Hostname: page-component-76fb5796d-dfsvx Total loading time: 0 Render date: 2024-04-26T17:40:14.246Z Has data issue: false hasContentIssue false

Full-plate modelling in pre-Jurassic time

Published online by Cambridge University Press:  19 December 2017

MATHEW DOMEIER*
Affiliation:
Centre for Earth Evolution and Dynamics (CEED), University of Oslo, Oslo, Norway
TROND H. TORSVIK
Affiliation:
Centre for Earth Evolution and Dynamics (CEED), University of Oslo, Oslo, Norway Helmholtz Centre Potsdam, GFZ, Potsdam, Germany Geodynamics Team, Geological Survey of Norway, Trondheim, Norway School of Geosciences, University of Witwatersrand, Johannesburg, South Africa
*
Author for correspondence: mathewd@geo.uio.no

Abstract

A half-century has passed since the dawning of the plate tectonic revolution, and yet, with rare exception, palaeogeographic models of pre-Jurassic time are still constructed in a way more akin to Wegener's paradigm of continental drift. Historically, this was due to a series of problems – the near-complete absence of in situ oceanic lithosphere older than 200 Ma, a fragmentary history of the latitudinal drift of continents, unconstrained longitudes, unsettled geodynamic concepts and a lack of efficient plate modelling tools – which together precluded the construction of plate tectonic models. But over the course of the last five decades strategies have been developed to overcome these problems, and the first plate model for pre-Jurassic time was presented in 2002. Following on that pioneering work, but with a number of significant improvements (most notably longitude control), we here provide a recipe for the construction of full-plate models (including oceanic lithosphere) for pre-Jurassic time. In brief, our workflow begins with the erection of a traditional (or ‘Wegenerian’) continental rotation model, but then employs basic plate tectonic principles and continental geology to enable reconstruction of former plate boundaries, and thus the resurrection of lost oceanic lithosphere. Full-plate models can yield a range of testable predictions that can be used to critically evaluate them, but also novel information regarding long-term processes that we have few (or no) alternative means of investigating, thus providing exceptionally fertile ground for new exploration and discovery.

Type
Original Articles
Copyright
Copyright © Cambridge University Press 2017 

Access options

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

References

Andersen, T. 2014. The detrital zircon record: supercontinents, parallel evolution – or coincidence? Precambrian Research 244, 279–87.Google Scholar
Andersen, T., Kristoffersen, M. & Elburg, M. A. 2016. How far can we trust provenance and crustal evolution information from detrital zircons? A South African case study. Gondwana Research 34, 129–48.Google Scholar
Berner, R. A. 2004. The Phanerozoic Carbon Cycle: CO2 and O2. Oxford: Oxford University Press.Google Scholar
Besse, J. & Courtillot, V. 2002. Apparent and true polar wander and the geometry of the geomagnetic field over the last 200 Myr. Journal of Geophysical Research: Solid Earth 107 (B11), 2300. doi: 10.1029/2000JB000050.Google Scholar
Boucot, A. J., Xu, C., Scotese, C. R. & Morley, R. J. 2013. Phanerozoic Paleoclimate: An Atlas of Lithologic Indicators of Climate. Tulsa, Oklahoma: Society of Economic Paleontologists and Mineralogists.Google Scholar
Bradley, D. C. 2008. Passive margins through earth history. Earth-Science Reviews 91 (1), 126.Google Scholar
Buiter, S. J. & 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 (2), 627–53.Google Scholar
Bull, A. L., Domeier, M. & Torsvik, T. H. 2014. The effect of plate motion history on the longevity of deep mantle heterogeneities. Earth and Planetary Science Letters 401, 172–82.Google Scholar
Bunge, H.-P., Richards, M. A., Lithgow-Bertelloni, C., Baumgardner, J. R., Grand, S. P. & Romanowicz, B. A. 1998. Time scales and heterogeneous structure in geodynamic Earth models. Science 280 (5360), 91–5.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 (1), 4960.Google 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 (3), 531–38.Google Scholar
Cawood, P. A., Kröner, A., Collins, W. J., Kusky, T. M., Mooney, W. D. & Windley, B. F. 2009. Accretionary orogens through Earth history. Geological Society, London, Special Publications 318 (1), 136.Google Scholar
Chandler, M. T., Wessel, P., Taylor, B., Seton, M., Kim, S.-S. & Hyeong, K. 2012. Reconstructing Ontong Java Nui: implications for Pacific absolute plate motion, hotspot drift and true polar wander. Earth and Planetary Science Letters 331, 140–51.Google Scholar
Chang, T. 1988. Estimating the relative rotation of two tectonic plates from boundary crossings. Journal of the American Statistical Association 83 (404), 1178–83.Google Scholar
Chase, C. 1972. The N plate problem of plate tectonics. Geophysical Journal International 29 (2), 117–22.Google Scholar
Chase, C. G. 1978. Plate kinematics: the Americas, East Africa, and the rest of the world. Earth and Planetary Science Letters 37 (3), 355–68.Google Scholar
Chopin, C. 2003. Ultrahigh-pressure metamorphism: tracing continental crust into the mantle. Earth and Planetary Science Letters 212 (1), 114.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 (6), 715–37.Google Scholar
Cocks, L. & Fortey, R. 1982. Faunal evidence for oceanic separations in the Palaeozoic of Britain. Journal of the Geological Society 139 (4), 465–78.Google Scholar
Cocks, L. & Torsvik, T. 2002. Earth geography from 500 to 400 million years ago: a faunal and palaeomagnetic review. Journal of the Geological Society 159 (6), 631–44.Google Scholar
Coltice, N., Seton, M., Rolf, T., Müller, R. & Tackley, P. J. 2013. Convergence of tectonic reconstructions and mantle convection models for significant fluctuations in seafloor spreading. Earth and Planetary Science Letters 383, 92100.Google Scholar
Conrad, C. P. & Behn, M. D. 2010. Constraints on lithosphere net rotation and asthenospheric viscosity from global mantle flow models and seismic anisotropy. Geochemistry, Geophysics, Geosystems 11 (5), Q05W05. doi: 10.1029/2009GC002970.Google Scholar
Conrad, C. P. & Lithgow-Bertelloni, C. 2002. How mantle slabs drive plate tectonics. Science 298 (5591), 207–9.Google Scholar
Conrad, C. P., Steinberger, B. & Torsvik, T. H. 2013. Stability of active mantle upwelling revealed by net characteristics of plate tectonics. Nature 498 (7455), 479–82.Google Scholar
Cox, A. & Hart, R. 1986. Plate Tectonics: How It Works. Malden, Massachussetts: Blackwell Science, 392 pp.Google Scholar
Dalziel, I. W. 1997. Overview: Neoproterozoic-Paleozoic geography and tectonics: review, hypothesis, environmental speculation. Geological Society of America Bulletin 109 (1), 1642.Google Scholar
Davies, J. H. & Stevenson, D. J. 1992. Physical model of source region of subduction zone volcanics. Journal of Geophysical Research: Solid Earth 97 (B2), 2037–70.Google Scholar
Dewey, J. F. & Bird, J. M. 1970. Mountain belts and the new global tectonics. Journal of Geophysical Research 75 (14), 2625–47.Google Scholar
Dewey, J. F. & Burke, K. C. 1973. Tibetan, Variscan, and Precambrian basement reactivation: products of continental collision. The Journal of Geology 81 (6), 683–92.Google Scholar
Dewey, J., Holdsworth, R. & Strachan, R. 1998. Transpression and transtension zones. In Continental Transpressional and Transtensional Tectonics (eds Holdsworth, R. E., Strachan, R. A. & Dewey, J. F.), pp. 114. Geological Society of London, Special Publication no. 135.Google Scholar
Dickinson, W. R. 1970. Relations of andesites, granites, and derivative sandstones to arc-trench tectonics. Reviews of Geophysics 8 (4), 813–60.Google Scholar
Dickinson, W. R. 1971. Plate tectonics in geologic history. Science 174 (4005), 107–13.Google Scholar
Dietz, R. S. 1961. Continent and ocean basin evolution by spreading of the sea floor. Nature 190 (4779), 854–7.Google Scholar
Di Giuseppe, E., Van Hunen, J., Funiciello, F., Faccenna, C. & Giardini, D. 2008. Slab stiffness control of trench motion: insights from numerical models. Geochemistry, Geophysics, Geosystems 9 (2), Q02014. doi: 10.1029/2007GC001776.Google Scholar
Dilek, Y. & Furnes, H. 2011. Ophiolite genesis and global tectonics: geochemical and tectonic fingerprinting of ancient oceanic lithosphere. Geological Society of America Bulletin 123 (3–4), 387411.Google Scholar
Domeier, M. 2016. A plate tectonic scenario for the Iapetus and Rheic Oceans. Gondwana Research 36, 275–95.Google Scholar
Domeier, M., Doubrovine, P. V., Torsvik, T. H., Spakman, W. & Bull, A. L. 2016. Global correlation of lower mantle structure and past subduction. Geophysical Research Letters 43, 4945. doi: 10.1002/2016GL068827.Google Scholar
Domeier, M. & Torsvik, T. H. 2014. Plate tectonics in the late Paleozoic. Geoscience Frontiers 5 (3), 303–50.Google Scholar
Domeier, M., Van der Voo, R. & Torsvik, T. H. 2012. Paleomagnetism and Pangea: the road to reconciliation. Tectonophysics 514, 1443.Google Scholar
Doubrovine, P. V., Steinberger, B. & Torsvik, T. H. 2012. Absolute plate motions in a reference frame defined by moving hot spots in the Pacific, Atlantic, and Indian oceans. Journal of Geophysical Research: Solid Earth 117 (B9), 21562202.Google Scholar
Doubrovine, P. V., Steinberger, B. & Torsvik, T. H. 2016. A failure to reject: testing the correlation between large igneous provinces and deep mantle structures with EDF statistics. Geochemistry, Geophysics, Geosystems 17 (3), 1130–63.Google Scholar
Ernst, W. 1971. Metamorphic zonations on presumably subducted lithospheric plates from Japan, California and the Alps. Contributions to Mineralogy and Petrology 34 (1), 4359.Google Scholar
Ernst, W. & Liou, J. 2008. High-and ultrahigh-pressure metamorphism: past results and future prospects. American Mineralogist 93 (11–12), 1771–86.Google Scholar
Evans, D. A. 2006. Proterozoic low orbital obliquity and axial-dipolar geomagnetic field from evaporite palaeolatitudes. Nature 444 (7115), 51–5.Google Scholar
Fisher, R. 1953. Dispersion on a sphere. Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences 217 (1130), 295305.Google Scholar
Flament, N., Williams, S., Müller, R., Gurnis, M. & Bower, D. J. 2017. Origin and evolution of the deep thermochemical structure beneath Eurasia. Nature Communications 8, 14164. doi: 10.1038/ncoms14164.Google Scholar
Forsyth, D. & Uyeda, S. 1975. On the relative importance of the driving forces of plate motion. Geophysical Journal International 43 (1), 163200.Google Scholar
Funiciello, F., Faccenna, C., Heuret, A., Lallemand, S., Di Giuseppe, E. & Becker, T. 2008. Trench migration, net rotation and slab–mantle coupling. Earth and Planetary Science Letters 271 (1), 233–40.Google Scholar
Gass, I. 1968. Is the Troodos massif of Cyprus a fragment of Mesozoic ocean floor? Nature 220 (5162), 3942.Google Scholar
Gold, T. 1955. Instability of the Earth's axis of rotation. Nature 175 (4456), 526–9.Google Scholar
Goldreich, P. & Toomre, A. 1969. Some remarks on polar wandering. Journal of Geophysical Research 74 (10), 2555–67.Google Scholar
Golonka, J. 2002. Plate-tectonic maps of the Phanerozoic. In Phanerozoic Reef Patterns (eds Kiessling, W., Flugel, E. & Golonka, J.), pp. 1120. SEPM Special Publication no. 72.Google Scholar
Gordon, R. G. 1998. The plate tectonic approximation: plate nonrigidity, diffuse plate boundaries, and global plate reconstructions. Annual Review of Earth and Planetary Sciences 26 (1), 615–42.Google Scholar
Gordon, W. A. 1975. Distribution by latitude of Phanerozoic evaporite deposits. The Journal of Geology 83 (6), 671–84.Google Scholar
Gordon, R. G. & Jurdy, D. M. 1986. Cenozoic global plate motions. Journal of Geophysical Research: Solid Earth 91 (B12), 12389–406.Google Scholar
Grove, T., Till, C., Lev, E., Chatterjee, N. & Médard, E. 2009. Kinematic variables and water transport control the formation and location of arc volcanoes. Nature 459 (7247), 694–7.Google Scholar
Gurnis, M. & Torsvik, T. H. 1994. Rapid drift of large continents during the late Precambrian and Paleozoic: paleomagnetic constraints and dynamic models. Geology 22 (11), 1023–6.Google Scholar
Gurnis, M., Turner, M., Zahirovic, S., DiCaprio, L., Spasojevic, S., Müller, R. D., Boyden, J., Seton, M., Manea, V. C. & Bower, D. J. 2012. Plate tectonic reconstructions with continuously closing plates. Computers & Geosciences 38 (1), 3542.Google Scholar
Harland, W. & Gayer, R. 1972. The Arctic Caledonides and earlier oceans. Geological Magazine 109 (4), 289314.Google Scholar
Hellinger, S. 1981. The uncertainties of finite rotations in plate tectonics. Journal of Geophysical Research: Solid Earth 86 (B10), 9312–8.Google Scholar
Hess, H. 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. Boulder, CO: Geological Society of America.Google Scholar
Hochard, C. 2008. GIS and geodatabases application to global scale plate tectonics modelling. Ph.D. thesis, University of Lausanne, Lausanne, Switzerland. Published thesis.Google Scholar
Hoffman, P. F., Kaufman, A. J., Halverson, G. P. & Schrag, D. P. 1998. A Neoproterozoic snowball earth. Science 281 (5381), 1342–6.Google Scholar
Holmes, A. 1931. XVIII. Radioactivity and earth movements. Transactions of the Geological Society of Glasgow 18 (3), 559606.Google Scholar
Huang, K., Opdyke, N. D., Peng, X. & Li, J. 1992. Paleomagnetic results from the Upper Permian of the eastern Qiangtang Terrane of Tibet and their tectonic implications. Earth and Planetary Science Letters 111 (1), 110.Google Scholar
Irving, E. & Briden, J. 1962. Palaeolatitude of evaporite deposits. Nature 196 (4853), 425–8.Google Scholar
Isacks, B., Oliver, J. & Sykes, L. R. 1968. Seismology and the new global tectonics. Journal of Geophysical Research 73 (18), 5855–99.Google Scholar
Jarrard, R. D. 1986. Relations among subduction parameters. Reviews of Geophysics 24 (2), 217–84.Google Scholar
Jurdy, D. & Van der Voo, R. 1975. True polar wander since the Early Cretaceous. Science 187 (4182), 1193–6.Google Scholar
Kirkwood, B. H., Royer, J.-Y., Chang, T. C. & Gordon, R. G. 1999. Statistical tools for estimating and combining finite rotations and their uncertainties. Geophysical Journal International 137 (2), 408–28.Google Scholar
Klootwijk, C. T., Gee, J. S., Peirce, J. W., Smith, G. M. & McFadden, P. L. 1992. An early India-Asia contact: paleomagnetic constraints from Ninetyeast ridge, ODP Leg 121. Geology 20 (5), 395–8.Google Scholar
Le Pichon, X. 1968. Sea-floor spreading and continental drift. Journal of Geophysical Research 73 (12), 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. 2008. Assembly, configuration, and break-up history of Rodinia: a synthesis. Precambrian Research 160 (1), 179210.Google Scholar
Lister, G., Etheridge, M. & Symonds, P. 1986. Detachment faulting and the evolution of passive continental margins. Geology 14 (3), 246–50.Google Scholar
Lithgow-Bertelloni, C. & Richards, M. A. 1998. The dynamics of Cenozoic and Mesozoic plate motions. Reviews of Geophysics 36 (1), 2778.Google Scholar
Lowry, D., Poulsen, C., Horton, D., Torsvik, T. & Pollard, D. 2014. Thresholds for Paleozoic ice sheet initiation. Geology 42 (7), 627–30.Google Scholar
Marty, B. & Tolstikhin, I. N. 1998. CO2 fluxes from mid-ocean ridges, arcs and plumes. Chemical Geology 145 (3), 233–48.Google Scholar
Matsuda, T. & Uyeda, S. 1971. On the Pacific-type orogeny and its model – extension of the paired belts concept and possible origin of marginal seas. Tectonophysics 11 (1), 527.Google Scholar
Matthews, K. J., Maloney, K. T., Zahirovic, S., Williams, S. E., Seton, M. & Müller, R. D. 2016. Global plate boundary evolution and kinematics since the late Paleozoic. Global and Planetary Change 146, 226–50.Google Scholar
McElhinny, M., McFadden, P. & Merrill, R. 1996. The time-averaged paleomagnetic field 0–5 Ma. Journal of Geophysical Research: Solid Earth 101 (B11), 25007–27.Google Scholar
McKenzie, D. P. & Parker, R. L. 1967. The North Pacific: an example of tectonics on a sphere. Nature 216, 1276–80.Google Scholar
McKerrow, W. & Cocks, L. 1986. Oceans, island arcs and olistostromes: the use of fossils in distinguishing sutures, terranes and environments around the Iapetus Ocean. Journal of the Geological Society 143 (1), 185–91.Google Scholar
Meert, J. G. & Torsvik, T. H. 2003. The making and unmaking of a supercontinent: Rodinia revisited. Tectonophysics 375 (1), 261–88.Google Scholar
Meert, J. G. & Van der Voo, R. 1997. The assembly of Gondwana 800–550 Ma. Journal of Geodynamics 23 (3–4), 223–35.Google Scholar
Meert, J. G., Van der Voo, R., Powell, C. M., Li, Z.-X., McElhinny, M. W., Chen, Z. & Symons, D. 1993. A plate-tectonic speed limit? Nature 363 (6426), 216–17.Google Scholar
Merdith, A. S., Collins, A. S., Williams, S. E., Pisarevsky, S., Foden, J. F., Archibald, D., Blades, M. L., Alessio, B. L., Armistead, S. & Plavsa, D. 2017. A full-plate global reconstruction of the Neoproterozoic. Gondwana Research 50, 84134.Google Scholar
Metcalf, R. V. & Shervais, J. W. 2008. Suprasubduction-zone ophiolites: is there really an ophiolite conundrum? In Opholites, Arcs and Batholiths: A Tribute to Cliff Hopson (eds Wright, J. E. & Shervais, J. W.), pp. 191222. Geological Society of America, Special Paper no. 438.Google Scholar
Minster, J. B. & Jordan, T. H. 1978. Present-day plate motions. Journal of Geophysical Research: Solid Earth 83 (B11), 5331–54.Google Scholar
Minster, J., Jordan, T., Molnar, P. & Haines, E. 1974. Numerical modelling of instantaneous plate tectonics. Geophysical Journal International 36 (3), 541–76.Google Scholar
Miyashiro, A. 1972. Metamorphism and related magmatism in plate tectonics. American Journal of Science 272 (7), 629–56.Google Scholar
Molnar, P. 1988. Continental tectonics in the aftermath of plate tectonics. Nature 335 (6186), 131–7.Google Scholar
Molnar, P. & Stock, J. M. 1985. A method for bounding uncertainties in combined plate reconstructions. Journal of Geophysical Research: Solid Earth 90 (B14), 12537–44.Google Scholar
Molnar, P. & Tapponnier, P. 1975. Cenozoic tectonics of Asia: effects of a continental collision. Science 189 (4201), 419–26.Google Scholar
Moores, E. 1982. Origin and emplacement of ophiolites. Reviews of Geophysics 20 (4), 735–60.Google Scholar
Morgan, W. J. 1968. Rises, trenches, great faults, and crustal blocks. Journal of Geophysical Research 73 (6), 1959–82.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 (3), 275–8.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 (4), Q04006. doi: 10.1029/2007GC001743.Google Scholar
Muttoni, G., Mattei, M., Balini, M., Zanchi, A., Gaetani, M. & Berra, F. 2009. The drift history of Iran from the Ordovician to the Triassic. In South Caspian to Central Iran Basins (eds Brunet, M.-F., Wilmsen, M. & Granath, J. W.), pp. 729. Geological Society of London, Special Publication no. 312.Google Scholar
O'Connell, R. J., Gable, C. W. & Hager, B. H. 1991. Toroidal-poloidal partitioning of lithospheric plate motions. In Glacial Isostasy, Sea-Level and Mantle Rheology (eds Sabadini, R., Lambeck, K. & Boschi, E.), pp. 535–51. Berlin: Springer.Google Scholar
O'Neill, C., Müller, D. & Steinberger, B. 2005. On the uncertainties in hot spot reconstructions and the significance of moving hot spot reference frames. Geochemistry, Geophysics, Geosystems 6 (4), Q04003. doi: 10.1029/2004GC000784.Google Scholar
Pearce, J. A., Lippard, S. & Roberts, S. 1984. Characteristics and tectonic significance of supra-subduction zone ophiolites. In Margined Basin Geology (eds Kokelaar, B. P. & Howells, M. F.), pp. 7794. Geological Society of London, Special Publication no. 16.Google Scholar
Peron-Pinvidic, G., Manatschal, G. & Osmundsen, P. T. 2013. Structural comparison of archetypal Atlantic rifted margins: a review of observations and concepts. Marine and Petroleum Geology 43, 2147.Google Scholar
Pisarevsky, S. A., Wingate, M. T., Powell, C. M., Johnson, S. & Evans, D. A. 2003. Models of Rodinia assembly and fragmentation. In Supercontinent Assembly and Breakup (eds Yoshida, M., Windley, B. F. and Dasgupta, S.), pp. 3555. Geological Society of London, Special Publication no. 206.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: Solid Earth 96 (B5), 8407–15.Google Scholar
Ringwood, A. E. 1974. The petrological evolution of island arc systems. Twenty-seventh William Smith Lecture. Journal of the Geological Society 130 (3), 183204.Google Scholar
Royden, L. H. & Husson, L. 2006. Trench motion, slab geometry and viscous stresses in subduction systems. Geophysical Journal International 167 (2), 881905.Google Scholar
Royer, D. 2014. Atmospheric CO2 and O2 during the Phanerozoic: tools, patterns, and impacts. Treatise on Geochemistry 6, 251–67.Google Scholar
Sano, Y. & Williams, S. N. 1996. Fluxes of mantle and subducted carbon along convergent plate boundaries. Geophysical Research Letters 23 (20), 2749–52.Google Scholar
Schellart, W., Freeman, J., Stegman, D., Moresi, L. & May, D. 2007. Evolution and diversity of subduction zones controlled by slab width. Nature 446 (7133), 308–11.Google Scholar
Schellart, W., Stegman, D. & Freeman, J. 2008. Global trench migration velocities and slab migration induced upper mantle volume fluxes: constraints to find an Earth reference frame based on minimizing viscous dissipation. Earth-Science Reviews 88 (1), 118–44.Google Scholar
Schult, F. R. & Gordon, R. G. 1984. Root mean square velocities of the continents with respect to the hot spots since the Early Jurassic. Journal of Geophysical Research: Solid Earth 89 (B3), 1789–800.Google Scholar
Scotese, C. R., Bambach, R. K., Barton, C., Van der Voo, R. & Ziegler, A. M. 1979. Paleozoic base maps. The Journal of Geology 87 (3), 217–77.Google Scholar
Scotese, R. C. & Barrett, S. F. 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. S. & Scotese, C. R.), pp. 7585. Geological Society of London, Memoir no. 12.Google Scholar
Scotese, C. R. & McKerrow, W. S. 1990. Revised world maps and introduction. In Palaeozoic Palaeogeography and Biogeography (eds McKerrow, W. S. & Scotese, C. R.), pp. 121. Geological Society of London, Memoir no. 12.Google Scholar
Scrutton, R. A. 1982. Crustal structure and development of sheared passive continental margins. In Dynamics of Passive Margins (ed. Scrutton, R. A.), pp. 133–46. Boulder, CO: American Geophysical Union.Google Scholar
Şengör, A. C. 1984. The Cimmeride orogenic system and the tectonics of Eurasia. Geological Society of America Special Paper 195, 74 pp.Google Scholar
Seton, M., Müller, R., Zahirovic, S., Gaina, C., Torsvik, T., Shephard, G., Talsma, A., Gurnis, M., Turner, M. & Maus, S. 2012. Global continental and ocean basin reconstructions since 200 Ma. Earth-Science Reviews 113 (3), 212–70.Google Scholar
Simpson, C. & De Paor, D. G. 1993. Strain and kinematic analysis in general shear zones. Journal of Structural Geology 15 (1), 120.Google Scholar
Sobel, D. 2007. Longitude: The True Story of a Lone Genius who Solved the Greatest Scientific Problem of his Time. New York: Bloomsbury Publishing USA.Google Scholar
Stampfli, G. M. 2000. Tethyan oceans. In Tectonics and Magmatism in Turkey and the Surrounding Area (eds Bozkurt, E., Winchester, J. A. & Piper, J. D. A.), pp. 163–85. Geological Society of London, Special Publication no. 173.Google Scholar
Stampfli, G. M. & 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 (1), 1733.Google Scholar
Stegman, D., Farrington, R., Capitanio, F. & Schellart, W. 2010. A regime diagram for subduction styles from 3-D numerical models of free subduction. Tectonophysics 483 (1), 2945.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 (6996), 167–73.Google Scholar
Steinberger, B. & Torsvik, T. H. 2008. Absolute plate motions and true polar wander in the absence of hotspot tracks. Nature 452 (7187), 620–3.Google Scholar
Steinberger, B. & Torsvik, T. 2010. Toward an explanation for the present and past locations of the poles. Geochemistry, Geophysics, Geosystems 11 (6), Q06W06. doi: 10.1929/2009GC002889.Google Scholar
Steinberger, B., Torsvik, T. H. & Becker, T. 2012. Subduction to the lower mantle-a comparison between geodynamic and tomographic models. Solid Earth 3 (2), 415–32.Google Scholar
Stern, R. J. 2002. Subduction zones. Reviews of Geophysics 40 (4), 1012. doi: 10.1029/2001RG000108.Google Scholar
Stern, R. J. 2005. Evidence from ophiolites, blueschists, and ultrahigh-pressure metamorphic terranes that the modern episode of subduction tectonics began in Neoproterozoic time. Geology 33 (7), 557–60.Google Scholar
Swanson-Hysell, N. L., Maloof, A. C., Weiss, B. P. & Evans, D. A. 2009. No asymmetry in geomagnetic reversals recorded by 1.1-billion-year-old Keweenawan basalts. Nature Geoscience 2 (10), 713–17.Google Scholar
Tauxe, L. 2010. Essentials of paleomagnetism. Berkeley, CA: University of California Press.Google Scholar
Teyssier, C., Tikoff, B. & Markley, M. 1995. Oblique plate motion and continental tectonics. Geology 23 (5), 447–50.Google Scholar
Thomas, W. A. 2011. Detrital-zircon geochronology and sedimentary provenance. Lithosphere 3 (4), 304–8.Google Scholar
Torsvik, T. H., Burke, K., Steinberger, B., Webb, S. J. & Ashwal, L. D. 2010b. Diamonds sampled by plumes from the core-mantle boundary. Nature 466 (7304), 352–5.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 161 (4), 555–72.Google Scholar
Torsvik, T. H. & Cocks, L. R. M. 2017. Earth History and Palaeogeography. Cambridge: Cambridge University Press.Google Scholar
Torsvik, T. H. & Domeier, M. 2017. Correspondence: The PERM anomaly, the Emeishan Large Igneous Province and numerical modelling in the Earth Sciences. Nature Communications 8. doi: 10.1038/s41467-017-00125-2.Google Scholar
Torsvik, T. H., Müller, R. D., Van der Voo, R., Steinberger, B. & Gaina, C. 2008. Global plate motion frames: toward a unified model. Reviews of Geophysics 46 (3), RG3004. doi: 10.1029/2007RG000227.Google Scholar
Torsvik, T. H., Steinberger, B., Gurnis, M. & Gaina, C. 2010a. Plate tectonics and net lithosphere rotation over the past 150 My. Earth and Planetary Science Letters 291 (1), 106–12.Google Scholar
Torsvik, T. H., Van der Voo, R., Doubrovine, P. V., Burke, K., Steinberger, B., Ashwal, L. D., Trønnes, R. G., Webb, S. J. & Bull, A. L. 2014. Deep mantle structure as a reference frame for movements in and on the Earth. Proceedings of the National Academy of Sciences 111 (24), 8735–40.Google Scholar
Torsvik, T. H., Van der Voo, R., Preeden, U., Mac Niocaill, C., Steinberger, B., Doubrovine, P. V., van Hinsbergen, D. J., Domeier, M., Gaina, C. & Tohver, E. 2012. Phanerozoic polar wander, palaeogeography and dynamics. Earth-Science Reviews 114 (3), 325–68.Google Scholar
Tsai, V. C. & Stevenson, D. J. 2007. Theoretical constraints on true polar wander. Journal of Geophysical Research: Solid Earth 112 (B5), B05415. doi: 10.1029/2005JB003923.Google Scholar
Van Der Meer, D. G., Spakman, W., Van Hinsbergen, D. J., Amaru, M. L. & Torsvik, T. H. 2010. Towards absolute plate motions constrained by lower-mantle slab remnants. Nature Geoscience 3 (1), 3640.Google Scholar
Van Der Meer, D. G., Zeebe, R. E., van Hinsbergen, D. J., Sluijs, A., Spakman, W. & Torsvik, T. H. 2014. Plate tectonic controls on atmospheric CO2 levels since the Triassic. Proceedings of the National Academy of Sciences 111 (12), 4380–5.Google Scholar
Vérard, C., Hochard, C., Baumgartner, P. O., Stampfli, G. M. & Liu, M. 2015. Geodynamic evolution of the Earth over the Phanerozoic: plate tectonic activity and palaeoclimatic indicators. Journal of Palaeogeography 4 (2), 167–88.Google Scholar
Vine, F. J. & Matthews, D. H. 1963. Magnetic anomalies over oceanic ridges. Nature 199 (4897), 947–9.Google Scholar
Wegener, A. 1912. Die Entstehung der Kontinente. Geologische Rundschau 3 (4), 276–92.Google Scholar
Whattam, S. A. & Stern, R. J. 2011. The ‘subduction initiation rule’: a key for linking ophiolites, intra-oceanic forearcs, and subduction initiation. Contributions to Mineralogy and Petrology 162 (5), 1031– 45.Google Scholar
Williams, S., Flament, N., Müller, R. D. & Butterworth, N. 2015. Absolute plate motions since 130 Ma constrained by subduction zone kinematics. Earth and Planetary Science Letters 418, 6677.Google Scholar
Wilson, J. T. 1965. A new class of faults and their bearing on continental drift. Nature 207 (4995), 343–7.Google Scholar
Wilson, J. T. 1966. Did the Atlantic close and then re-open? Nature 211 (5050), 676–81.Google Scholar
Zahirovic, S., Müller, R. D., Seton, M. & Flament, N. 2015. Tectonic speed limits from plate kinematic reconstructions. Earth and Planetary Science Letters 418, 4052.Google Scholar
Zhong, S. & Rudolph, M. L. 2015. On the temporal evolution of long-wavelength mantle structure of the Earth since the early Paleozoic. Geochemistry, Geophysics, Geosystems 16 (5), 1599–615.Google Scholar
Zhong, S., Zhang, N., Li, Z.-X. & Roberts, J. H. 2007. Supercontinent cycles, true polar wander, and very long-wavelength mantle convection. Earth and Planetary Science Letters 261 (3), 551–64.Google Scholar