Hostname: page-component-8448b6f56d-c4f8m Total loading time: 0 Render date: 2024-04-23T16:04:46.194Z Has data issue: false hasContentIssue false

Plate tectonic modelling: review and perspectives

Published online by Cambridge University Press:  14 February 2018

CHRISTIAN VÉRARD*
Affiliation:
Chemin de Servasse, F – 74930 Reignier-Ésery, France
*
*Author for correspondence: xian_verard@hotmail.com

Abstract

Since the 1970s, numerous global plate tectonic models have been proposed to reconstruct the Earth's evolution through deep time. The reconstructions have proven immensely useful for the scientific community. However, we are now at a time when plate tectonic models must take a new step forward. There are two types of reconstructions: those using a ‘single control’ approach and those with a ‘dual control’ approach. Models using the ‘single control’ approach compile quantitative and/or semi-quantitative data from the present-day world and transfer them to the chosen time slices back in time. The reconstructions focus therefore on the position of tectonic elements but may ignore (partially or entirely) tectonic plates and in particular closed tectonic plate boundaries. For the readers, continents seem to float on the Earth's surface. Hence, the resulting maps look closer to what Alfred Wegener did in the early twentieth century and confuse many people, particularly the general public. With the ‘dual control’ approach, not only are data from the present-day world transferred back to the chosen time slices, but closed plate tectonic boundaries are defined iteratively from one reconstruction to the next. Thus, reconstructions benefit from the wealth of the plate tectonic theory. They are physically coherent and are suited to the new frontier of global reconstruction: the coupling of plate tectonic models with other global models. A joint effort of the whole community of geosciences will surely be necessary to develop the next generation of plate tectonic models.

Type
Review Article
Copyright
Copyright © Cambridge University Press 2018 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Adriasola Munoz, A., Harris, J., Glover, C., Goodrich, M., Hudson, L. & Ady, B. 2010. Modelling continental margin extension using combined rigid/deformable plate tectonic reconstructions. Search and Discovery Article #40603. Poster at AAPG Annual Convention and Exhibition, New Orleans, April 11–14.Google Scholar
Attewell, P. & Farmer, I. 1976. Principles of Engineering Geology. London: Chapman and Hall, 1045 pp.Google Scholar
Bachtadse, V. & Briden, J. 1990. Paleomagnetic constraints on the position of Gondwana during Ordovician to Devonian times. In Palaeozoic Palaeogeography and Biogeography (eds McKerrow, W. & Scotese, C.), pp. 43–8. Geological Society of London, Memoir no. 12.Google Scholar
Bachtadse, V. & Briden, J. 1991. Palaeomagnetism of Devonian ring complexes from the Bayuda Desert – new constraints on the Apparent Polar Wander path for Gondwanaland. Geophysical Journal International 104, 635–46.Google Scholar
Bacon, F. 1620. Novum Organum Scientiarum. Londini, apud Joannem Billium, Typographum Regium, 404 pp.Google Scholar
Baes, M., Govers, R. & Wortel, R. 2011. Switching between alternative responses of the lithosphere to continental collision. Geophysical Journal International 187, 1151–74.Google Scholar
Baines, G., Cheadle, M., John, B. & Schwartz, J. 2008. The rate of oceanic detachment faulting at Atlantis Bank, SW Indian Ridge. Earth and Planetary Science Letters 273, 105–14.Google Scholar
Barker, P. 2001. Scotia Sea regional tectonic evolution: implications for mantle flow and palaeocirculation. Earth-Science Reviews 55, 139.Google Scholar
Bartol, J. & Govers, R. 2014. A single cause for uplift of the Central and Eastern Anatolian plateau? Tectonophysics 637, 116–36.Google Scholar
Baumgartner, P. O. 2013. Mesozoic radiolarites – accumulation as a function of sea surface fertility on Tethyan margins and in ocean basins. Sedimentology 60, 292318.Google Scholar
Beaumont, C., Jamieson, R., Nguyen, M. & Medvedev, S. 2004. Crustal channel flows: 1. Numerical models with applications to the tectonics of the Himalayan-Tibetan orogen. Journal of Geophysical Research 109, B06406. doi: 10.1029/2003JB002809, 29 pp.Google Scholar
Benioff, H. 1955. Earthquake seismographs and associated instruments. Advances in Geophysics 2, 219–75.Google Scholar
Berner, R. 1998. The carbon cycle and CO2 over Phanerozoic time: the role of land plants. Philosophical Transactions of the Royal Society of London B 353, 7582.Google Scholar
Berner, R. 2006. GeoCarbSulf: a combined model for Phanerozoic atmospheric O2 and CO2. Geochimica et Cosmochimica Acta 70, 5653–64.Google Scholar
Bird, P. 2003. An updated digital model of plate boundaries. Geochemistry, Geophysics, Geosystems (G3) 4, 1027–52.Google Scholar
Blakey, R. 2008. Gondwana palaeogeography from assembly to breakup – a 500 m.y. odyssey. In Resolving the Late Paleozoic Ice Age in Time and Space (eds Fielding, C., Frank, T. & Isbell, J.), pp. 128. Geological Society of America, Special Paper no. 441.Google Scholar
Boucot, A. 1999. Southern African Phanerozoic marine invertebrates: biogeography, palaeoeology, climatology and comments on adjacent regions. Journal of African Earth Sciences 28, 129–43.Google Scholar
Boucot, A., Xu, C., Scotese, C. & Morley, R. 2013. Phanerozoic palaeoclimate: an atlas of lithologic indicators of climate. SEPM Concepts in Sedimentology and Paleontology 11, Map Folio, 30 pp.Google Scholar
Bower, D., Gurnis, M. & Flament, N. 2015. Assimilating lithosphere and slab history in 4-D Earth models. Physics of the Earth and Planetary Interiors 238, 822.Google Scholar
Bower, D., Gurnis, M. & Seton, M. 2013. Lower mantle structure from paleogeographically constrained dynamic Earth models. Geochemistry, Geophysics, Geosystems (G3) 14, 4463.Google Scholar
Braun, J., Simon-Labric, T., Murray, K. & Reiners, P. 2014. Topographic relief driven by variations in surface rock density. Nature Geoscience 7, 534–40.Google Scholar
Braun, J., van der Beek, P., Valla, P., Robert, X., Herman, F., Glotzbach, C., Pedersen, V., Perry, C., Simon-Labric, T. & Prigent, C. 2012. Quantifying rates of landscape evolution and tectonic processes by thermochronology and numerical modelling of crustal heat transport using Pecube. Tectonophysics 524–525, 128.Google Scholar
Breuer, M., Wessling, S., Schmalzl, J. & Hansen, U. 2004. Effect of inertia in Rayleigh-Bénard convection. Physical Review E 69, 026302. doi: 10.1103/PhysRevE.69.026302, 10 pp.Google Scholar
Brongniart, A. 1828. Prodomes d'une Histoire des Végétaux Fossiles. Paris: F. G. Levrault, 223 pp.Google Scholar
Brune, S., Popov, A. & Sobolev, S. 2013. Quantifying the thermo-mechanical impact of plume arrival on continental break-up. Tectonophysics 604, 51–9.Google Scholar
Brunetti, M., Vérard, C. & Baumgartner, P. O. 2015. Modelling the Middle Jurassic Ocean Circulation. Journal of Palaeogeography 4, 371–83.Google Scholar
Bullard, E., Everett, J. & Smith, G. 1965. The fit of the continents around the Atlantic. Philosophical Transactions of the Royal Society of London, Series A, Mathematical and Physical Sciences 258 (1088), 4151.Google Scholar
Bunge, H.-P., Richards, M. & Baumgardner, J. 1997. A sensitivity study of 3-D spherical mantle convection at 10exp8 Rayleigh Number: effects of depth dependent viscosity, heating mode and endothermic phase change. Journal of Geophysical Research 102, 1199112007.Google Scholar
Čadek, O. & Fleitout, L. 2003. Effect of lateral viscosity variations in the top 300 km on the geoid and dynamic topography. Geophysical Journal International 152, 566–80.Google Scholar
Cande, S. & Stegman, D. 2011. Indian and African plate motions driven by the push force of the Réunion plume head. Nature 475, 4752.Google Scholar
Carlson, R. L. & Raskin, G. S. 1984. Density of the ocean crust. Nature 311, 555–8.Google Scholar
Cavazza, W., Roure, F., Spakman, W., Stampfli, G. & Ziegler, P. 2004. The Transmed Atlas; the Mediterranean region from crust to mantle. In The Mediterranean Consortium for the 32nd International Geological Congress, XXIII, 141 pp. with CD-Rom.Google Scholar
Chandler, M., Rind, D. & Ruedy, R. 1992. Pangaean climate during the Early Jurassic: GCM simulations and the sedimentary record of palaeoclimate. Geological Society of America Bulletin 104, 543–59.Google Scholar
Chumakov, N. 2004. Trends in global climate changes inferred from geological data. Stratigraphy and Geological Correlation 12, 117–38.Google Scholar
Chumakov, N. & Zharkov, M. 2002. Climate during Permian–Triassic biosphere reorganisations, Article 1: climate of the Early Permian. Stratigraphy and Geological Correlation 10, 586602.Google Scholar
Chumakov, N. & Zharkov, M. 2003. Climate during Permian–Triassic biosphere reorganisations, Article 2: climate of the Late Permian and Early Triassic: general inferences. Stratigraphy and Geological Correlation 11, 361–75.Google Scholar
Cloetingh, S., Wortel, R. & Vlaar, N. 1989. On the initiation of subduction zones. Pure and Applied Geophysics 129, 725.Google Scholar
Cloos, M. 1993. Lithospheric buoyancy and collisional orogenesis: subduction of oceanic plateaus, continental margins, island arcs, spreading ridges, and seamounts. Geological Society of America Bulletin 105, 715–37.Google Scholar
Cocks, R. & Fortey, R. 1982. Faunal evidence for oceanic separations in the Palaeozoic of Britain. Journal of the Geological Society, London 139, 465–78.Google Scholar
Cocks, R. & Torsvik, T. 2011. The Palaeozoic geography of Laurentia and western Laurussia: a stable craton with mobile margins. Earth-Science Reviews 106, 151.Google Scholar
Cocks, R. & Torsvik, T. 2013. The dynamic evolution of the Palaeozoic geography of eastern Asia. Earth-Science Reviews 117, 4079.Google Scholar
Conrad, C. & Lithgow-Bertelloni, C. 2002. How mantle slabs drive plate tectonics. Science 298, 207–9.Google Scholar
Cope, J., Ingham, J. & Rawson, P. (eds) 1992. Atlas of Palaeogeography and Lithofacies. Geological Society of London, Memoir 13.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., Jaupart, C., Manighetti, I., Tapponnier, P. & Besse, J. 1999. On causal links between flood basalts and continental breakup. Earth and Planetary Science Letters 166, 177–95.Google Scholar
Crameri, F., Tackley, P., Meilick, I., Gerya, T. & Kaus, J. 2012. A free plate surface and weak oceanic crust produce single-sided subduction on Earth. Geophysical Research Letters 39, L03306. doi: 10.1029/2011GL050046, 7 pp.Google Scholar
Creer, K., Irving, E. & Runcorn, S. 1954. The direction of the geomagnetic field in remote epochs in Great Britain. Journal of Geomagnetism and Geoelectricity 6, 163–8.Google Scholar
Cruciani, C., Carminati, E. & Doglioni, C. 2005. Slab dip vs. lithosphere age: no direct function. Earth and Planetary Science Letters 238, 298310.Google Scholar
Dalziel, I. 1997. Neoproterozoic–Paleozoic geography and tectonics: review, hypothesis, environmental speculation. Geological Society of America Bulletin 109, 1642.Google Scholar
Dana, J. 1880. Manual of Geology. New York: Taylor & Co, 798 pp.Google Scholar
D'Aubuisson de Voisins, J.-F. 1828. Traité de Géognosie (nouvelle édition revue et corrigée). Strasbourg, 496 pp.Google Scholar
Davies, J. H. 2013. Global map of solid Earth surface heat flow. Geochemistry, Geophysics, Geosystems (G3) 14, 4608–23.Google Scholar
Davies, J. H. & Davies, D. R. 2010. Earth's surface heat flux. Solid Earth 1, 524.Google Scholar
Davies, D. R., Davies, J. H., Bollada, P., Hassan, O., Morgan, K. & Nithiarasu, P. 2013. A hierarchical mesh refinement technique for global 3-D spherical mantle convection modelling. Geoscientific Model Development 6, 1095–7.Google Scholar
Davies, R., Goes, S., Davies, H., Schubert, B., Bunge, H.-P. & Ritsema, J. 2012. Reconciling dynamic and seismic models of Earth's lower mantle: the dominant role of thermal heterogeneity. Earth and Planetary Science Letters 353–354, 253–69.Google Scholar
De Melo Garcia, S.-F. 2012. Restauração estrutural da haltectônica na porção central da bacia de Santos e implicações para os sistemas petrolíferos. Ph.D. thesis, Universidade Federal de Ouro Preto, Escola de Minas, Brasil, and Université de Cergy-Pontoise, Ecole doctorale Sciences et Ingénérie, France, 237 pp. Published thesis.Google Scholar
DeMets, C., Gordon, R., Argus, D. & Stein, S. 1994. Effect of recent revisions to the geomagnetic reversal time scale on estimates of current plate motions. Geophysical Research Letters 21, 2191–4.Google Scholar
Dera, G. & Donnadieu, Y. 2012. Modeling evidences for global warming, Arctic seawater freshening, and sluggish oceanic circulation during the Early Toarcian anoxic event. Paleoceanography 27, PA2211. doi: 10.1029/2012PA002283, 15 pp.Google Scholar
Dercourt, J. 2000. Atlas Peri-Tethys Palaeogeographical Maps. Paris: Gauthier-Villars/CCGM, 268 pp., 24 maps.Google Scholar
Dercourt, J., Ricou, L. & Vrielynck, B. (eds) 1993. Atlas Tethys Palaeoenvironmental Maps. Atlas and Explanatory Notes. Paris: Gauthier Villars, 307 pp., 14 maps.Google Scholar
Dewey, J. & Bird, J. 1970. Plate tectonics and geosynclines. Tectonophysics 10, 625–38.Google Scholar
Dick, H., Lin, J. & Schouten, H. 2003. An ultraslow-spreading class of ocean ridge. Nature 426, 405–12.Google Scholar
Dietz, R. 1961. Continent and ocean basin evolution by spreading of the sea floor. Nature 190, 854–7.Google Scholar
Direen, N., Stagg, H., Symonds, P. & Norton, I. 2012. Variations in rift symmetry: cautionary examples from the Southern Rift System (Australia–Antarctica). In Conjugate Divergent Margins (eds Mohriak, W., Danforth, A., Post, P., Brown, D., Tari, G., Nemčok, A. & Sinha, S.), pp. 453–75. Geological Society of London, Special Publication no. 369.Google Scholar
Doglioni, C., Carminati, E., Cuffaro, M., Scrocca, D. 2007. Subduction kinematics and dynamic constraints. Earth-Science Reviews 83, 125–75.Google Scholar
Domeier, M. & Torsvik, T. 2014. Plate tectonics in the late Paleozoic. Geoscience Frontiers 5, 303–50.Google Scholar
Domeier, M., Van der Voo, R. & Torsvik, T. H. 2012. Paleomagnetism and Pangea: the road to reconciliation. Tectonophysics 514–517, 1443.Google Scholar
Donnadieu, Y., Goddéris, Y. & Bouttes, N. 2009. Exploring the climatic impact of the continental vegetation on the Mesozoic atmospheric CO2 and climate history. Climate Past 5, 8596.Google Scholar
Donnadieu, Y., Goddéris, Y., Pierrehumbert, R., Dromart, G., Fluteau, F. & Jacob, R. 2006 a. A GeoClim simulation of climatic and biogeochemical consequences of Pangea breakup. Geochemistry, Geophysics, Geosystems (G3) 7, 121.Google Scholar
Donnadieu, Y., Pierrehumbert, R., Jacob, R. & Fluteau, F. 2006 b. Modelling the primary control of paleogeography on Cretaceous climate. Earth and Planetary Science Letters 248, 426–37.Google Scholar
Duarte, J., Rosas, F., Terrinha, P., Schellart, W., Boutelier, D., Gutscher, M.-A. & Ribeiro, A. 2013. Are subduction zones invading the Atlantic? Evidence from the southwest Iberia margin. Geology 41, 839– 42.Google Scholar
Duretz, T. & Gerya, T. 2013. Slab detachment during continental collision: influence of crustal rheology and interaction with lithospheric delamination. Tectonophysics 602, 124–40.Google Scholar
Du Toit, A. 1937. Our Wandering Continents: An Hypothesis of Continental Drifting: ‘Africa Forms the Key’. Edinburgh: Oliver & Boyd, 366 pp.Google Scholar
Eagles, G. & Jokat, W. 2014. Tectonic reconstructions for paleobathymetry in Drake Passage. Tectonophysics 611, 2850.Google Scholar
Eglington, B. M., Reddy, S. M. & Evans, D. A. D. 2009. The IGCP 509 database system: design and application of a tool to capture and illustrate the litho- and chrono-statigraphic information for Palaeoproterozoic tectonic domains, large igneous provinces, and ore deposits; with examples from southern Africa. In Palaeoproterozoic Supercontinents and Global Evolution (eds Reddy, S. M., Mazumder, R., Evans, D. A. D. & Collins, A. S.), pp. 2747. Geological Society of London, Special Publication no. 323.Google Scholar
Ferrari, O., Hochard, C. & Stampfli, G. 2008. An alternative plate tectonic model for the Palaeozoic – Early Mesozoic Palaeotethyan evolution of Southeastern Asia (Northern Thailand – Burma). Tectonophysics 451, 346–65.Google Scholar
Flament, N. 2014. Linking plate tectonics and mantle flow to Earth's topography. Geology 42, 927–8.Google Scholar
Flament, N., Gurnis, M., Williams, S., Seton, M., Skogseid, J., Heine, C. & Müller, R. D. 2014. Topographic asymmetry of the South Atlantic from global models of mantle flow and lithosphere stretching. Earth and Planetary Science Letters 387, 107–19.Google Scholar
Flögel, S., Wold, C. & Hay, W. 2000. Evolution of sediments and ocean salinity. In Abstracts Volume, 31st International Geological Congress, Rio de Janeiro – Brazil, August 6–17, 2000, 4 pp. and CD-Rom.Google Scholar
Fluteau, F. 2001. The Late Permian climate. What can be inferred from climate modelling concerning Pangea scenarios and Hercynian range altitude? Palaeogeography, Palaeoclimatology, Palaeoecology 167, 3971.Google Scholar
Fluteau, F. 2013. L’évolution des climats à l’échelle des temps géologiques; Le rôle des changements paléogéographiques. In Paléoclimatologie: Enquête sur les climats anciens – Tome II (eds Duplessy, J.-C. & Ramstein, G.), pp. 79138. Paris: EDP Sciences & CNRS Editions.Google Scholar
Fluteau, F., Besse, J., Broutin, J. & Ramstein, G. 2001. The Late Permian climate. What can be inferred from climate modelling concerning Pangea scenarios and Hercynian range altitude? Palaeogeography, Palaeoclimatology, Palaeoecology 167, 3971.Google Scholar
Fluteau, F., Ramstein, G., Besse, J., Guiraud, R. & Masse, J.-P. 2007. Impacts of palaeogeography and sea level changes on Mid-Cretaceous climate. Palaeogeography, Palaeoclimatology, Palaeoecology 247, 357–81.Google Scholar
Follows, M. & Dutkiewicz, S. 2011. Modeling diverse communities of marine microbes. Annual Review of Marine Science 3, 427–51.Google Scholar
Forsyth, D. & Uyeda, S. 1975. On the relative importance of the driving forces of plate motion. Geophysical Journal of the Royal Astronomical Society 43, 163200.Google Scholar
François, L. & Walker, J. 1992. Modelling the Phanerozoic carbon cycle and climate: constraints from the 87Sr/86Sr isotopic ratio of seawater. American Journal of Science 292, 81135.Google Scholar
Frohlich, C. 1987. Kiyoo Wadati and early research on deep focus earthquakes: introduction to special section on deep and intermediate focus earthquakes. Journal of Geophysical Research 92 (B13), 13777–88.Google Scholar
Fryer, P. 2002. Recent studies of serpentinite occurrences in the oceans: mantle-ocean interactions in the plate tectonic cycle. Chemie der Erde, Geochemistry 62, 257302.Google Scholar
Gaetani, M., Dercourt, J. & Vrielynck, B. 2003. The Peri-Tethys Programme: achievements and results. Episodes 26, 7993.Google Scholar
Gahagan, L., Scotese, C., Royer, J.-Y., Sandwell, D., Winn, J., Tomlins, R., Ross, M., Newman, J., Müller, D., Mayes, C., Lawver, L. & Heubeck, C. 1988. Tectonic fabric map of the ocean basins from satellite altimetry data. Tectonophysics 155, 126.Google Scholar
Gaina, C., Müller, D., Royer, J.-Y., Stock, J., Hardebeck, J. & Symonds, P. 1998 a. The tectonic history of the Tasman Sea: a puzzle with 13 pieces. Journal of Geophysical Research 103 (B6), 12413–33.Google Scholar
Gaina, C., Roest, W., Müller, D. & Symonds, P. 1998 b. The opening of the Tasman Sea: a gravity anomaly animation. Earth Interactions 2 (4). doi: 10.1175/1087-3562(1998)002<0001:TOOTTS>2.3.CO; 2, 23 pp.2.3.CO;+2,+23+pp.>Google Scholar
Gerya, T. 2013. Three-dimensional thermomechanical modelling of oceanic spreading initiation and evolution. Physics of the Earth and Planetary Interiors 214, 3552.Google Scholar
Gerya, T. 2011. Future directions in subduction modelling. Journal of Geodynamics 52, 344–78.Google Scholar
Goddéris, Y., Donnadieu, Y., Lefebvre, V., Le Hir, G. & Nardin, E. 2012. Tectonic control of continental weathering, atmospheric CO2, and climate over Phanerozoic times. Comptes Rendus Geoscience 344, 568–85.Google Scholar
Golonka, J. 2000. Cambrian-Neogene Plate Tectonic Maps. Wydawn. Uniwersytetu Jagiellonskiego, 125 pp. and 36 maps.Google Scholar
Golonka, J. 2007 a. Phanerozoic paleoenvironment and paleolithofacies maps: Early Paleozoic. Geologia 35, 589654.Google Scholar
Golonka, J. 2007 b. Phanerozoic paleoenvironment and paleolithofacies maps: Late Paleozoic. Geologia 33, 145209.Google Scholar
Golonka, J. 2007 c. Phanerozoic paleoenvironment and paleolithofacies maps: Mesozoic. Geologia 33, 211–64.Google Scholar
Golonka, J. 2009. Phanerozoic paleoenvironment and paleolithofacies maps: Cenozoic. Geologia 35, 507–85.Google Scholar
Golonka, J., Ross, M. & Scotese, C. 1994. Phanerozoic paleogeographic and paleoclimatic modelling maps. In Pangea: Global Environment and Resources (eds Embry, A., Beauchamp, B. & Glass, D.), pp. 147. Memoir of the Canadian Society of Petroleum Geologists 17.Google Scholar
Govers, R. & Wortel, R. 2005. Lithosphere tearing at Step faults: response to edges of subduction zones. Earth and Planetary Science Letters 236, 505–23.Google Scholar
Grand, S., van der Hilst, R. & Widiyantoro, S. 1997. Global seismic tomography: a snapshot of convection in the earth. GSA Today 7, 17.Google Scholar
Grunow, A. 1999. Gondwana events and palaeogeography: a palaeomagnetic review. Journal of African Earth Sciences 28, 5369.Google Scholar
Gurnis, M., Turner, M., Zahirovic, S., DiCaprio, L., Spasojevic, S., Müller, D., Boyden, J., Seton, M., Constantin Manea, V. & Bower, D. 2012. Plate tectonic reconstructions with continuously closing plates. Computers & Geosciences 38, 3542.Google Scholar
Hafkenscheid, E., Warners-Ruckstuhl, K., van Oosterhoot, C., Bergman, S., Davies, H., Govers, R., Hochard, C., Kennan, L., Ross, M., Stampfli, G., Vérard, C., Webb, P. & Wortel, R. 2013. Integrating plate tectonic reconstruction and mantle dynamics: a valuable aid in frontier exploration. Poster #EGU2013-3204 at the EGU General Assembly, Vienna.Google Scholar
Hafkenscheid, E., Wortel, R. & Spakman, W. 2006. Subduction history of the Tethyan region derived from seismic tomography and tectonic reconstructions. Journal of Geophysical Research 111, B08401. doi: 10.1029/2005JB003791, 26 pp.Google Scholar
Hall, R. 2012. Late Jurassic-Cenozoic reconstructions of the Indonesian region and the Indian Ocean. Tectonophysics 570–571, 141.Google Scholar
Hallam, A. 1985. A review of Mesozoic climates. Journal of the Geological Society, London 142, 433–45.Google Scholar
Harris, J., Ashley, A., Otto, S., Valdes, P., Crossley, R., Preston, R., Watson, J., Goodrich, M. & Team, Merlin + Project. 2017. Paleogeography and paleo-Earth systems in the modeling of marine paleoproductivity: A prerequisite for the prediction of petroleum source rocks. In Petroleum System Case Studies: AAPG Memoir (eds AbuAli, M. & Moretti, I.). vol. 114, pp. 3760.Google Scholar
Hay, W. W., DeConto, R. M., Wold, C. N., Wilson, K. M., Voigt, S., Schulz, M., Rossby-Wold, A., Dullo, W.-C., Ronov, A. B., Balukhovsky, A. N. & Söding, E. 1999. Alternative global Cretaceous paleogeography. In Evolution of the Cretaceous Ocean-Climate System (eds Barrera, E. & Johnson, C. C.), pp. 147. Geological Society of America, Special Paper 332.Google Scholar
Hay, W. & Flögel, S. 2012. New thoughts about the Cretaceous climate and oceans. Earth-Science Reviews 115, 262–72.Google Scholar
Hay, W., Wold, C., Söding, E. & Flögel, S. 2001. Evolution of sediment fluxes and ocean salinity. In Geologic Modelling and Simulation: Sedimentary Systems (eds Merriam, D. & Davies, J.), pp. 153–67. Dordrecht, The Netherlands: Kluwer Academic/Plenum Publishers.Google Scholar
Haywood, A., Valdes, P. & Markwick, P. 2004. Cretaceous (Wealden) climates: a modelling perspective. Cretaceous Research 25, 303–11.Google Scholar
Heidbach, O., Tingay, M., Barth, A., Reinecker, J., Kurfeß, D. & Müller, B. 2009. The World Stress Map Based on the Database Release 2008, Equatorial Scale 1:46,000,000. Paris: Commission for the Geological Map of the World. doi: 10.1594/GFZ.WSM.Map2009.Google Scholar
Hellinger, S. 1981. The uncertainties in finite rotations in plate tectonics. Journal of Geophysical Research 86, 9312–8.Google Scholar
Hess, H. 1962. History of ocean basins. In Petrologic Studies: A Volume to Honor A. F. Buddington (eds Engel, A. E. J., James, H. L. & Leonard, B. F.), pp. 599620. New York: Geological Society of America.Google Scholar
Hewitt, P., Lyons, S., Suchocki, J. & Yeh, J. 2007. Conceptual Integrated Science. San Francisco: Pearson Education Inc., 668 pp.Google Scholar
Hochard, C., Vérard, C. & Baumgartner, P. 2011. Geodynamic evolution of the Earth over 600 Ma: implications for palaeo-climatic indicators. Poster #AGU2011-pp13d-1849 at the AGU Fall Meeting, San Francisco.Google Scholar
Hoffman, P. 1991. Did the breakout of Laurentia turn Gondwanaland inside-out? Science 252 (5011), 1409–12.Google Scholar
Holmes, A. 1913. The Age of the Earth. London: Harper, 196 pp.Google Scholar
Holmes, A. 1944. Principles of Physical Geology. London: Thomas Nelson & Sons, Ltd, 628 pp.Google Scholar
Howell, D. 1995. Principles of Terrane Analysis: New Applications for Global Tectonics. 2nd edition. Topics in the Earth Sciences 8. London: Chapman & Hall, 224 pp.Google Scholar
Howell, D., Jones, D. & Schermer, E. 1985. Tectonostratigraphic Terranes of the Circum-Pacific Region. CircumPacific Council for Energy and Mineral Resources, Earth Science Series vol. 1, pp. 3–30.Google Scholar
Hughes, N. F. (ed.) 1973. Organisms and Continents Through Time: A Symposium. Special Papers in Palaeontology 12. London: The Palaeontological Association, 334 pp.Google Scholar
Huismans, R. & Beaumont, C. 2007. Roles of lithospheric strain softening and heterogeneity in determining the geometry of rifts and continental margins. In Imaging, Mapping and Modelling Continental Lithosphere Extension and Breakup (eds. Karner, G., Manatschal, G. & Pinheiro, L.), pp. 111–38. Geological Society of London, Special publication no. 282.Google Scholar
Isacks, B., Oliver, J. & Sykes, L. 1968. Seismology and the new global tectonics. Journal of Geophysical Research 73, 5855–99.Google Scholar
Jamieson, R., Beaumont, C., Medvedev, S. & Nguyen, M. 2004. Crustal channel flows: 2. Numerical models with applications for metamorphism of the Himalayan-Tibetan orogen. Journal of Geophysical Research 109, B06407. doi: 10.1029/2003JB002811, 24 pp.Google Scholar
Jordan, T. 1978. Composition and development of the continental tectosphere. Nature 274, 544–8.Google Scholar
Kamesh Raju, K., Samudrala, K., Drolia, R., Amarnath, D., Ramachandran, R. & Mudholkar, A. 2012. Segmentation and morphology of the Central Indian Ridge between 3°S and 11°S, Indian Ocean. Tectonophysics 554–557, 114–26.Google Scholar
Kaplan, J., Bigelow, N., Prentice, I., Harrison, S., Bartlein, P., Christensen, T., Cramer, W., Matveyeva, N., McGuire, A., Murray, D., Razzhivin, V., Smith, B., Walker, D., Anderson, P., Andreev, A., Brubaker, L., Edwards, M. & Lozhkin, A. 2003. Climate change and Arctic ecosystems: 2. Modeling, paleodata-model comparisons, and future projections. Journal of Geophysical Research 108, 8171. doi: 10.1029/2002JD002559, 17 pp.Google Scholar
Kent, D. & Van der Voo, R. 1990. Palaeozoic palaeogeography from palaeomagnetism of the Atlantic-bordering continents. In Palaeozoic Palaeogeography and Biogeography (eds McKerrow, W. & Scotese, C.), pp. 4956. Geological Society of London, Memoir no. 12.Google Scholar
Kergoat, G., Bouchard, P., Clamens, A.-L., Abbate, J., Jourdan, H., Jabbour-Zahab, R., Genson, G., Soldati, L. & Condamine, F. 2014. Cretaceous environmental changes led to high extinction rates in a hyperdiverse beetle family. BMC Evolutionary Biology 14, 220. doi: 10.1186/s12862-014-0220-1.Google Scholar
Klein, G. (ed.) 1994. Pangea: Paleoclimate, Tectonics, and Sedimentation During Accretion, Zenith, and Breakup of a Supercontinent. The Geological Society of America, Special Paper 288.Google Scholar
Kneller, E., Johnson, C., Karner, G., Einhorn, J. & Queffelec, T. 2012. Inverse methods for modeling non-rigid plate kinematics: application to Mesozoic plate reconstructions of the Central Atlantic. Computers & Geosciences 49, 217–30.Google Scholar
Lagabrielle, Y., Chauvet, A., Ulrich, M. & Guillot, S. 2013. Passive obduction and gravity-driven emplacement of large ophiolitic sheets: the New Caledonia ophiolite (SW Pacific) as a case study? Bulletin de la Société géologique de France 6, 373–84.Google Scholar
Laskar, J., Fienga, A., Gastineau, M. & Manche, H. 2011. La2010: a new orbital solution for the long-term motion of the Earth. Astronomy & Astrophysics 532, A89. doi: 10.1051/0004-6361/201116836, 15 pp.Google Scholar
Le Grand, H. 2002. Plate tectonics, terranes and continental geology. In The Earth Inside and Out; Some Major Contributions to Geology in the Twentieth Century (ed. Oldroyd, D.), pp. 199213. London: Geological Society of London.Google Scholar
Le Hir, G., Donnadieu, Y., Goddéris, Y., Meyer-Berthaud, B., Ramstein, G. & Blakey, R. 2011. The climate change caused by the land plant invasion in the Devonian. Earth and Planetary Science Letters 310, 203–12.Google Scholar
Le Pichon, X. 1968. Sea floor spreading and continental drift. Journal of Geophysical Research 73, 3661– 97.Google Scholar
Li, Z.-X., Bogdanova, S., Collins, A., Davidson, A., De Waele, B., Ernst, R., Fitzsimons, I., Fuck, R., Gladkochub, D., Jacobs, J., Karlstrom, K., Lu, S., Natapov, L., Pease, V., Pisarevsky, S., Thrane, K. & Vernikovsky, V. 2008. Assembly, configuration, and break-up history of Rodinia: a synthesis. Precambrian Research 160, 179210.Google Scholar
Li, Z.-X. & Powell, C. 2001. An outline of the palaeogeographic evolution of the Australasian region since the beginning of the Neoproterozoic. Earth-Science Reviews 53, 237–77.Google Scholar
Manatschal, G. & Müntener, O. 2009. A type sequence across an ancient magma-poor ocean-continent transition: the example of the western Alpine Tethys ophiolites. Tectonophysics 473, 419.Google Scholar
Markwick, P. 2007. The palaeogeographic and palaeoclimatic significance of climate proxies for data-model comparisons. In Deep-Time Perspectives on Climate Change: Marrying the Signal from Computer Models and Biological Proxies (eds Williams, M., Haywood, A., Gregory, J. & Schmidt, D.), pp. 251312. The Micropalaeontological Society, Special Publication. The Geological Society of London.Google Scholar
Markwick, P. & Valdes, P. 2004. Palaeo-digital elevation models for use as boundary conditions in coupled ocean-atmosphere GCM experiments: a Maastrichian (late Cretaceous) example. Palaeogeography, Palaeoclimatology, Palaeoecology 213, 3763.Google Scholar
Matthews, K., Maloney, K. T., Zahirovic, S., Williams, S. E. & Müller, R. D. 2016. Global plate boundary evolution and kinematics since the late Paleozoic. Global and Planetary Change 146, 226–50.Google Scholar
May, D., Schellart, W. & Moresi, L. 2013. Overview of adaptative finite element analysis in computational geodynamics. Journal of Geodynamics 70, 120.Google Scholar
McElhinny, M., Powell, C. & Pisarevsky, S. 2003. Paleozoic terranes of eastern Australia and the drift history of Gondwana. Tectonophysics 362, 4165.Google Scholar
McKenzie, D. & Parker, R. 1967. The North Pacific: an example of tectonics on a sphere. Nature 216, 1276– 80.Google Scholar
McKerrow, W. & Scotese, C. (eds) 1990. Palaeozoic, Palaeogeography and Biogeography. Geological Society of London, Memoir no. 12, 435 pp.Google Scholar
McMenamin, M. & McMenamin, D. 1990. The Emergence of Animals: The Cambrian Breakthrough. New York: Columbia University Press, 217 pp.Google Scholar
Meert, J. 2012. What's in a name? The Columbia (Paleopangaea/Nuna) supercontinent. Gondwana Research 21, 987–93.Google Scholar
Meert, J. G., van der Voo, R., Powell, C. McA., Li, Z.-X., McElhinny, M. W., Chen, Z. & Symons, D. T. A. 1993. A plate-tectonic speed limit? Nature 363, 216–7.Google Scholar
Metcalfe, I. 2011. Tectonic framework and Phanerozoic evolution of Sundaland. Gondwana Research 19, 321.Google Scholar
Michael, P., Langmuir, C., Dick, H., Snow, J., Goldstein, S., Graham, D., Lehnert, K., Kurras, G., Jokat, W., Muhe, R. & Edmonds, H. 2003. Magmatic and amagmatic seafloor generation at the ultraslow-spreading Gakkel ridge, Arctic Ocean. Nature 423, 956–61.Google Scholar
Morel, P. & Irving, E. 1978. Tentative paleocontinental maps for the Early Phanerozoic and Proterozoic. The Journal of Geology 86, 535–61.Google Scholar
Morgan, J. 1968. Rises, trenches, great faults, and crustal blocks. Journal of Geophysical Research 73, 1959–82.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
Mueller, S. & Phillips, R. 1991. On the initiation of subduction. Journal of Geophysical Research 96, 651–65.Google Scholar
Müller, R. D. 2010. Sinking continents. Nature Geoscience 3, 7980.Google Scholar
Müller, R. D., Dutkiewicz, A., Seton, M. & Gaina, C. 2013. Seawater chemistry driven by supercontinent assembly, breakup and dispersal. Geology 41, 907–10.Google Scholar
Müller, R. D., Dyksterhuis, S. & Rey, P. 2012. Australian paleo-stress fields and tectonic reactivation over the past 100 Ma. Australian Journal of Earth Sciences 59, 1328.Google Scholar
Müller, R. D., Roest, W., Royer, J.-Y., Gahagan, L. & Sclater, J. 1996. Age of the Ocean Floor. World Dara Center-A for Marine Geology and Geophysics Report MGG-12, Data Announcement 96-MGG-04. Boulder: National Geophysical Data Center, 1 pp.Google Scholar
Müller, R. D., Roest, W., Royer, J.-Y., Gahagan, L. & Sclater, J. 1997. Digital isochrons of the World's ocean floor. Journal of Geophysical Research 102, 3211–4.Google Scholar
Müller, R. D., Royer, J.-Y. & Lawver, L. 1993. Revised plate motion relative to the hotspots from combined Atlantic and Indian Ocean hotspot tracks. Geology 21, 275–8.Google Scholar
Müller, R. D., Sdrolias, M., Gaina, C. & Roest, W. 2008. Age, spreading rates, and spreading asymmetry of the world's ocean crust. Geochemistry, Geophysics, Geosystems (G3) 9 (4), Q04006. doi: 10.1029/2007GC001743, 19 pp.Google Scholar
Oliver, J. & Isacks, B. 1967. Deep earthquake zones, anomalous structures in the upper mantle, and the lithosphere. Journal of Geophysical Research 72, 4259–75.Google Scholar
Ortelius, A. 1570. Thesaurus Geographicus. Antwerpen: Plantin, 1587, 736 pp.Google Scholar
Parrish, J. 1985. Latitudinal Distribution of Land and Shelf and Absorbed Solar Radiation During the Phanerozoic. United States Geological Survey, Open-File Report, 85-31, 21 pp.Google Scholar
Parrish, J., Ziegler, A. & Scotese, C. 1982. Rainfall patterns and the distribution of coals and evaporates in the Mesozoic and Cenozoic. Palaeogeography, Palaeoclimatology, Palaeoecology 40, 67101.Google Scholar
Perroud, M., Brunetti, M. & Vérard, C. 2015. Sensitivity of the ocean system to the bathymetry in numerical simulations of climate. Poster #EGU2015-5336, European Geophysical Union meeting, Vienna, April 14th.Google Scholar
Pindell, J. & Kennan, L. 2009. Tectonic evolution of the Gulf of Mexico, Caribbean and northern South America in the mantle reference frame: an update. In The Origin and Evolution of the Caribbean Plate (eds James, K., Lorente, M. & Pindell, J.), pp. 155. Geological Society of London, Special Publication no. 328.Google Scholar
Pohl, A., Donnadieu, Y., Le Hir, G., Buoncristiani, J.-F. & Vennin, E. 2014. Effect of the Ordovician paleogeography on the (in)stability of the climate. Climate of the Past – Discussions 10, 2767–804.Google Scholar
Püthe, C. & Gerya, T. 2014. Dependence of mid-ocean ridge morphology on spreading rate in numerical 3-D models. Gondwana Research 25, 270–83.Google Scholar
Renaut, R. 1994. Carbonate and evaporite sedimentation at Clinton Lake, British Columbia, Canada. In Palaeoclimate and Basin Evolution of Playa Systems (ed. Rosen, R.), pp. 4968. Geological Society of America, Special Paper no. 289.Google Scholar
Ribeiro, A. 2002. Soft Plate and Impact Tectonics. Berlin: Springer, 324 pp.Google Scholar
Ricard, Y., Doglioni, C. & Sabadini, R. 1991. Differential rotation between lithosphere and mantle: a consequence of lateral mantle viscosity variations. Journal of Geophysical Research B 96, 8407–15.Google Scholar
Rogers, J. & Santosh, M. 2002. Configuration of Columbia, a Mesoproterozoic supercontinent. Gondwana Research 5, 522.Google Scholar
Rogers, J. & Santosh, M. 2003. Supercontinent in Earth History. Gondwana Research 6, 357–68.Google Scholar
Ronov, A., Khain, V. & Balukhovsky, A. 1989. Atlas of Lithological–Paleogeographical Maps of the World, Mesozoic and Cenozoic of Continents and Oceans. Leningrad: U.S.S.R Academy of Sciences, 79 pp.Google Scholar
Ronov, A., Khain, V., Blukhovsky, A. & Seslavinsky, K. 1980. Quantitative analysis of Phanerozoic sedimentation. Sedimentary Geology 25, 311–25.Google Scholar
Ronov, A., Khain, V. & Seslavinsky, K. 1984. Atlas of Lithological–Paleogeographical Maps of the World, Late Precambrian and Paleozoic of Continents. Leningrad: U.S.S.R Academy of Sciences, 70 pp.Google Scholar
Ross, M. I. & Scotese, C. R. 1988. A hierarchical model of the Gulf of Mexico and Caribbean region. Tectonophysics 155, 139–68.Google Scholar
Ross, M. & Scotese, C. 2000. PaleoGIS/Arcview 3.5. Paleomap Project. Arlington: University of Texas.Google Scholar
Ross, M., Scotese, C. & Otto-Bliesner, B. 1992. Phanerozoic paleoclimate simulations: a comparison of the parametric climate model and the low resolution climate model. The Geological Society of America (GSA), Abstracts A89, Annual Meeting 1992, Cincinnati.Google Scholar
Rowley, D., Raymond, A., Totman Parrish, J., Lottes, A., Scotese, C. & Ziegler, A. 1985. Carboniferous paleogeographic, phytogeographic, and paleoclimatic reconstructions. International Journal of Coal Geology 5, 742.Google Scholar
Royer, J.-Y., Müller, D., Gahagan, L., Lawver, L., Mayes, C., Nürnberg, D. & Sclater, J. 1992. A Global Isochron Chart. University of Texas, Institute for Geophysics Technical Report, 117, 38 pp.Google Scholar
Schardt, H. 1893. Sur l'origine des Préalples romandes. Archives de physique et des sciences naturelles de Genève 3, 570–83.Google Scholar
Schettino, A. & Scotese, C. 2005. Apparent polar wander paths for the major continents (200 Ma to the present day): a palaeomagnetic reference frame for global plate tectonic reconstructions. Geophysical Journal International 163, 727–59.Google Scholar
Schettino, A. & Turco, E. 2011. Tectonic history of the western Tethys since the Late Triassic. GSA Bulletin 123, 89105.Google Scholar
Schmalzl, J., Houseman, G. & Hansen, U. 1996. Mixing in vigorous, time-dependent three-dimensional convection and application to Earth's mantle. Journal of Geophysical Research 101, 21847–58.Google Scholar
Schmidt, P., Powell, C., Li, Z.-X. & Thrupp, G. 1990. Reliability of Palaeozoic palaeomagnetic poles and APWP of Gondwanaland. In Reliability of Palaeomagnetic Data (eds Schmidt, P. & Van der Voo, R.). Tectonophysics 184, 87100.Google Scholar
Schreiber, T. 2000. Measuring information transfer. Physical Review Letters 85, 461–4.Google Scholar
Scotese, C. 1976. A continental drift ‘flip book’. Computers and Geology 2, 13116.Google Scholar
Scotese, C. 2015. Plate tectonics driving mechanisms: some simple rules that explain why the plates move the way we do. Technical Report. doi: 10.13140/2.1.3376. 8325.Google Scholar
Scotese, C. & Baker, C. 1975. Continental drift reconstructions and animation. Journal of Geological Education 23, 167–71.Google Scholar
Scotese, C., Bambach, R., Barton, C., Van der Voo, R. & Ziegler, A. 1979. Paleozoic base maps. Journal of Geology 87, 217–77.Google Scholar
Scotese, C. & Barrett, S. 1990. Gondwana's movement over the South Pole during the Palaeozoic: evidence from lithological indicators of climate. In Palaeozoic Palaeogeography and Biogeography (eds McKerrow, W. & Scotese, C.), pp. 7585. Geological Society of London, Memoir no. 12.Google Scholar
Scotese, C. R., Gahagan, L. M., & Larson, R. L. 1988. Plate tectonic reconstructions of the Cretaceous and Cenozoic ocean basins. In The 8th Geodynamics Symposium, Mesozoic and Cenozoic Plate Reconstructions (eds Scotese, C. R. & Sager, W. W.). Tectonophysics 155, 261–83.Google Scholar
Sellwood, B. & Valdes, P. 2008. Jurassic climates. Proceedings of the Geologists’ Association 119, 517.Google Scholar
Sengör, A., Dewey, F. & Robertson, A. 1990. Terranology: vice or virtue? Philosophical Transactions of the Royal Society of London A331, 457–77.Google Scholar
Seton, M., Müller, D., Zahirovic, S., Gaina, C., Torsvik, T., Shephard, G., Talsma, A., Gurnis, M., Turner, M., Maus, S. & Chandler, M. 2012. Global continental and ocean basin reconstructions since 200 Ma. Earth-Science Reviews 113, 212–70.Google Scholar
Seton, M., Whittaker, J., Wessel, P., Müller, D., DeMets, C., Merkouriev, S., Cande, S., Gaina, C., Eagles, G., Granot, R., Stock, J., Wright, N. & Williams, S. 2014. Community infrastructure and repository for marine magnetic identifications. Geochemistry, Geophysics, Geosystems 15, 1629–41.Google Scholar
Shaviv, N. & Veizer, J. 2003. Celestial driver of Phanerozoic climate? GSA Today 13, 410.Google Scholar
Shemenda, A. 1985. Modelling of the opening mechanism for certain types of back arc basins. Oceanography 25, 204–10.Google Scholar
Shemenda, A. 1993. Subduction of the lithosphere and back arc dynamics: insights from physical modelling. Journal of Geophysical Research 98, 16167–85.Google Scholar
Shephard, G. E., Liu, L., Müller, R. D. & Gurnis, M. 2012. Dynamic topography and anomalously negative residual depth of the Argentine Basin. Gondwana Research 22, 658–63.Google Scholar
Smith, A. 1999. Gondwana: its shape, size and position from Cambrian to Triassic times. Journal of African Earth Sciences 28, 7197.Google Scholar
Smith, A. G., Briden, J. C. & Drewry, G. E. 1973. Phanerozoic world maps. In Organisms and Continents Through Time (ed. Hughes, N. F.), pp. 142. Special Papers in Palaeontology 12. London: The Palaeontological Association, 334 pp.Google Scholar
Smith, M., Kurtz, J., Richards, S., Forster, M. & Lister, G. 2007. A re-evaluation of the break-up of South America and Africa using deformable mesh reconstruction software. Journal of the Virtual Explorer 25, doi: 10.3809/jvirtex.2007.00163, 13 pp.Google Scholar
Smith, A. G., Smith, D. G. & Funnell, B. M. 1994. Atlas of Mesozoic and Cenozoic Coastlines. Cambridge: Cambridge University Press, 456 pp. & 31 maps.Google Scholar
Stampfli, G. & Borel, G. 2002. A plate tectonic model for the Paleozoic and Mesozoic constrained by dynamic plate boundaries and restored synthetic oceanic isochrons. Earth and Planetary Science Letters 196, 1733.Google Scholar
Stampfli, G. & Borel, G. 2004. The Transmed transects in space and time: constraints on the paleotectonic evolution of the Mediterranean Domain. In The TRANSMED Atlas: The Mediterranean Region from Crust to Mantle (eds Cavazza, W., Roure, F., Sparkman, W., Stampfli, G., Ziegler, P.), pp. 5380. Berlin: Springer.Google Scholar
Stampfli, G. M., Borel, G. D., Marchant, R. & Mosar, J. 2002. Western Alps geological constraints on western Tethyan reconstructions. In Reconstruction of the Evolution of the Alpine–Himalayan Orogen (eds Rosenbaum, G. & Lister, G. S.). Journal of the Virtual Explorer 7, doi: 10.3809/jvirtex.2002.00057.Google Scholar
Stampfli, G., Hochard, C., Vérard, C., Wilhem, C. & von Raumer, J. 2013. The formation of Pangea. Tectonophysics 593, 119.Google Scholar
Suess, E. 1892. Das Antlitz der Erde I–III. Prag & Wien: F. Tempsky / Leipzig: G. Freytag, 778 pp.Google Scholar
Sugihara, G., May, R., Ye, H., Hsieh, C.-H., Deyle, E., Fogarty, M. & Munch, S. 2012. Detecting causality in complex ecosystems. Science 338 (6106), 496500.Google Scholar
Thomas, P. 2010. La convection, moteur du manteau. In La Terre à Cœur Ouvert. Dossier Pour la Science 67, Avril-Juin, 3845.Google Scholar
Tindall, J., Flecker, R., Valdes, P., Schmidt, D., Markwick, P. & Harris, J. 2010. Modelling the oxygen isotope distribution of ancient seawater using a coupled ocean-atmosphere GCM: implications for reconstructing early Eocene climate. Earth and Planetary Science Letters 292, 265–73.Google Scholar
Torsvik, T. & Cocks, R. 2011. The Palaeozoic palaeogeography of central Gondwana. In The Formation and Evolution of Africa: A Synopsis of 3.8 Ga of Earth History (eds Hinsbergen, D. van, Buiter, S., Torsvik, T., Gaina, C. & Webb, S.), pp. 137–66. Geological Society of London, Special Publication no. 357.Google Scholar
Torsvik, T. & Cocks, R. 2013. Gondwana from top to base in space and time. Gondwana Research 24, 9991030.Google Scholar
Torsvik, T. & Cocks, R. 2017. Earth History and Palaeogeography. Cambridge: Cambridge University Press, 317 pp.Google Scholar
Torsvik, T., Müller, D., Van der Voo, R., Steinberger, B. & Gaina, C. 2008. Global plate motion frames: toward a unified model. Reviews of Geophysics 46, RG3004. doi: 10.1029/2007RG000227, 44 pp.Google Scholar
Torsvik, T. & Smethrust, M. 1999. Plate tectonic modelling: virtual reality with GMap. Computers & Geosciences 25, 395402.Google Scholar
Torsvik, T., Steinberger, B., Gurnis, M. & Gaina, C. 2010. Plate tectonics and net lithosphere rotation over the past 150 My. Earth and Planetary Science Letters 291, 106–12.Google Scholar
Torsvik, T. & Van der Voo, R. 2002. Refining Gondwana and Pangea palaeogeography: estimates of Phanerozoic non-dipole (octupole) fields. Geophysical Journal International 151, 771–94.Google Scholar
Torsvik, T., Van der Voo, R., Preeden, U., Mac Niocaill, C., Steinberger, B., Doubrovine, P., van Hinsbergen, D., Domeier, M., Gaina, C., Tohver, E., Meert, J., McCausland, Ph. & Cocks, R. 2012. Phanerozoic polar wander, palaeogeography and dynamics. Earth-Science Reviews 114, 325–68.Google Scholar
Turcotte, D. L. & Schubert, G. 2002. Geodynamics. 2nd Edition. Cambridge: Cambridge University Press, 456 pp.Google Scholar
Van der Voo, R., Scotese, C. & Bonhommet, N. (eds) 1984. Plate Reconstruction from Paleozoic Paleomagnetism. American Geophysical Union, Geodynamics Series vol. 12, Washington, DC, USA, 136 pp.Google Scholar
Vérard, C. In press. Panalesis: towards global synthetic palaeogeographies using integration and coupling of manifold models. Geological Magazine.Google Scholar
Vérard, C., Flores, K., Stampfli, G. 2012 b. Geodynamic reconstructions of the South America – Antarctica plate system. Journal of Geodynamics 53, 4360.Google Scholar
Vérard, C., Hochard, C., Baumgartner, P. O. & Stampfli, G. 2015 a. Geodynamic evolution of the Earth over the Phanerozoic: plate tectonic activity and palaeo-climatic indicators. Journal of Palaeogeography 4, 167–88.Google Scholar
Vérard, C., Hochard, C., Baumgartner, P. O. & Stampfli, G. 2015 b. 3D palaeogeographic reconstructions of the Phanerozoic versus sea-level and Sr-ratio variations. Journal of Palaeogeography 4, 6484.Google Scholar
Vérard, C., Hochard, C. & Stampfli, G. 2012 a. Non-random distribution of Euler poles: is plate tectonics subject to rotational effects? Terra Nova 24, 467–76.Google Scholar
Vérard, C. & Stampfli, G. 2013 a. Geodynamic reconstructions of the Australides – 1: Palaeozoic. Geosciences 3, 311–30.Google Scholar
Vérard, C. & Stampfli, G. 2013 b. Geodynamic reconstructions of the Australides – 2: Mesozoic and Cainozoic. Geosciences, 3, 331–53.Google Scholar
Vine, F. & Hess, H. 1968. Sea floor spreading. Princeton University Technical Report. The Sea 4, 35 pp.Google Scholar
Vine, F. & Matthews, D. 1963. Magnetic anomalies over oceanic ridges. Nature 4897, 947–9.Google Scholar
Vrielynck, B. & Bouysse, P. 2003. Le Visage Changeant de la Terre: L’éclatement de la Pangée et la Mobilité des Continents au Cours des Derniers 250 Millions d'Années. Paris: Commission de la Carte Géologique du Monde, UNESCO Edition, 32 pp.Google Scholar
Wadati, K. 1928. Shallow and deep earthquakes [1st paper]. Geophysical Magazine 1, 162202.Google Scholar
Wadati, K. 1929. Shallow and deep earthquakes [2nd paper]. Geophysical Magazine 2, 136.Google Scholar
Wadati, K. 1931. Shallow and deep earthquakes [3rd paper]. Geophysical Magazine 4, 231–83.Google Scholar
Warners-Ruckstuhl, K., Govers, R. & Wortel, R. 2012. Lithosphere-mantle coupling and the dynamics of the Eurasian plate. Geophysical Journal International 189, 1253–76.Google Scholar
Warners-Ruckstuhl, K., Govers, R. & Wortel, R. 2013. Dynamics and stress field of the Eurasian plate. Poster #EGU2013-9094 at the EGU General Assembly, Vienna.Google Scholar
Wegener, A. 1912. Die Enstehung der Kontinente. Geologische Rundschau 2, 276–92.Google Scholar
Wegener, A. 1915. Die Enstehung der Kontinente und Ozeane. Braunschweig: Vieweg F. & Sohn, 94 pp.Google Scholar
Wegener, A. 1929. Die Entstehung der Kontinente und Ozeane. Braunschweig: Vieweg F & Sohn, 231 pp.Google Scholar
Wessel, P. & Kroenke, L. 2009. Observations of geometry and ages constrain relative motion of Hawaii and Louisville plumes. Earth and Planetary Science Letters 284, 467–72.Google Scholar
Wilhem, C., Windley, B. & Stampfli, G. 2012. The Altaids of Central Asia: a tectonic and evolutionary innovative review. Earth-Science Reviews 113, 303–41.Google Scholar
Williams, S., Müller, R. D., Landgrebe, T. & Whittaker, J. 2012. An open-source software environment for visualizing and refining plate tectonic reconstructions using high-resolution geological and geophysical data sets. GSA Today 22 (4/5), 49.Google Scholar
Wilson, J. T. 1965. A new class of faults and their bearing on continental drift. Nature 207 (4995), 343–7.Google Scholar
Wu, B., Conrad, C., Heuret, A., Lithgow-Bertelloni, C. & Lallemand, S. 2008. Reconciling strong slab pull and weak plate bending: the plate motion constraint on the strength of mantle slabs. Earth Planetary Science Letters 272, 412–21.Google Scholar
Zharkov, M. & Chumakov, N. 2001. Paleogeography and sedimentation settings during Permian–Triassic reorganizations in biosphere. Stratigraphy and Geological Correlation 9, 340–63.Google Scholar
Zharkov, M., Murdmaa, I. & Filatova, N. 1995. Paleogeography of the mid-Cretaceous period. Stratigraphy and Geological Correlation 3, 216–40.Google Scholar
Zharkov, M., Murdmaa, I. & Filatova, N. 1998 a. Paleogeography of the Berriasian–Barremian ages of the Early Cretaceous. Stratigraphy and Geological Correlation 6, 4769.Google Scholar
Zharkov, M., Murdmaa, I. & Filatova, N. 1998 b. Paleogeography of the Coniacian–Maastrichian ages of the Early Cretaceous. Stratigraphy and Geological Correlation 6, 209–21.Google Scholar
Ziegler, P. 1982. Geological Atlas of Western and Central Europe. Shell International Petroleum Mij. B.V. and Amsterdam: Elsevier Science Publishers, 130 pp.Google Scholar
Ziegler, A., Scotese, C. & Barrett, S. 1983. Mesozoic and Cenozoic paleogeographic maps. In Tidal Friction and the Earth's Rotation, II (eds Brosche, P. & Sündermann, J.), pp. 240–52. Berlin: Springer-Verlag.Google Scholar
Zonenshain, L. & Kuzmin, M. 1992. Paleogeodynamics; The Plate Tectonic Evolution of the Earth. Moscow: Nauka. English version published by American Geophysical Union, Washington, DC, USA, 1997, 218 pp.Google Scholar