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:

Widespread phytoplankton blooms triggered by 2019–2020 Australian wildfires

Abstract

Droughts and climate-change-driven warming are leading to more frequent and intense wildfires1,2,3, arguably contributing to the severe 2019–2020 Australian wildfires4. The environmental and ecological impacts of the fires include loss of habitats and the emission of substantial amounts of atmospheric aerosols5,6,7. Aerosol emissions from wildfires can lead to the atmospheric transport of macronutrients and bio-essential trace metals such as nitrogen and iron, respectively8,9,10. It has been suggested that the oceanic deposition of wildfire aerosols can relieve nutrient limitations and, consequently, enhance marine productivity11,12, but direct observations are lacking. Here we use satellite and autonomous biogeochemical Argo float data to evaluate the effect of 2019–2020 Australian wildfire aerosol deposition on phytoplankton productivity. We find anomalously widespread phytoplankton blooms from December 2019 to March 2020 in the Southern Ocean downwind of Australia. Aerosol samples originating from the Australian wildfires contained a high iron content and atmospheric trajectories show that these aerosols were likely to be transported to the bloom regions, suggesting that the blooms resulted from the fertilization of the iron-limited waters of the Southern Ocean. Climate models project more frequent and severe wildfires in many regions1,2,3. A greater appreciation of the links between wildfires, pyrogenic aerosols13, nutrient cycling and marine photosynthesis could improve our understanding of the contemporary and glacial–interglacial cycling of atmospheric CO2 and the global climate system.

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: Maps of black carbon AOD and [Chla] anomalies and their historical records.
Fig. 2: Temporal patterns of black carbon AOD and satellite [Chla] in two regions denoted in Fig. 1 during the 2019–2020 Australian wildfire season.
Fig. 3: Plankton blooms observed by in situ measurements from BGC-Argo floats and satellites.
Fig. 4: Enhancement in marine phytoplankton productivity during the 2019–2020 Australian wildfires.

Similar content being viewed by others

Data availability

The ESA’s chlorophyll-a products can be accessed at http://www.esa-oceancolour-cci.org/. Satellite aerosol data are available from the Giovanni online data system (https://giovanni.gsfc.nasa.gov/giovanni/). The Copernicus Atmosphere Monitoring Service (CAMS) aerosol reanalysis datasets can be downloaded from the CAMS Atmosphere Data Store (ADS; https://ads.atmosphere.copernicus.eu/cdsapp#!/dataset/cams-global-reanalysis-eac4?tab=overview). The Argo float data are openly available on the Ifremer ftp-server (ftp://ftp.ifremer.fr/ifremer/argo/dac/). The net primary production estimates are available from the Ocean Productivity website (http://sites.science.oregonstate.edu/ocean.productivity/index.php). Access to datasets analysed in this study is also provided in the Methods section. Datasets generated in this study are provided as Source data and at https://doi.org/10.5281/zenodo.4895657Source data are provided with this paper.

References

  1. Bowman, D. M. J. S. et al. Vegetation fires in the Anthropocene. Nat. Rev. Earth Environ. 1, 500–515 (2020).

    Article  ADS  Google Scholar 

  2. Abatzoglou, J. T., Williams, A. P. & Barbero, R. Global emergence of anthropogenic climate change in fire weather indices. Geophys. Res. Lett. 46, 326–336 (2019).

    Article  ADS  Google Scholar 

  3. Huang, Y., Wu, S. & Kaplan, J. O. Sensitivity of global wildfire occurrences to various factors in the context of global change. Atmos. Environ. 121, 86–92 (2015).

    Article  ADS  CAS  Google Scholar 

  4. van Oldenborgh, G. J. et al. Attribution of the Australian bushfire risk to anthropogenic climate change. Nat. Hazards Earth Syst. Sci. 21, 941–960 (2021).

    Article  ADS  Google Scholar 

  5. Ward, M. et al. Impact of 2019–2020 mega-fires on Australian fauna habitat. Nat. Ecol. Evol. 4, 1321–1326 (2020).

    Article  PubMed  Google Scholar 

  6. Kablick III, G. P., Allen, D. R., Fromm, M. D. & Nedoluha, G. E. Australian PyroCb smoke generates synoptic-scale stratospheric anticyclones. Geophys. Res. Lett. 47, e2020GL088101 (2020).

    Article  ADS  Google Scholar 

  7. Hirsch, E. & Koren, I. Record-breaking aerosol levels explained by smoke injection into the stratosphere. Science 371, 1269–1274 (2021).

    Article  ADS  CAS  PubMed  Google Scholar 

  8. Schlosser, J. S. et al. Analysis of aerosol composition data for western United States wildfires between 2005 and 2015: Dust emissions, chloride depletion, and most enhanced aerosol constituents. J. Geophys. Res. Atmos. 122, 8951–8966 (2017).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  9. Barkley, A. E. et al. African biomass burning is a substantial source of phosphorus deposition to the Amazon, Tropical Atlantic Ocean, and Southern Ocean. Proc. Natl Acad. Sci. USA 116, 16216–16221 (2019).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  10. Guieu, C., Bonnet, S., Wagener, T. & Loÿe-Pilot, M.-D. Biomass burning as a source of dissolved iron to the open ocean? Geophys. Res. Lett. 32, L19608 (2005).

    Article  ADS  Google Scholar 

  11. Ito, A. Mega fire emissions in Siberia: potential supply of bioavailable iron from forests to the ocean. Biogeosciences 8, 1679–1697 (2011).

    Article  ADS  CAS  Google Scholar 

  12. Abram, N. J., Gagan, M. K., McCulloch, M. T., Chappell, J. & Hantoro, W. S. Coral reef death during the 1997 Indian Ocean Dipole linked to Indonesian wildfires. Science 301, 952–955 (2003).

    Article  ADS  CAS  PubMed  Google Scholar 

  13. Ito, A. et al. Pyrogenic iron: the missing link to high iron solubility in aerosols. Sci. Adv. 5, eaau7671 (2019).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  14. Jia, G. et al. in Climate Change and Land: an IPCC Special Report on Climate Change, Desertification, Land Degradation, Sustainable Land Management, Food Security, and Greenhouse Gas Fluxes in Terrestrial Ecosystems Ch. 2 (IPCC, in the press).

  15. Jiang, Y. et al. Impacts of wildfire aerosols on global energy budget and climate: the role of climate feedbacks. J. Clim. 33, 3351–3366 (2020).

    Article  ADS  Google Scholar 

  16. Bowman, D. et al. Wildfires: Australia needs national monitoring agency. Nature 584, 188–191 (2020).

    Article  ADS  CAS  PubMed  Google Scholar 

  17. New WWF report: 3 billion animals impacted by Australia’s bushfire crisis. WWF https://www.wwf.org.au/news/news/2020/3-billion-animals-impacted-by-australia-bushfire-crisis#gs.ebzve2 (2020).

  18. van der Velde, I. R. et al. Vast CO2 release from Australian fires in 2019–2020 constrained by satellite. Nature https://doi.org/10.1038/s41586-021-03712-y (2021).

  19. National Greenhouse Gas Inventory Report: 2018 (Australian Government, 2020); https://www.industry.gov.au/data-and-publications/national-greenhouse-gas-inventory-report-2018.

  20. Mahowald, N. M. et al. Aerosol impacts on climate and biogeochemistry. Annu. Rev. Environ. Res. 36, 45–74 (2011).

    Article  Google Scholar 

  21. Boyd, P. W. et al. Mesoscale iron enrichment experiments 1993–2005: synthesis and future directions. Science 315, 612–617 (2007).

    Article  ADS  CAS  Google Scholar 

  22. Jickells, T. et al. Global iron connections between desert dust, ocean biogeochemistry, and climate. Science 308, 67–71 (2005).

    Article  ADS  CAS  PubMed  Google Scholar 

  23. Martin, J. H. Glacial‐interglacial CO2 change: the iron hypothesis. Paleoceanography 5, 1–13 (1990).

    Article  ADS  Google Scholar 

  24. Tagliabue, A. et al. Surface-water iron supplies in the Southern Ocean sustained by deep winter mixing. Nat. Geosci. 7, 314–320 (2014).

    Article  ADS  CAS  Google Scholar 

  25. Cassar, N. et al. The Southern Ocean biological response to aeolian iron deposition. Science 317, 1067–1070 (2007).

    Article  ADS  CAS  PubMed  Google Scholar 

  26. Gabric, A. J., Cropp, R., Ayers, G. P., McTainsh, G. & Braddock, R. Coupling between cycles of phytoplankton biomass and aerosol optical depth as derived from SeaWiFS time series in the Subantarctic Southern Ocean. Geophys. Res. Lett. 29, 16-11–16-14 (2002).

    Article  Google Scholar 

  27. Ardyna, M. et al. Hydrothermal vents trigger massive phytoplankton blooms in the Southern Ocean. Nat. Commun. 10, 2451 (2019).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  28. Duprat, L. P. A. M., Bigg, G. R. & Wilton, D. J. Enhanced Southern Ocean marine productivity due to fertilization by giant icebergs. Nat. Geosci. 9, 219–221 (2016).

    Article  ADS  CAS  Google Scholar 

  29. Bixby, R. J. et al. Fire effects on aquatic ecosystems: an assessment of the current state of the science. Freshwater Sci. 34, 1340–1350 (2015).

    Article  Google Scholar 

  30. Inness, A. et al. The CAMS reanalysis of atmospheric composition. Atmos. Chem. Phys. 19, 3515–3556 (2019).

    Article  ADS  CAS  Google Scholar 

  31. Shafeeque, M., Sathyendranath, S., George, G., Balchand, A. N. & Platt, T. Comparison of seasonal cycles of phytoplankton chlorophyll, aerosols, winds and sea-surface temperature off Somalia. Front. Marine Sci. 4, 384 (2017).

    Article  Google Scholar 

  32. Cassar, N. et al. The influence of iron and light on net community production in the Subantarctic and Polar Frontal zones. Biogeosciences 8, 227–237 (2011).

    Article  ADS  CAS  Google Scholar 

  33. Mitchell, B. G. & Holm-Hansen, O. Observations of modeling of the Antartic phytoplankton crop in relation to mixing depth. Deep Sea Res. Part A 38, 981–1007 (1991).

    Article  ADS  CAS  Google Scholar 

  34. Longo, A. F. et al. Influence of atmospheric processes on the solubility and composition of iron in Saharan dust. Environ. Sci. Technol. 50, 6912–6920 (2016).

    Article  ADS  CAS  PubMed  Google Scholar 

  35. Meskhidze, N., Nenes, A., Chameides, W. L., Luo, C. & Mahowald, N. Atlantic Southern Ocean productivity: fertilization from above or below? Global Biogeochem. Cycles 21, GB2006 (2007).

    Article  ADS  Google Scholar 

  36. Sarmiento, J. L., Slater, R. D., Dunne, J., Gnanadesikan, A. & Hiscock, M. R. Efficiency of small scale carbon mitigation by patch iron fertilization. Biogeosciences 7, 3593–3624 (2010).

    Article  ADS  CAS  Google Scholar 

  37. Brzezinski, M. A., Jones, J. L. & Demarest, M. S. Control of silica production by iron and silicic acid during the Southern Ocean Iron Experiment (SOFeX). Limnol. Oceanogr. 50, 810–824 (2005).

    Article  ADS  CAS  Google Scholar 

  38. Lovenduski, N. S. & Gruber, N. Impact of the Southern Annular Mode on Southern Ocean circulation and biology. Geophys. Res. Lett. 32, L11603 (2005).

    Article  ADS  Google Scholar 

  39. Cai, W., Cowan, T. & Raupach, M. Positive Indian Ocean Dipole events precondition southeast Australia bushfires. Geophys. Res. Lett. 36, L19710 (2009).

    Article  ADS  Google Scholar 

  40. Chen, Y. et al. A pan-tropical cascade of fire driven by El Niño/Southern Oscillation. Nat. Climate Change 7, 906–911 (2017).

    Article  ADS  CAS  Google Scholar 

  41. Lim, E.-P. et al. Australian hot and dry extremes induced by weakenings of the stratospheric polar vortex. Nat. Geosci. 12, 896–901 (2019).

    Article  ADS  CAS  Google Scholar 

  42. Cai, W. et al. Increased frequency of extreme Indian Ocean Dipole events due to greenhouse warming. Nature 510, 254–258 (2014).

    Article  ADS  CAS  PubMed  Google Scholar 

  43. Cropp, R. A. et al. The likelihood of observing dust-stimulated phytoplankton growth in waters proximal to the Australian continent. J. Mar. Syst. 117–118, 43–52 (2013).

    Article  Google Scholar 

  44. Hamilton, D. S. et al. Impact of changes to the atmospheric soluble iron deposition flux on ocean biogeochemical cycles in the anthropocene. Global Biogeochem. Cycles 34, e2019GB006448 (2020).

    Article  ADS  CAS  Google Scholar 

  45. Duce, R. et al. Impacts of atmospheric anthropogenic nitrogen on the open ocean. Science 320, 893–897 (2008).

    Article  ADS  CAS  PubMed  Google Scholar 

  46. Han, Y. et al. Asian inland wildfires driven by glacial-interglacial climate change. Proc. Natl Acad. Sci. USA 117, 5184–5189 (2020).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  47. van der Werf, G. R. et al. Global fire emissions estimates during 1997–2016. Earth Sys. Sci. Data 9, 697–720 (2017).

    Article  ADS  Google Scholar 

  48. Orsi, A. H., Whitworth, T. & Nowlin, W. D. On the meridional extent and fronts of the Antarctic Circumpolar Current. Deep Sea Res. Part I 42, 641–673 (1995).

    Article  Google Scholar 

  49. Sathyendranath, S. et al. An ocean-colour time series for use in climate studies: the experience of the Ocean-Colour Climate Change Initiative (OC-CCI). Sensors 19, 4285 (2019).

    Article  ADS  CAS  PubMed Central  Google Scholar 

  50. Morcrette, J.-J. et al. Aerosol analysis and forecast in the European Centre for Medium-Range Weather Forecasts Integrated Forecast System: forward modeling. J. Geophys. Res. Atmospheres 114, D06206 (2009).

    Article  ADS  Google Scholar 

  51. Levy, R. C. et al. Exploring systematic offsets between aerosol products from the two MODIS sensors. Atmos. Meas. Tech. 11, 4073–4092 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Benedetti, A. et al. Aerosol analysis and forecast in the European Centre for Medium-Range Weather Forecasts Integrated Forecast System: 2. Data assimilation. J. Geophys. Res. 114, D13 (2009).

    Google Scholar 

  53. Kaiser, J. W. et al. Biomass burning emissions estimated with a global fire assimilation system based on observed fire radiative power. Biogeosciences 9, 527–554 (2012).

    Article  ADS  CAS  Google Scholar 

  54. Y. Bennouna et al. Validation Report of the CAMS Global Reanalysis of Aerosols and Reactive Gases, Years 2003–2019 (Copernicus Atmosphere Monitoring Service, 2020).

  55. Ito, A. et al. Evaluation of aerosol iron solubility over Australian coastal regions based on inverse modeling: implications of bushfires on bioaccessible iron concentrations in the Southern Hemisphere. Prog. Earth Planet. Sci. 7, 42 (2020).

    Article  ADS  Google Scholar 

  56. Khaykin, S. et al. The 2019/20 Australian wildfires generated a persistent smoke-charged vortex rising up to 35 km altitude. Commun. Earth Environ. 1, 22 (2020).

  57. Haëntjens, N., Boss, E. & Talley, L. D. Revisiting Ocean Color algorithms for chlorophyll a and particulate organic carbon in the Southern Ocean using biogeochemical floats. J. Geophys. Res. Oceans 122, 6583–6593 (2017).

    Article  ADS  Google Scholar 

  58. Boss, E. et al. The characteristics of particulate absorption, scattering and attenuation coefficients in the surface ocean; contribution of the Tara Oceans expedition. Methods Oceanogr. 7, 52–62 (2013).

    Article  Google Scholar 

  59. de Boyer Montégut, C., Madec, G., Fischer, A. S., Lazar, A. & Iudicone, D. Mixed layer depth over the global ocean: an examination of profile data and a profile‐based climatology. J. Geophys. Res. Oceans 109, C12003 (2004).

    Article  ADS  Google Scholar 

  60. Dong, S., Sprintall, J., Gille, S. T. & Talley, L. Southern Ocean mixed-layer depth from Argo float profiles. J. Geophys. Res. Oceans 113, C06013 (2008).

    Article  ADS  Google Scholar 

  61. Cutter, G. A. et al. Sampling and Sample-handling Protocols for GEOTRACES Cruises, version 3.0 (2017).

    Google Scholar 

  62. Morton, P. L. et al. Methods for the sampling and analysis of marine aerosols: results from the 2008 GEOTRACES aerosol intercalibration experiment. Limnol. Oceanogr. Methods 11, 62–78 (2013).

    Article  CAS  Google Scholar 

  63. Perron, M. M. G. et al. Assessment of leaching protocols to determine the solubility of trace metals in aerosols. Talanta 208, 120377 (2020).

    Article  CAS  PubMed  Google Scholar 

  64. Shelley, R. U., Landing, W. M., Ussher, S. J., Planquette, H. & Sarthou, G. Regional trends in the fractional solubility of Fe and other metals from North Atlantic aerosols (GEOTRACES cruises GA01 and GA03) following a two-stage leach. Biogeosciences 15, 2271–2288 (2018).

    Article  ADS  CAS  Google Scholar 

  65. Sanz Rodriguez, E. et al. Analysis of levoglucosan and its isomers in atmospheric samples by ion chromatography with electrospray lithium cationisation—triple quadrupole tandem mass spectrometry. J. Chromatogr. A 1610, 460557 (2020).

    Article  CAS  PubMed  Google Scholar 

  66. McLennan, S. M. Relationships between the trace element composition of sedimentary rocks and upper continental crust. Geochem. Geophys. Geosyst. 2, 1201 (2001).

    Article  Google Scholar 

  67. Shelley, R. U. et al. Quantification of trace element atmospheric deposition fluxes to the Atlantic Ocean (>40°N; GEOVIDE, GEOTRACES GA01) during spring 2014. Deep Sea Res. Part I 119, 34–49 (2017).

    Article  CAS  Google Scholar 

  68. Sholkovitz, E. R., Sedwick, P. N., Church, T. M., Baker, A. R. & Powell, C. F. Fractional solubility of aerosol iron: synthesis of a global-scale data set. Geochim. Cosmochim. Acta 89, 173–189 (2012).

    Article  ADS  CAS  Google Scholar 

  69. Stein, A. F. et al. NOAA’s HYSPLIT atmospheric transport and dispersion modeling system. Bull. Am. Meteorol. Soc. 96, 2059–2077 (2016).

    Article  ADS  Google Scholar 

  70. Kalnay, E. et al. The NCEP/NCAR 40-year reanalysis project. Bull. Am. Meteorol. Soc. 77, 437–471 (1996).

    Article  ADS  Google Scholar 

  71. Tatlhego, M., Bhattachan, A., Okin, G. S. & D’Odorico, P. Mapping areas of the Southern Ocean where productivity likely depends on dust‐delivered Iron. J. Geophys. Res. Atmospheres 125, e2019JD030926 (2020).

    Article  ADS  CAS  Google Scholar 

  72. Stein, A. F., Rolph, G. D., Draxler, R. R., Stunder, B. & Ruminski, M. Verification of the NOAA smoke forecasting system: model sensitivity to the injection height. Weather Forecast. 24, 379–394 (2009).

    Article  ADS  Google Scholar 

  73. Behrenfeld, M. J. & Falkowski, P. G. Photosynthetic rates derived from satellite‐based chlorophyll concentration. Limnol. Oceanogr. 42, 1–20 (1997).

    Article  ADS  CAS  Google Scholar 

  74. Behrenfeld, M. J., Boss, E., Siegel, D. A. & Shea, D. M. Carbon-based ocean productivity and phytoplankton physiology from space. Global Biogeochem. Cycles 19, GB1006 (2005).

    Article  ADS  Google Scholar 

  75. Westberry, T., Behrenfeld, M. J., Siegel, D. A. & Boss, E. Carbon-based primary productivity modeling with vertically resolved photoacclimation. Global Biogeochem. Cycles 22, GB2024 (2008).

    Article  ADS  Google Scholar 

  76. Silsbe, G. M., Behrenfeld, M. J., Halsey, K. H., Milligan, A. J. & Westberry, T. K. The CAFE model: a net production model for global ocean phytoplankton. Global Biogeochem. Cycles 30, 1756–1777 (2016).

    Article  ADS  CAS  Google Scholar 

  77. Laws, E. A., D’Sa, E. & Naik, P. Simple equations to estimate ratios of new or export production to total production from satellite‐derived estimates of sea surface temperature and primary production. Limnol. Oceanogr. Methods 9, 593–601 (2011).

    Article  Google Scholar 

  78. Dunne, J. P., Armstrong, R. A., Gnanadesikan, A. & Sarmiento, J. L. Empirical and mechanistic models for the particle export ratio. Global Biogeochem. Cycles 19, GB4026 (2005).

    Article  ADS  Google Scholar 

  79. Li, Z. & Cassar, N. Satellite estimates of net community production based on O2/Ar observations and comparison to other estimates. Global Biogeochem. Cycles 30, 735–752 (2016).

    Article  ADS  CAS  Google Scholar 

  80. Siegel, D. A. et al. Global assessment of ocean carbon export by combining satellite observations and food‐web models. Global Biogeochem. Cycles 28, 181–196 (2014).

    Article  ADS  CAS  Google Scholar 

  81. Marshall, G. J. Trends in the Southern Annular Mode from observations and reanalyses. J. Climate 16, 4134–4143 (2003).

    Article  ADS  Google Scholar 

  82. Saji, N. H. & Yamagata, T. Possible impacts of Indian Ocean Dipole mode events on global climate. Climate Res. 25, 151–169 (2003).

    Article  ADS  Google Scholar 

Download references

Acknowledgements

Analyses of satellite aerosol observations used in this study were produced with the Giovanni online data system, developed and maintained by the NASA GES DISC. We thank SeaWiFS and MODIS mission scientists and associated NASA personnel for the production of the data used in this research effort. The BGC-Argo data were collected and made freely available by the International Argo Program and the national programs that contribute to it (http://www.argo.ucsd.edu, http://argo.jcommops.org). The Argo Program is part of the Global Ocean Observing System (https://doi.org/10.17882/42182). W.T. is supported by the Harry H. Hess Postdoctoral Fellowship from Princeton University. N.C. is supported by the “Laboratoire d’Excellence” LabexMER (ANR‐10‐LABX‐19) and co-funded by a grant from the French government under the program “Investissements d’Avenir”. S.B. acknowledges the AXA Research Fund for the support of the long-term research line on Sand and Dust Storms at the Barcelona Supercomputing Center (BSC) and CAMS Global Validation (CAMS-84). P.G.S., J.L., M.M.G.P. and A.R.B. are supported by the Australian Research Council Discovery Projects scheme (DP190103504). P.G.S. and J.W. are supported by the Australian Research Council Centre of Excellence for Climate Extremes (CLEX: CE170100023). J.L. is supported by the European Union’s Horizon 2020 research and innovation programme under the Marie Skłodowska-Curie grant agreement no. 754433. A.R.B. is supported by the Australian Research Council Future Fellowship scheme (FT130100037). R.M. is supported by the CSIRO Decadal Climate Forecasting Project. We thank M. Strzelec, M. East, T. Holmes, M. Corkill, S. Meyerink and the Wellington Park Management Trust for help with installation and sampling the Tasmanian aerosol time-series station; A. Townsend for iron aerosol analyses by ICPMS at the University of Tasmania; and A. Benedetti and S. Remy for providing insights on the validation of aerosol reanalysis.

Author information

Authors and Affiliations

Authors

Contributions

This study was conceived by N.C., J.L. and R.M. W.T. and N.C. wrote the manuscript with contribution from co-authors. J.L. and W.T. analysed the spatial distribution and time-series of AOD, aerosol deposition and [Chla], and coordinated the interdisciplinary approach. J.W., C.S. and P.G.S. conducted the analysis of BGC-Argo float observations. S.B. and J.L. conducted the AOD decomposition reanalysis. Z.L. calculated MLD from Argo floats and estimated marine production with W.T. S.S. and T.J. provided and helped with interpretation of satellite observations of [Chla]. M.M.G.P., B.C.P. and A.R.B. collected the aerosol samples and analysed the aerosol Fe content and solubility. E.S.R. analysed levoglucosan in the aerosol samples. All authors contributed to the interpretation of the results.

Corresponding authors

Correspondence to Richard Matear or Nicolas Cassar.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature thanks the anonymous reviewers 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 Emission and deposition of aerosols and response of phytoplankton.

a, Monthly aerosol optical depth (AOD) at 550 nm observed by MODIS satellite during the 2019–2020 Australian wildfires from November 2019 to February 2020. b, Monthly chlorophyll-a absolute anomaly. c, Monthly cumulative black carbon aerosol deposition. Contour lines indicate the monthly maximum black carbon AOD (black = 0.05, grey = 0.1, light-grey = 0.15). All: Subtropical, Subantarctic and Polar Fronts are indicated with a dotted, a dashed and a solid black line, respectively48.

Extended Data Fig. 2 Forward trajectories tracking the emission and transport of aerosols from major fire events during the 2019–2020 Australian wildfires.

a, Seven-day trajectories (grey lines) launched every 6 h and originated from wildfires during the period of 26 October to 4 November 2019 and black carbon AOD from the same period shown as the background map. The trajectory origins are depicted by black circles. The distribution of trajectories generally follows the black AOD pattern. b, Seven-day trajectories (grey lines) originated from wildfires during the period of 26 November to 5 December 2019. c, The spatial distribution of 7-day trajectory endpoints frequencies in 2° by 2° grid over the period of November 2019 to January 2020. The trajectory origins are depicted by black circles to represent the major fires’ locations. The 7-day air parcel forward trajectories were launched daily. The red contours depict regions where [Chla] more than doubled during the same period compared with their climatologies. The large [Chla] anomalies generally occurred in regions over which the trajectories passed.

Extended Data Fig. 3 Identification of regions of interest with potential aerosol fertilization.

a, Austral summer (DJF) 2019–2020 averaged chlorophyll-a relative anomaly map with cyan contour lines indicating where the anomaly is equal to 100%. b, Austral summer (DJF) 2019–2020 cumulative deposition of dust and black-carbon with black contour line indicating where deposition is equal to 150 mg m−2. c, Pixels where both [Chla] relative anomaly exceeds 100% and cumulative deposition exceeds 150 mg m−2 are marked in green. Black boxes indicate the South of Australia and Pacific Southern Ocean regions defined in this study. All: Subtropical, Subantarctic and Polar Fronts are indicated with a dotted, a dashed and a solid black line, respectively.

Extended Data Fig. 4 Large chlorophyll-a ([Chla]) anomaly in a big box region of the South Pacific and Southern Ocean during 2019–2020 Australian wildfires.

a, [Chla] anomaly map from December 2019 to February 2020 in comparison to their climatologies. A large portion of the ocean basin (solid black box) was selected to calculate [Chla] time-series. STF, Subtropical Front; SAF, Subantarctic Front. b, Time-series of average [Chla] in the selected box region. Monthly climatological values shown in solid black line. Red and blue areas denote monthly data higher or lower than climatological values, respectively. c, Monthly average [Chla] in individual years in the selected box region. Grey lines, historical years; solid black line, monthly climatologies; dashed black line, 2002 Australian wildfire season; red line, 2019–2020 Australian wildfire season.

Extended Data Fig. 5 Large chlorophyll-a ([Chla]) anomaly in numerous small box regions during 2019–2020 Australian wildfires.

a, [Chla] time-series was calculated in 4,681 of 10° by 10° boxes from 1997 to 2020 in the broad South Pacific and Southern Ocean (20° S–60° S; 120° E–80° W). Yellow circles and yellow dashed boxes are examples to show the center and coverage of each box region. Box moves by 1° eastward and southward sequentially illustrated by the black arrows. Box position 1, 151, 4,531 and 4,681 denoting the edge of the study region are shown as examples on the map of annual [Chla] climatology. The ratio of monthly [Chla] to its monthly climatology is calculated for each 10° by 10° box starting from October 1997 to May 2020. Black circles: centre locations of 10° by 10° boxes where \(\frac{{\rm{m}}{\rm{o}}{\rm{n}}{\rm{t}}{\rm{h}}{\rm{l}}{\rm{y}}[{\rm{C}}{\rm{h}}{\rm{l}}{\rm{a}}]}{[{\rm{C}}{\rm{h}}{\rm{l}}{\rm{a}}]{\rm{c}}{\rm{l}}{\rm{i}}{\rm{m}}{\rm{a}}{\rm{t}}{\rm{o}}{\rm{l}}{\rm{o}}{\rm{g}}{\rm{y}}} > 2.5\) before the 2019–2020 wildfires (from October 1997 to August 2019); red circles: centre locations of 10° by 10° boxes where \(\frac{{\rm{monthly}}[{\rm{Chla}}]}{[{\rm{Chla}}]{\rm{climatology}}} > 2.5\) during or after the 2019–2020 wildfires (from September 2019 to May 2020). Historically, regions with a large anomaly (black circles) are mostly located in coastal waters (for example, east coast of Australia). In contrast, during the 2019–2020 Australian wildfires (red circles), large areas of the open ocean show a high [Chla] anomaly (for example, south of Australia and Pacific sector of the Southern Ocean). Oceanic [Chla] anomalies of this magnitude are unprecedented in the historical record. Some of the black and red circles are on land because a fraction of the 10° by 10° box around these circles covers the ocean. b, Ratio of monthly [Chla] to its corresponding monthly climatologies for each box region from 1997 to 2020. c, Frequency distributions of the monthly [Chla] to monthly climatology ratios over the historical and 2019–2020 austral summers.

Extended Data Fig. 6 Maps of bbp anomalies and comparison between calibrated and uncalibrated BGC Argo in situ bbp measurements.

a, Satellite backscatter bbp relative anomaly for the 2019–2020 austral summer. Bloom region and BGC-Argo float trajectories superimposed on the map. Float positions from September 2019 through March 2020 highlighted. The southern float (red) was in a biologically more active region of the bloom than the two northern floats (blue and yellow). This corroborates the stronger bloom signal shown by the southern float. Dotted, dot-dashed, and solid lines in a represent the climatological positions of the Subtropical Front, Subantarctic Front and Polar Front, respectively48. b, Satellite bbp averaged over two sub-regions encompassing the float paths. The solid lines are 2019–2020 observations and the dotted lines with coloured standard deviation envelopes are the climatology. This analysis corroborates the stronger bloom signal shown by the southern float compared with the two northern floats. c, Comparison between uncalibrated (dashed lines) and calibrated (solid lines) in situ bbp measured by the three BGC-Argo floats. Surface bbp estimates were calculated as the median bbp between 0 and 20 m depth and then calibrated using a linear regression (see Methods for details). The calibration was applied to allow for comparison between float bbp and the satellite-based climatology. The general trend of the float signal is not altered by the calibration.

Extended Data Fig. 7 Iron (Fe) concentration and origin of aerosols collected at an aerosol time-series sampling station in Tasmania during the 2019–2020 Australian wildfires.

a, Total Fe concentration (blue line) during the 2019–2020 Australian wildfire season is compared with the historical median value from 2016–2019 (dashed black line). High levoglucosan concentration (green bar) indicates wildfire-derived aerosols. Grey shaded areas represent samples influenced by anthropogenic sources. See Methods for the use of tracers to track the sources of aerosols. b, Labile Fe concentration (blue line) during the 2019–2020 Australian wildfire season is compared with the historical median value from 2016–2019 (dashed black line). The aerosols with high Fe content collected around 15 January 2020 are likely to have originated from wildfires, indicated by the high concentration of levoglucosan concentrations and low concentration of anthropogenic tracers.

Extended Data Fig. 8 Tracking the origins of aerosols with high iron content collected at an aerosol time-series sampling station in Tasmania during 15–17 January 2020.

a, High black carbon AOD plume passing the sampling station (cyan star). b, Five-day backward trajectories were launched every 6 h from the sampling station (cyan star) during 15–17 January 2020. Both the distribution of trajectories and the frequency of trajectories’ endpoints confirm that the majority of the aerosols came from southeastern Australia where wildfires were raging.

Extended Data Fig. 9 Anomalies in marine phytoplankton productivity during 2019–2020 Australian wildfires.

a, b, Net primary production (NPP) (a) and export production (EP) (b) anomalies in 2019–2020 austral summer relative to their climatologies. Black boxes denote the basin-scale regions (20° S–55° S, 120° E–90° W) used to estimate changes in marine production during the 2019–2020 Australian wildfires.

Extended Data Fig. 10 Relations of large-scale climate patterns to the occurrence of wildfires and to chlorophyll a distribution.

ac, Time-series of climate indices Indian Ocean Dipole (IOD) (a), Southern Annular Mode (SAM) (b) and Oceanic Niño Index (c). Historical Australian mega-wildfire periods shaded in orange (>1 million hectares of land burned). d, [Chla] anomaly predicted by IOD index during the 2019–2020 Australian wildfires. e, [Chla] anomaly predicted by SAM index during the 2019–2020 Australian wildfires. The [Chla] anomaly potentially induced by the climate patterns are substantially smaller than the observed [Chla] anomaly (Fig. 1d).

Supplementary information

Supplementary Information

This file contains Supplementary Discussion, Tables 1–3, Figs. 1–9, and additional references.

Supplementary Video 1

Animation of daily black carbon aerosol optical depth (BC AOD) derived from CAMS reanalysis during the 2019–2020 Australian wildfires.

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tang, W., Llort, J., Weis, J. et al. Widespread phytoplankton blooms triggered by 2019–2020 Australian wildfires. Nature 597, 370–375 (2021). https://doi.org/10.1038/s41586-021-03805-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-021-03805-8

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