Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Dynamics of the abrupt change in Pacific Plate motion around 50 million years ago

Abstract

A drastic change in plate tectonics and mantle convection occurred around 50 Ma as exemplified by the prominent Hawaiian–Emperor Bend. Both an abrupt Pacific Plate motion change and a change in mantle plume dynamics have been proposed to account for the Hawaiian–Emperor Bend, but debates surround the relative contribution of the two mechanisms. Here we build kinematic plate reconstructions and high-resolution global dynamic models to quantify the amount of Pacific Plate motion change. We find Izanagi Plate subduction, followed by demise of the Izanagi–Pacific Ridge and Izu–Bonin–Mariana subduction initiation alone, is incapable of causing a sudden change in plate motion, challenging the conventional hypothesis on the mechanisms of Pacific Plate motion change. Instead, Palaeocene slab pull from Kronotsky intraoceanic subduction in the northern Pacific exerts a northward pull on the Pacific Plate, while its Eocene demise leads to a sudden 30–35° change in plate motion, accounting for about half of the Hawaiian–Emperor Bend. We suggest the Pacific Plate motion change and hotspot drift due to plume dynamics could have contributed nearly equally to the formation of the Hawaiian–Emperor Bend. Such a scenario is consistent with available constraints from global plate circuits, palaeomagnetic data and geodynamic models.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Alternative plate reconstruction models from Late Cretaceous to Eocene.
Fig. 2: Geodynamic models used to predict plate motions.
Fig. 3: Change of the azimuth of the Pacific Plate motion with time.
Fig. 4: Computed tracks of the Pacific Plate motion and the residual hotspot drift.

Similar content being viewed by others

Data availability

The raw data for the paper have been deposited on Caltech Data (https://doi.org/10.22002/D1.2150), including the digitized alternative plate reconstruction and the predicted plate motion. The authors declare that all other data supporting the findings of this study are available within the paper and its Supplementary Information files with their sources annotated in the text.

Code availability

The plate kinematic tool GPlates and its python version can be accessed at www.gplates.org/. The original version of CitcomS is available at www.geodynamics.org/cig/software/citcoms/. The adaptive nonlinear Stokes solver (Rhea) used to predict plate motion is available upon request.

References

  1. Morgan, W. J. Deep mantle convection plumes and plate motions. Bull. Am. Assoc. Pet. Geol. 56, 203–213 (1972).

    Google Scholar 

  2. Torsvik, T. H. et al. Pacific plate motion change caused the Hawaiian–Emperor Bend. Nat. Commun. 8, 15660 (2017).

    Article  Google Scholar 

  3. Tarduno, J. A. et al. The Emperor seamounts: southward motion of the Hawaiian hotspot plume in Earth’s mantle. Science 301, 1064–1069 (2003).

    Article  Google Scholar 

  4. Steinberger, B., Sutherland, R. & O’Connell, R. J. Prediction of Emperor–Hawaii seamount locations from a revised model of global plate motion and mantle flow. Nature 430, 167–173 (2004).

    Article  Google Scholar 

  5. Hassan, R., Müller, R. D., Gurnis, M., Williams, S. E. & Flament, N. A rapid burst in hotspot motion through the interaction of tectonics and deep mantle flow. Nature 533, 239–242 (2016).

    Article  Google Scholar 

  6. Wessel, P. & Kroenke, L. W. Pacific absolute plate motion since 145 Ma: an assessment of the fixed hot spot hypothesis. J. Geophys. Res. Solid Earth 113, B6 (2008).

    Article  Google Scholar 

  7. Torsvik, T. H. et al. Pacific–Panthalassic reconstructions: overview, errata and the way forward. Geochem. Geophys. Geosyst. 20, 3659–3689 (2019).

    Article  Google Scholar 

  8. Bono, R. K., Tarduno, J. A. & Bunge, H.-P. Hotspot motion caused the Hawaiian–Emperor Bend and LLSVPs are not fixed. Nat. Commun. 10, 3370 (2019).

    Article  Google Scholar 

  9. Müller, R. D. et al. Global-scale plate reorganization events since Pangea breakup. Annu. Rev. Earth Planet. Sci. 44, 107–138 (2016).

    Article  Google Scholar 

  10. Müller, R. D. et al. A global plate model including lithospheric deformation along major rifts and orogens since the Triassic. Tectonics 38, 1884–1907 (2019).

    Article  Google Scholar 

  11. Reagan, M. et al. Forearc ages reveal extensive short-lived and rapid seafloor spreading following subduction initiation. Earth Planet. Sci. Lett. 506, 520–529 (2019).

    Article  Google Scholar 

  12. Sutherland, R. et al. Continental scale of geographic change across Zealandia during subduction zone initiation. Geology 48, 419–424 (2020).

    Article  Google Scholar 

  13. Whittaker, J. et al. Major Australian–Antarctic plate reorganization at Hawaiian–Emperor Bend time. Science 318, 83–86 (2007).

    Article  Google Scholar 

  14. Patriat, P. & Achache, J. India–Eurasia collision chronology has implications for crustal shortening and driving mechanism of plates. Nature 311, 615–621 (1984).

    Article  Google Scholar 

  15. Richards, M. A. & Lithgow-Bertelloni, C. Plate motion changes, the Hawaiian–Emperor Bend, and the apparent success and failure of geodynamic models. Earth Planet. Sci. Lett. 137, 19–27 (1996).

    Article  Google Scholar 

  16. Müller, R. D., Sdrolias, M., Gaina, C., Steinberger, B. & Heine, C. Long-term sea-level fluctuations driven by ocean basin dynamics. Science 319, 1357–1362 (2008).

    Article  Google Scholar 

  17. Faccenna, C., Becker, T. W., Lallemand, S. & Steinberger, B. On the role of slab pull in the Cenozoic motion of the Pacific Plate. Geophys. Res. Lett. 39, L03305 (2012).

    Article  Google Scholar 

  18. Conrad, C. P. & Lithgow-Bertelloni, C. How mantle slabs drive plate tectonics. Science 298, 207–209 (2002).

    Article  Google Scholar 

  19. Buffett, B. A. Plate force due to bending at subduction zones. J. Geophys. Res. Solid Earth 111, B9 (2006).

    Article  Google Scholar 

  20. Forsyth, D. & Uyeda, S. On the relative importance of the driving forces of plate motion. Geophys. J. Int. 43, 163–200 (1975).

    Article  Google Scholar 

  21. Billen, M. I. Modeling the dynamics of subducting slabs. Annu. Rev. Earth Planet. Sci. 36, 325–356 (2008).

    Article  Google Scholar 

  22. Stadler, G. et al. The dynamics of plate tectonics and mantle flow: from local to global scales. Science 329, 1033–1038 (2010).

    Article  Google Scholar 

  23. Taylor, B. The single largest oceanic plateau: Ontong Java–Manihiki–Hikurangi. Earth Planet. Sci. Lett. 241, 372–380 (2006).

    Article  Google Scholar 

  24. Levashova, N. M., Shapiro, M. N., Beniamovsky, V. N. & Bazhenov, M. L. Paleomagnetism and geochronology of the Late Cretaceous–Paleogene island arc complex of the Kronotsky Peninsula, Kamchatka, Russia: kinematic implications. Tectonics 19, 834–851 (2000).

    Article  Google Scholar 

  25. Konstantinovskaya, E. in Arc–Continent Collision (eds Brown D. & Ryan, P. D.) 247–277 (Springer, 2011).

  26. Rudi, J. et al. An extreme-scale implicit solver for complex PDEs: highly heterogeneous flow in Earth’s mantle. In Proc. International Conference for High Performance Computing, Networking, Storage, and Analysis (Association for Computing Machinery, 2015).

  27. Bakhteev, M., Morozov, O. & Tikhomirova, S. Structure of the eastern Kamchatka ophiolite-free collisional suture–Grechishkin thrust. Geotectonics 31, 236–246 (1997).

    Google Scholar 

  28. Domeier, M. et al. Intraoceanic subduction spanned the Pacific in the Late Cretaceous–Paleocene. Sci. Adv. 3, eaao2303 (2017).

    Article  Google Scholar 

  29. Vaes, B., Van Hinsbergen, D. J. & Boschman, L. M. Reconstruction of subduction and back-arc spreading in the NW Pacific and Aleutian basin: clues to causes of Cretaceous and Eocene plate reorganizations. Tectonics 38, 1367–1413 (2019).

    Article  Google Scholar 

  30. Seton, M. et al. Melanesian back-arc basin and arc development: constraints from the eastern Coral Sea. Gondwana Res. 39, 77–95 (2016).

    Article  Google Scholar 

  31. Leng, W. & Gurnis, M. Subduction initiation at relic arcs. Geophys. Res. Lett. 42, 7014–7021 (2015).

    Article  Google Scholar 

  32. Faccenna, C., Holt, A. F., Becker, T. W., Lallemand, S. & Royden, L. H. Dynamics of the Ryukyu/Izu–Bonin–Marianas double subduction system. Tectonophysics 746, 229–238 (2018).

    Article  Google Scholar 

  33. Tarduno, J., Bunge, H.-P., Sleep, N. & Hansen, U. The bent Hawaiian–Emperor hotspot track: inheriting the mantle wind. Science 324, 50–53 (2009).

    Article  Google Scholar 

  34. Matthews, K. J., Seton, M. & Müller, R. D. A global-scale plate reorganization event at 105–100 Ma. Earth Planet. Sci. Lett. 355, 283–298 (2012).

    Article  Google Scholar 

  35. Steinberger, B. & Gaina, C. Plate-tectonic reconstructions predict part of the Hawaiian hotspot track to be preserved in the Bering Sea. Geology 35, 407–410 (2007).

    Article  Google Scholar 

  36. Seton, M. et al. Global continental and ocean basin reconstructions since 200 Ma. Earth Sci. Rev. 113, 212–270 (2012).

    Article  Google Scholar 

  37. Matthews, K. J. et al. Geologic and kinematic constraints on Late Cretaceous to mid Eocene plate boundaries in the southwest Pacific. Earth Sci. Rev. 140, 72–107 (2015).

    Article  Google Scholar 

  38. Whittaker, J. M., Williams, S. E. & Müller, R. D. Revised tectonic evolution of the eastern Indian Ocean. Geochem. Geophys. Geosyst. 14, 1891–1909 (2013).

    Article  Google Scholar 

  39. Granot, R., Cande, S., Stock, J. & Damaske, D. Revised Eocene–Oligocene kinematics for the West Antarctic rift system. Geophys. Res. Lett. 40, 279–284 (2013).

    Article  Google Scholar 

  40. Doubrovine, P. V. & Tarduno, J. A. A revised kinematic model for the relative motion between Pacific oceanic plates and North America since the Late Cretaceous. J. Geophys. Res. Solid Earth 113, 2 (2008).

    Article  Google Scholar 

  41. Mortimer, N. Evidence for a pre-Eocene proto-Alpine Fault through Zealandia. N. Z. J. Geol. Geophys. 61, 251–259 (2018).

    Article  Google Scholar 

  42. Sutherland, R. et al. Widespread compressional faulting associated with forced Tonga–Kermadec subduction initiation. Geology 45, 355–358 (2017).

    Article  Google Scholar 

  43. Karlsen, K. S., Domeier, M., Gaina, C. & Conrad, C. P. A tracer-based algorithm for automatic generation of seafloor age grids from plate tectonic reconstructions. Comput. Geosci. 140, 104508 (2020).

  44. Hu, J. & Gurnis, M. Subduction duration and slab dip. Geochem. Geophys. Geosyst. 21, e2019GC008862 (2020).

    Article  Google Scholar 

  45. Coney, P. J. & Reynolds, S. J. Cordilleran Benioff zones. Nature 270, 403–406 (1977).

    Article  Google Scholar 

  46. Bower, D. J., Gurnis, M. & Flament, N. Assimilating lithosphere and slab history in 4-D Earth models. Phys. Earth Planet. Inter. 238, 8–22 (2015).

    Article  Google Scholar 

  47. Zhong, S., Zuber, M. T., Moresi, L. N. & Gurnis, M. The role of temperature-dependent viscosity and surface plates in spherical shell models of mantle convection. J. Geophys. Res. 105, 11063–11082 (2000).

    Article  Google Scholar 

  48. Hu, J., Liu, L. & Zhou, Q. Reproducing past subduction and mantle flow using high-resolution global convection models. Earth Planet. Phys. 2, 189–207 (2018).

    Article  Google Scholar 

  49. Campbell, I. & Griffiths, R. Implications of mantle plume structure for the evolution of flood basalts. Earth Planet. Sci. Lett. 99, 79–93 (1990).

    Article  Google Scholar 

  50. Mao, X., Gurnis, M. & May, D. A. Subduction initiation with vertical lithospheric heterogeneities and new fault formation. Geophys. Res. Lett. https://doi.org/10.1002/2017GL075389 (2017).

  51. Zhong, X. & Li, Z. Forced subduction initiation at passive continental margins: velocity-driven versus stress-driven. Geophys. Res. Lett. 46, 11054–11064 (2019).

    Article  Google Scholar 

  52. Burstedde, C., Wilcox, L. C. & Ghattas, O. p4est: scalable algorithms for parallel adaptive mesh refinement on forests of octrees. SIAM J. Sci. Comput. 33, 1103–1133 (2011).

    Article  Google Scholar 

  53. Sundar, H. et al. Parallel geometric–algebraic multigrid on unstructured forests of octrees in SC12. In Proc. International Conference for High Performance Computing, Networking, Storage and Analysis (Association for Computing Machinery, 2012).

  54. Rudi, J., Stadler, G. & Ghattas, O. Weighted BFBT preconditioner for Stokes flow problems with highly heterogeneous viscosity. SIAM J. Sci. Comput. 39, S272–S297 (2017).

    Article  Google Scholar 

  55. Rudi, J., Shih, Y.-h. & Stadler, G. Advanced Newton methods for geodynamical models of Stokes flow with viscoplastic rheologies. Geochem. Geophys. Geosyst. https://doi.org/10.1029/2020GC009059 (2020).

Download references

Acknowledgements

J.H. and M.G. were partially supported by the US National Science Foundation (NSF) through awards EAR-1645775 and EAR-2009935. J.R. was supported by the US Department of Energy, Office of Science, under contract DE-AC02-06CH11357. G.S. was partially supported by NSF grants EAR-1646337 and DMS-1723211. Computations were carried out on the NSF-supported Stampede-2 and Frontera supercomputers at the Texas Advanced Computer Center under allocations TG-EAR160027, TG-DPP130002 and FTA-SUB-CalTech. The funders had no role in study design, data collection and analysis, decision to publish or preparation of the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

J.H. and M.G. designed the study. J.H. carried out the numerical experiments. J.R. and G.S. expanded the functionality of the adaptive nonlinear Stokes solver Rhea and provided expertise in scientific computing. R.D.M. helped with plate reconstruction. All authors participated in result interpretation and manuscript preparation.

Corresponding author

Correspondence to Jiashun Hu.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Geoscience thanks Claudio Faccenna and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Primary Handling Editors: Stefan Lachowycz and Rebecca Neely in collaboration with the Nature Geoscience team.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 Representative data showing construction of a paleo slab with pyGplates.

Representative data showing construction of a paleo slab with pyGplates. a, The thermal age of the lithosphere within the slab. b, Depth of the slab. c, A cross section of the slab with plates labeled as NAM for North America, FAR for Farallon, VAN for Vancouver and PAC for Pacific. The location is labeled in b.

Extended Data Fig. 2 Representative global thermal structure.

Representative global thermal structure. This example shows the input for the geodynamic model based on the traditional plate reconstruction9 at 47 Ma.

Extended Data Fig. 3 Predicted plate motion with the traditional plate reconstruction that incorporates the Izanagi-Pacific Ridge subduction.

Predicted plate motion at 60 (a), 50 (b) and 47 (c) Ma, based on the traditional plate reconstruction that incorporates the Izanagi-Pacific Ridge subduction.

Extended Data Fig. 4 Model with plume-head push.

Model with plume-head push. a, b and c show the thermal structures of the forward CitcomS model with plume-head. a and b show a cross section (see the location of the cross section in the inserted panel) and a map view (at 200 km) of the initial plume head at 120 Ma. c represents the thermal structure of the forward CitcomS model at 50 Ma. d represents the thermal structure of the CitcomS model at 50 Ma without imposing the plume head at 120 Ma. The difference between c and d represents the contribution from the plume head, which is highlighted with the red circle in c. The difference is superimposed on the thermal structure of the forward model at 50 Ma based on the traditional model, to test the effect of plume-head push, which is shown in e. In e, the green, orange, black and red arrows represent the geodynamically predicted velocities without plume-head push, the geodynamically predicted velocities with plume-head push, the reconstructed velocities from ref. 9 and the reconstructed velocities from ref. 10, respectively.

Extended Data Fig. 5 Comparison of the models with and without asthenosphere.

Comparison of the models with (b) and without (c) asthenosphere. In a, the green, orange, black and red arrows represent the geodynamically predicted velocities without asthenosphere, the geodynamically predicted velocities with asthenosphere, the reconstructed velocities from ref. 9 and the reconstructed velocities from ref. 10, respectively.

Extended Data Fig. 6 Testing the effect of the nascent IBM slab at 50 Ma.

Testing the effect of the nascent IBM slab at 50 Ma. a shows the predicted plate motion for the models with the nascent IBM slab (green arrows) and without the nascent IBM slab (orange arrows). b, c and d show the viscosity along the cross sections labeled in a for the model without the IBM slab. e, f and g show the viscosity along the cross sections labeled in a for the model with the IBM slab.

Extended Data Fig. 7 Testing the effect of nascent IBM slab at 47 Ma.

Same as Extended Data Fig. 6, but testing the effect of nascent IBM slab at 47 Ma.

Extended Data Fig. 8 Computations of the Hawaiian-Emperor Seamount hotspot tracks and plume motion.

Computations of the Hawaiian-Emperor Seamount hotspot tracks and plume motion. We combine the Pacific Plate motion change and the hotspot drift path to reconstruct the Hawaiian-Emperor Seamount chain. For the Pacific Plate motion, we use the preferred geodynamic Kronotsky-SUB track that combines the Pacific Plate motion from Müller et al.9 after 47 Ma and from the geodynamic models that incorporates the Kronotsky subduction before 47 Ma. The residual hotspot drift path in Fig. 4 is used in a, while the hotspot drift in Hassan et al.5 is used in b. To first order, the two hotspot drift paths are similar.

Extended Data Fig. 9 Geodynamically predicted horizontal velocity in the vicinity of Hawaiian plume as a function of depth for different models.

Geodynamically predicted horizontal velocity in the vicinity of Hawaiian plume as a function of depth for different models. For each depth, the velocities are averaged over the region from 190°E to 220°E and from 10°N to 35°N. The blue arrows represent the averaged velocities at 60 Ma, the red arrows at 50 Ma and the black arrows at 47 Ma.

Supplementary information

Supplementary Information

Supplementary Tables 1–3.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hu, J., Gurnis, M., Rudi, J. et al. Dynamics of the abrupt change in Pacific Plate motion around 50 million years ago. Nat. Geosci. 15, 74–78 (2022). https://doi.org/10.1038/s41561-021-00862-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41561-021-00862-6

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing