Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-25wd4 Total loading time: 0 Render date: 2024-04-26T17:17:29.412Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  24 November 2016

Trond H. Torsvik
Affiliation:
Universitetet i Oslo
L. Robin M. Cocks
Affiliation:
Natural History Museum, London
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2016

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

Abrajevitch, A., Van der Voo, R., Levashova, N.M. & Bazhenov, M.L. (2007). Paleomagnetic constraints on the paleogeography and oroclinal bending of the Devonian volcanic arc in Kazakhstan orocline. Tectonophysics, 441, 6784.CrossRefGoogle Scholar
Abrajevitch, A., Van der Voo, R., Bazhenov, M.L. et al. (2008). The role of the Kazakhstan orocline in the late Paleozoic amalgamation of Eurasia. Tectonophysics, 455, 6176.Google Scholar
Allen, M.B., Alsop, G.I. & Zhemchuzhnikov, V.G. (2001). Dome and basin refolding and transpressive inversion along the Karatau Fault System, southern Kazakhstan. Journal of the Geological Society, London, 158, 8395.Google Scholar
Almalki, K.A., Betts, P.G. & Ailleres, L. (2015). The Red Sea – 50 years of geological and geophysical research. Earth-Science Reviews, 147, 109140.Google Scholar
Álvaro, J.J., Elicki, O., Rushton, A.W.A. & Shergold, J.H. (2003). Palaeogeographical controls on the Cambrian immigration and evolutionary patterns reported in the western Gondwana margin. Palaeogeography, Palaeoclimatology, Palaeoecology, 195, 535.Google Scholar
Alvey, A., Gaina, C., Kusznir, N.J. & Torsvik, T.H. (2008). Integrated crustal thickness mapping and plate reconstructions for the high Arctic. Earth and Planetary Science Letters, 274, 310321.Google Scholar
Andersen, M.B., Elliott, T., Freymuth, H. et al. (2015). The terrestrial uranium isotope cycle. Nature, 517, 356359.Google Scholar
Andersen, T.B., Jamtveit, B., Dewey, J.F. & Swensson, E. (1991). Subduction and eduction of continental crust: major mechanism during continent–continent collision and orogenic extensional collapse, a model based on the south Caledonides. Terra Nova, 3, 303310.Google Scholar
Angiolini, L., Gaetani, M., Muttoni, G. et al. (2007). Tethyan oceanic currents and climate gradients 300 m.y. ago. Geology, 35, 10711074.Google Scholar
Arenas, R., Fernández, R.D., Martínez, S.S. et al. (2014). Two-stage collision: exploring the birth of Pangea in the Variscan terranes. Gondwana Research, 25, 756763.Google Scholar
Armijo, R., Lacassin, R., Coudurier-Curveur, A. & Carrizo, D. (2015). Coupled tectonic evolution of Andean orogeny and global climate. Earth-Science Reviews, 143, 135.CrossRefGoogle Scholar
Ashwal, L.D., Demaiffe, D. & Torsvik, T.H. (2002). Petrogenesis of Neoproterozoic granitoids and related rocks from the Seychelles: evidence for an Andean arc origin. Journal of Petrology, 43, 4583.Google Scholar
Assumpção, M., Feng, M. Tassara, A. & Julia, J. (2013). Models of crustal thickness for South America from seismic refraction, receiver functions and surface wave tomography. Tectonophysics, 609, 8296.CrossRefGoogle Scholar
Astashkin, V.A., Pegel, T.V., Repina, L.N. et al. (1995). The Cambrian System of the Foldbelts of Russia and Mongolia. International Union of Geological Sciences Publications, 32.Google Scholar
Aubrey, M.P., Lucas, S.G. & Berggren, W.A. (eds.) (1998). Late Paleocene–Early Eocene Climatic and Biotic Events in the Marine and Terrestrial Records. New York: Columbia University Press.Google Scholar
Baarli, G.B., Johnson, M.E. & Antoshkina, A.L. (2003). Silurian stratigraphy and paleogeography of Baltica. New York State Museum Bulletin, 493, 334,Google Scholar
Badarch, G., Cunningham, W.D., & Windley, B.F. (2002). A new terrane subdivision for Mongolia: implications for the Phanerozoic crustal growth of central Asia. Journal of Asian Earth Sciences, 21, 87110.CrossRefGoogle Scholar
Bassett, M.G. & Cocks, L.R.M. (1974). A review of Silurian brachiopods from Gotland. Fossils and Strata, 3, 156.Google Scholar
Batkhishig, B, Noriyoshi, T. & Greg, B. (2010). Magmatism of the Shuteen Complex and Carboniferous subduction of the Gurvansaihan terrane, South Mongolia. Journal of Asian Earth Sciences, 37, 399411.Google Scholar
Bazhenov, M.L., Collins, A.Q., Degtyarev, K.E. et al. (2003). Paleozoic northward drift of the North Tien Shan (Central Asia) as revealed by Ordovician and Carboniferous paleomagnetism. Tectonophysics, 366, 113141.Google Scholar
Beck, M.E. Jr. & Housen, B.A. (2003). Absolute velocity of North America during the Mesozoic from paleomagnetic data. Tectonophysics, 377, 3354.CrossRefGoogle Scholar
Becker, T.P., Thomas, W.A. & Gehrels, G.E. (2006). Linking Late Paleozoic sedimentary provenance in the Appalachian Basin to the history of the Alleghanian deformation. American Journal of Science, 306, 777798.CrossRefGoogle Scholar
Becker, T.W. & Boschi, I. (2002). A comparison of tomographic and geodynamic mantle models. Geochemistry, Geophysics, Geosystems, 3, doi:10.1029/2001GC000168.Google Scholar
Belasky, P., Stevens, C.H. & Hanger, R.A. (2002). Early Permian location of western North American terranes based on brachiopod, fusulinid and coral biogeography. Palaeogeography, Palaeoclimatology, Palaeoecology, 179, 245266.Google Scholar
Belousov, V.I. (2007). The Upper Palaeozoic preflysch and overthrusting in the Türkstan–Alay ranges, southern Fergana. Geotektonika, 2007(5), 6375 (in Russian).Google Scholar
Benedetto, J.L. (1998). Early Palaeozoic brachiopods and associated shelly faunas from western Gondwana: their bearing on the geodynamic history of the pre-Andean margin. In Pankhurst, R.J. & Rapela, C.W. (eds.), The Proto-Andean Margin of Gondwana. Geological Society, London, Special Publications, 142, pp. 5783.Google Scholar
Benton, M.J. (1995). Diversification and extinction in the history of life. Science, 268, 5258.Google Scholar
Benton, M.J. (2005). Vertebrate Palaeontology, 3rd edn. Oxford: Blackwell.Google Scholar
Benton, M.J. (2008). The end-Permian mass extinction events on land in Russia. Proceedings of the Geologists’ Association, 119, 119136.Google Scholar
Berner, R.A. (1997). The rise of plants and their effect on weathering and atmospheric CO2. Science, 276, 544546.CrossRefGoogle Scholar
Berner, R.A., Lasaga, A.C. & Garrels, R.M. (1983). The carbonate– silicate geochemical cycle and its effect on atmospheric carbon dioxide over the past 100 million years. American Journal of Science, 283, 641683.CrossRefGoogle Scholar
Berra, F. & Angiolini, L. (2014). The evolution of the Tethys Region throughout the Phanerozoic: a brief tectonic reconstruction. In Marlow, L., Kendall, C.C.G. & Yose, L.A. (eds.), Petroleum Systems of the Tethyan Region. AAPG Memoir, 106, pp. 127.Google Scholar
Beuf, S., Bijou-Duval, V., De Charpal, O. et al. (1971). Les Grés du Paléozoïque au Sahara. Publications de l’institut français du pétrole, 18.Google Scholar
Biggin, A.J., Steinberger, B., Aubert, J. et al. (2012). Long term geomagnetic variations and whole-mantle convection processes. Nature Geoscience, 5, 526533.Google Scholar
Bird, P. (2003). An updated digital model of plate boundaries. Geochemistry, Geophysics, Geosystems, 4, 1027, doi:10.1029/2001GC000252.Google Scholar
Biske, Y.S. & Seltmann, R. (2010). Paleozoic Tian-Shan as a transitional region between the Rheic and Urals–Turkestan oceans. Gondwana Research, 17, 602613.Google Scholar
Blieck, A. & Cloutier, R. (2000). Biostratigraphical correlations of Early Devonian vertebrate assemblages of the Old Red Sandstone continent. Courier Forschungsinstitut Senckenberg, 223, 223269.Google Scholar
Blodgett, R.B. & Stanley, G.D. (eds.) (2008). The Terrane Puzzle: New Perspectives on Paleontology and Stratigraphy from the North American Cordillera. Geological Society of America Special Paper, 442.CrossRefGoogle Scholar
Bonev, N. (2006). Cenozoic tectonic evolution of the eastern Rhodope Massif (Bulgaria): basement structure and kinematics of syn- to postcollisional extensional deformation. In Dilek, Y. & Pavlides, S. (eds.), Postcollisional Tectonics and Magmatism in the Mediterranean Region and Asia. Geological Society of America Special Paper, 409, pp. 211235.Google Scholar
Boschman, L.M., van Hinsbergen, D.J.J., Torsvik, T.H. et al. (2014). Kinematic reconstruction of the Caribbean region since the Early Jurassic. Earth-Science Reviews, 138, 102136.CrossRefGoogle Scholar
Boucot, A.J. (1975). Evolution and Extinction Rate Controls. Amsterdam: Elsevier.Google Scholar
Boucot, A.J. & Blodgett, R.B. (2001). Silurian–Devonian biogeography. In Brunton, C.H.C., Cocks, L.R.M. & Long, S.L. (eds.), Brachiopods Past and Present. London: Taylor and Francis, pp. 335344.Google Scholar
Boucot, A.J., Johnson, J.G. & Talent, J.A. (1969). Early Devonian Brachiopod Zoogeography. Geological Society of America Special Paper, 119.Google Scholar
Boucot, A.J., Xu, C. & Scotese, C.R. (2013). Phanerozoic Paleoclimate: An Atlas of Lithologic Indicators of Climate. SEPM Concepts in Sedimentology and Paleontology, 11.Google Scholar
Bowring, S.A., Erwin, D.H., Jin, Y.G. et al. (1998). U/Pb zircon geochronology and tempo of the end-Permian mass extinction. Science, 280, 10391045.Google Scholar
Bowring, S.A. & Williams, I.S. (1999). Priscoan (4.00 ± 4.03 Ga) orthogneisses from northwestern Canada. Contributions to Mineralogy and Petrology, 134, 316.Google Scholar
Bradley, D.C. (2008). Passive margins through earth history. Earth-Science Reviews, 91, 126.Google Scholar
Braitenberg, C. (2015). Exploration of tectonic structures with GOCE in Africa and across-continents. International Journal of Applied Earth Observation and Geoinformation, 35, 8895.Google Scholar
Brenchley, P.J. & Cocks, L.R.M. (1982). Ecological associations in a regressive sequence: the latest Ordovician of the Oslo–Asker district, Norway. Palaeontology, 25, 783815.Google Scholar
Brenchley, P.J. & Rawson, P.F. (eds.) (2006). The Geology of England and Wales. The Geological Society, London.Google Scholar
Brew, G., Barazangi, M., Al-Maleh, A.K. & Sawaf, F. (2001). Tectonic and geologic evolution of Syria. GeoArabia, 6, 573615.Google Scholar
Brown, D., Herrington, R. & Alvarez-Marron, J. (2011). Processes of arc–continent collision in the Uralides. In Brown, D. & Ryan, P.D. (eds.), Arc–Continent Collision. Berlin: Springer-Verlag, pp. 311340.Google Scholar
Buiter, S.J.H. & Torsvik, T.H. (2007). Horizontal movements in the Eastern Barents Sea constrained by numerical models and plate reconstructions. Geophysical Journal International, 171, 13761389.CrossRefGoogle Scholar
Buiter, S.J.H. & Torsvik, T.H. (2014). A review of Wilson Cycle plate margins: a role for mantle plumes in continental break-up along sutures? Gondwana Research, 26, 627653, doi:10.1016/j.gr.2014.02.007.Google Scholar
Bullard, E.C., Everett, J.E. & Smith, A.G. (1965). The fit of the continents around the Atlantic. Philosophical Transactions of the Royal Society, A258, 4151.Google Scholar
Burke, K. (2011). Plate tectonics, the Wilson Cycle, and mantle plumes: geodynamics from the top. Annual Review of Earth and Planetary Sciences, 39, 129.Google Scholar
Burke, K., Steinberger, B., Torsvik, T.H. & Smethurst, M.A. (2008). Plume Generation Zones at the margins of Large Low Shear Velocity Provinces on the core–mantle boundary. Earth and Planetary Science Letters, 265, 4960.CrossRefGoogle Scholar
Burke, K. & Torsvik, T.H. (2004). Derivation of Large Igneous Provinces of the past 200 million years from long-term heterogeneities in the deep mantle. Earth and Planetary Science Letters, 227, 531538.Google Scholar
Burtman, V.S. (2008). Nappes of the southern Tien Shan. Russian Journal of Earth Sciences, 10(ES1006), 135.CrossRefGoogle Scholar
Bussien, D., Gombojav, N., Winkler, W. et al. (2011). The Mongol–Okhotsk belt in Mongolia. Tectonophysics, 510, 132150.Google Scholar
Cai, J.X. & Zhang, K.J. (2009). A new model for the Indochina and South China collision during the Late Permian to the Middle Triassic. Tectonophysics, 467, 3543.Google Scholar
Calvès, G., Schwab, A.M., Huuse, M. et al. (2011). Seismic volcanostratigraphy of the western Indian rifted margin: the pre‐Deccan igneous province. Journal of Geophysical Research, 116, B01101, doi:10.1029/2010JB000862.CrossRefGoogle Scholar
Came, R.E., Eiler, J.M., Veizer, J. et al. (2007). Coupling of surface temperatures and atmospheric CO2 concentrations during the Palaeozoic era. Nature, 449, 198201.Google Scholar
Cavazza, W., Roure, F., Spakman, W., Stampfli, G.M. & Ziegler, P.A. (eds.) (2004). The TRANSMED Atlas: The Mediterranean Region from Crust to Mantle. Berlin: Springer.Google Scholar
Charvet, J., Shu, L. & Laurent-Charvet, S. (2007). Paleozoic structural and geodynamic evolution of eastern Tianshan (NW China): welding of the Tarim and Junggar plates. Episodes, 30, 162186.Google Scholar
Chen, X., Zhou, Z. & Fan, J. (2010). Ordovician paleogeography and tectonics of the major paleoplates of China. In Finney, S.C. and Berry, W.B.N. (eds.), The Ordovician Earth System. Geological Society of America Special Paper, 466, pp. 85104.Google Scholar
Chen, Z.Q., Shi, G.R. & Zhan, L.P. (2003). Early Carboniferous athyridid brachiopods from the Qaidam Basin, northwest China. Journal of Paleontology, 77, 844862.2.0.CO;2>CrossRefGoogle Scholar
Christiansen, J.L. & Stouge, S. (1999). Oceanic circulation as an element in palaeogeographical reconstructions: the Arenig (early Ordovician) as an example. Terra Nova, 11, 7378.Google Scholar
Chulick, G.S., Detweiler, S. & Mooney, W.D. (2013). Seismic structure of the crust and uppermost mantle of South America and surrounding oceanic basins, Journal of South American Earth Sciences, 42, 260276.Google Scholar
Clack, J.A. (2002). Gaining Ground: The Origin and Evolution of Tetrapods. Bloomington: Indiana University Press.Google Scholar
Cocks, L.R.M. (1972). The origin of the Silurian Clarkeia shelly fauna of South America, and its extension to West Africa. Palaeontology, 15, 623630.Google Scholar
Cocks, L.R.M. (ed.) (1981). The Evolving Earth. Cambridge: Cambridge University Press.Google Scholar
Cocks, L.R.M. (2011). There’s no place like home: Cambrian to Devonian brachiopods critically useful for analysing palaeogeography. Memoir of the Association of Australasian Palaeontologists, 41, 135148.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, 138, 465478.Google Scholar
Cocks, L.R.M. & Fortey, R.A. (1988). Lower Palaeozoic facies and faunas round Gondwana. In Audley-Charles, M.G. & Hallam, A. (eds.), Gondwana and Tethys. Geological Society, London, Special Publications, 37, pp. 183200.Google Scholar
Cocks, L.R.M. & Fortey, R.A. (2009). Avalonia: a long-lived terrane in the Lower Palaeozoic? In Bassett, M.G. (ed.), Early Palaeozoic Peri-Gondwana Terranes. Geological Society, London, Special Publications, 325, pp. 141155.Google Scholar
Cocks, L.R.M., Fortey, R.A. & Lee, C.P. (2005). A review of Lower and Middle Palaeozoic biostratigraphy in west peninsula Malaya and southern Thailand in its context within the Sibumasu Terrane. Journal of Asian Earth Sciences, 34, 703717.Google Scholar
Cocks, L.R.M. & Rong, J. (2008). Earliest Silurian faunal survival and recovery after the end-Ordovician glaciation: evidence from the brachiopods. Transactions of the Royal Society of Edinburgh Earth and Environmental Sciences, 98, 291301.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, 631644.CrossRefGoogle Scholar
Cocks, L.R.M. & Torsvik, T.H. (2005). Baltica from the Late Precambrian to mid-Palaeozoic times: the gain and loss of a terrane’s identity. Earth-Science Reviews, 72, 3966.Google Scholar
Cocks, L.R.M. & Torsvik, T.H. (2007). Siberia, the wandering northern terrane, and its changing geography through the Palaeozoic. Earth-Science Reviews, 82, 2974.Google Scholar
Cocks, L.R.M. & Torsvik, T.H. (2011). The Palaeozoic geography of Laurentia and western Laurussia: a stable craton with mobile margins. Earth-Science Reviews, 106, 151.Google Scholar
Cocks, L.R.M. & Torsvik, T.H. (2013). The dynamic evolution of the Palaeozoic geography of eastern Asia. Earth-Science Reviews, 117, 4079.Google Scholar
Cocks, L.R.M. & Verniers, J. (2000). Applicability of planktic and nektic fossils to palaeogeographic reconstructions. Acta Universitatis Carolinae – Geologica, 42, 399400.Google Scholar
Cody, S., Richardson, J.E., Ruli, V. et al. (2010). The Great American Biotic Interchange revisited. Ecography, 33, 326332.Google Scholar
Coffin, M.F. & Eldholm, O. (1994). Large igneous provinces: crustal structure, dimensions, and external consequences. Reviews of Geophysics, 32, 136.Google Scholar
Cohen, K.M., Finney, S.C., Gibbard, P.L. & Fan, J.-X. (2013; updated). The ICS International Chronostratigraphic Chart. Episodes, 36, 199204.Google Scholar
Collier, J.S., Minshull, T.A., Haqmmond, J.O.S. et al. (2009). Factors influencing magmatism during continental breakup: new insights from a wide-angle seismic experiment across the conjugate Seychelles–Indian margins. Journal of Geophysical Research, 114, B03101.Google Scholar
Collinson, M.E. & Hooker, J.J. (2003). Paleogene vegetation of Eurasia: framework for mammalian faunas. Deinsea, 10, 4184.Google Scholar
Colpron, M. & Nelson, J.L. (2006). Palaeozoic Evolution and Metallogeny of Pericratonic Terranes at the Ancient Pacific Margin of North America. Geological Association of Canada Special Paper, 45.Google Scholar
Colpron, M. & Nelson, J.L. (2009). A Palaeozoic Northwest Passage: incursion of Caledonian, Baltican and Siberian terranes into eastern Panthalassa and the early evolution of the North American Cordillera. In Cawood, P.A. & Kröner, A. (eds.), Earth Accretionary Systems in Space and Time, Geological Society, London, Special Publications, 318, pp. 273307.Google Scholar
Connelly, J.N., Bizzarro, M., Krot, A.N. et al. (2012). Absolute chronology and thermal processing of solids in the solar protoplanetary disk. Science, 338, 651655.Google Scholar
Conrad, C.P., Steinberger, B. & Torsvik, T.H. (2014). Dynamic topography and sea level change inferred from dipole and quadrupole moments of plate tectonic reconstructions. American Geophysical Union Fall Meeting, San Francisco, Abstract.Google Scholar
Cook, H.E., Zhemchuzhnikov, V.G., Zempolich, W.G. et al. (2002). Devonian and Carboniferous platform facies in the Bolshoi Karatau, southern Kazakhstan: outcrop analogs for coeval carbonate oil and gas fields in the North Caspian Basin, western Kazakhstan. In Zempolich, W.G. & Cook, H.E. (eds.), Paleozoic Carbonates of the Commonwealth of Independent States. SEPM Special Publication, 74, pp. 81122.Google Scholar
Cope, T., Ritts, B.D., Darby, B.J. et al. (2005). Late Paleozoic sedimentation on the northern margin of the North China Block: implications for regional tectonics and climate changes. International Geology Review, 47, 270296.Google Scholar
Copper, P. (2002). Silurian and Devonian reefs. In Kiessling, W., Flügel, E. & Golonka, J. (eds.), Phanerozoic Reef Patterns. SEPM Special Publication, 72, pp. 181238.Google Scholar
Copper, P. & Jin, J. (2015). Tracking the early Silurian post-extinction faunal recovery in the Jupiter Formation of Anticosti Island, eastern Canada: a stratigraphical revision. Newsletters on Stratigraphy, 48, 221240.Google Scholar
Corfu, F., Polteau, S., Planke, S. et al. (2013). U–Pb geochronology of Cretaceous magmatism on Svalbard and Franz Josef Land, Barents Sea Large Igneous Province. Geological Magazine, 150, 11271135.Google Scholar
Courtillot, V., Davaille, A., Besse, J. & Stock, J. (2003). Three distinct types of hotspots in the Earth’s mantle. Earth and Planetary Science Letters, 205, 295308.Google Scholar
Courtillot, V., Gallet, Y., Le Mouël, J.-L., Fluteau, F. & Genevey, A. (2007). Are there connections between the Earth’s magnetic field and climate? Earth and Planetary Science Letters, 253, 328339.Google Scholar
Courtillot, V.E. & Renne, P.-R. (2003). On the ages of flood basalt events. Comptes Rendus Geoscience, 335, 113140.Google Scholar
Dabard, M.P., Loi, A., Paris, F. et al. (2015). Sea-level curve for the Middle to early Late Ordovician in the Armorican Massif (western France): icehouse third-order glacio-eustatic cycles. Palaeogeography, Palaeoclimatology, Palaeoecology, 436, 96111.CrossRefGoogle Scholar
Dal Corso, J., Marzoli, A., Tateo, F. et al. (2014). The dawn of CAMP volcanism and its bearing on the end-Triassic carbon cycle disruption. Journal of the Geological Society, London, doi.org/10.1144/jgs2013-063.Google Scholar
Dalhquist, J.A., Pankhurst, R.J., Gaschnig, R.M. et al. (2013). Hf and Nd isotopes in Early Ordovician to Early Carboniferous granites in the Proto-Andean margin of Gondwana. Gondwana Research, 23, 16171630.Google Scholar
Dalziel, I.W.D. (1997). Neoproterozoic–Paleozoic geography and tectonics: review, hypothesis, environmental speculation. Geological Society of America Bulletin, 109, 1642.Google Scholar
Darwin, C. (1859). The Origin of Species. London: John Murray.Google Scholar
Daukeev, S.Z., Uzhkenov, B.S., Miletenko, N.V. et al. (eds.) (2002). Atlas of Lithology – Paleogeographical, Structural, Palinspastic and Geoenvironmental Maps of Central Eurasia. Almaty: Scientific Research Institute of Natural Resources (in Russian).Google Scholar
Dawes, P.R. (2009). Precambrian–Palaeozoic geology of Smith Sound, Canada and Greenland: key constraint to palaeogeographical reconstructions of northern Laurentia and the North Atlantic region. Terra Nova, 21, 113.CrossRefGoogle Scholar
Dean, W.T., Monod, O., Rickards, R.B. et al. (2000). Lower Palaeozoic stratigraphy and palaeontology, Karadire-Zirze area, Pontus Mountains, northern Turkey. Geological Magazine, 137, 555582.Google Scholar
de Freitas, T.A. & Dixon, O.A. (1995). Silurian microbial buildups, Canadian Arctic. In Monty, C.L.V., Bosence, D.W.J., Bridges, P.H. & Pratt, B.R. (eds.), Carbonate Mud-Mounds: Their Origin and Evolution. International Association of Sedimentologists, Special Publication, 23, pp. 151169.Google Scholar
Degtyarev, K.Y. & Ryazantsev, A.V. (2007). Cambrian arc–continent collision in the Palaeozoides of Kazakhstan. Geotectonics, 43, 6386.Google Scholar
de Jong, K, Xiao, W., Windley, B.F. et al. (2006). Ordovician 40Ar/39Ar phengite ages from the blueschist-facies Ondor Sum subduction–accretion complex (Inner Mongolia) and implications for the Early Paleozoic history of continental blocks in China and adjacent areas. American Journal of Science, 306, 799845.CrossRefGoogle Scholar
Dewing, K., Harrison, J.C., Pratt, B.R. & Mayr, U. (2004). A probable late Neoproterozoic age for the Kennedy Channel and Ella Bay formations, northeastern Ellesmere Island and its implications for passive margin history in the Canadian Arctic. Canadian Journal of Earth Sciences, 41, 10131025.CrossRefGoogle Scholar
Dhuime, B., Hawkesworth, C.J., Cawood, P.A. & Storey, C.D. (2012). A change in the geodynamics of continental growth 3 billion years ago. Science, 335,13341336.Google Scholar
Dhuime, B., Wuestefeld, A. & Hawkesworth, C.J. (2015). Emergence of modern continental crust about 3 billion years ago. Nature Geoscience, 8, doi:10.1038/NGEO2466.Google Scholar
Dickinson, W.R. (2000). Geodynamic interpretation of Paleozoic tectonic trends oriented oblique to the Mesozoic Klamath–Sierran continental margin in California. In Soreghan, M.J. & Gehrels, G.E. (eds.), Paleozoic and Triassic paleogeography and tectonics of western Nevada and Northern California. Geological Society of America Special Paper, 347, pp. 200245.Google Scholar
Dickinson, W.R. (2009). Anatomy and global context of the North American Cordillera. In Kay, S. Mahlburg, Ramos, V.A. & Dickinson, W.R. (eds.), Backbone of the Americas: Shallow Subduction, Plateau Uplift, and Ridge and Terrane Collision. Geological Society of America Memoir, 204, pp. 129.Google Scholar
Dickinson, W.R. & Lawton, T.F. (2001). Carboniferous to Cretaceous assembly and fragmentation of Mexico. Geological Society of America Bulletin, 113, 11421160.Google Scholar
DiMichele, W.A., Montanez, I.P., Poulsen, C.J. & Tabor, N.J. (2009). Climate and vegetational regime shifts in the late Paleozoic ice age earth. Geobiology, 7, 200226.Google Scholar
DiMichele, W.A., Gastaldo, R.A. & Pfefferkorn, H.W. (2005). Plant biodiversity partitioning in the Late Carboniferous and Early Permian and its implications for ecosystem assembly. Proceedings of the California Academy of Sciences, 56, Supplement 1(4), 3249.Google Scholar
Dobretsov, N.L., Berzin, N.A., Buslov, M.M. (1995). Opening and tectonic evolution of the Paleo-Asian ocean. International Geology Review, 35, 335360.Google Scholar
Dobretsov, N.L., Buslov, M.M. & Vernikovsky, V.A. (2003). Neoproterozoic to Early Ordovician evolution of the Paleo-Asian Ocean: implications to the break-up of Rodinia. Gondwana Research, 6, 143159.Google Scholar
Dobretsov, N.L., Buslov, M.M., Zhimulev, F.I. et al. (2006). Vendian–early Ordovician geodynamic evolution and model for exhumation of ultrahigh- and high-pressure rocks from the Kokchetav subduction–collision zone. Geologiya i Geofizika, 47, 428444 [in Russian].Google Scholar
Dodd, S.C., MacNiocaill, C. & Muxworthy, A.R. (2015). Long duration (>4 Ma) and steady-state volcanic activity in the early Cretaceous Paraná–Etendeka Large Igneous Province: new palaeomagnetic data from Namibia. Earth and Planetary Science Letters, 414, 1629.Google Scholar
Domeier, M. (2015). A plate tectonic scenario in the Iapetus and Rheic oceans. Gondwana Research, doi:10.1016/j.gr.2015.08.003.Google Scholar
Domeier, M. & Torsvik, T.H. (2014). Plate tectonics in the late Paleozoic. Geoscience Frontiers, 5, 303350.Google Scholar
Domeier, M., Van der Voo, R. & Torsvik, T.H. (2012). Paleomagnetism and Pangea: the Road to reconciliation. Tectonophysics, 514, 1443.Google Scholar
Dornbos, S.Q. & Bottjer, D.J. (2000). Evolutionary paleoecology of the earliest echinoderms: helicoplacoids and the Cambrian substrate revolution. Geology, 28, 839842.Google Scholar
Doubrovine, P.V., Steinberger, B. & Torsvik, T.H. (2012). Absolute plate motions in a reference frame defined by moving hotspots in the Pacific, Atlantic and Indian oceans. Journal of Geophysical Research, 117, B09101, doi:10.1029/2011JB009072.Google Scholar
Doubrovine, P.V. & Tarduno, J.A. (2004). Late Cretaceous paleolatitude of the Hawaiian Hot Spot: new paleomagnetic data from Detroit Seamount (ODP Site 883). Geochemistry, Geophysics, Geosystems, 5, Q11L04, doi:10.1029/2004GC000745.Google Scholar
Edwards, D., Cherns, L. & Raven, J.A. (2015). Could land-based early photosynthesizing ecosystems have bioengineered the planet in mid-Palaeozoic times? Palaeontology, 58, 803837.Google Scholar
Egan, S.S., Mosar, J., Brunet, M.F. & Kangarli, T. (2009). Subsidence and uplift mechanisms within the South Caspian Basin: insights from the onshore and offshore Azerbaijan region. In Brunet, M.-F., Wilmsenj, M. & Granath, W. (eds.), South Caspian to Central Iran Basins. Geological Society, London, Special Publications, 312, pp. 219240.Google Scholar
Eide, E.A. & Torsvik, T.H. (1996). Paleozoic supercontinent assembly, mantle flushing and genesis of the Kiaman Superchrons. Earth and Planetary Science Letters, 144, 389402.Google Scholar
Eldholm, O. & Myhre, A.M. (1977). Hovgaard Fracture Zone. In Årbok 1976 Oslo: Norsk Polarinstitutt, 195208.Google Scholar
Elliott, D. (2013). The geological and tectonic evolution of the Transantarctic Mountains: a review. In Hambrey, M.J. et al. (eds.), Antarctic Palaeoenvironments and Earth-Surface Processes. Geological Society, London, Special Publications, 381, pp. 735.Google Scholar
Embry, A.F. (1991). Middle–Upper Devonian clastic wedge of the Arctic islands. In Trettin, H.P. (ed.) Geology of the Innuitian Orogen and Arctic Platform of Canada and Greenland. Ottawa: Geological Survey of Canada, Geology of Canada, 3, pp. 263279.Google Scholar
Engen, Ø., Faleide, J.I. & Dyreng, T.K. (2008). Opening of the Fram Strait gateway: a review of plate tectonic constraints. Tectonophysics, 450, 5169, doi:10.1016/j.tecto.2008.01.002.Google Scholar
Escayola, M.P., van Staal, C.R. & Davis, W.J. (2011). The age and tectonic setting of the Punoviscana Formation in northwestern Argentina: an accretionary complex related to early Cambrian closure of the Punoviscana Ocean and accretion of the Arequipa–Antofalla blocks. Journal of South American Earth Sciences, 32, 438450.CrossRefGoogle Scholar
Evans, D.A.D. (2013). Reconstructing pre-Pangean supercontinents. Geological Society of America Bulletin, 125, 17351751.Google Scholar
Evenchick, C.A., Davis, W.J., Bédard, J.H., Hayward, N. & Friedman, R.N. (2015). Evidence for protracted High Arctic Large Igneous Province magmatism in the central Sverdrup Basin from stratigraphy, geochronology, and paleodepths of saucer-shaped sills. Geological Society of America Bulletin, 127, 13661390, doi.org/10.1130/B31190.1Google Scholar
Faccenna, C., Becker, T.W., Auer, L. et al. (2014). Mantle dynamics in the Mediterranean, Reviews of Geophysics, 52, 283332.Google Scholar
Faleide, J.I., Bjørlykke, K. & Gabrielsen, R.H. (2010). Geology of the Norwegian Continental Shelf. In Bjørlykke, K. (ed.), Petroleum Geoscience: From Sedimentary Environments to Rock Physics, Springer Science Business Media, pp. 467499.Google Scholar
Fergusson, C.L. & Henderson, R.A. (2015). Early Palaeozoic continental growth in the Tasmanides of northeast Gondwana and its implications for Rodinia assembly and rifting. Gondwana Research, 28, 933953.Google Scholar
Fielding, C.R., Frank, T.D. & Isbell, J.L. (eds.) (2008). Resolving the Late Paleozoic Ice Age in Time and Space. Geological Society of America Special Paper, 441, pp. 7182.Google Scholar
Flecker, R., Krijgsman, W., Capella, W. et al. (2015). Evolution of the Late Miocene Mediterranean–Atlantic gateways and their impact on regional and global environmental change. Earth-Science Reviews, 150, 365392.Google Scholar
Fortey, R.A. & Cocks, L.R.M. (2003). Palaeontological evidence bearing on global Ordovician–Silurian continental reconstructions. Earth-Science Reviews, 61, 245307.CrossRefGoogle Scholar
Fortey, R.A. & Cocks, L.R.M. (2005). Late Ordovician global warming: the Boda Event. Geology, 33, 405408.Google Scholar
Francis, J.E., Ashworth, A., Cantrill, D.J. et al. (2008). 100 million years of Antarctic climate evolution: evidence from fossil plants. In Cooper, A.K., Barrett, P.J. et al. (eds.), Antarctica: A Keystone in a Changing World. Washington, DC: National Academies Press, pp. 1927,Google Scholar
Franke, W. (2006). The Variscan orogen in central Europe: construction and collapse. In Gee, D.G. & Stephenson, R.A. (eds.), European Lithosphere Dynamics. Geological Society, London, Memoir, 32, pp. 333343.Google Scholar
Franke, W., Cocks, L.R.M. & Torsvik, T.H. (2016). Fresh insights into an old orogeny: the Variscan revisited. Gondwana Research (in press).Google Scholar
French, S.W. & Romanowicz, B. (2015). Broad plumes rooted at the base of the Earth’s mantle beneath major hotspots. Nature, 525, 9599.Google Scholar
Friend, P.F. & Williams, B.P.J. (eds.) (2000). New Perspectives on the Old Red Sandstone. Geological Society, London, Special Publications, 180.Google Scholar
Friis, E.M., Crane, P.R. & Pedersen, K.R. (2011). Early Flowers and Angiosperm Evolution. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Froitzheim, N., Plašienka, D. & Schuster, R. (2008). Alpine tectonics of the Alps and western Carpathians. In McCann, T. (ed.), The Geology of Central Europe. The Geological Society, London, pp. 11411232.Google Scholar
Gabrielse, H. & Yorath, C.J. (1992). Geology of the Cordilleran Orogeny in Canada. The Geology of North America, Vol. G-2. Ottawa: Geological Survey of Canada.Google Scholar
Gaetani, M. (1997). The Karakorum Block in Central Asia, from Ordovician to Cretaceous. Sedimentary Geology, 109, 339359.Google Scholar
Gaetani, M., Angiolini, L, Ueno, K. et al. (2009). Pennsylvanian–Early Triassic stratigraphy in the Alborz Mountains (Iran). In Brunet, M.-F., Wilmsenj, M. & Granath, W. (eds.), South Caspian to Central Iran Basins. Geological Society, London, Special Publications, 312, pp. 79128.Google Scholar
Gaina, C., Gernigon, L. & Ball, P. (2009). Palaeocene–Recent plate boundaries in the NE Atlantic and the formation of the Jan Mayen microcontinent. Journal of the Geological Society, London, 166, 601616.Google Scholar
Gaina, C., Müller, D.R., Royer, J.-Y. et al. (1998). The tectonic history of the Tasman Sea: a puzzle with 13 pieces. Journal of Geophysical Research, 103, 1241312433.Google Scholar
Gaina, C., Müller, D.R., Royer, J.-Y. & Symonds, P. (1999). Evolution of the Louisiade triple junction. Journal of Geophysical Research, 104, 1292712939.Google Scholar
Gaina, C., Medvedev, S., Torsvik, T.H., Koulakov, I.Yu & Werner, S.C. (2013a). 4D Arctic: a glimpse into the structure and evolution of the Arctic in the light of new geophysical maps, plate tectonics and tomographic models. Surveys in Geophysics, 35, 10951122.Google Scholar
Gaina, C., Roest, W.R. & Muller, R.D. (2002). Late Cretaceous–Cenozoic deformation of northeast Asia. Earth and Planetary Science Letters, 197, 273286.Google Scholar
Gaina, C., Torsvik, T.H., van Hinsbergen, D. et al. (2013b). The African Plate: a history of oceanic crust accretion and subduction since the Jurassic. Tectonophysics, 604, 425.Google Scholar
Gaina, C., van Hinsbergen, D. & Spakman, W. (2015). Tectonic interactions between India and Arabia since the Jurassic reconstructed from marine geophysics, ophiolite geology, and seismic tomography. Tectonics, 34(5), 875906.Google Scholar
Garnero, E.J., Lay, T. & McNamara, A. (2007). Implications of lower-mantle structural heterogeneity for existence and nature of whole-mantle plumes. In Foulger, G.R. & Jurdy, D.M. (eds.), Plates, Plumes and Planetary Processes. Geological Society of America Special Paper, 430, pp. 79101.Google Scholar
Gee, D.G. & Pease, V.I. (eds.) (2005). The Neoproterozoic Timanide Orogen of Eastern Baltica. Geological Society, London, Memoir, 30.Google Scholar
Gee, J.S. & Kent, D.V. (2007). Source of oceanic magnetic anomalies and the geomagnetic polarity time scale. Treatise on Geophysics, 5, 455507.Google Scholar
Ghienne, J.F., Le Heron, D.P., Moreau, J. et al. (2007). The Late Ordovician sedimentary system of the North Gondwana platform. In Hambrey, M. J. et al. (eds.), Glacial Sedimentary Processes and Products. International Association of Sedimentologists Special Publication, 39, pp. 297319.Google 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, 101146.Google Scholar
Gibbons, W. & Moreno, T. (eds.) (2002). The Geology of Spain. The Geological Society, London.Google Scholar
Gibbons, A.D., Whittaker, J.M. & Müller, R.D. (2013). The breakup of East Gondwana: assimilating constraints from Cretaceous ocean basins around India into a best-fit tectonic model. Journal of Geophysical Research, 118, 115.Google Scholar
Gibling, M.R., Davies, N.S., Falcon-Lang, H.J. et al. (2014). Palaeozoic co-evolution of rivers and vegetation: a synthesis of current knowledge. Proceedings of the Geologists’ Association, 125, 524533.CrossRefGoogle Scholar
Glass, L.M. & Phillips, D. (2006). The Kalkarindji continental flood basalt province: a new Cambrian large igneous province in Australia with possible links to faunal extinctions. Geology, 34, 461464.CrossRefGoogle Scholar
Glen, R.A. (2005). The Tasmanides of eastern Australia. In Vaughan, A.P.M., Leat, P.T. & Pankhurst, R.J. (eds.), Terrane Processes at the Margins of Gondwana. Geological Society, London, Special Publications, 246, pp. 2396.Google Scholar
Glennie, K. (ed.) (2006). Oman’s Geological Heritage, 2nd edn. Muscat: Petroleum Development Oman.Google Scholar
Goldreich, P. & Toomre, A. (1969). Some remarks on polar wandering. Journal of Geophysical Research, 74, 25552569.Google Scholar
Goodfellow, W.D., Cecile, M.P. & Leybourne, M.I. (1995). Geochemistry, petrogenesis and tectonic setting of Lower Paleozoic alkalic and potassic volcanic rocks, northern Canadian Cordilleran Miogeocline. Canadian Journal of Earth Sciences, 32, 12361254.Google Scholar
Greb, S.F., Pashin, J.C., Martino, R.L. & Eble, C.F. (2008). Appalachian sedimentary cycles during the Pennsylvanian: changing influences of sea level, climate, and tectonics. In Fielding, C.R., Frank, T.D. & Isbell, J.L. (eds.), Resolving the Late Paleozoic Ice Age in Time and Space. Geological Society of America Special Paper, 441, pp. 235248.Google Scholar
Gulbranson, E.L., Ryberg, P.E., Decobeix, A.-L. et al. (2014). Leaf habit of Late Permian Glossopteris trees from high-palaeolatitude forests. Journal of the Geological Society, London, 171, 493507.Google Scholar
Hall, R. (2012). Late Jurassic–Cenozoic reconstructions of the Indonesian margin and the Indian Ocean. Tectonophysics, 570–571, 141.Google Scholar
Hall, R. & Holloway, J.D. (eds.) (1998). Biogeography and Geological Evolution of SE Asia. Leiden: Backhuys.Google Scholar
Hallam, A. (1988). A reevaluation of Jurassic eustasy in the light of new data and the revised Exxon curve. In Wilgus, C.K. et al. (eds.), Sea-Level Changes: An Integrated Approach. SPEM Special Publication, 42, pp. 261273.CrossRefGoogle Scholar
Haq, B.U. & Al-Qahtani, A.M. (2005). Phanerozoic cycles of sea-level change on the Arabian Platform. GeoArabia, 10, 127160.Google Scholar
Haq, B.U. & Shutter, S.R. (2008). A chronology of Paleozoic sea-level changes. Science, 322, 6468.Google Scholar
Harper, D.A.T., Mac Niocaill, C. & Williams, S.H. (1996). The palaeogeography of the early Ordovician Iapetus terranes: an integration of faunal and palaeomagnetic constraints. Palaeogeography, Palaeoclimatology, Palaeoecology, 121, 297312.Google Scholar
Harper, D.A.T. & Servais, T. (eds.) (2013). Early Palaeozoic Biogeography and Palaeogeography. Geological Society, London, Memoirs, 38.Google Scholar
Hartz, E.H. & Torsvik, T.H. (2002). Baltica upside down: a new plate tectonic model for Rodinia and the Iapetus Ocean. Geology, 30, 255258.Google Scholar
Hatcher, R.D., Thomas, W.A. & Viele, G.W. (eds.) (1989). The Appalachian–Ouachita Orogeny in the United States: The Geology of North America, Vol. F-2. Boulder: Geological Society of America.Google Scholar
Havlíček, V., Vaněk, J. & Fatka, O. (1994). Perunica microcontinent in the Ordovician (its position within the Mediterranean Province, series divisions, benthic and pelagic associations). Sborník geologickych věd Geologie, 46, 2556.Google Scholar
Hawkesworth, C., Dhuime, B., Pietranik, A. et al. (2010). The generation and evolution of the continental crust. Journal of the Geological Society, London, 167, 229248.Google Scholar
Hawkins, T., Smith, M.P., Herrington, R.J. et al. (2016). The geology and genesis of the iron skarns of the Turgai belt, northwestern Kazakhstan. Ore Geology Reviews (in press).Google Scholar
Heine, C., Zoethout, J. & Muller, R.D. (2013). Kinematics of the South Atlantic rift. Solid Earth, 4, 215253.Google Scholar
Helbing, H. & Tiepolo, M. (2005). Age determination of Ordovician magmatism in NE Sardinia and its bearing on Variscan basement evolution. Journal of the Geological Society, London, 162, 689700.Google Scholar
Hellinger, S.J. (1981). The uncertainties of finite rotations in plate tectonics. Journal of Geophysical Research, 86, 93129318.Google Scholar
Henriksen, N. (2008). Geological History of Greenland: Four Billion Years of Earth Evolution. Copenhagen: Geological Survey of Denmark.Google Scholar
Hervé, F., Calderón, M., Fanning, C.M. et al. (2013). Provenance variations in the Late Paleozoic accretionary complex of central Chile as indicated by detrital zircons. Gondwana Research, 23, 11221135.Google Scholar
Higgins, A.K., Gilotti, J.A. & Smith, M.P. (eds.) (2008). The Greenland Caledonides: Evolution of the Northeast Margin of Laurentia. Geological Society of America Memoir, 202.Google Scholar
Hildebrand, R.S. (2014). Geology, mantle tomography, and inclination corrected paleogeographic trajectories support westward subduction during Cretaceous orogenesis in the North American Cordillera. Geoscience Canada, 41, doi.org/10.12789/geocanj.2014.41.032.Google Scholar
Hoepffer, C., Soulaimani, A. & Piqué, A. (2005). The Moroccan Hercynides. Journal of African Earth Sciences, 43, 144165.Google Scholar
Hoffman, P.F., Kaufman, A.J., Halverson, G.P. & Schrag, D.P. (1998). A Neoproterozoic Snowball Earth. Science, 281, 13421346.Google Scholar
Holmer, L.E., Popov, L.E., Koneva, S.P. & Bassett, M.G. (2001). Cambrian–Early Ordovician Brachiopods from Malyi Karatau, the Western Balkash Region, and Tien Shan, Central Asia. The Palaeontological Society, Special Papers in Palaeontology, 65.Google Scholar
Holz, M., França, A.B., Sousa, P.A. et al. (2010). A stratigraphic chart of the Late Carboniferous/Permian succession of the eastern border of the Paraná Basin, Brazil. Journal of South American Earth Sciences, 29, 381389.Google Scholar
Husson, L., Conrad, C.P. & Faccenna, C. (2012). Plate motions, Andean orogeny, and volcanism above the South Atlantic convection cell. Earth and Planetary Science Letters, 317–318, 126135.Google Scholar
Huyghe, D., Lartaud, F., Emmanuel, L., Merle, D. & Renard, M. (2015). Palaeogene climate evolution in the Paris Basin from oxygen stable isotope (δ18O) compositions of marine molluscs. Journal of the Geological Society, London, 172, 576587.Google Scholar
Ikeda, M., Hori, R.S., Okada, Y. & Nakada, A. (2015). Volcanism and deep-ocean acidification across the end-Triassic extinction event. Palaeogeography, Palaeoclimatology, Palaeoecology, 440, 725733.Google Scholar
IPCC (2013). Summary for Policymakers. In Stocker, T.F., Qin, D., Plattner, G.-K. et al. (eds.), Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge: Cambridge University Press.Google Scholar
Isozaki, Y., Aoki, K., Nakama, T. & Yanai, S. (2010). New insight into a subduction-related orogeny: a reappraisal of the geotectonic framework and evolution of the Japanese islands. Gondwana Research, 18, 82105.Google Scholar
Jaanusson, V. (1973). Aspects of carbonate sedimentation in the Ordovician of Baltoscandia. Lethaia, 6, 1134.Google Scholar
James, K.H., Lorente, M.A. & Pindell, J.L. (eds.) (2009). The Origin and Evolution of the Caribbean Plate. Geological Society, London, Special Publications, 328.Google Scholar
Jenkyns, H.C. (2010). Geochemistry of oceanic anoxic events. Geochemistry, Geophysics, Geosystems, 11, Q03004, doi:10.1029/2009GC002788.Google Scholar
Jian, P., Liu, D., Kröner, A. et al. (2008). Time scale of an early to mid-Palaeozoic orogenic cycle of the long-lived Central Asian Orogenic Belt, Inner Mongolia of China: implications for continental growth. Lithos, 101, 233259.Google Scholar
Jian, P., Liu, D., Kröner, A. et al. (2009a). Devonian to Permian plate tectonic cycle of the Paleo-Tethys Orogen in southwest China (I): geochemistry of ophiolites, arc/back-arc assemblages and within-plate igneous rocks. Lithos, 113, 748766.Google Scholar
Jian, P., Liu, D., Kröner, A. et al. (2009b). Devonian to Permian plate tectonic cycle of the Paleo-Tethys Orogen in southwest China (II): insights from zircon ages of ophiolites, arc/back-arc assemblages and within plate igneous rocks and generation of the Emeishan CFB province. Lithos, 113, 767784.Google Scholar
Joachimski, M.M., Breizig, S., Buggisch, W. et al. (2009). Devonian climate and reef evolution: insights from oxygen isotopes in apatite. Earth and Planetary Science Letters, 284, 599609.Google Scholar
Johnson, D.P., Maillet, P.C. & Price, R. (1993). Regional setting of a complex backarc: New Hebrides Arc, northern Vanuatu–eastern Solomon Islands. Geo-Marine Letters, 13, 8289.Google Scholar
Johnston, S.T. (2008). The Cordilleran Ribbon Continent of North America. Annual Review of Earth and Planetary Sciences, 36, 495530.Google Scholar
Kay, S.M., Mpodozis, C. & Ramos, V.A. (2005). Andes. In Selley, R.C., Cocks, L.R.M. & Plimer, I.R. (eds.), Encyclopedia of Geology, Volume 1, Amsterdam: Elsevier, pp. 118131.Google Scholar
Keller, B.M. & Predtechensky, N.N. (eds.) (1968). Atlas of Lithology – Paleogeographical Maps of the U.S.S.R. Precambrian, Cambrian, Ordovician, Silurian, Volume 1. Moscow: Ministry of Geology of the USSR (in Russian).Google Scholar
Kennan, L. & Pindell, J.L. (2009). Dextral shear, terrane accretion and basin formation in the Northern Andes: best explained by interaction with a Pacific-derived Caribbean Plate? In James, K.H., Lorente, M.A. & Pindell, J.L. (eds.), The Origin and Evolution of the Caribbean Plate. Geological Society, London, Special Publications, 328, pp. 487531.Google Scholar
Kenrick, P. & Davis, P. (2004). Fossil Plants. London: The Natural History Museum.Google Scholar
Kenrick, P., Wellman, C.H., Schneider, H. & Edgecombe, G.D. (2012). A time-line for terrestrialization: consequences for the carbon cycle in the Palaeozoic. Philosophical Transactions of the Royal Society, B367, 519536.Google Scholar
Kent, D.V. & Tauxe, L. (2005). Corrected Late Triassic latitudes for continents adjacent to the North Atlantic. Science, 307, 240247.Google Scholar
Keppie, J.D. (2004). Terranes of Mexico revisited: a 1.3 billion year odyssey. International Geology Review, 46, 765794.Google Scholar
Keppie, J.D., Dostal, J., Murphy, J.B. & Nance, R.D. (2008). Synthesis and tectonic interpretation of the westernmost Variscan orogeny in southern Mexico: from rifted Rheic margin to active Pacific margin. Tectonophysics, 461, 277290.Google Scholar
Keppie, J.D., Nance, R.D., Dostal, J., Lee, J.K.W. & Ortega-Rivera, A. (2012). Constraints on the subduction erosion/extrusion cycle in the Paleozoic Acatlán Complex of southern Mexico: geochemistry and geochronology of the type Plaxtia Suite. Gondwana Research, 21, 10501065.Google Scholar
Kheraskova, T.N., Didenko, A.N., Bush, V.A. & Volozh, Y.A. (2003). The Vendian–Early Paleozoic history of the continental margin of eastern Paleogondwana, Paleoasian Ocean, and Central Asia Foldbelt. Russian Journal of Earth Sciences, 5, 165184.Google Scholar
Khozyem, H., Adatte, T., Spangenberg, J.E. et al. (2013). Palaeoenvironmental and climatic changes during the Palaeocene–Eocene Thermal Maximum (PETM) at the Wadi Nukhul Section, Sinai, Egypt. Journal of the Geological Society, London, 170, 341352.Google Scholar
Kirschvink, J.L. (1992). Late Proterozoic low-latitude global glaciation: the snowball Earth. In Schopf, J.W., Klein, C. & Des Maris, D. (eds.), The Proterozoic Biosphere: A Multidisciplinary Study. Cambridge: Cambridge University Press, pp. 5152.Google Scholar
Knudsen, M.F. & Riisager, P. (2009). Is there a link between earth’s magnetic field and low-latitude precipitation? Geology, 37, 7174.Google Scholar
Kodama, K.P. (2009). Simplification of the anisotropy-based inclination correction technique for magnetite- and hematite-bearing rocks: a case study for the Carboniferous Glenshaw and Mauch Chunk formations, North America. Geophysical Journal International, 176, 467477.Google Scholar
Kohn, M.J. (2014). Himalayan metamorphism and its tectonic implications. Annual Review of Earth and Planetary Sciences, 42, 381419.Google Scholar
Koppers, A.A.P., Yamazaki, T., Geldmacher, J. et al. (2012). Limited latitudinal mantle plume motion for the Louisville hotspot. Nature Geoscience, 5, 911917, doi:10.1038/NGEO1638.CrossRefGoogle Scholar
Korte, C., Hesselbo, S.P., Jenkyns, H.C. et al. (2009). Palaeoenvironmental significance of carbon- and oxygen-isotope stratigraphy of marine Triassic–Jurassic boundary sections in SW Britain. Journal of the Geological Society, London, 166, 431445.Google Scholar
Kravchinsky, V.A., Konstantinov, K.M. & Cogné, J.P. (2001). Palaeomagnetic study of Vendian and Early Cambrian rocks of South Siberia and Central Mongolia: was the Siberian platform assembled at the time? Precambrian Research, 110, 6192.Google Scholar
Kröner, A. (ed.) (2015). The Central Asian Orogenic Belt. Stuttgart: Borntraeger.Google Scholar
Krstić, B., Maslarević, L., Ercegovać, M. & Dajić, S. (1999). Ordovician of the East-Serbian South Carpathians. Acta Universitatis Carolinae – Geologica, 43, 101114.Google Scholar
Labails, C., Olivet, J.L., Aslanian, D. & Roest, W.R. (2010). An alternative early opening scenario for the Central Atlantic Ocean. Earth and Planetary Science Letters, 297, 355368.Google Scholar
Landing, E., Rushton, A.A., Fortey, R.A. & Bowring, S.A. (2015). Improved geochronologic accuracy and precision for the ICS Chronostratigraphic Charts: examples from the late Cambrian–early Ordovician. Episodes, 38, 154161.Google Scholar
Lapworth, C. (1879). On the tripartite classification of the Lower Palaeozoic rocks. Geological Magazine, 6, 115.Google Scholar
Lee, S., Choi, D.R. & Shi, G.R. (2010). Pennsylvanian brachiopods from the Geumcheon-Jangseong Formation, Pyeongan Supergroup, Taebaeksan Basin, Korea. Journal of Paleontology, 84, 417443.Google Scholar
Lefebvre, B. & Fatka, O. (2003). Palaeogeographical and palaeoecological aspects of the Cambro-Ordovician radiation of echinoderms. Palaeogeography, Palaeoclimatology, Palaeoecology, 195, 7397.Google Scholar
Lekic, V., Cottar, S., Dziewonski, A. & Romanowicz, B. (2012). Cluster analysis of global lower mantle tomography: a new class of structure and implications for chemical heterogeneity. Earth and Planetary Science Letters, 357, 6877.Google Scholar
Leroy, S., Razin, P., Autin, J. et al. (2012). From rifting to oceanic spreading in the Gulf of Aden: a synthesis. Arabian Journal of Geosciences, 5, 859901, doi:10.1007/s12517-011-0475-4.CrossRefGoogle Scholar
Lethiers, F. & Crasquin-Soleau, S. (1995). Distributions des ostracodes et paléocurrantologie au Carbonifère terminal– Permien. Geobios, 18, 257272.Google Scholar
Levashova, N.M., Van der Voo, R., Abrajevitch, A. & Bazhenov, M.L. (2009). Paleomagnetism of mid-Paleozoic subduction-related volcanics from the Chingiz Ridge in NE Kazakhstan: the evolving paleogeography of the MALGAMATING Eurasian composite continent. Geological Society of America Bulletin, 121, 555573.Google Scholar
Leveridge, B.E & Shail, R.K. (2011). The marine Devonian stratigraphy of Great Britain. Proceedings of the Geologists’ Association, 122, 540567.Google Scholar
Li, J.Y. (2006). Permian geodynamic setting of Northeast China and adjacent regions: closure of the Paleo-Asian Ocean and subduction of the Paleo-Pacific Plate. Journal of Asian Earth Sciences, 26, 207224.Google Scholar
Li, Z.X., Bogdanova, S.V., Collins, A.S. et al. (2008). Assembly, configuration, and break-up history of Rodinia: a synthesis. Precambrian Research, 160, 179210.Google Scholar
Liu, L., Gurnis, M., Seton, M. et al. (2010). The role of oceanic plateau subduction in the Laramide orogeny. Nature Geoscience, 3, 353357, doi:10.1038/NGEO829.Google Scholar
Liu, L. & Stegman, D.R. (2012). Origin of Columbia River flood basalt controlled by propagating rupture of the Farallon slab. Nature, 482, 386390.Google Scholar
Lorenz, H., Männik, P., Gee, D. G. & Proskurnin, V. (2008). Geology of the Severnaya Zemlya Archipelago and new tectonic interpretation for the North Kara Terrane in the Russian high Arctic. International Journal of Earth Sciences, 97, 519547.Google Scholar
Lowry, D.P, Poulsen, C.J., Horton, D.E. et al. (2014). Thresholds for Paleozoic ice sheet initiation. Geology, 42(7), 627630.Google Scholar
Lyons, T.W., Reinhard, C.T. & Planavsky, N.J. (2014). The rise of oxygen in Earth’s early ocean and atmosphere. Nature, 506, 307315.Google Scholar
MacLeod, N., Rawson, P.E., Forey, P.L. et al. (1997). The Cretaceous–Tertiary biotic transition. Journal of the Geological Society, London, 154, 265292.Google Scholar
Mac Niocaill, C., van de Pluijm, B.A. & Van der Voo, R. (1997). Ordovician paleogeography and evolution of the Iapetus Ocean. Geology, 25, 159162.Google Scholar
Maloney, K.T., Clarke, G.L., Klepeis, K.A. & Quevedo, L. (2013). The Late Jurassic to present evolution of the Andean margin: drivers and the geological record. Tectonics, 32, 10491065.Google Scholar
Manankov, I.N., Shi, G.R. & Shen, S. (2006). An overview of Permian marine stratigraphy and biostratigraphy of Mongolia. Journal of Asian Earth Sciences, 26, 294303.Google Scholar
Mander, L., Kürschner, W.M. & McElwain, J.C. (2013). Palynostratigraphy and vegetation history of the Triassic–Jurassic transition in East Greenland. Journal of the Geological Society, London, 170, 3746.Google Scholar
Marcussen, C., Knudsen, C., Hopper, J.R. et al. (2015). Age and origin of the Lomonosov Ridge: a key continental fragment in Arctic Ocean reconstructions. Geophysical Research Abstracts, 17, EGU2015-10207-1.Google Scholar
Marshall, P.E., Widdowson, M. & Murphy, D.T. (2016). The Giant Lavas of Kalkarindji: rubbly pāhoehoe lava in an ancient continental flood basalt province. Palaeogeography, Palaeoclimatology, Palaeoecology, 441, 2237.Google Scholar
Martindale, R.C., Corsetti, F.A., James, N.P. & Bottjer, D.J. (2015). Paleogeographic trends in Late Triassic reef ecology from northeastern Panthalassa. Earth-Science Reviews, 142, 1837.Google Scholar
McCall, G.J.H. (2006). The Vendian (Ediacaran) in the geological record: enigmas in geology’s prelude to the Cambrian explosion. Earth-Science Reviews, 77, 1229.Google Scholar
McCann, T. (ed.) (2008). The Geology of Central Europe. London: Geological Society.Google Scholar
McHone, J.G. (2002). Volatile emissions of Central Atlantic Magmatic Province basalts: mass assumptions and environmental consequences. In Hames, W.E., McHone, J.G., Renne, P.R. & Ruppel, C. (eds.), The Central Atlantic Magmatic Province. American Geophysical Union, Geophysical Monograph, 136, pp. 241254.Google Scholar
McKenzie, P.M., Hughes, N.C., Myrow, P.M. et al. (2011). Trilobites and zircons link north China with the eastern Himalaya during the Cambrian. Geology, 39, 591594.Google Scholar
McKerrow, W.S. & Cocks, L.R.M. (1976). Progressive faunal migration across the Iapetus Ocean. Nature, 263, 304306.Google Scholar
McKerrow, W.S., Mac Niocaill, C., Ahlberg, P.E. et al. (2000). The Late Palaeozoic relations between Gondwana and Laurussia. In Franke, W., Haak, V., Oncken, O. & Tanner, D. (eds.), Orogenic Processes: Quantification and Modelling in the Variscan Belt. Geological Society, London, Special Publications, 179, 920.Google Scholar
McQuarrie, N. & van Hinsbergen, D.J.J. (2013). Retrodeforming the Arabia–Eurasia collision zone: age of collision versus magnitude of continental subduction. Geology, 41, 315318.Google Scholar
Meert, J.G. (2003). A synopsis of events related to the assembly of eastern Gondwana, Tectonophysics, 362, 140.Google Scholar
Meert, J.G. (2012). What’s in a name? The Columbia (Palaeopangea/Nuna) Supercontinent. Gondwana Research, 21, 987993.Google Scholar
Meert, J.G. (2014). Strange attractors, spiritual interlopers and lonely wanderers: the search for pre-Pangæan supercontinents. Geoscience Frontiers, 5, 155166.Google Scholar
Mei, S. & Henderson, C.M. (2001). Evolution of Permian conodont provincialism and its significance in global correlation and paleoclimate implication. Palaeogeography, Palaeoclimatology, Palaeoecology, 270, 217260.Google Scholar
Meng, J. & McKenna, M.C. (1998). Faunal turnover of Palaeogene mammals from the Mongolian Plateau. Nature, 394, 364367.Google Scholar
Metcalfe, I. (2006). Palaeozoic and Mesozoic tectonic evolution and palaeogeography of East Asian crustal fragments: the Korean Peninsula in context. Gondwana Research, 9, 2446.Google Scholar
Metcalfe, I. (2011). Palaeozoic–Mesozoic history of SE Asia. In Hall, R., Cottamm, M.A. & Wilson, E.J. (eds.), The SE Asian Gateway: History and Tectonics of the Australia–Asia Collision. Geological Society, London, Special Publications, 355, pp. 735.Google Scholar
Michard, A., Saddiqi, O., Chalouan, A. & Lamotte, D. F. (eds.) (2008). Continental Evolution: The Geology of Morocco. Berlin: Springer-Verlag.Google Scholar
Mojzsis, S.J., Cates, N.L., Bleeker, W. et al. (2014). Component geochronology of the ca. 3960 Ma Acasta Gneiss. Geochimica et Cosmochimica Acta, 133, 6896.Google Scholar
Montelli, R., Nolet, G., Dahlen, F. & Masters, G. (2006). A catalogue of deep mantle plumes: new results from finite-frequency tomography. Geochemistry, Geophysics, Geosystems, 7, Q11007, doi:10.1029/2006GC001248.Google Scholar
Moratti, G. & Chalouan, A. (eds.) (2006). Tectonics of the Western Mediterranean and North Africa. Geological Society, London, Special Publications, 262.Google Scholar
Moreno, T. & Gibbons, W. (eds.) (2007). The Geology of Chile. London: Geological Society.Google Scholar
Morgan, W.J. (1971). Convection plumes in the lower mantle. Nature, 230, 4243.Google Scholar
Mortimer, N., Herzer, R.H., Gans, P.B., Parkinson, D.L. & Seward, D. (1998). Basement geology from Three Kings Ridge to West Norfolk Ridge, southwest Pacific Ocean: evidence from petrology, geochemistry and isotopic dating of dredge samples. Marine Geology, 148, 135162.Google Scholar
Mortimore, R.N. (2011). A chalk revolution – what have we done to the chalk of England? Proceedings of the Geologists’ Association, 122, 232297.Google Scholar
Moulin, M., Aslanian, D. & Unternehr, P. (2010). A new starting point for the South and Equatorial Atlantic Ocean. Earth-Science Reviews, 98, 137.Google Scholar
Moullade, M. & Nairn, A.E.M. (eds.) (1983). The Mesozoic, B. The Phanerozoic Geology of the World II. Amsterdam: Elsevier.Google Scholar
Müller, R.D., Royer, J.-Y. & Lawver, L.A. (1993). Revised plate motions relative to the hotspots from combined Atlantic and Indian Ocean hotspot tracks. Geology, 21, 275278.Google Scholar
Müller, R.D., Sdrolias, M., Gaina, C. & Roest, W.R. (2008). Age, spreading rates, and spreading asymmetry of the world’s ocean crust. Geochemistry, Geophysics, Geosystems, 9, Q04006, doi:10.1029/2007GC001743.Google Scholar
Murphy, J.B, van Staal, C.R. & Keppie, J.D. (1999). Middle to Late Paleozoic Acadian orogeny in the northern Appalachians: a Laramide-style plume-related orogeny? Geology, 27, 653656.Google Scholar
Musteikis, P. & Cocks, L.R.M. (2004). Strophomenide and orthotetide Silurian brachiopods from the Baltic region, with particular reference to Lithuanian boreholes. Acta Palaeontologica Polonica, 49. 455482.Google Scholar
Myhre, A.M., Eldholm, O. & Sundvor, E. (1982). The margin between Senja and Spitsbergen fracture zones: implications from plate tectonics. Tectonophysics, 89(1–3), 3350.Google Scholar
Nance, R.D., Keppie, J.D., Miller, B.V. et al. (2009). Palaeozoic palaeogeography of Mexico: constraints from detrital zircon age data. In Murphy, J.B., Keppie, J.D. & Hynes, A. (eds.), Ancient Orogens and Modern Analogues. Geological Society, London, Special Publications, 327, pp. 239269.Google Scholar
Natal’in, B.A. & Şengör, A.M.C. (2005). Late Palaeozoic to Triassic evolution of the Turan and Scythian platforms: the pre-history of the Palaeo-Tethyan closure. Tectonophysics, 404, 175202.Google Scholar
Nelson, J. & Colpron, M. (2007). Tectonics and metallogeny of the British Columbia, Yukon and Alaskan Cordillera: 1.8 Ga to the present. In Goodfellow, W.D. (ed.), Mineral Deposits of Canada. Geological Association of Canada, Mineral Deposits Division, Special Publication, 5, pp. 755791.Google Scholar
Nielsen, K.C. (2005). Ouachitas. In Selley, R.C., Cocks, L.R.M. & Plimer, I.R. (eds.), Encyclopedia of Geology, Volume 4, Amsterdam: Elsevier, pp. 6171.Google Scholar
Nikishin, A.M., Ziegler, P.A., Stephenson, R.A. et al. (1996). Late Precambrian to Triassic history of the East European Craton: dynamics of sedimentary basin evolution. Tectonophysics, 268, 2363.Google Scholar
Nokleberg, W.J., Parfenov, I.M., Monger, J.W.H. et al. (2000). Phanerozoic Tectonic Evolution of the Circum-North Pacific. US Geological Survey Professional Paper, 1626.Google Scholar
Oakey, G.N. & Damaske, D. (2006). Continuity of basement structures and dyke swarms in the Kane Basin region of central Nares Strait constrained by aeromagnetic data. Polarforschung, 74, 5162.Google Scholar
Olierook, H.K.H., Merle, R.E., Jourdan, F. et al. (2015). Age and geochemistry of magmatism of the oceanic Wallaby Plateau and implications for the opening of the Indian Ocean. Geology, 43, 971974.Google Scholar
O’Neill, C., Lenardic, A., Moresi, L. et al. (2007). Episodic Precambrian Subduction. Earth and Planetary Science Letters, 262, 552562.Google Scholar
O’Neill, C., Müller, R.D. & Steinberger, B. (2005). On the uncertainties in hot spot reconstructions and the significance of moving hot spot reference frames. Geochemistry, Geophysics, Geosystems, 6, Q04003, doi:10.1929/2004GC000784.Google Scholar
Owen-Smith, T.M., Ashwal, L.D., Torsvik, T.H. et al. (2013). Seychelles alkaline suite records the culmination of Deccan Traps continental flood volcanism. Lithos, 182–183, 3347.Google Scholar
Oxman, V.S. (2003). Tectonic evolution of the Mesozoic Verkhoyansk–Kolyma belt (NE Asia). Tectonophysics, 365, 4576.Google Scholar
Parman, S.W. (2015). Time-lapse zirconography: imaging punctuated continental evolution. Geochemical Perspective Letters, 1, 4352.Google Scholar
Parrish, J.T. (1982). Upwelling and petroleum source beds, with reference to the Paleozoic. Bulletin – American Association of Petroleum Geologists, 66, 750774.Google Scholar
Pausata, F.S.R., Chafik, L., Caballero, R. & Battisti, D.S. (2015). Impacts of high-latitude volcanic eruptions on ENSO and AMOC. Proceedings of the National Academy of Science, 112, 1378413788.Google Scholar
Pegel, T.V. (2000). Evolution of trilobite biofacies in Cambrian basins of the Siberian Platform. Journal of Paleontology, 74, 10001017.Google Scholar
Percival, I.G. (1991). Late Ordovician articulate brachiopods from central New South Wales. Memoirs of the Society of Australasian Palaeontologists, 12, 107177.Google Scholar
Percival, I.G. & Glenn, R.A. (2007). Ordovician to earliest Silurian history of the Macquarie Arc, Lachlan Orogen, New South Wales. Australian Journal of Earth Sciences, 54, 143165.Google Scholar
Peron-Pinvidic, G., Gernigon, L., Gaina, C. & Ball, P. (2012). Insights from the Jan Mayen system in the Norwegian-Greenland Sea: II. Architecture of a microcontinent. Geophysics Journal International, 191, 413435.Google Scholar
Petterson, M.G., Babbs, T., Neal, C.R. et al. (1999). Geological–tectonic framework of Solomon Islands, SW Pacific: crustal accretion and growth within an intra-oceanic setting. Tectonophysics, 301, 3560.Google Scholar
Pirajno, F., Mao, J., Zhang, Z. & Chai, F. (2008). The association of mafic–ultramafic intrusions and A-type magmatism in the Tian Shan and Altay orogens, NW China: implications for geodynamic evolution and potential for the discovery of new ore deposits. Journal of Asian Earth Sciences, 32, 165183.Google Scholar
Plafker, G. & Berg, H.C. (eds.) (1994). The Geology of Alaska. The Geology of North America, Vol. G-1. Boulder: The Geological Society of America.Google Scholar
Popov, L.E., Bassett, M.G., Zhemchuzhnikov, V.G. et al. (2009). Gondwanan faunal signatures from early Palaeozoic terranes of Kazakhstan and central Asia. In Bassett, M.G. (ed.), Early Palaeozoic Peri-Gondwana Terranes. Geological Society, London, Special Publications, 325, pp. 2364.Google Scholar
Popov, L.E. & Cocks, L.R.M. (2017). Late Ordovician brachiopods from Kazakhstan, and their palaeogeography. Acta Geologica Polonica (in press).Google Scholar
Potter, A.W., Boucot, A.J., Bergström, S.M. et al. (1990). Early Paleozoic stratigraphic, paleogeographic, and biogeographic relations of the eastern Klamath belt, northern California. In Harwood, D. S. & Miller, M.M. (eds.), Paleozoic and Early Mesozoic Paleogeographic Relations; Sierra Nevada, Klamath Mountains, and Related Terranes. Geological Society of America Special Paper, 255, pp. 5774.Google Scholar
Pownall, J.M., Hall, R. & Watkinson, I.M. (2013). Extreme extension across Seram and Ambon, eastern Indonesia: evidence for Banda slab rollback. Solid Earth, 4, 277314.Google Scholar
Puchkov, V.N. (2009). The evolution of the Uralian orogen. In Murphy, J.B., Keppie, J.D. & Hynes, A. (eds.), Ancient Orogens and Modern Analogues. Geological Society, London, Special Publications, 327, pp. 161195.Google Scholar
Qiao, L. & Shen, S. (2014). Global paleobiogeography of brachiopods during the Mississippian – response to the lobal tectonic configuration, ocean circulation, and climate changes. Gondwana Research, 26, 11731185.Google Scholar
Quintaville, M., Tongiorgi, M. & Gaetani, M. (2000). Lower to Middle Ordovician acritarchs and chitinozoans from northern Karakorum Mountains, Pakistan. Rivista Italiana di Paleontologia e Stratigrafia, 106, 318.Google Scholar
Radley, J.D. & Allen, P. (2012). The non-marine Lower Cretaceous Wealden strata of southern England. Proceedings of the Geologists’ Association, 123, 235385.Google Scholar
Ramos, V.A. & Folguera, A. (2009). Andean flat-slab subduction through time. In Murphy, J.B., Keppie, J.D. & Hynes, A. (eds.), Ancient Orogens and Modern Analogues, Geological Society, London, Special Publications, 327, pp. 3154.Google Scholar
Rasmussen, C.M.Ø., Ullmann, C.V., Jakobsen, K.G. et al. (2016). Onset of main Phanerozoic marine radiation sparked by emerging Mid Ordovician icehouse. Nature, Scientific Reports, doi:10.1038/srep18884.Google Scholar
Raymo, M.E., Ruddiman, W.F. & Froelich, P.N. (1988). Influence of late Cenozoic mountain building on ocean geochemical cycles. Geology, 16, 649653.Google Scholar
Rees, P.M. (2002). Land-plant diversity and the end-Permian mass extinction. Geology, 30, 827830.Google Scholar
Rees, P.M., Noto, C.R., Parrish, J.M. & Parrish, J.T. (2004). Late Jurassic climates, vegetation and dinosaur distributions. Journal of Geology, 112, 643653.Google Scholar
Reidel, S.P., Camp, V.E., Tolan, T.L. & Martin, B.S. (2013). The Columbia River flood basalt province: stratigraphy, areal extent, volume, and physical volcanology. In Reidel, S.P. et al. (eds.), The Columbia River Flood Basalt Province, Geological Society of America Special Paper, 497, pp. 143.Google Scholar
Retallack, G.J. (2015). Silurian vegetation structure and density inferred from fossil soils and plants in Pennsylvania, USA. Journal of the Geological Society, London, 172, 693709.Google Scholar
Ricci, J., Quidelleur, X., Pavlov, V. et al. (2013). New 40Ar/39Ar and K–Ar ages of the Viluy traps (Eastern Siberia): further evidence for a relationship with the Frasnian–Famennian mass extinction. Palaeogeography, Palaeoclimatology, Palaeoecology, 386, 531540.Google Scholar
Ridd, M.F., Barber, A.J. & Crowe, M.J. (eds.) (2011). The Geology of Thailand. London: Geological Society.Google Scholar
Riefstahl, F., Estrada, S., Geissler, W.H. et al. (2013). Provenance and characteristics of rocks from the Yermak Plateau, Arctic Ocean: petrographic, geochemical and geochronological constraints. Marine Geology, 343, 125145.Google Scholar
Ritsema, J. & Allen, R.M. (2003). The elusive mantle plume. Earth and Planetary Science Letters, 207, 112.Google Scholar
Roberts, N.M. & Spencer, C.J. (2014). The zircon archive of continent formation through time. In Roberts, N.M.W. et al. (eds.), Continent Formation through Time. Geological Society, London, Special Publications, 389, pp. 197225.Google Scholar
Rocha-Campos, A.C., Santos, P.R.D. & Canuto, J.R. (2008). Late Paleozoic glacial deposits of Brazil: Paraná Basin. In Fielding, C.R., Frank, T.D. & Isbell, J.L. (eds.), Resolving the Late Paleozoic Ice Age in Time and Space. Geological Society of America Special Paper, 441, pp. 97114.Google Scholar
Rong, J., Boucot, A.J., Su, Y. & Strusz, D.L. (1995). Biogeographical analysis of Late Silurian brachiopod faunas, chiefly from Asia and Australia. Lethaia, 28, 3960.Google Scholar
Rong, J., Chen, X., Su, Y. et al. (2003). Silurian paleogeography of China. New York State Museum Bulletin, 493, 243298.Google Scholar
Rong, J. & Harper, D.A.T. (1988). A global synthesis of the latest Ordovician Hirnantian brachiopod fauna. Transactions of the Royal Society of Edinburgh, Earth Sciences, 79, 383401.Google Scholar
Ross, C.A. & Ross, J.R.P. (1983). Late Paleozoic accreted terranes of western North America. In Stevens, C.H. (ed.), Pre-Jurassic Rocks in Western American Suspect Terranes, Los Angeles: SEPM Pacific Section, pp. 722.Google Scholar
Royer, D.L. (2006). CO2-forced climate thresholds during the Phanerozoic. Geochimica et Cosmochimica Acta, 70, 56655675.Google Scholar
Royer, D.L., Berner, R.A. & Park, J. (2007). Climate sensitivity constrained by CO2 concentrations over the past 420 million years. Nature, 446, 530532.Google Scholar
Rozman, K.S. (1978). Brachiopods of the Obikolon Beds. USSR Academy of Sciences Siberian Branch Institute of Geology and Geophysics Transactions, 397, 75101 (in Russian).Google Scholar
Ruban, D.A., al-Husseini, M.L. & Iwasaki, Y. (2007). Review of Middle East plate tectonics. GeoArabia, 12, 3556.Google Scholar
Ruddiman, W.F. (2014). Earth’s Climate Past and Future, 3rd edn. New York: W.H. Freeman.Google Scholar
Rushton, A.W.A., Cocks, L.R.M. & Fortey, R.A. (2002). Upper Cambrian trilobites and brachiopods from Severnaya Zemlya, Arctic Russia, and their implications for correlation and biogeography. Geological Magazine, 139, 281290.Google Scholar
Sahney, S., Benton, M.J. & Falcon-Lang, H.J. (2010). Rainforest collapse triggered Carboniferous tetrapod diversification in Euramerica. Geology, 38, 10791082.Google Scholar
Savage, R.J.G. & Long, M.R. (1986). Mammal Evolution: An Illustrated Guide. London: British Museum (Natural History).Google Scholar
Scarrow, J.H., Ayala, C. & Kimball, G.S. (2002). Insights into orogenesis: getting to the root of a continental-ocean–continent collision. Journal of the Geological Society, London, 159, 659671.Google Scholar
Schallreuter, R. & Siveter, D.J. (1985). Ostracodes across the Iapetus Ocean. Palaeontology, 28, 577598.Google Scholar
Schandelmeier, H. & Reynolds, P.O. (1997). Palaeogeographic–Palaeotectonic Atlas of North-Eastern \Africa, Arabia, and Adjacent Areas. Rotterdam: Balkema.Google Scholar
Schmid, S.M., Bernoulli, D., Fügenschuh, B. et al. (2008). The Alpine–Carpathian–Dinaridic orogenic system: correlation and evolution of tectonic units. Swiss Journal of Geosciences, 101, 139183.Google Scholar
Scotese, C.R. & Barrett, S.F. (1990). Gondwana’s movement over the South Pole during the Palaeozoic: evidence from lithological indicators of climate. In McKerrow, W.S. & Scotese, C.R. (eds.), Palaeozoic Palaeogeography and Biogeography. Geological Society, London, Memoir, 12, pp. 7585.Google Scholar
Searle, M.P. (2013). Colliding Continents: A Geological Exploration of the Himalaya, Karakoram, and Tibet. Oxford: Oxford University Press.Google Scholar
Searle, M.P., Cherry, A.G., Ali, M.Y. & Cooper, D.J.W. (2014). Tectonics of the Musandam Peninsula and northern Oman Mountains: from ophiolite obduction to continental collision. GeoArabia, 19, 135174.Google Scholar
Sedgwick, A. & Murchison, R.I. (1835). On the Cambrian and Silurian systems, exhibiting the order in which the older sedimentary strata succeed each other in England and Wales. The London and Edinburgh Philosophical Magazine and Journal of Science, 7, 483–5.Google Scholar
Sedgwick, A. & Murchison, R.I. (1837). A classification of the old slate rocks of the north of Devonshire. Report of the British Association for the Advancement of Science (for 1836), 95–96.Google Scholar
Sedlock, R.L. (2003). Geology and tectonics of the Baja California peninsula and adjacent areas. In Johnson, S.E., Paterson, S.R., Fletcher, J.M. et al. (eds.), Tectonic Evolution of Northwestern Mexico and the Southwestern USA. Geological Society of America Special Paper, 374, pp. 142.Google Scholar
Selley, R.C., Cocks, L.R.M. & Plimer, I.R. (eds.) (2005). Encyclopedia of Geology. Amsterdam: Elsevier.Google Scholar
Şengör, A.M.C. & Atayman, S. (2009). The Permian Extinction and the Tethys. Geological Society of America Special Paper, 448.Google Scholar
Şengör, A.M.C. & Natal’in, B.A. (1996). Paleotectonics of Asia: fragments of a synthesis. In Yin, A. & Harrison, M. (eds.), The Tectonic Evolution of Asia. Cambridge: Cambridge University Press, pp. 486646.Google Scholar
Sennikov, N.V. (2003). Ordovician events in Altai–Sayan–Kuznesty and Tuva basins and their influence on the sedimentary facies and marine biota (Siberia, Russia). INSUGEO Serie Correlación Geológica, 17, 461465.Google Scholar
Seton, M., Gaina, C., Müller, R.D. & Heine, C. (2009). Mid-Cretaceous seafloor spreading pulse: fact or fiction? Geology, 37, 687690.Google Scholar
Seton, M., Flament, N., Whittaker, J. et al. (2015). Ridge subduction sparked reorganization of the Pacific plate–mantle system 60–50 million years ago. Geophysical Research Letters, 42, 17321740.Google Scholar
Seton, M., Müller, R.D., Zahirovic, S. et al. (2012). Global continental and ocean basin reconstructions since 200 Ma. Earth-Science Reviews, 113, 212270.Google Scholar
Shao, L., Zhang, P., Gayer, R.A. et al. (2003). Coal in a carbonate sequence stratigraphic framework: the Upper Permian Heshan Formation in central Guangxi, southern China. Journal of the Geological Society, London, 160, 285298.Google Scholar
Shaw, J., Johnston, S.T., Gutiérrez-Alonso, G., et al. (2012). Oroclines of the Variscan orogeny of Iberia: paleocurrent analysis and paleogeographic implications. Earth and Planetary Science Letters, 329–330, 6070.Google Scholar
Shelley, D. & Bossière, G. (2000). A new model for the Hercynian orogeny of Gondwanan France and Iberia. Journal of Structural Geology, 22, 757776.Google Scholar
Shellnutt, J.G., Bhat, G.M., Brookfield, M.E., & Jahn, B.M. (2011). No link between the Panjal Traps (Kashmir) and the Late Permian mass extinctions. Geophysical Research Letters, 38, L19308.Google Scholar
Shen, S., Xie, J., Zhang, H. & Shi, G.R. (2009). Roadian–Wordian (Guadalupian, Middle Permian) global palaeobiogeography. Global and Planetary Change, 65, 166181.Google Scholar
Shephard, G., Flament, N., Williams, S. et al. (2014). Circum-Arctic mantle structure and long-wavelength topography since the Jurassic. Journal of Geophysical Research, 119, 78897908, doi:10.1002/2014JB011078.Google Scholar
Shephard, G., Müller, R.D. & Seton, M. (2013). The tectonic evolution of the Arctic since Pangea breakup: integrating constraints from surface geology and geophysics with mantle structure. Earth-Science Reviews, 124, 148183.Google Scholar
Shergold, J.H. (1988). Review of trilobite biofacies distributions at the Cambrian–Ordovician boundary. Geological Magazine, 125, 363380.Google Scholar
Shergold, J.H. (1991). Late Cambrian (Payntonian) and Early Ordovician (Late Warendian) trilobite faunas of the Amadeus Basin Central Australia. Bulletin of the Bureau of Mineral Resources, Geology and Geophysics, 237, 1575.Google Scholar
Shi, G.R. (2006). The marine Permian of east and northeast Asia: an overview of biostratigraphy, palaeobiogeography and palaeogeographical implications. Journal of Asian Earth Sciences, 26, 175206.Google Scholar
Shirey, S.B. & Richardson, S.H. (2011). Start of the Wilson cycle at 3 Ga shown by diamonds from subcontinental mantle. Science, 333, 434436.Google Scholar
Sigloch, K. & Mihalynuk, M. G. (2013). Intra-oceanic subduction shaped the assembly of Cordilleran North America. Nature, 496, 5056, doi:10.1038/nature12019.Google Scholar
Silva, D.R.A., Mizusaki, A.M.P., Milani, E. & Pimentel, M. (2012). Determination of depositional age of Paleozoic and pre-rift supersequences of the Recôncavo Basin in northeastern Brazil by applying Rb–Sr radiometric dating technique to sedimentary rocks. Journal of South American Earth Sciences, 37, 1324.Google Scholar
Sleep, N.H. (1997). Lateral flow and ponding of starting plume material. Journal of Geophysical Research, 102, 1000110012.Google Scholar
Smelror, M., Petrov, O.V., Larssen, G.B. & Werner, S. (eds.) (2009). Geological History of the Barents Sea. Trondheim: Geological Survey of Norway.Google Scholar
Smith, M.P. & Rasmussen, J.A. (2008). Cambro-Silurian development of the Laurentian margin of the Iapetus Ocean in Greenland and related areas. In Higgins, A.K., Gilotti, J.A. & Smith, M.P. (eds.), The Greenland Caledonides: Evolution of the Northeast Margin of Laurentia. Geological Society of America Memoir, 202, pp. 137167.Google Scholar
Sone, M. & Metcalfe, I. (2008). Parallel Tethyan sutures in mainland Southeast Asia: new insights for Palaeo-Tethys closure and implications for the Indosinian orogeny. Comptes Rendus Geoscience, 340, 166179.Google Scholar
Song, S., Su, L., Niu, Y. et al. (2009b). Tectonic evolution of Early Paleozoic HP metamorphic rocks in the North Qilian Mountains, NW China: new perspectives. Journal of Asian Earth Sciences, 35, 334353.Google Scholar
Spencer, A.M., Embry, A.F., Gautier, D.L. et al. (eds.) (2011). Arctic Petroleum Geology. Geological Society, London, Memoirs, 35.Google Scholar
Speranza, F., Minelli, L., Pignatelli, A. & Chiappini, M. (2012). The Ionian Sea: the oldest in situ ocean fragment of the world? Journal of Geophysical Research, 117, B12101, doi:10.1029–2012JB009475.Google Scholar
Stampfli, G.M. & Borel, G.D. (2004). The TRANSMED transects in space and time: constraints on the paleotectonic evolution of the Mediterranean domain. In Cavazza, W., Roure, F., Spakman, W., Stampfli, G.G. & Ziegler, P.A. (eds.), The TRANSMED Atlas: The Mediterranean Region from Crust to Mantle. Berlin: Springer, pp. 5390.Google Scholar
Steinberger, B. (2000). Plumes in a convecting mantle: models and observations for individual hotspots. Journal of Geophysical Research, 105, 11,12711,152.Google Scholar
Steinberger, B. & Gaina, C. (2007). Plate-tectonic reconstructions predict part of the Hawaiian hotspot track to be preserved in the Bering Sea. Geology, 35, 407410.Google Scholar
Steinberger, B., Spakman, W., Japsen, P. & Torsvik, T.H. (2015). The key role of global solid-Earth processes in preconditioning Greenland’s glaciation since the Pliocene. Terra Nova, 27, 18.Google Scholar
Steinberger, B., Sutherland, R. & O’Connell, R.J. (2004). Prediction of Emperor–Hawaii seamount locations from a revised model of global plate motion and mantle flow. Nature, 430, 167173.Google Scholar
Steinberger, B. & Torsvik, T.H. (2008). Absolute plate motions and true polar wander. Nature, 452, 620623.Google Scholar
Steinberger, B. & Torsvik, T.H. (2010). Toward an explanation for the present and past locations of the poles. Geochemistry, Geophysics, Geosystems, 11, Q06W06, doi:10.1929/2009GC002889.Google Scholar
Stemmerik, B. (2000). Late Palaeozoic evolution of the North Atlantic margin of Pangea. Palaeogeography, Palaeoclimatology, Palaeoecology, 161, 95126.Google Scholar
Stern, R.J. (2008). Modern-style plate tectonics began in Neoproterozoic time: an alternative interpretation of Earth’s tectonic history. In Condie, K. & Pease, V. (eds.), When Did Plate Tectonics Begin? Geological Society of America Special Paper, 440, pp. 265280.Google Scholar
Stevens, C.H. & Stone, P. (2007). The Pennsylvanian–Early Permian Bird Spring Carbonate Shelf, Southeastern California: Fusulinid Biostratigraphy, Paleogeographic Evolution, and Tectonic Implications. Geological Society of America Special Paper, 429.Google Scholar
Stevens, L.G., Hilton, J., Bond, D.P.G. et al. (2011). Radiation and extinction patterns in Permian floras from North China as indicators for environmental and climate change. Journal of the Geological Society, London, 168, 607619.Google Scholar
Stringer, C.B. (2002). Modern human origin: progress and prospects. Philosophical Transactions of the Royal Society, London, B357, 563579.Google Scholar
Svensen, H., Planke, S. & Corfu, F. (2010). Zircon dating ties NE Atlantic sill emplacement to initial Eocene global warming. Journal of the Geological Society, London, 167, 433436.Google Scholar
Svensen, H., Planke, S., Malthe-Sorenssen, A. et al. (2004). Release of methane from a volcanic basin as a mechanism for initial Eocene global warming. Nature, 429, 542545.Google Scholar
Svensen, H., Hammer, Ø. & Corfu, F. (2015). Astronomically forced cyclicity in the Upper Ordovician and U–Pb ages of interlayered tephra, Oslo Region, Norway. Palaeogeography, Palaeoclimatology, Palaeoecology, 418, 150159.Google Scholar
Szederkényi, T., Haas, N. & Hámor, G. (2012). Geology and history of evolution of the Tisza Unit. In Haas, J., (ed.), Geology of Hungary. Berlin: Springer-Verlag.Google Scholar
Tankard, A.J., Suárez-Soruco, R. & Welsink, H.J. (eds.) (1995). Petroleum Basins of South America. American Association of Petroleum Geologists Memoir, 6.Google Scholar
Tarduno, J., Bunge, H.-P., Sleep, N. & Hansen, U. (2009). The bent Hawaiian–Emperor hotspot track: inheriting the mantle wind. Science, 324, 5053.Google Scholar
Tarduno, J.A., Duncan, R.A., Scholl, D.W. et al. (2003). The Emperor seamounts: southward motion of the Hawaiian hotspot plume in Earth’s mantle. Science, 301, 10641069.Google Scholar
Tarduno, J.A., Brinkman, D.B., Renne, P.R. et al. (1998). Late Cretaceous Arctic volcanism: tectonic and climatic connections. In American Geophysical Union Spring Meeting Abstracts, Washington, DC: American Geophysical Union.Google Scholar
Tauxe, L. & Kent, D.V. (2004). A simplified statistical model for the geomagnetic field and the detection of shallow bias in paleomagnetic inclinations: was the ancient magnetic field dipolar? In Channell, J.E.T. et al. (eds.), Timescales of the Paleomagnetic Field, 145, Washington, DC: American Geophysical Union, pp. 101116.Google Scholar
Taylor, T.N., Taylor, E.L. & Krings, M. (2009). Paleobotany: The Biology and Evolution of Fossil Plants. Amsterdam: Academic Press.Google Scholar
Tejada-Lara, J.V., Salas-Gismondi, R., Pujos, F. et al. (2015). Life in Proto-Amazonia: Middle Miocene mammals from the Fitzcarrald Arch (Peruvian Amazonia). Palaeontology, 58, 341378.Google Scholar
Tessensohn, F. & Henjes-Kunst, F. (2005). Northern Victoria Land terranes, Antarctica: far-travelled or local products? In Vaughan, A.P.M., Leat, P.T. & Pankhurst, R.J. (eds.), Terrane Processes at the Margins of Gondwana. Geological Society, London, Special Publications, 246, 275291.Google Scholar
Torsvik, T.H. (2003). The Rodinia jigsaw puzzle. Science, 300, 13791381.Google Scholar
Torsvik, T.H., Amundsen, H., Hartz, E.H., et al. (2013). A Precambrian microcontinent in the Indian Ocean. Nature Geoscience, 6, 223227.Google Scholar
Torsvik, T.H., Amundsen, H.E.F., Trønnes, R.G. et al. (2015). Continental crust beneath southeast Iceland. Proceedings of the National Academy of Sciences, 112, E1818–E1827, doi:10.1073/pnas.1423099112.Google Scholar
Torsvik, T.H. & Andersen, T.B. (2002). The Taimyr fold belt, Arctic Siberia: timing of pre-fold remagnetization and regional tectonics. Tectonophysics, 352, 335348.Google Scholar
Torsvik, T.H., Burke, K., Steinberger, B. et al. (2010). Diamonds sourced by plumes from the core–mantle boundary. Nature, 466, 352355.Google Scholar
Torsvik, T.H., Carter, L.M., Ashwal, L.D. et al. (2001). Rodinia refined or obscured: palaeomagnetism of the Malani Igneous Suite (NW India). Precambrian Research, 108, 319333.Google Scholar
Torsvik, T.H. & Cocks, L.R.M. (2004). Earth geography from 400 to 250 Ma: a palaeomagnetic, faunal and facies review. Journal of the Geological Society, London, 161, 555572.Google Scholar
Torsvik, T.H. & Cocks, L.R.M. (2005). Norway in space and time: a centennial cavalcade. Norwegian Journal of Geology, 85, 7386.Google Scholar
Torsvik, T.H. & Cocks, L.R.M. (2009). The Lower Palaeozoic palaeogeographical evolution of the northeastern and eastern peri-Gondwanan margin from Turkey to New Zealand. In Bassett, M.G. (ed.), Early Palaeozoic Peri-Gondwana Terranes. Geological Society, London, Special Publications, 325, pp. 321.Google Scholar
Torsvik, T.H. & Cocks, L.R.M. (2011). The Palaeozoic geography of central Gondwana. In van Hinsbergen, D.J.J., Buiter, S.J.H., Torsvik, T.H. et al. (eds.), The Formation and Evolution of Africa: A Synopsis of 3.8 Ga of Earth History. Geological Society, London, Special Publications, 357, pp. 137166.Google Scholar
Torsvik, T.H. & Cocks, L.R.M. (2012). From Wegener until now: the development of our understanding of Earth’s Phanerozoic evolution. Geologica Belgica, 15, 181192.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., Gaina, C. & Redfield, T.F. (2008a). Antarctica and global paleogeography: from Rodinia, through Gondwanaland and Pangea, to the birth of the Southern Ocean and the opening of gateways. In Cooper, A.K., Barrett, P., Stagg, H. et al. (eds.), Antarctica, a Keystone in a Changing World, Washington DC: National Academies Press, pp. 125129.Google Scholar
Torsvik, T.H., Müller, R.D., Van der Voo, R. et al. (2008b). Global plate motion frames: towards a unified model. Reviews of Geophysics, 46, RG3004, doi:10.1029/2007RG000227.Google Scholar
Torsvik, T.H. & Rehnström, E.F. (2003). The Tornquist Sea and Baltica–Avalonia docking. Tectonophysics, 362, 6782.Google Scholar
Torsvik, T.H., Rousse, S., Labails, C. & Smethurst, M.A. (2009). A new scheme for the opening of the South Atlantic Ocean and the dissection of an Aptian salt basin. Geophysical Journal International, 177, 13151333.Google Scholar
Torsvik, T.H., Smethurst, M.A., Burke, K. & Steinberger, B. (2006). Large Igneous Provinces generated from the margins of the Large Low Velocity Provinces in the deep mantle. Geophysical Journal International, 167, 14471460.Google Scholar
Torsvik, T.H., Smethurst, M.A., Meert, J.G. et al. (1996). Continental break-up and collisions in the Neoproterozoic and Palaeozoic: a tale of Baltica and Laurentia. Earth-Science Reviews, 40, 229258.Google Scholar
Torsvik, T.H., Steinberger, B., Cocks, L.R.M. & Burke, K. (2008c). Longitude: linking Earth’s ancient surface to its deep interior. Earth and Planetary Science Letters, 276, 273282.Google Scholar
Torsvik, T.H., Van der Voo, R., Preeden, V. et al. (2012). Phanerozoic polar wander, palaeogeography, and dynamics. Earth-Science Reviews, 114, 325368.Google Scholar
Torsvik, T.H., Van der Voo, R., Doubrovine, P.V. et al. (2014). Deep mantle structure as a reference frame for movements in and on the Earth. Proceedings of the National Academy of Sciences, 111, 24, 87358740.Google Scholar
Trettin, H.P. (1998). Pre-Carboniferous geology of the northern part of the Arctic Islands. Geological Survey of Canada Bulletin, 425, 1401.Google Scholar
Tucker, R.D., Ashwal, L.D. & Torsvik, T.H. (2001). U–Pb geochronology of Seychelles granitoid: Neoproterozoic construction of a Rodinia continental fragment. Earth and Planetary Science Letters, 187, 2738.Google Scholar
Tucker, R.D. & McKerrow, W.S. (1995). Early Palaeozoic chronology: a review in light of new U–Pb zircon ages from Newfoundland and Britain. Canadian Journal of Earth Sciences, 32, 368379.Google Scholar
Tull, J.F., Barineau, C.L., Mueller, P.A. & Wooden, J.L. (2007). Volcanic arc emplacement onto the southernmost Appalachian Laurentian shelf: characteristics and constraints. Geological Society of America Bulletin, 119, 261274.Google Scholar
Tomurtogoo, O., Windley, B.F., Kröner, A., Badarch, G., Liu, D.Y. (2005). Zircon age and occurrence of the Adaatsag ophiolite and Muron shear zone, central Mongolia: constraints on the evolution of the Mongol–Okhotsk ocean, suture and orogen. Journal of the Geological Society, London, 162, 125134.Google Scholar
Ustaszewski, K., Schmid, S.M., Fügenschuh, B. et al. (2008). A map-view restoration of the Alpine–Carpathian–Dinaridic system for the Early Miocene. Swiss Journal of Geosciences, 101, 273294.Google Scholar
van der Meer, D.G., Spakman, W., van Hinsbergen, D.J.J. et al. (2010). Towards absolute plate motions constrained by lower mantle slab remnants. Nature Geoscience, 3, 3640.Google Scholar
van der Meer, D.G., Torsvik, T.H., Spakman, W. et al. (2012). Intra-Panthalassa Ocean subduction zones revealed by fossil arcs and mantle structure. Nature Geoscience, 5, 215219, doi:10.1038/NGEO1401.Google Scholar
van der Meer, D.G., Zeebe, R., van Hinsbergen, D.J.J. et al. (2014). Long-term trends in atmospheric CO2 levels over the past 250 million years driven by plate tectonic volcanic degassing. Proceedings of the National Academy of Sciences, 111, 43804385, doi:10.1073/pnas.1315657111.Google Scholar
Van der Voo, R. (1993). Paleomagmatism of the Atlantic, Tethys and Iapetus Oceans. Cambridge: Cambridge University Press.Google Scholar
Van der Voo, R., van Hinsbergen, D.J.J., Domeier, M. et al. (2015). Latest Jurassic–earliest Cretaceous oroclinal closure of the Mongol–Okhotsk Ocean and implications for Mesozoic Central Asian plate reconstructions. In Anderson, T.H., Didenko, A.N., Johnson, C.L., Khanchuk, A.I. & MacDonald, J.H. Jr. (eds.), Late Jurassic Margin of Laurasia: A Record of Faulting Accommodating Plate Rotation, Geological Society of America Special Paper, 513, doi:10.1130/2015.2513(19).Google Scholar
van Hinsbergen, D.J.J., Buiter, S.J.H., Torsvik, T.H., et al. (eds.) (2011). The Formation and Evolution of Africa: A Synopsis of 3.8 Ga of Earth History. Geological Society, London, Special Publications, 357.Google Scholar
van Hinsbergen, D.J.J., Lippert, P.C., Dupont-Nivet, G. et al. (2012). Greater India Basin hypothesis and a two-stage Cenozoic collision between India and Asia. Proceedings of the National Academy of Sciences, 109, 76597664, doi:10.1073/pnas.1117262109.Google Scholar
Van Roy, P., Daley, A.C., Briggs, D.E.G. (2015). Anomalocaridid trunk limb homology revealed by a giant filter-feeder with paired flaps. Nature, 522, 7780.Google Scholar
van Staal, C.R., Whalen, J.B., McNicol, V.J. et al. (2007). The Notre Dame Arc and the Taconic Orogeny in Newfoundland. In Hatcher, R.D. et al. (eds.), 4-D Framework of Continental Crust. Geological Society of America Memoir, 200, pp. 511552.Google Scholar
van Staal, C.R., Whalen, J.B., Vaquero, P.V. et al. (2009). Pre-Carboniferous episodic accretion-related orogenesis along the Laurentian margin of the northern Appalachians. In Murphy, J.B., Keppie, J.D. & Hynes, A. (eds.), Ancient Orogens and Modern Analogues. Geological Society, London, Special Publications, 327, pp. 271316.Google Scholar
Van Wagoner, N.A., Leybourne, N.I., Dadd, K.A. et al. (2002). Late Silurian bimodal volcanism of southwestern New Brunswick, Canada: products of continental extension. Geological Society of America Bulletin, 114, 400418.Google Scholar
Vaughan, A.P.M., Leat, P.J. & Pankhurst, R.J. (eds.) (2005). Terrane Processes at the Margins of Gondwana. Geological Society, London, Special Publications, 246.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
Veizer, J., Godderis, Y. & François, L.M. (2000). Evidence for decoupling of atmospheric CO2 and global climate during the Phanerozoic eon. Nature, 408, 698701.Google Scholar
Villas, E., Vizcaïno, D., Álvaro, J.J., Destombes, J. & Vennin, E. (2006). Biostratigraphic control of the latest-Ordovician glaciogenic unconformity in Alnif (Eastern Anti-Atlas, Morocco), based on brachiopods. Geobios, 39, 727737.Google Scholar
Vine, F.J. & Matthews, D.H. (1963). Magnetic anomalies over oceanic ridges. Nature, 199, 947949.Google Scholar
Vissers, R.L.M. &. Meijer, P.Th. (2012). Iberian Plate kinematics and Alpine collision in the Pyrenees. Earth-Science Reviews, 114, 6183.Google Scholar
Walderhaug, H.J., Eide, E.A., Scott, R.A., Inger, S. & Golionko, E.G. (2005). Palaeomagnetism and 40Ar/39Ar geochronology from the South Taimyr igneous complex, Arctic Russia: a Middle–Late Triassic magmatic pulse after Siberian flood-basalt volcanism. Geophysical Journal, 163, 501517.Google Scholar
Waldron, J.W.F., Schofield, D.I, White, C.E. & Barr, S.M. (2013). 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. & van Staal, C.R. (2001). Taconian orogeny and the accretion of the Dashwoods block: a peri-Laurentian microcontinent in the Iapetus Ocean. Geology, 29, 811814.Google Scholar
Wang, B., Chen, Y., Zhan, S. et al. (2007). Primary Carboniferous and Permian paleomagnetic results from the Yili Block (NW China) and their implications on the geodynamic evolution of Chinese Tianshan Belt. Earth and Planetary Science Letters, 263, 288308.Google Scholar
Watkins, R. (1994). Evolution of Silurian pentamerid communities in Wisconsin. Palaios, 9, 488499.Google Scholar
Webby, B.D., Paris, F., Droser, M.L. & Percival, I.G. (eds.) (2004). The Great Ordovician Biodiversification Event. New York: Columbia University Press.Google Scholar
Wegener, A. (1912). Die Entstehung der Kontinente. Dr. A. Petermanns Mitteilungen aus Justus Perthes geographischer Anstalt, 58, 185195, 253–256, 305–309.Google Scholar
Wegener, A. (1915). Die Entstehung der Kontinente und Ozeane. Brunswick: Vieweg.Google Scholar
Wellman, C.H. & Strother, P.K. (2015). The terrestrial biota prior to the origin of land plants (Embryophytes): a review of the evidence. Palaeontology, 58, 601627.Google Scholar
Wignall, P.B. (2007). The end-Permian mass extinction: how bad did it get? Geobiology, 5, 303309.Google Scholar
Wignall, P.B. & Bond, D.P.G. (2008). The end-Triassic and Early Jurassic mass extinction records of the British Isles. Proceedings of the Geologists’ Association, 119, 7384.Google Scholar
Wilde, S.A., Valley, J.W., Peck, W.H., Graham, C.M. (2001). Evidence from detrital zircons for the existence of continental crust and oceans on the Earth 4.4 Gyr ago. Nature, 409, 175178.Google Scholar
Willem, C., Windley, B.F. & Stampfli, G.M. (2012). The Altaids of Central Asia: a preliminary innovative review. Earth-Science Reviews, 113, 303341.Google Scholar
Wilson, J.T. (1963). A possible origin of the Hawaiian islands. Canadian Journal of Physics, 41, 863870.Google Scholar
Windley, B.F., Alexeiev, D., Xiao, W. et al. (2007). Tectonic models for accretion of the Central Asian Orogenic Belt. Journal of the Geological Society, London, 164, 3147.Google Scholar
Wright, A.J., Young, G.C., Talent, J.A. & Laurie, J.R., (2000). Paleobiogeography of Australasian faunas and floras. Memoirs of the Association of Australasian Palaeontologists, 23, 1515.Google Scholar
Wright, J.E. & Wyld, S.J. (2006). Gondwanan, Iapetan, Cordilleran interactions: a geodynamic model for the Paleozoic tectonic evolution of the North American Cordillera. In Haggart, J., Enkin, R.J. & Monger, J.W.H. (eds.), Paleogeography of the North American Cordillera. Geological Association of Canada Special Paper, 46, pp. 377408.Google Scholar
Wu, F.Y., Sun, D.Y., Ge, W.C. et al. (2011). Geochronology of the Phanerozoic granitoids in northeastern China. Journal of Asian Earth Sciences, 41, 130.Google Scholar
Xiao, W., Han, C., Uan, C., et al. (2008). Middle Cambrian to Permian subduction-related accretionary orogenesis of Northern Xinjiang, NW China: implications for the tectonic evolution of central Asia. Journal of Asian Earth Sciences, 32, 102117.Google Scholar
Xiao, W., Windley, B.F., Hao, J. & Li, J. (2002). Arc-ophiolite obduction in the Western Kunlun Range (China): implications for the Palaeozoic crustal evolution of central Asia. Journal of the Geological Society, London, 159, 517528.Google Scholar
Xiao, W., Windley, B.F., Yong, Y. et al. (2009a). Early Palaeozoic to Devonian multiple-accretionary model for the Qilian-Shan, NW China. Journal of Asian Earth Sciences, 35, 323333.Google Scholar
Xiao, W., Windley, B.F., Yuan, C., et al. (2009b). Paleozoic multiple subduction–accretion processes of the southern Altaids. American Journal of Science, 309, 221270.Google Scholar
Yan, Z., Xiao, W., Windley, B.F., Wang, Z.Q. & Li, J.L. (2010). Silurian clastic sediments in the North Qilian Shan, NW China: chemical and isotopic constraints on their forearc provenance with implications for the Paleozoic evolution of the Tibetan Plateau. Sedimentary Geology, 231, 98114.Google Scholar
Yanev, S., Göncüoğlu, M.C., Gedik, I. et al. (2006). Stratigraphy, correlations and palaeogeography of Palaeozoic terranes of Bulgaria and NW Turkey: a review of recent data. In Robertson, A.H.F. & Mountrakis, D. (eds.), Tectonic Development of the Eastern Mediterranean Region. Geological Society, London, Special Publications, 260, pp. 421430.Google Scholar
Yang, J., Cawood, P.A., Du, Y. et al. (2014). A sedimentary archive of tectonic switching from Emeishan Plume to Indosinian orogenic sources in SW China. Journal of the Geological Society, London, 171, 269280.Google Scholar
Yolkin, E.A., Sennikov, N.V., Bakharev, N.K. et al. (2003). Silurian paleogeography along the southwest margin of the Siberian continent: Altai–Sayan folded area. New York State Museum Bulletin, 493, 299322.Google Scholar
Young, G.C. (1990). Devonian vertebrate distribution patterns and cladistics analysis of palaeogeographic hypotheses. In McKerrow, W.S. & Scotese, C.R. (eds.), Palaeozoic Palaeogeography and Biogeography. Geological Society, London, Memoir, 12, pp. 243255.Google Scholar
Young, G.C. & Janvier, P. (1999). Early–middle Palaeozoic vertebrate faunas in relation to Gondwana dispersion and Asian accretion. In Metcalfe, I. (ed.), Gondwana Dispersion and Accretion. Rotterdam: Balkema, pp. 115140.Google Scholar
Yue, Y., Liao, J.G. & Graham, S.A. (2001). Tectonic correlation of Beishan and Inner Mongolia orogens and its implications for the palinspastic reconstruction of North China. In Hendrix, M.S. & Davis, G.A. (eds.), Paleozoic and Mesozoic Tectonic Evolution of Central and Eastern Asia: From Continental Assembly to Intracontinental Deformation. Geological Society of America Memoir, 194, pp. 101116.Google Scholar
Zachos, J.C., Dickens, G.R. & Zeebe, R.E. (2008). An early Cenozoic perspective on greenhouse warming and carbon-cycle dynamics. Nature, 451, 279283.Google Scholar
Zachos, J., Pagani, M., Sloan, L., Thomas, E. & Billups, K. (2001). Trends, rhythms, and aberrations in global climate 65 Ma to present. Science, 292, 688693.Google Scholar
Zakharov, Y.D., Popov, A.M. & Blakov, A.S. (2008). Late Permian to Middle Triassic palaeogeographic differentiation of key ammonoid groups: evidence from the former USSR. Polar Research, 27, 441468.Google Scholar
Zanchi, A., Poli, S., Fumagalli, P. & Gaetani, M. (2000). Mantle exhumation along the Trich Mir Fault Zone NW Pakistan: pre-mid-Cretaceous accretion of the Karakorum terrane to the Asian margin. In Khan, M.A. et al. (eds.), Tectonics of the Nanga Parbat Syntaxis and the Western Himalaya. Geological Society, London, Special Publications, 170, pp. 237252.Google Scholar
Zhan, R., Rong, Y., Percival, I.G. & Liang, Y. (2011). Brachiopod biogeographic change during the Early to Middle Ordovician in South China. Memoirs of the Society of Australasian Palaeontologists, 41, 273287.Google Scholar
Zhang, L., Ai, Y., Li, X. et al. (2007). Triassic collision of western Tianshan orogenic belt, China: evidence from SHRIMP U–Pb dating of zircon from HP/UHP eclogitic rocks. Lithos, 96, 266280.Google Scholar
Zhang, L., Qin, K. & Xian, W. (2008). Multiple mineralisation events in the eastern Tienshan district, NW China: isotopic geochronology and geological significance. Journal of Asian Earth Sciences, 32, 236246.Google Scholar
Zhang, W., Chen, P. & Palmer, A.R. (2003). Biostratigraphy of China. Beijing: Science Press.Google Scholar
Zhao, G.C., Sun, M., Wilde, S.A., Li, S.Z. (2004). A Paleo-Mesoproterozoic supercontinent: assembly, growth and breakup. Earth-Science Reviews, 67, 91123.Google Scholar
Zhou, D., Graham, S.A., Chang, E.Z., Wang, B. & Hacker, B. (2001). Paleozoic tectonic amalgamation of the Chinese Tian Shan: evidence from a transect along the Dushanzi–Kuqa Highway. In Hendrix, M.S. & Davis, G.A. (eds.), Paleozoic and Mesozoic Tectonic Evolution of Central and Eastern Asia: From Continental Assembly to Intracontinental Deformation. Geological Society of America Memoir, 194, pp. 2346.Google Scholar
Zhou, J., Wilde, S.A., Zhao, G.C. et al. (2010). Was the easternmost segment of the Central Asian Orogenic Belt derived from Gondwana or Siberia: an intriguing dilemma? Journal of Geodynamics, 50, 300317.Google Scholar
Zhou, Z. & Dean, W.T. (eds.) (1996). Phanerozoic Geology of Northwest China. Beijing: Science Press.Google Scholar
Zhu, D., Zhao, Z., Niu, Y. et al. (2013). The origin and pre-Cenozoic evolution of the Tibetan Plateau. Gondwana Research, 23, 14291454.Google Scholar
Zhu, Y., Guo, X., Song, B., et al. (2009). Petrology, Sr–Nd–Hf isotopic geochemistry and zircon chronology of the Late Palaeozoic volcanic rocks in the southwestern Tianshan Mountains, Xinjiang, NW China. Journal of the Geological Society, London, 166, 10851099.Google Scholar
Ziegler, A.M., Cocks, L.R.M. & Bambach, R.K. (1968). The composition and structure of Lower Silurian marine communities. Lethaia, 1, 127.Google Scholar
Ziegler, A.M., Hulvey, M.I. & Rowley, D.B. (1997). Permian world topography and climate. In Martini, L.P. (ed.), Late Glacial and Post-Glacial Environmental Changes: Quaternary, Carboniferous, Proterozoic. Oxford: Oxford University Press, pp. 111146.Google Scholar
Ziegler, M.A. (2001). Late Permian to Holocene paleofacies evolution of the Arabian Plate and its hydrocarbon occurrences. GeoArabia, 6, 445504.Google Scholar
Ziegler, P.A. (1989). Evolution of Laurussia: A Study in Late Palaeozoic Plate Tectonics. Dordrecht: Kluwer.Google Scholar
Ziegler, P.A. (1990). Geological Atlas of Western and Central Europe, 2nd edn. The Hague: Shell and London: Geological Society.Google Scholar
Žigaitė, Ž. & Blieck, A. (2006). Palaeobiogeographical significance of early Silurian thelodonts from central Asia and southern Siberia. Geologiska Föreingens i Stockholm Förhandlingar, 128, 203206.Google Scholar
Zonenshain, L.P., Kuzmin, M.I. & Natapov, L.M. (1990). Geology of the USSR: A Plate Tectonic Synthesis. American Geophysical Union Geodynamics Series, 21.Google Scholar
Zurevinski, S.E., Heaman, L.M. & Creaser, R.A. (2011). The origin of Triassic/Jurassic kimberlite magmatism, Canada: two mantle sources revealed from the Sr–Nd isotopic composition of groundmass perovskite. Geochemistry, Geophysics, Geosystems, 12, Q09005, doi:10.1029/2011GC003659.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • References
  • Trond H. Torsvik, Universitetet i Oslo, L. Robin M. Cocks, Natural History Museum, London
  • Book: Earth History and Palaeogeography
  • Online publication: 24 November 2016
  • Chapter DOI: https://doi.org/10.1017/9781316225523.022
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • References
  • Trond H. Torsvik, Universitetet i Oslo, L. Robin M. Cocks, Natural History Museum, London
  • Book: Earth History and Palaeogeography
  • Online publication: 24 November 2016
  • Chapter DOI: https://doi.org/10.1017/9781316225523.022
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • References
  • Trond H. Torsvik, Universitetet i Oslo, L. Robin M. Cocks, Natural History Museum, London
  • Book: Earth History and Palaeogeography
  • Online publication: 24 November 2016
  • Chapter DOI: https://doi.org/10.1017/9781316225523.022
Available formats
×