotopes in zircon from the Mira terrane, Canada, and the Stavelot–Venn Massif, Belgium. Journal of the Geological Society, London 170, 769–84.Google Scholar
Yao, J., Shu, L., Santosh, M. & Li, J. 2012. Precambrian crustal evolution of the South China Block and its relation to supercontinent history: constraints from U–Pb ages, Lu–Hf isotopes and REE geochemistry of zircons from sandstones and granodiorite. Precambrian Research 208–211, 1948.Google Scholar
Young, T. P. 1990. Ordovician sedimentary facies and faunas of southwest Europe: palaeogeographic and tectonic implications. In Palaeozoic Palaeogeography and Biogeography (eds McKerrow, W. S. & Scotese, C. R.), pp. 421–30. Geological Society of London, Memoirs no. 12.Google Scholar
Zagorevski, A., Van Staal, C. R., Rogers, N., McNicoll, V. J. & Pollock, J. 2010. Middle Cambrian to Ordovician arc-backarc development on the leading edge of Ganderia, Newfoundland Appalachians. In From Rodinia to Pangea: The Lithotectonic Record of the Appalachian Region (eds Tollo, R. P., Bartholomew, M. J., Hibbard, J. P. & Karabinos, P. M.), pp. 367–96. Geological Society of America, Memoirs no. 206.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. 20 pp.Google Scholar
Zeh, A. & Gerdes, A. 2010. Baltica- and Gondwana-derived sediments in the Mid-German Crystalline Rise (Central Europe): implications for the closure of the Rheic ocean. Gondwana Research 17, 254–63.Google Scholar
Zimmermann, U., Andersen, T., Madland, M. V. & Larsen, I. S. 2015. The role of U–Pb ages of detrital zircons in sedimentology – an alarming case study for the impact of sampling for provenance interpretation. Sedimentary Geology 320, 3850.Google Scholar
Zlatkin, O., Avigad, D. & Gerdes, A. 2014. Peri-Amazonian provenance of the Proto-Pelagonian basement (Greece), from zircon U–Pb geochronology and Lu–Hf isotopic geochemistry. Lithos 184–187, 379–92.Google Scholar
Zlatkin, O., Avigad, D. & Gerdes, A. 2017. The Pelagonian terrane of Greece in the peri-Gondwanan mosaic of the Eastern Mediterranean: implications for the geological evolution of Avalonia. Precambrian Research 290, 163–83.Google Scholar
Zurbriggen, R. 2015. Ordovician orogeny in the Alps: a reappraisal. International Journal of Earth Sciences 104, 335–50.Google Scholar
Supplementary material: File

Stephan et al. supplementary material 1

Stephan et al. supplementary material

Download Stephan et al. supplementary material 1(File)
File 2.7 MB
Supplementary material: File

Stephan et al. supplementary material 2

Stephan et al. supplementary material

Download Stephan et al. supplementary material 2(File)
File 2.7 MB
Supplementary material: File

Stephan et al. supplementary material 3

Stephan et al. supplementary material

Download Stephan et al. supplementary material 3(File)
File 2.7 MB
Supplementary material: File

Stephan et al. supplementary material 4

Stephan et al. supplementary material

Download Stephan et al. supplementary material 4(File)
File 54.5 KB
Supplementary material: File

Stephan et al. supplementary material 5

Supplementary Table

Download Stephan et al. supplementary material 5(File)
File 57.5 KB
Supplementary material: File

Stephan et al. supplementary material 6

Supplementary Table

Download Stephan et al. supplementary material 6(File)
File 14.2 MB
Supplementary material: PDF

Stephan et al. supplementary material 7

Stephan et al. supplementary material

Download Stephan et al. supplementary material 7(PDF)
PDF 315.3 KB
Supplementary material: File

Stephan et al. supplementary material 8

Supplementary Table

Download Stephan et al. supplementary material 8(File)
File 31.1 KB