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:

Continental configuration controls ocean oxygenation during the Phanerozoic

Abstract

The early evolutionary and much of the extinction history of marine animals is thought to be driven by changes in dissolved oxygen concentrations ([O2]) in the ocean1,2,3. In turn, [O2] is widely assumed to be dominated by the geological history of atmospheric oxygen (pO2)4,5. Here, by contrast, we show by means of a series of Earth system model experiments how continental rearrangement during the Phanerozoic Eon drives profound variations in ocean oxygenation and induces a fundamental decoupling in time between upper-ocean and benthic [O2]. We further identify the presence of state transitions in the global ocean circulation, which lead to extensive deep-ocean anoxia developing in the early Phanerozoic even under modern pO2. Our finding that ocean oxygenation oscillates over stable thousand-year (kyr) periods also provides a causal mechanism that might explain elevated rates of metazoan radiation and extinction during the early Palaeozoic Era6. The absence, in our modelling, of any simple correlation between global climate and ocean ventilation, and the occurrence of profound variations in ocean oxygenation independent of atmospheric pO2, presents a challenge to the interpretation of marine redox proxies, but also points to a hitherto unrecognized role for continental configuration in the evolution of the biosphere.

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: Earth system model results for simulations at 2,240 ppm CO2 (series #1).
Fig. 2: Benthic oxygen concentrations for simulations at 2,240 ppm CO2 (series #1).
Fig. 3: Earth system model results for simulations in which we varied pCO2 to approximately ‘correct’ for the palaeogeographical impacts on climate (series #2).

Similar content being viewed by others

Code availability

The version of the cGENIE code used in this paper is tagged as release v0.9.31 and is available at https://doi.org/10.5281/zenodo.6823664. Necessary boundary condition files are included as part of the code release. Configuration files for the specific experiments presented in the paper can be found in the installation subdirectory: genie-userconfigs/PUBS/published/Pohl_et_al.2022. Details of the experiments, plus the command line needed to run each one, are given in the readme.txt file in that directory. A manual describing code installation, basic model configuration and an extensive series of tutorials is provided (https://doi.org/10.5281/zenodo.5500696). The FOAM output is hosted on Zenodo (https://doi.org/10.5281/zenodo.5780096).

References

  1. Payne, J. L. et al. Two-phase increase in the maximum size of life over 3.5 billion years reflects biological innovation and environmental opportunity. Proc. Natl Acad. Sci. USA 106, 24–27 (2009).

    Article  ADS  CAS  PubMed  Google Scholar 

  2. Cole, D. B. et al. On the co-evolution of surface oxygen levels and animals. Geobiology 18, 260–281 (2020).

    Article  PubMed  Google Scholar 

  3. Sperling, E. A., Knoll, A. H. & Girguis, P. R. The ecological physiology of Earth’s second oxygen revolution. Annu. Rev. Ecol. Evol. Syst. 46, 215–235 (2015).

    Article  Google Scholar 

  4. Krause, A. J. et al. Stepwise oxygenation of the Paleozoic atmosphere. Nat. Commun. 9, 4081 (2018).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  5. Tostevin, R. & Mills, B. J. Reconciling proxy records and models of Earth’s oxygenation during the Neoproterozoic and Palaeozoic. Interface Focus 10, 20190137 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  6. Kocsis, Á. T., Reddin, C. J., Alroy, J. & Kiessling, W. The R package divDyn for quantifying diversity dynamics using fossil sampling data. Methods Ecol. Evol. 10, 735–743 (2019).

    Article  Google Scholar 

  7. Penn, J. L., Deutsch, C., Payne, J. L. & Sperling, E. A. Temperature-dependent hypoxia explains biogeography and severity of end-Permian marine mass extinction. Science 362, eaat1327 (2018).

    Article  ADS  PubMed  Google Scholar 

  8. Edwards, C. T., Saltzman, M. R., Royer, D. L. & Fike, D. A. Oxygenation as a driver of the Great Ordovician Biodiversification Event. Nat. Geosci. 10, 925–929 (2017).

    Article  ADS  CAS  Google Scholar 

  9. Dahl, T. W., Hammarlund, E. U. & Anbar, A. D. Devonian rise in atmospheric oxygen correlated to the radiations of terrestrial plants and large predatory fish. Proc. Natl Acad. Sci. 107, 17911–17915 (2010).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  10. Zou, C. et al. Ocean euxinia and climate change “double whammy” drove the Late Ordovician mass extinction. Geology 46, 535–538 (2018).

    Article  ADS  CAS  Google Scholar 

  11. Bond, D., Wignall, P. B. & Racki, G. Extent and duration of marine anoxia during the Frasnian–Famennian (Late Devonian) mass extinction in Poland, Germany, Austria and France. Geol. Mag. 141, 173–193 (2004).

    Article  ADS  Google Scholar 

  12. Lau, K. V. et al. Marine anoxia and delayed Earth system recovery after the end-Permian extinction. Proc. Natl Acad. Sci. 113, 2360–2365 (2016).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  13. Lu, W. et al. Late inception of a resiliently oxygenated upper ocean. Science 5, eaar5372 (2018).

    Article  Google Scholar 

  14. Sperling, E. A. et al. A long-term record of early to mid-Paleozoic marine redox change. Sci. Adv. 7, eabf4382 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  15. Lenton, T. M. Earliest land plants created modern levels of atmospheric oxygen. Proc. Natl Acad. Sci. 113, 9704–9709 (2016).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  16. Valdes, P., Scotese, C. & Lunt, D. Deep ocean temperatures through time. Clim. Past 17, 1483–1506 (2021).

    Article  Google Scholar 

  17. Farnsworth, A. et al. Climate sensitivity on geological timescales controlled by nonlinear feedbacks and ocean circulation. Geophys. Res. Lett. 46, 9880–9889 (2019).

    Article  ADS  Google Scholar 

  18. Scotese, C. R., Song, H., Mills, B. J. & van der Meer, D. G. Phanerozoic paleotemperatures: the earth’s changing climate during the last 540 million years. Earth Sci. Rev. 215, 103503 (2021).

    Article  CAS  Google Scholar 

  19. Monteiro, F. M., Pancost, R. D., Ridgwell, A. & Donnadieu, Y. Nutrients as the dominant control on the spread of anoxia and euxinia across the Cenomanian-Turonian oceanic anoxic event (OAE2): model-data comparison. Paleoceanography 27, PA4209 (2012).

    Article  ADS  Google Scholar 

  20. Ridgwell, A. et al. Marine geochemical data assimilation in an efficient Earth system model of global biogeochemical cycling. Biogeosciences 4, 87–104 (2007).

    Article  ADS  CAS  Google Scholar 

  21. Pohl, A. et al. Vertical decoupling in Late Ordovician anoxia due to reorganization of ocean circulation. Nat. Geosci. 14, 868–873 (2021).

    Article  ADS  CAS  Google Scholar 

  22. Ward, B. A. et al. EcoGEnIE 1.0: plankton ecology in the cGENIE Earth system model. Geosci. Model Dev. 11, 4241–5267 (2018).

    Article  ADS  CAS  Google Scholar 

  23. Crichton, K. A., Wilson, J. D., Ridgwell, A. & Pearson, P. N. Calibration of temperature-dependent ocean microbial processes in the cGENIE.muffin (v0.9.13) Earth system model. Geosci. Model Dev. 14, 125–149 (2021).

    Article  ADS  Google Scholar 

  24. Scotese, C. R. & Wright, N. PALEOMAP Paleodigital Elevation Models (PaleoDEMS) for the Phanerozoic. PALEOMAP Project. https://www.earthbyte.org/paleodem-resource-scotese-and-wright-2018/ (2018).

  25. Pohl, A. et al. Quantifying the paleogeographic driver of Cretaceous carbonate platform development using paleoecological niche modeling. Palaeogeogr. Palaeoclimatol. Palaeoecol. 514, 222–232 (2019).

    Article  Google Scholar 

  26. Hülse, D. et al. End-Permian marine extinction due to temperature-driven nutrient recycling and euxinia. Nat. Geosci. 14, 862–867 (2021).

    Article  ADS  Google Scholar 

  27. Baudin, F. & Riquier, L. The late Hauterivian Faraoni ‘oceanic anoxic event’: an update. Bull. Soc. Géol. Fr. 185, 359–377 (2014).

    Article  Google Scholar 

  28. Laugié, M. et al. Exploring the impact of Cenomanian paleogeography and marine gateways on oceanic oxygen. Paleoceanogr. Paleoclimatol. 36, e2020PA004202 (2021).

    Article  Google Scholar 

  29. De Vleeschouwer, D. et al. Timing and pacing of the Late Devonian mass extinction event regulated by eccentricity and obliquity. Nat. Commun. 8, 2268 (2017).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  30. Ruvalcaba Baroni, I. et al. Ocean circulation in the Toarcian (Early Jurassic): a key control on deoxygenation and carbon burial on the European Shelf. Paleoceanogr. Paleoclimatol. 33, 994–1012 (2018).

    Article  ADS  Google Scholar 

  31. Torsvik, T. H. BugPlates: Linking Biogeography and Palaeogeography (2009).

  32. Ferreira, D., Marshall, J., Ito, T. & McGee, D. Linking glacial-interglacial states to multiple equilibria of climate. Geophys. Res. Lett. 45, 9160–9170 (2018).

    Article  ADS  Google Scholar 

  33. Jaccard, S. L. & Galbraith, E. D. Large climate-driven changes of oceanic oxygen concentrations during the last deglaciation. Nat. Geosci. 5, 151–156 (2012).

    Article  ADS  CAS  Google Scholar 

  34. Weijer, W. & Dijkstra, H. A. Multiple oscillatory modes of the global ocean circulation. J. Phys. Oceanogr. 33, 2197–2213 (2003).

    Article  ADS  MathSciNet  Google Scholar 

  35. Sirkes, Z. & Tziperman, E. Identifying a damped oscillatory thermohaline mode in a general circulation model using an adjoint model. J. Phys. Oceanogr. 31, 2297–2306 (2001).

    Article  ADS  Google Scholar 

  36. Meissner, K. J., Eby, M., Weaver, A. J. & Saenko, O. A. CO2 threshold for millennial-scale oscillations in the climate system: Implications for global warming scenarios. Clim. Dyn. 30, 161–174 (2008).

    Article  Google Scholar 

  37. Haarsma, R. J., Opsteegh, J. D., Selten, F. M. & Wang, X. Rapid transitions and ultra-low frequency behaviour in a 40 kyr integration with a coupled climate model of intermediate complexity. Clim. Dyn. 17, 559–570 (2001).

    Article  Google Scholar 

  38. Stolper, D. A. & Keller, C. B. A record of deep-ocean dissolved O2 from the oxidation state of iron in submarine basalts. Nature 553, 323–327 (2018).

    Article  ADS  CAS  PubMed  Google Scholar 

  39. Brand, U. et al. Atmospheric oxygen of the Paleozoic. Earth Sci. Rev. 216, 103560 (2021).

    Article  CAS  Google Scholar 

  40. Dahl, T. W. et al. Reorganisation of Earth’s biogeochemical cycles briefly oxygenated the oceans 520 Myr ago. Geochem. Perspect. Lett. 3, 210–220 (2019).

    Google Scholar 

  41. Wei, G. Y. et al. Global marine redox evolution from the late Neoproterozoic to the early Paleozoic constrained by the integration of Mo and U isotope records. Earth Sci. Rev. 214, 103506 (2021).

    Article  CAS  Google Scholar 

  42. Wei, G. Y. et al. Marine redox fluctuation as a potential trigger for the Cambrian explosion. Geology 46, 587–590 (2018).

    Article  ADS  CAS  Google Scholar 

  43. Kendall, B. et al. Uranium and molybdenum isotope evidence for an episode of widespread ocean oxygenation during the late Ediacaran Period. Geochim. Cosmochim. Acta 156, 173–193 (2015).

    Article  ADS  CAS  Google Scholar 

  44. Dahl, T. W. et al. Brief oxygenation events in locally anoxic oceans during the Cambrian solves the animal breathing paradox. Sci Rep. 9, 11669 (2019).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  45. Payne, J. L., Bachan, A., Heim, N. A., Hull, P. M. & Knope, M. L. The evolution of complex life and the stabilization of the Earth system. Interface Focus 10, 20190106 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  46. Wilson, J. D., Monteiro, F. M., Schmidt, D. N., Ward, B. A. & Ridgwell, A. Linking marine plankton ecosystems and climate: a new modeling approach to the warm early Eocene climate. Paleoceanogr. Paleoclimatol. 33, 1439–1452 (2018).

    Article  ADS  Google Scholar 

  47. Cao, L. et al. The role of ocean transport in the uptake of anthropogenic CO2. Biogeosciences 6, 375–390 (2009).

    Article  ADS  CAS  Google Scholar 

  48. Reinhard, T. C. et al. Oceanic and atmospheric methane cycling in the cGENIE Earth system model – release v0.9.14. Geosci. Model Dev. 13, 5687–5706 (2020).

    Article  ADS  CAS  Google Scholar 

  49. van de Velde, S. J., Hülse, D., Reinhard, C. T. & Ridgwell, A. Iron and sulfur cycling in the cGENIE.muffin Earth system model (v0.9.21). Geosci. Model Dev. 14, 2713–2745 (2021).

    Article  ADS  Google Scholar 

  50. 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  ADS  PubMed  Google Scholar 

  51. Jacob, R. L. Low Frequency Variability in a Simulated Atmosphere-Ocean System.Thesis, Univ. Wisconsin (1997).

  52. Crichton, K. A., Ridgwell, A., Lunt, D., Farnsworth, A. & Pearson, P. Data-constrained assessment of ocean circulation changes since the middle Miocene in an Earth system model. Clim. Past 17, 2223–2254 (2021).

    Article  Google Scholar 

  53. Lê, S., Josse, J. & Husson, F. FactoMineR: an R package for multivariate analysis. J. Stat. Softw. 25, 1–18 (2008).

    Article  Google Scholar 

  54. Weijer, W. et al. Stability of the Atlantic Meridional Overturning Circulation: a review and synthesis. J. Geophys. Res. Oceans 124, 5336–5375 (2019).

    Article  ADS  Google Scholar 

  55. Ferreira, D., Marshall, J. & Campin, J. M. Localization of deep water formation: role of atmospheric moisture transport and geometrical constraints on ocean circulation. J. Clim. 23, 1456–1476 (2010).

    Article  ADS  Google Scholar 

  56. Garcia, H. E. et al. World Ocean Atlas 2018, Volume 3: Dissolved Oxygen, Apparent Oxygen Utilization, and Dissolved Oxygen Saturation (National Oceanic and Atmospheric Administration, 2019).

  57. Marsh, R. et al. Bistability of the thermohaline circulation identified through comprehensive 2-parameter sweeps of an efficient climate model. Clim. Dyn. 23, 761–777 (2004).

    Article  Google Scholar 

  58. DeVries, T. & Holzer, M. Radiocarbon and helium isotope constraints on deep ocean ventilation and mantle-3He sources. J. Geophys. Res. Oceans 124, 3036–3057 (2019).

    Article  ADS  Google Scholar 

  59. Song, H. et al. The onset of widespread marine red beds and the evolution of ferruginous oceans. Nat. Commun. 8, 399 (2017).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  60. Melchin, M. J., Mitchell, C. E., Holmden, C. & Štorch, P. Environmental changes in the Late Ordovician–early Silurian: review and new insights from black shales and nitrogen isotopes. Geol. Soc. Am. Bull. 125, 1635–1670 (2013).

    Article  ADS  CAS  Google Scholar 

  61. Pohl, A., Nardin, E., Vandenbroucke, T. R. A. & Donnadieu, Y. High dependence of Ordovician ocean surface circulation on atmospheric CO2 levels. Palaeogeogr. Palaeoclimatol. Palaeoecol. 458, 39–51 (2016).

    Article  Google Scholar 

  62. Meyer, K. M., Ridgwell, A. & Payne, J. L. The influence of the biological pump on ocean chemistry: implications for long-term trends in marine redox chemistry, the global carbon cycle, and marine animal ecosystems. Geobiology 14, 207–219 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank J. Mossinger for editorial handling. This project has received financing from the European Union’s Horizon 2020 research and innovation programme under Marie Skłodowska-Curie grant agreement no. 838373. Calculations were partly performed using HPC resources from DNUM CCUB (Centre de Calcul de l’Université de Bourgogne). A.R. acknowledges support from NSF grants 1736771 and EAR-2121165, as well as from the Heising-Simons Foundation. This is a contribution to UNESCO project IGCP 735 ‘Rocks and the Rise of Ordovician Life (Rocks n’ ROL)’.

Author information

Authors and Affiliations

Authors

Contributions

A.P. and A.R. designed the study and wrote the manuscript, with input from all co-authors. A.P. and A.R. conducted the FOAM and cGENIE experiments. A.P., A.R. and A.K. led the analysis of the model results. C.R.S. produced the continental reconstructions.

Corresponding author

Correspondence to Alexandre Pohl.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature thanks Tais Dahl, Shuhai Xiao, Sabin Zahirovic and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available.

Additional information

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

Extended data figures and tables

Extended Data Fig. 1 Comparison of modern observations of seafloor [O2] with the cGENIE model.

Results are shown for different (modern) continental grids, boundary conditions and biogeochemical cycling parameterization assumptions. Projection is equal-area rectangular and the colour scale is chosen to approximately match that in Fig. 2. a, Present-day distribution of [O2] globally at the seafloor, for which we re-grid the World Ocean Atlas 2018 (ref. 56) data to the modern continental grid of ref. 47 and show the oxygen concentration in the deepest model grid point. Apparent is both the production and southward propagation (via North Atlantic Deep Water (NADW)) of highly oxygenated waters from the North Atlantic and oxygenated deep-water production and mixing around Antarctica. b, Benthic [O2] distribution for the ‘standard’ modern ocean circulation of ref. 47 run at 278 ppm CO2, plus simplified biological export scheme. The large-scale patterns of benthic [O2] are reasonably reproduced, with the exception of too weak mixing of oxygen around the Southern Ocean, itself caused by a weak simulated Antarctic Circumpolar Current (ACC) and the difficulty in adequately representing the Drake Passage at this resolution. The slightly too low compared with observations values in the North Pacific is a further consequence of this. c, Test of substituting the simplified biological export scheme of ref. 47 with the explicit ecosystem model used here in the Phanerozoic series of simulations (but still at 278 ppm CO2). The slightly greater export simulated by the ecological model reduces benthic [O2] by about 10–20 µmol kg−1 while leaving the large-scale patterns largely unaltered. d, Deep-sea oxygenation in the 0 Ma Phanerozoic simulation (as per Fig. 2, 0 Ma) run at 2,240 ppm CO2. Without flux adjustment applied in the simplified cGENIE 2D EMBM atmosphere (see ref. 57), there is virtually no NADW, explaining the relatively poor and south-to-north oxygenation of the Atlantic. The Indian Ocean is also too poorly oxygenated, although the general pattern in the Pacific is reproduced. There are several main reasons for this model–data mismatch. First, directly re-gridding from a relatively coarse resolution GCM (FOAM) creates a highly restricted and shallow Drake Passage (Extended Data Fig. 7a, 0 Ma), precluding a strong ACC forming. More pertinently, the Phanerozoic simulation series (Fig. 2) are all run at 2,240 ppm. The resulting much-warmer-than-modern ocean is associated with the absence of any sea-ice formation and lower seawater oxygen solubility, probably explaining at least some of the spatial pattern and much of the lower global mean [O2] inventory (Extended Data Fig. 8a, 0 Ma). In terms of ventilation and water mass idealized mean age (not shown), the Atlantic, lacking an Atlantic MOC is far too old, whereas the Indian and Pacific oceans are similar to the inverse modelling of ref. 58, despite the aforementioned issues with the Drake Passage and that the 0 Ma simulation is run at 2,240 ppm CO2.

Extended Data Fig. 2 Selected redox proxy data versus the corresponding model oxygenation realization.

Benthic oxygen concentrations for simulations at 2,240 ppm CO2 (series #1). Eckert IV projection. Emerged continental masses are shaded white. Results are averaged over the last 5,000 years. Black (white) dots represent anoxic (oxic) conditions and grey points represent possible or intermittent anoxia, for 100 Ma after ref. 28, for 120 Ma after ref. 27, for 180 Ma after ref. 30, for 260 Ma after ref. 26, for 380 and 500 Ma after ref. 59 and for 440 Ma after ref. 60. These time slices have been chosen to represent regularly spaced periods during the Phanerozoic, typified by OAEs. The last 100 Myr are deliberately omitted because (1) the ocean is, overall, well oxygenated during this time interval and (2) our constant boundary conditions (deliberately chosen to isolate the role of tectonics from climate, see main text) do not (and, indeed, do not intend to) reproduce the pronounced cooling trend through the Cenozoic (especially from Early Eocene Climatic Optimum (ca. 50 Ma) onwards), precluding direct model–data comparison.

Extended Data Fig. 3 Deep-ocean circulation for simulations at 2,240 ppm CO2 (series #1).

a, Meridional overturning stream function, in Sv (sverdrup, 1 Sv = 106 m3 s−1). A negative (blue) stream function corresponds to an anticlockwise circulation. b, Annual distribution of convective adjustments across the water column. Emerged continental masses are shaded white. Eckert IV projection. Results are averaged over the last 5,000 years.

Extended Data Fig. 4 Sensitivity of benthic [O2] to the continental reconstruction.

Benthic oxygen concentrations for simulations at 2,240 ppm CO2 (such as series #1) at 440 Ma (a) and 460 Ma (b), using the continental reconstructions of BugPlates31, with topography/bathymetry after ref. 61. Eckert IV projection. Emerged continental masses are shaded white. Results are averaged over the last 5,000 years. Panels a and b are identical to the 440 Ma and 460 Ma panels of Fig. 2, except that simulations have been conducted using another continental reconstruction. Note that, although we simulate sulphate reduction in cGENIE, with SO42− being used as the electron acceptor for the remineralization of organic matter in the ocean interior once dissolved O2 has become depleted (see ref. 48), small negative O2 concentrations can arise when several geochemical reactions compete simultaneously for the same depleted oxygen pool. However, because the product of sulphate reduction—hydrogen sulphide (H2S)—has fast oxidation kinetics in the presence of free oxygen, the transport of H2S from anoxic to oxic areas closely mirrors the transport and fate of ‘negative oxygen’ (see ref. 62) and the overall redox landscape is largely independent of this small modelled oxygen overconsumption.

Extended Data Fig. 5 Sensitivity to the remineralization scheme.

Earth system model results for simulations at 2,240 ppm CO2 (such as #1) but with no dependence of remineralization on temperature. a, Same as Fig. 1. b, Same as Fig. 2.

Extended Data Fig. 6 Sensitivity of benthic [O2] to deep-ocean bathymetry.

Benthic oxygen concentrations for simulations at 2,240 ppm CO2 (such as #1) but with no mid-ocean ridges (see Extended Data Fig. 7). Results are averaged over the last 5,000 years. Emerged continental masses are shaded white. Eckert IV projection.

Extended Data Fig. 7 Bathymetric reconstructions.

Bathymetry with mid-ocean ridges used in series #1 and #2 (a) and flat-bottomed bathymetric reconstructions used in simulations with no mid-ocean ridges (b) (see Extended Data Fig. 6). Only reconstructions for 0 to 140 Ma (both included) differ (see Methods). Emerged continental masses are shaded white. Eckert IV projection.

Extended Data Fig. 8 Time evolution of benthic [O2].

a, Simulations at 2,240 ppm CO2 with temperature-dependent remineralization (series #1). b, Simulations in which we varied pCO2 to approximately ‘correct’ for the palaeogeographical impacts on climate (series #2).

Extended Data Fig. 9 Sensitivity of benthic [O2] to atmospheric forcing (pCO2).

Benthic oxygen concentrations for simulations at 1,120 ppm CO2, with temperature-dependent remineralization and mid-ocean ridges. Emerged continental masses are shaded white. Results are averaged over the last 5,000 years. Eckert IV projection.

Extended Data Fig. 10 Ocean circulation regimes in the cGENIE model.

a, Envelope of benthic [O2] values simulated at various atmospheric CO2 levels using Drake world (in blue, see maps in panel b) and ridge world (in black) (Methods). Regimes of stable equilibria and regimes of stable oscillations for Drake world simulations are numbered and labelled using a blue and a red background, respectively. b, Meridional overturning stream function in Sv (sverdrup, 1 Sv = 106 m3 s−1) and map of benthic ventilation age for each stable equilibrium and the two extreme states of each stable oscillatory regime identified in panel a, using the same numbering and colour coding. Lambert cylindrical equal-area projection. A negative (blue) stream function corresponds to an anticlockwise circulation. c, Tentative comparison with the latest Ordovician simulations of Pohl et al.21. Results are shown at 6,720 ppm and 2,380 ppm CO2, corresponding to the warm and cold states of scenario #1 in ref. 21.

Supplementary information

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pohl, A., Ridgwell, A., Stockey, R.G. et al. Continental configuration controls ocean oxygenation during the Phanerozoic. Nature 608, 523–527 (2022). https://doi.org/10.1038/s41586-022-05018-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-022-05018-z

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing Microbiology

Sign up for the Nature Briefing: Microbiology newsletter — what matters in microbiology research, free to your inbox weekly.

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