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:

Indo-Pacific Walker circulation drove Pleistocene African aridification

A Publisher Correction to this article was published on 23 February 2022

This article has been updated

Abstract

Today, the eastern African hydroclimate is tightly linked to fluctuations in the zonal atmospheric Walker circulation1,2. A growing body of evidence indicates that this circulation shaped hydroclimatic conditions in the Indian Ocean region also on much longer, glacial–interglacial timescales3,4,5, following the development of Pacific Walker circulation around 2.2–2.0 million years ago (Ma)6,7. However, continuous long-term records to determine the timing and mechanisms of Pacific-influenced climate transitions in the Indian Ocean have been unavailable. Here we present a seven-million-year-long record of wind-driven circulation of the tropical Indian Ocean, as recorded in Mozambique Channel Throughflow (MCT) flow-speed variations. We show that the MCT flow speed was relatively weak and steady until 2.1 ± 0.1 Ma, when it began to increase, coincident with the intensification of the Pacific Walker circulation6,7. Strong increases during glacial periods, which reached maxima after the Mid-Pleistocene Transition (0.9–0.64 Ma; ref. 8), were punctuated by weak flow speeds during interglacial periods. We provide a mechanism explaining that increasing MCT flow speeds reflect synchronous development of the Indo-Pacific Walker cells that promote aridification in Africa. Our results suggest that after about 2.1 Ma, the increasing aridification is punctuated by pronounced humid interglacial periods. This record will facilitate testing of hypotheses of climate–environmental drivers for hominin evolution and dispersal.

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: Regional oceanographic settings of IODP Site U1476 (2,166 m water depth).
Fig. 2: Climatic and oceanographic records over the past seven million years.
Fig. 3: MCT and eastern African aridity records with distinct glacial and interglacial intervals over the past 3.3 Myr.

Similar content being viewed by others

Data availability

All new benthic oxygen isotopic and lithogenic grain-size data of the spliced record of IODP Site U1476 are available via www.pangaea.de at https://doi.org/10.1594/PANGAEA.933831 and https://doi.org/10.1594/PANGAEA.933833, respectively.

Change history

References

  1. Saji, N., Goswami, B., Vinayachandran, P. & Yamagata, T. A dipole mode in the tropical Indian Ocean. Nature 401, 360–363 (1999).

    Article  ADS  CAS  Google Scholar 

  2. Tierney, J. E., Smerdon, J. E., Anchukaitis, K. J. & Seager, R. Multidecadal variability in East African hydroclimate controlled by the Indian Ocean. Nature 493, 389–392 (2013).

    Article  ADS  CAS  PubMed  Google Scholar 

  3. Kaboth-Bahr, S. et al. Paleo-ENSO influence on African environments and early modern humans. Proc. Natl Acad. Sci. USA 118, e2018277118 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Maslin, M. A. et al. East African climate pulses and early human evolution. Quat. Sci. Rev. 101, 1–17 (2014).

    Article  ADS  Google Scholar 

  5. Trauth, M. H., Larrasoana, J. C. & Mudelsee, M. Trends, rhythms and events in Plio–Pleistocene African climate. Quat. Sci. Rev. 28, 399–411 (2009).

    Article  ADS  Google Scholar 

  6. Etourneau, J., Schneider, R., Blanz, T. & Martinez, P. Intensification of the Walker and Hadley atmospheric circulations during the Pliocene–Pleistocene climate transition. Earth Planet. Sci. Lett. 297, 103–110 (2010).

    Article  ADS  CAS  Google Scholar 

  7. Ravelo, A. C., Andreasen, D. H., Lyle, M., Olivarez Lyle, A. & Wara, M. W. Regional climate shifts caused by gradual global cooling in the Pliocene epoch. Nature 429, 263–267 (2004).

    Article  ADS  CAS  PubMed  Google Scholar 

  8. Mudelsee, M. & Schulz, M. The Mid-Pleistocene climate transition: onset of 100 ka cycle lags ice volume build-up by 280 ka. Earth Planet. Sci. Lett. 151, 117–123 (1997).

    Article  ADS  CAS  Google Scholar 

  9. Hermes, J. & Reason, C. Annual cycle of the South Indian Ocean (Seychelles‐Chagos) thermocline ridge in a regional ocean model. J. Geophys. Res. Oceans 113, 1–10 (2008).

    Article  Google Scholar 

  10. Rao, S. A. & Behera, S. K. Subsurface influence on SST in the tropical Indian Ocean: structure and interannual variability. Dyn. Atmos. Oceans 39, 103–135 (2005).

    Article  ADS  Google Scholar 

  11. Nicholson, S. E. Climate and climatic variability of rainfall over eastern Africa. Rev. Geophys. 55, 590–635 (2017).

    Article  ADS  Google Scholar 

  12. deMenocal, P. B. African climate change and faunal evolution during the Pliocene–Pleistocene. Earth Planet. Sci. Lett. 220, 3–24 (2004).

    Article  ADS  CAS  Google Scholar 

  13. Levin, N. E. Environment and climate of early human evolution. Annu. Rev. Earth Planet. Sci.43, 405–429 (2015).

    Article  ADS  CAS  Google Scholar 

  14. Hall I. R. et al. South African Climates (Agulhas LGM Density Profile) (International Ocean Discovery Program, 2017).

  15. Ullgren, J., van Aken, H., Ridderinkhof, H., De, Ruijter, W. The hydrography of the Mozambique Channel from six years of continuous temperature, salinity, and velocity observations. Deep Sea Res. I 69, 36–50 (2012).

    Article  Google Scholar 

  16. Backeberg, B. C. & Reason, C. J. C. A connection between the South Equatorial Current north of Madagascar and Mozambique Channel eddies. Geophys. Res. Lett. 37, 1–6 (2010).

    Article  Google Scholar 

  17. Makarim, S. et al. Previously unidentified Indonesian Throughflow pathways and freshening in the Indian Ocean during recent decades. Sci. Rep. 9, 7364 (2019).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  18. Backeberg, B. C., Penven, P. & Rouault, M. Impact of intensified Indian Ocean winds on mesoscale variability in the Agulhas system. Nat. Clim. Change 2, 608–612 (2012).

    Article  ADS  Google Scholar 

  19. Paris, M. L., Subrahmanyam, B., Trott, C. B. & Murty, V. S. N. Influence of ENSO events on the Agulhas leakage region. Remote Sens. Earth Syst. Sci. 1, 79–88 (2018).

    Article  Google Scholar 

  20. Palastanga, V., Van Leeuwen, P. & De, Ruijter, W. A link between low‐frequency mesoscale eddy variability around Madagascar and the large‐scale Indian Ocean variability. J. Geophys. Res. Oceans 111, 1–15 (2006).

    Article  Google Scholar 

  21. Ridderinkhof, H. et al Seasonal and interannual variability in the Mozambique Channel from moored current observations. J. Geophys. Res. Oceans 115, 1–18 (2010).

    Article  Google Scholar 

  22. Ridderinkhof, W., Le Bars, D., Von der Heydt, A. & De, Ruijter, W. Dipoles of the south east Madagascar Current. Geophys. Res. Lett. 40, 558–562 (2013).

    Article  ADS  Google Scholar 

  23. Yamagami, Y. & Tozuka, T. Interannual variability of South Equatorial Current bifurcation and western boundary currents along the Madagascar coast. J. Geophys. Res. Oceans 120, 8551–8570 (2015).

    Article  ADS  Google Scholar 

  24. Jin, X. et al. Influences of Pacific climate variability on decadal subsurface ocean heat content variations in the Indian Ocean. J. Clim. 31, 4157–4174 (2018).

    Article  ADS  Google Scholar 

  25. van Oldenborgh, G. J., Doblas-Reyes, F. J., Wouters, B. & Hazeleger, W. Decadal prediction skill in a multi-model ensemble. Clim. Dyn. 38, 1263–1280 (2012).

    Article  Google Scholar 

  26. Herbert, T. D. et al. Late Miocene global cooling and the rise of modern ecosystems. Nat. Geosci. 9, 843–847 (2016).

    Article  ADS  CAS  Google Scholar 

  27. Groeneveld, J. et al. Australian shelf sediments reveal shifts in Miocene Southern Hemisphere westerlies. Sci. Adv. 3, 1–9 (2017).

    Article  Google Scholar 

  28. Liu, J. et al. Eastern equatorial Pacific cold tongue evolution since the late Miocene linked to extratropical climate. Sci. Adv. 5, 1–7 (2019).

    ADS  Google Scholar 

  29. Fedorov, A. et al. Patterns and mechanisms of early Pliocene warmth. Nature 496, 43–49 (2013).

    Article  ADS  CAS  PubMed  Google Scholar 

  30. Haug, G. H., Tiedemann, R., Zahn, R. & Ravelo, A. C. Role of Panama uplift on oceanic freshwater balance. Geology 29, 207–210 (2001).

    Article  ADS  CAS  Google Scholar 

  31. Brierley, C. M. et al. Greatly expanded tropical warm pool and weakened Hadley circulation in the Early Pliocene. Science 323, 1714–1718 (2009).

    Article  ADS  CAS  PubMed  Google Scholar 

  32. Tierney, J. E., Haywood, A. M., Feng, R., Bhattacharya, T. & Otto-Bliesner, B. L. Pliocene warmth consistent with greenhouse gas forcing. Geophys. Res. Lett. 46, 9136–9144 (2019).

    Article  ADS  Google Scholar 

  33. Cane, M. A. & Molnar, P. Closing of the Indonesian seaway as a precursor to east African aridification around 3–4 million years ago. Nature 411, 157–162 (2001).

    Article  ADS  CAS  PubMed  Google Scholar 

  34. Petrick, B. et al. Glacial Indonesian Throughflow weakening across the Mid-Pleistocene climatic transition. Sci. Rep. 9, 16995 (2019).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  35. Karas, C. et al. Mid-Pliocene climate change amplified by a switch in Indonesian subsurface throughflow. Nat. Geosci. 2, 434–438 (2009).

    Article  ADS  CAS  Google Scholar 

  36. Karas, C., Nürnberg, D., Tiedemann, R. & Garbe-Schönberg, D. Pliocene Indonesian Throughflow and Leeuwin Current dynamics: implications for Indian Ocean polar heat flux. Paleoceanography 26, PA2217 (2011).

    Article  ADS  Google Scholar 

  37. Christensen, B. A. et al. Indonesian Throughflow drove Australian climate from humid Pliocene to arid Pleistocene. Geophys. Res. Lett. 44, 6914–6925 (2017).

    Article  ADS  Google Scholar 

  38. Chen, Z., Wu, L., Qiu, B., Sun, S. & Jia, F. Seasonal variation of the South Equatorial Current bifurcation off Madagascar. J. Phys. Oceanogr. 44, 618–631 (2014).

    Article  ADS  Google Scholar 

  39. Rippert, N., Baumann, K.-H. & Pätzold, J. Thermocline fluctuations in the western tropical Indian Ocean during the past 35 ka. J. Quat. Sci. 30, 201–210 (2015).

    Article  Google Scholar 

  40. Ahn, S., Khider, D., Lisiecki, L. E. & Lawrence, C. A probabilistic Pliocene–Pleistocene stack of benthic δ18O using a profile hidden Markov model. Dyn. Stat. Clim. Syst. 2, 1–16 (2017).

    Google Scholar 

  41. Zachos, J., Pagani, M., Sloan, L., Thomas, E. & Billups, K. Trends, rhythms, and aberrations in global climate 65 Ma to present. Science 292, 686–693 (2001).

    Article  ADS  CAS  PubMed  Google Scholar 

  42. Ravelo, A. C., Lawrence, K. T., Fedorov, A. & Ford, H. L. Comment on “A 12-million-year temperature history of the tropical Pacific Ocean”. Science 346, 1467 (2014).

    Article  ADS  CAS  PubMed  Google Scholar 

  43. Wara, M. W., Ravelo, A. C. & Delaney, M. L. Permanent El Niño-like conditions during the Pliocene warm period. Science 309, 758–761 (2005).

    Article  ADS  CAS  PubMed  Google Scholar 

  44. Bjerknes, J. Atmospheric teleconnections from the equatorial Pacific. Mon. Weather Rev. 97, 163–172 (1969).

    Article  ADS  Google Scholar 

  45. Emeis, K.-C., Anderson, D. M., Doose, H., Kroon, D., Schulz-Bull, D. Sea-surface temperatures and the history of monsoon upwelling in the northwest Arabian Sea during the last 500,000 years. Quat. Res. 43, 355–361 (1995).

    Article  Google Scholar 

  46. Herbert, T. D., Peterson, L. C., Lawrence, K. T. & Liu, Z. Tropical ocean temperatures over the past 3.5 million years. Science 328, 1530–1534 (2010).

    Article  ADS  CAS  PubMed  Google Scholar 

  47. Bobe, R. & Behrensmeyer, A. K. The expansion of grassland ecosystems in Africa in relation to mammalian evolution and the origin of the genus Homo. Palaeogeogr. Palaeoclimatol. Palaeoecol. 207, 399–420 (2004).

    Article  Google Scholar 

  48. Kiefer, T., McCave, I. N. & Elderfield, H. Antarctic control on tropical Indian Ocean sea surface temperature and hydrography. Geophys. Res. Lett. 33, 1–5 (2006).

    Article  Google Scholar 

  49. Mohtadi, M., Prange, M., Schefuß, E. & Jennerjahn, T. C. Late Holocene slowdown of the Indian Ocean Walker circulation. Nat. Commun. 8, 1015 (2017).

    Article  CAS  Google Scholar 

  50. Faith J. T. et al. Rethinking the ecological drivers of hominin evolution. Trends Ecol. Evol. 36, 797–807 (2021).

  51. Hall I. R., Hemming S. R. & LeVay L. J. South African Climates (Agulhas LGM Density Profile) (2016).

  52. Konert, M. & Vandenberghe, J. Comparison of laser grain size analysis with pipette and sieve analysis: a solution for the underestimation of the clay fraction. Sedimentology 44, 523–535 (1997).

    Article  ADS  CAS  Google Scholar 

  53. Van der Lubbe, H. et al. Sedimentation patterns off the Zambezi River over the last 20,000 years. Mar. Geol. 355, 189–201 (2014).

    Article  ADS  Google Scholar 

  54. McCave, I. & Andrews, J. Distinguishing current effects in sediments delivered to the ocean by ice. I. Principles, methods and examples. Quat. Sci. Rev. 212, 92–107 (2019).

    Article  ADS  Google Scholar 

  55. Weltje, G. J. & Tjallingii, R. Calibration of XRF core scanners for quantitative geochemical logging of sediment cores: theory and application. Earth Planet. Sci. Lett. 274, 423–438 (2008).

    Article  ADS  CAS  Google Scholar 

  56. Miramontes, E. et al. The influence of bottom currents on the Zambezi Valley morphology (Mozambique Channel, SW Indian Ocean): in situ current observations and hydrodynamic modelling. Mar. Geol. 410, 42–55 (2019).

    Article  ADS  Google Scholar 

  57. Starr, A. et al. Antarctic icebergs reorganize ocean circulation during Pleistocene glacials. Nature 589, 236-241 (2021).

    Article  ADS  CAS  PubMed  Google Scholar 

  58. Anthonissen, D. E. & Ogg, J. G. in The Geologic Time Scale (eds Gradstein, F. M. et al.) 1083–1127 (Elsevier, 2012).

  59. Tangunan, D. N. et al. The last 1 million years of the extinct genus Discoaster: Plio–Pleistocene environment and productivity at Site U1476 (Mozambique Channel). Palaeogeogr. Palaeoclimatol. Palaeoecol. 505, 187–197 (2018).

    Article  Google Scholar 

  60. Lougheed, B. C. & Obrochta, S. P. A rapid, deterministic age-depth modeling routine for geological sequences with inherent depth uncertainty. Paleoceanogr. Paleoclimatol. 34, 122–133 (2019).

    Article  ADS  Google Scholar 

  61. Mudelsee, M. Ramp function regression: a tool for quantifying climate transitions. Comput. Geosci. 26, 293–307 (2000).

    Article  ADS  Google Scholar 

  62. Huang, Y., Clemens, S. C., Liu, W., Wang, Y. & Prell, W. L. Large-scale hydrological change drove the late Miocene C4 plant expansion in the Himalayan foreland and Arabian Peninsula. Geology 35, 531–534 (2007).

    Article  ADS  CAS  Google Scholar 

  63. Schott, F. A., Xie, S. P. & McCreary, J. P. Jr Indian Ocean circulation and climate variability. Rev. Geophys. 47, 1–46 (2009).

    Article  Google Scholar 

  64. McCreary, J. P., Kundu, P. K. & Molinari, R. L. A numerical investigation of dynamics, thermodynamics and mixed-layer processes in the Indian Ocean. Prog. Oceanogr. 31, 181–244 (1993).

    Article  ADS  Google Scholar 

  65. Centenary Edition of the GEBCO Digital Atlas (IOC, IHO, BODC, 2003).

  66. Mulet, S. et al. The new CNES-CLS18 global mean dynamic topography. Ocean Sci. Discuss. 2021, 1–31 (2021).

    Google Scholar 

  67. Putrasahan, D., Kirtman, B. P. & Beal, L. M. Modulation of SST interannual variability in the Agulhas leakage region associated with ENSO. J. Clim. 29, 7089–7102 (2016).

    Article  ADS  Google Scholar 

  68. Abram, N. J. et al. Coupling of Indo-Pacific climate variability over the last millennium. Nature 579, 385–392 (2020).

    Article  ADS  CAS  PubMed  Google Scholar 

  69. Rayner, N. A. et al. Global analyses of sea surface temperature, sea ice, and night marine air temperature since the late nineteenth century. J. Geophys. Res. Atmos. 108, 1–29 (2003).

    Article  Google Scholar 

  70. McCave, I., Thornalley, D. & Hall, I. Relation of sortable silt grain-size to deep-sea current speeds: calibration of the ‘mud current meter’. Deep Sea Res. I 127, 1–2 (2017).

    Article  Google Scholar 

  71. McCave, I. N., Manighetti, B. & Robinson, S. G. Sortable silt and fine sediment size/composition slicing: parameters for palaeocurrent speed and palaeoceanography. Paleoceanography 10, 593–610 (1995).

    Article  ADS  Google Scholar 

  72. van Aken, H. M., Ridderinkhof, H. & de Ruijter, W. P. North Atlantic deep water in the south-western Indian Ocean. Deep Sea Res. I 51, 755–776 (2004).

    Article  Google Scholar 

  73. Ridderinkhof, H. & De Ruijter, W. Moored current observations in the Mozambique Channel. Deep Sea Res. II 50, 1933–1955 (2003).

    Article  ADS  Google Scholar 

  74. Lisiecki, L. E., Raymo, M. E. A Pliocene–Pleistocene stack of 57 globally distributed benthic δ18O records. Paleoceanography 20, PA1003 (2005).

    ADS  Google Scholar 

  75. Dekens, P. S., Lea, D. W., Pak, D. K., Spero, H. J. Core top calibration of Mg/Ca in tropical foraminifera: refining paleotemperature estimation. Geochem. Geophys. Geosyst. 3, 1–29 (2002).

    Article  Google Scholar 

Download references

Acknowledgements

We are grateful to the government of Mozambique for their permission for deep-sea drilling operations in the context of palaeoclimate and palaeoceanography research. This research used samples and data provided by the Integrated Ocean Drilling Program (IODP). We appreciate the entire shipboard party of IODP Expedition 361 ‘South African Climates’ for their tireless efforts at sea. We thank S. Conn, M. Hagen, S. Lordsmith, A. Nederbraght, L. Owen, M. P. Prins, S. Rumping, S. Slater, C. W. Nooitgedacht, F. van Bakel, S. de Bie and C. van Eijbergen for laboratory assistance. Funding for this work was provided in part by the Dutch Research Council (NWO) (Open Programme grant number 824.01.005 to H.J.L.v.d.L., and Talent Programme grant number 016.Vidi.171.049 to J.C.A.J.) and the UK Natural Environment Research Council (grant number NE/P000037/1 to I.R.H. and grant number NE/P000878/1 to S.B.). We acknowledge R. D. Norris and N. Mantke for X-ray fluorescence scanning the spliced record of Site U1476.

Author information

Authors and Affiliations

Authors

Contributions

IODP Expedition 361 was led by I.R.H. and S.R.H. S.B. produced the benthic oxygen isotope record and J.J. produced the X-ray fluorescence bulk chemistry records. H.J.L.v.d.L. and I.R.H. designed the research, H.J.L.v.d.L. performed the grain-size analysis with input from T.F.B. and performed further data analyses with input from I.R.H. and A.S. H.J.L.v.d.L. conducted oceanographic and climatic analyses with input from B.C.B. H.J.L.v.d.L. wrote the manuscript with contributions from I.R.H., J.C.A.J. and A.S. All authors contributed to the data interpretation and commented on the final manuscript. T.F.B. edited the figure layout for publication.

Corresponding authors

Correspondence to H. J. L. van der Lubbe or I. R. Hall.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature thanks John Andrews, Clara Bolton, Matt O’Regan, Benjamin Petrick and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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 Coupled oceanographic and atmospheric circulation of the Indian Ocean region.

a, Surface circulation indicated by mean geostrophic velocities and directions. IODP Site U1476 is situated in the Mozambique Channel, which experiences a net southward flow of the Mozambique Channel Throughflow (MCT). The tropical Indian gyre receives and redistributes inflow from the Indonesian Throughflow (ITF). The main components of the tropical Indian gyre: East African Coastal Current (EACC), Southern Equatorial Current (SEC), Southern Equatorial Counter Current (SECC) and Northeast Madagascar Current (NEMC) are highlighted with a dark grey outline marking the mean extension of the tropical gyre63,64. At ~60°E, the SEC bifurcates into two main branches feeding into the NEMC and the Southeastern Madagascar Current (SEMC) as it crosses bathymetric highs including the Mascarene Plateau, highlighted by the solid grey bathymetric contour at 2,200 m water depth. The SEMC is outside the outline of the tropical gyre as it is part of the anti-cyclonic subtropical gyre. b, Mean wind stress (gray arrows) and wind stress curl of the Indian Ocean indicate the basin-wide negative (positive) wind stress curl forcing the tropical (subtropical) gyre. A black line indicates the zero wind stress curl. c, Mean dynamic sea surface topography indicating a sea-level low between 10-5°S at the centre of the tropical Indian gyre; the Seychelles–Chagos thermocline ridge (SCTR)9. The thicker contour of 1.05 m denotes the northern extent of the subtropical gyre, whereby the blues (reds) highlight lower (higher) sea surfaces, which are associated with the cyclonic (anti-cyclonic) circulation of the tropical (subtropical) gyres. The bathymetric, oceanic (1993-2012) and surface wind data (1979-2019) are derived from: General Bathymetric Chart of the Oceans (GEBCO)65, CNES-CLS18 MDT66, ERA5 monthly-averaged data on single levels (DOI: 10.24381/cds.6860a573), respectively. These maps are generated with MATLAB and Mapping Toolbox, version 9.8.0.1323502 (R2020a, Natick, Massachusetts: The MathWorks Inc., United States).

Extended Data Fig. 2 Dynamic sea surface topography in meters superimposed by the mean of monthly Indian Ocean sea surface height (SSH) maps.

a, Composite of all monthly SSH between 1999-2019, superimposed on dynamic sea surface topography, marks sea-level low of the SCTR (5°S–12°S, 45°E–90°E)9. Solid red line refers to cross section along longitude 52°E between latitudes 25 to 5°S, across the SEC, see b. b, Mean dynamic sea surface gradient at 52°E (red thick line) as marked by solid black line in a, c, d, e and f. Meridional surface height gradient and associated near-surface pressure gradient south of the SCTR drives (influenced by Coriolis force) the westward deep-reaching South Equatorial Current (SEC). Positive (negative) SSH anomalies formed during ENSO and IOD propagate westward as downwelling (upwelling) Rossby waves in ca. 6 months10,19,24,67. The dynamic sea surface topography averaged over 6 months following positive Indian Ocean Dipole (+IOD), negative Indian Dipole (-IOD), positive El Niño Southern Oscillation (ENSO) - El Niño, and negative ENSO - La Niña are shown in c, d, e and f, respectively. g, IOD and ENSO time series with solid vertical lines indicating long-term mean with ±1SD and ±2SD (dashed lines). Red dots mark the months that are selected after a positive IOD (c) and ENSO (e) events, while blue dots mark the months after negative IOD (d) and ENSO (f) events. SSH anomalies induced by ENSO and IOD often reinforce each other10, since they are linked at interannual1 to decadal time-scales63, at least over the last millennium68. Sea surface topography, SSH maps, IOD SST index1, El Niño 3.4 SST Index69 are from Ssalto/Duacs-Cnes; https://www.aviso.altimetry.fr, https://psl.noaa.gov/gcos_wgsp/Timeseries/DMI/, https://psl.noaa.gov/gcos_wgsp/Timeseries/Nino34. These maps are generated with MATLAB® and Mapping Toolbox, version 9.8.0.1323502 (R2020a, Natick, Massachusetts: The MathWorks Inc., United States).

Extended Data Fig. 3

3d representations of the mean sea surface topography. a, b, Averaged over 6 months after (a) warm (positive) and (b) cold (negative) ENSO-IOD phases, respectively induced by regional anti-cyclonic (AC) positive and cyclonic (C) negative wind stress curls along the equatorial Indian Ocean west of 100°E, which is coupled to the atmospheric Pacific Walker Cell circulation10,24. In contrast, sea surface height (SSH) variability in the eastern Indian Ocean is derived from the western Pacific Ocean via the Indonesian Throughflow (ITF; Extended Data Figs. 1a, 2c-f). Black arrows indicate schematic representation of the zonal Walker circulation. During positive ENSO phases, the center of atmospheric deep convection shifts eastward, resulting in anomalous descending air masses over the western Pacific Ocean and Maritime Continent63. The corresponding anomalous easterlies induce down-welling Rossby waves in the central Indian Ocean that propagate westward as positive SSH anomalies, increasing the thermocline depth at the SCTR while decreasing the meridional SSH gradient and corresponding SEC in the western Indian Ocean10,24 (Extended Data Fig. 2b). Conversely, the SSH gradient and in turn the SEC flow increases during negative ENSO phases. The associated westerlies/easterlies induce upwelling and thermocline shoaling (dark blue) in the western/eastern Indian Ocean, and in turn promote deep atmospheric convection and excess rainfall over the Maritime Continent/eastern Africa via sea atmospheric interactions10,17. The Mozambique Channel Throughflow (MCT) that is related to the westward flow of the SEC north of Madagascar therefore increases following a negative-cold ENSO-IOD phase (b) and vice versa (a).

Extended Data Fig. 4 Precipitation difference (mm/day) of mean precipitation between 2000-2012 minus that of 1979-1999.

The precipitation difference indicates systematically drier conditions in eastern Africa during the last decade17, which coincides with intensified Southern Equatorial Current (SEC)17 and Mozambique Channel Throughflow (MCT)70. Green triangles indicate the eastern African hominin sites (1. Afar, 2. Omo-Turkana Basin, 3. Baringo and Tugen Hills, 4. Kanjera, 5. Olorgesailie, 6. Laetoli and Olduvai), from which long-term carbon isotope records of soil carbonates are available serving as long-term proxy of eastern African aridity. The rainfall data is from Global Precipitation Climatology Project (GPCP) monthly precipitation analysis (1979–present; DOI 10.7289/V56971M6) and visualized with MATLAB and Mapping Toolbox, version 9.8.0.1323502 (R2020a, Natick, Massachusetts: The MathWorks Inc., United States).

Extended Data Fig. 5 Flow speed reconstructions using sortable silt mean grain size and abundance in the Mozambique Channel.

a, 3-part diagram of sortable silt properties of flow speed in the Mozambique Channel (Clock-wise). The relationship between sortable silt abundance (% lithogenic fraction between 10-63 μm; SS%) and mean grain size (\(\overline{{\rm{SS}}}\))54,71,72 for the spliced record from IODP Site U147614 and modern surface sediments (64PE304-47, −51, −56, −63, −66 and −68)53. The strong correlation (Pearsons correlation coefficient = 0.79) indicates that sortable silt deposition was subject to current sorting and selective transport71. The \(\overline{{\rm{SS}}}\) of U1476 and nearby surface sediments are well within the calibration range for universal near-bottom current flow speed reconstructions71. The inferred flow speeds for the surface sediment samples using the universal \(\overline{{\rm{SS}}}\) flow speed calibration correspond to the +2 s.d. from the mean near-bottom currents obtained from nearby mooring stations. The mean flow directions are southwards except for those of the Mozambique undercurrent (MUC) that is confined to the eastern African margin below 1.5 km water depth15. The \(\overline{{\rm{SS}}}\) is likely somewhat biased towards to higher near bottom flow speeds, as the finest fractions that may have selectively deposited during slow near-bottom current conditions were preferentially removed under intra- and interannually increased near-bottom currents. Additionally, slight deviations from the mean flow speed might be further attributed to the regional nature of the lithogenic sediments, as most \(\overline{{\rm{SS}}}\) flow speed calibrations are defined for the northern Atlantic Ocean71. However, the sensitivity is comparable amongst the local calibrations and therefore the inferred relative changes in flow speed are also accurate for the Mozambique Channel, albeit the absolute values may slightly differ. In this study, the coarsest \(\overline{{\rm{SS}}}\) and in turn highest flow speeds are obtained from near and within the topographic depression, where the continuation of the MUC passes through the Davie Ridge73. b, W-E transect across the Mozambique Channel with long-term flow contours (cm s−1)15 as in Fig. 1c, but superimposed by the mean flow speeds of individual moorings of mean near-bottom currents across the Mozambique Channel (cm s−1)2173, whereby negative (positive) values reflect southward (northward) flow. c, Mozambique Channel bathymetric map with transect across Site IODP U1476 (S2) and the mooring transect (S1) with station (small solid black squares). The mean near-bottom flow speeds vectors recorded at the mooring transect are indicated by dashed73 and solid21 lines. The locations of the surface sediment samples are marked by white dots with black outlines (64PE304-47, −51, −56, −63, −66 and −68)53.

Extended Data Fig. 6 Age model of the spliced record of IODP Site U1476.

a, From top to bottom: The chronology of spliced core of IODP Site U1476 is based on the correlation of the 𝛿18O benthic record with the global 𝛿18O benthic probabilistic stack (Prob-stack)40 using 38 tie points for the last ~2 Ma. Benthic 𝛿18O record of the spliced record of IODP Site U1476 over the last ~2 Ma (blue) and Prob-stack (black)40. Minimal number of tie-points that are used to tune the 𝛿18O benthic record of U1476 to Prob-stack are indicated by vertical lines in a following a similar approach as57. The Prob-stack overlain by the tuned benthic 𝛿18O record of the spliced record of IODP Site U1476 demonstrates the similarity between both records. b, The age-depth relationship agrees with the biochronology of calcareous nannofossils and planktonic foraminifera with only some minor deviations from the 𝛿18O bayesian age model. Bayesian age-modeling60 of U1476 with ±1SD (dark grey) and ±2SD (light grey) based on the benthic 𝛿18O tuning points and biostratigraphy of calcareous nannofossils over the last 3 Ma with the accumulation rates. The shipboard biostratigraphy of calcareous nannofossils14, which has been further refined by Tangunan et al.(2018)59 displaying internal agreement. c, Idem as b for the last ~7 Ma, Deviations of the shipboard biochronology of planktonic foraminifera might be partly attributable to the use of general low-latitude calibrations rather than specific calibrations for the tropical western Indian Ocean14,51.

Extended Data Fig. 7 7-Ma long records of sortable silt mean (\(\overline{{\rm{SS}}}\)) and derived flow speed changes together with lithogenic properties.

The accompanying elemental compositions are obtained through X-ray fluorescence (XRF) analyses of sediments at Site U1476 (Methods). The glacial periods are indicated for the last 5.3 Ma74 by vertical light blue bars. The 10-point running means of the lithogenic ln (Zr/Rb) record reflect the relative deposition of dense Zr grains that are sorted, via selective deposition, together with silt-sand sized terrigenous particles, and Rb that is mainly present in clay minerals as substitution for K. Additionally, the 10-point running means of XRF bulk record (ln) Ca/(Ca/Ti) record that represents the relative deposition of carbonates (including foraminiferal shells) versus the terrigenous fraction53. Reconstructed enhancements in flow speed after 2.1 ± 0.1 Ma correspond to increases in coarse-grained lithogenic sediments together with increases in marine carbonates, which suggests selective deposition and removal of the fine-grained lithogenic sediment fractions.

Extended Data Fig. 8 Long-term Sea Surface Temperature (SST) records and Indo-Pacific Walker Cell circulation.

a, Long-term SST records from the Indian and Pacific Ocean. The Mg/Ca-based SST records of DSDP 21435, ODP 709C36, 806 and 84743 that are calcuated and corrected for dissolution at depth75 are mainly derived from Globigerinoides sacculifer, therefore recording temperatures about 20-30 m below the surface. Divergence of the SST records at ~2.1 Ma reflects the onset of the modern Indo-Pacfiic Walker cell circulation. b, Representation of present-day coupled Indo-Pacific Walker cell circulation, which is characterized by climatological low-level westerlies11 and easterlies over the equatorial Indian and Pacific Oceans, respectively. The corresponding moisture-laden air masses of both oceans ascend over the Maritime Continent in southeastern Asia and the associated atmospheric deep convection induces excess rainfall. In contrast, the subsiding dry air masses over the cool western Indian Ocean cause arid conditions in eastern Africa. c, Ocean map color gradients show climatological mean Sea Surface Temperature (SST) and black arrows represent the atmospheric surface circulation of the Indian and Pacific Ocean basins (ERA5 monthly-averaged data on single levels; https://doi.org/10.24381/cds.6860a573). The low-level Pacific easterlies and Indian westerlies are driven by the temperature contrast between the center of the Indo-Pacific Warm Pool (IPWP; SST >28 °C) and upwelling areas in the eastern Pacific Ocean (Cold tongue) and western Indian Ocean (Seychelles-Chagos Thermocline Ridge (SCTR), western Indian Ocean and Arabian Sea). The yellow dots indicate IODP Site U1476 (this study), and U133728, DSDP Site 21435, ODP Sites 709C36, 721-72212,46,62, 806 and 84743 providing long-term SST records, as well as sites GeoB12610-239 and GeoB10038-449 that date back to the Last Glacial Maximum (LGM). The SST map has been derived from NOAA Extended Reconstructed Sea Surface Temperature (ERSST) (Version 5, NOAA National Centers for Environmental Information, DOI:10.7289/V5T72FNM) is plotted with MATLAB and Mapping Toolbox, version 9.8.0.1323502 (R2020a, Natick, Massachusetts: The MathWorks Inc., United States).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

van der Lubbe, H.J.L., Hall, I.R., Barker, S. et al. Indo-Pacific Walker circulation drove Pleistocene African aridification. Nature 598, 618–623 (2021). https://doi.org/10.1038/s41586-021-03896-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-021-03896-3

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

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