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:

Chicxulub impact winter sustained by fine silicate dust

Abstract

The Chicxulub impact is thought to have triggered a global winter at the Cretaceous-Palaeogene (K-Pg) boundary 66 million years ago. Yet the climatic consequences of the various debris injected into the atmosphere following the Chicxulub impact remain unclear, and the exact killing mechanisms of the K-Pg mass extinction remain poorly constrained. Here we present palaeoclimate simulations based on sedimentological constraints from an expanded terrestrial K-Pg boundary deposit in North Dakota, United States, to evaluate the relative and combined effects of impact-generated silicate dust and sulfur, as well as soot from wildfires, on the post-impact climate. The measured volumetric size distribution of silicate dust suggests a larger contribution of fine dust (~0.8–8.0 μm) than previously appreciated. Our simulations of the atmospheric injection of such a plume of micrometre-sized silicate dust suggest a long atmospheric lifetime of 15yr, contributing to a global-average surface temperature falling by as much as 15°C. Simulated changes in photosynthetic active solar radiation support a dust-induced photosynthetic shut-down for almost 2 yr post-impact. We suggest that, together with additional cooling contributions from soot and sulfur, this is consistent with the catastrophic collapse of primary productivity in the aftermath of the Chicxulub impact.

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: Conceptual model of the Chicxulub impact plume showing different stages of production, transport and deposition of coarse and fine-grained impact-generated ejecta (not to scale).
Fig. 2: Volume-weighted K-Pg boundary grain-size data.
Fig. 3: Number-weighted K-Pg boundary grain-size data in relation to atmospheric settling processes.
Fig. 4: Temporal evolution of the Chicxulub impact-generated global climatic responses.
Fig. 5: Temporal evolution of PAR flux reconstructions after the Chicxulub impact.

Similar content being viewed by others

Data availability

The palaeoclimate general circulation model (GCM) output data as well as the silicate dust grain-size data from the Tanis K-Pg site that support the findings of this study are publicly available in the OSF repository via https://doi.org/10.17605/OSF.IO/2CDQG. The proxy-based latest Cretaceous temperature reconstruction data are available online in the PANGEA repository: https://doi.org/10.1594/PANGAEA.879763.

Code availability

The Python and Matlab source codes developed for reproducing the figures in this study are publicly available at the GitHub repository via github.com/cem-berk-senel/naturegeoscience-chicxulub/. The PlanetWRF model is available upon request from https://planetwrf.com/.

References

  1. Smit, J. & Hertogen, J. An extraterrestrial event at the Cretaceous–Tertiary boundary. Nature 285, 198–200 (1980).

    Article  Google Scholar 

  2. Alvarez, L. W. et al. Extraterrestrial cause for the Cretaceous–Tertiary extinction. Science 208, 1095–1108 (1980).

    Article  Google Scholar 

  3. Kring, D. A. The Chicxulub impact event and its environmental consequences at the Cretaceous–Tertiary boundary. Palaeogeogr. Palaeoclimatol. Palaeoecol. 255, 4–21 (2007).

    Article  Google Scholar 

  4. Schulte, P. et al. The Chicxulub asteroid impact and mass extinction at the Cretaceous–Paleogene boundary. Science 327, 1214–1218 (2010).

    Article  Google Scholar 

  5. Collins, G. S. et al. A steeply-inclined trajectory for the Chicxulub impact. Nat. Commun. 11, 1480 (2020).

    Article  Google Scholar 

  6. Goderis, S. et al. Globally distributed iridium layer preserved within the Chicxulub impact structure. Sci. Adv. 7, eabe3647 (2021).

    Article  Google Scholar 

  7. Chiarenza, A. A. et al. Asteroid impact, not volcanism, caused the end-Cretaceous dinosaur extinction. Proc. Natl Acad. Sci. USA 117, 17084–17093 (2020).

    Article  Google Scholar 

  8. Pierazzo, E., Kring, D. A. & Melosh, H. J. Hydrocode simulation of the Chicxulub impact event and the production of climatically active gases. J. Geophys. Res. Planets 103, 28607–28625 (1998).

    Article  Google Scholar 

  9. Alvarez, W., Claeys, P. & Kieffer, S. W. Emplacement of Cretaceous–Tertiary boundary shocked quartz from Chicxulub crater. Science 269, 930–935 (1995).

    Article  Google Scholar 

  10. Morgan, J. V. et al. The Chicxulub impact and its environmental consequences. Nature Rev. Earth Environ. 3, 338–354 (2022).

  11. Morgan, J. V. et al. The formation of peak rings in large impact craters. Science 354, 878–882 (2016).

    Article  Google Scholar 

  12. López-Ramos, E. in Ocean Basins and Margins, the Gulf of Mexico and Caribbean (eds Nairn, A. E. M. & Stehli, F. G.) 257–282 (Plenum Press, 1975).

  13. Robertson, D. S. et al. K–Pg extinction patterns in marine and freshwater environments: the impact winter model. J. Geophys. Res. Biogeosci. 118, 1006–1014 (2013).

    Article  Google Scholar 

  14. Vellekoop, J. et al. Rapid short-term cooling following the Chicxulub impact at the Cretaceous–Paleogene boundary. Proc. Natl Acad. Sci. USA 111, 7537–7541 (2014).

    Article  Google Scholar 

  15. Brugger, J., Feulner, G. & Petri, S. Baby, it’s cold outside: climate model simulations of the effects of the asteroid impact at the end of the Cretaceous. Geophys. Res. Lett. 44, 419–427 (2017).

    Article  Google Scholar 

  16. Artemieva, N., Morgan, J. & Party, E. S. Quantifying the release of climate-active gases by large meteorite impacts with a case study of Chicxulub. Geophys. Res. Lett. 44, 10–180 (2017).

    Article  Google Scholar 

  17. Brugger, J. et al. A pronounced spike in ocean productivity triggered by the Chicxulub impact. Geophys. Res. Lett. 48, e2020GL092260 (2021).

    Article  Google Scholar 

  18. Tabor, C. R. et al. Causes and climatic consequences of the impact winter at the Cretaceous–Paleogene boundary. Geophys. Res. Lett. 47, e60121 (2020).

    Article  Google Scholar 

  19. Senel, C. B. et al. Relative roles of impact-generated aerosols on photosynthetic activity following the Chicxulub asteroid impact. GSA Connects 53, 6 (2021).

    Google Scholar 

  20. Wolbach, W. S., Lewis, R. S. & Anders, E. Cretaceous extinctions: evidence for wildfires and search for meteoritic material. Science 230, 167–170 (1985).

    Article  Google Scholar 

  21. Belcher, C. M. et al. Geochemical evidence for combustion of hydrocarbons during the KT impact event. Proc. Natl Acad. Sci. USA 106, 4112–4117 (2009).

    Article  Google Scholar 

  22. Kaiho, K. et al. Global climate change driven by soot at the K–Pg boundary as the cause of the mass extinction. Sci. Rep. 6, 28427 (2016).

    Article  Google Scholar 

  23. Bardeen, C. G. et al. On transient climate change at the Cretaceous-Paleogene boundary due to atmospheric soot injections. Proc. Natl Acad. Sci. USA 114, E7415–E7424 (2017).

    Article  Google Scholar 

  24. Sharpton, V. et al. in Global Catastrophes in Earth History (eds Virgil L. Sharpton, V. L. & Ward, P. D.) 349–357 (GSA, 1990).

  25. Kaskes, P. et al. High-resolution chemostratigraphy of the Cretaceous-Paleogene (K-Pg) boundary interval in the US western interior: Implications for Chicxulub impact ejecta dynamics. In 53rd Lunar and Planetary Science Conference, Vol. 2678, 2708 (2022).

  26. Henehan, M. J. et al. Rapid ocean acidification and protracted Earth system recovery followed the end-Cretaceous Chicxulub impact. Proc. Natl Acad. Sci. USA 116, 22500–22504 (2019).

    Article  Google Scholar 

  27. Pope, K. O. Impact dust not the cause of the Cretaceous–Tertiary mass extinction. Geology 30, 99–102 (2002).

    Article  Google Scholar 

  28. Pope, K. O. et al. Impact winter and the Cretaceous/Tertiary extinctions: results of a Chicxulub asteroid impact model. Earth Planet. Sci. Lett. 128, 719–725 (1994).

    Article  Google Scholar 

  29. Pierazzo, E., Hahmann, A. N. & Sloan, L. C. Chicxulub and climate: radiative perturbations of impact-produced S-bearing gases. Astrobiology 3, 99–118 (2003).

    Article  Google Scholar 

  30. Wolbach, W. S. et al. in Global Catastrophes in Earth History; An Interdisciplinary Conference on Impacts, Volcanism, and Mass Mortality Vol. 247 (eds Sharpton, V. L. & Ward, P. D.) 219–220 (GSA, 1990).

  31. Kring, D. A. & Durda, D. D. Trajectories and distribution of material ejected from the Chicxulub impact crater: implications for postimpact wildfires. J. Geophys. Res. Planets https://doi.org/10.1029/2001JE001532 (2002).

  32. Morgan, J., Artemieva, N. & Goldin, T. Revisiting wildfires at the K–Pg boundary. J. Geophys. Res. Biogeosci. 118, 1508–1520 (2013).

    Article  Google Scholar 

  33. Goldin, T. J. & Melosh, H. J. Self-shielding of thermal radiation by Chicxulub impact ejecta: firestorm or fizzle? Geology 37, 1135–1138 (2009).

    Article  Google Scholar 

  34. Harvey, M. C. et al. Combustion of fossil organic matter at the Cretaceous–Paleogene (KP) boundary. Geology 36, 355–358 (2008).

    Article  Google Scholar 

  35. Lyons, S. L. et al. Organic matter from the Chicxulub crater exacerbated the K–Pg impact winter. Proc. Natl Acad. Sci. USA 117, 25327–25334 (2020).

    Article  Google Scholar 

  36. Smit, J. The global stratigraphy of the Cretaceous–Tertiary boundary impact ejecta. Annu. Rev. Earth Planet. Sci. 27, 75–113 (1999).

    Article  Google Scholar 

  37. Bostwick, J. A. & Kyte, F. T. in The Cretaceous–Tertiary Event and Other Catastrophes in Earth History (eds Ryder, G., Fastovsky, D. E. & Gartner, S.) 403–415 (GSA, 1996).

  38. During, M. A. D. et al. The Mesozoic terminated in boreal spring. Nature 603, 91–94 (2022).

    Article  Google Scholar 

  39. Toon, O. B., Bardeen, C. & Garcia, R. Designing global climate and atmospheric chemistry simulations for 1 and 10 km diameter asteroid impacts using the properties of ejecta from the K–Pg impact. Atmos. Chem. Phys. 16, 13185–13212 (2016).

    Article  Google Scholar 

  40. Belza, J. et al. Petrography and geochemistry of distal spherules from the K–Pg boundary in the Umbria–Marche region (Italy) and their origin as fractional condensates and melts in the Chicxulub impact plume. Geochim. Cosmochim. Acta 202, 231–263 (2017).

    Article  Google Scholar 

  41. DePalma, R. A. et al. A seismically induced onshore surge deposit at the KPg boundary, North Dakota. Proc. Natl Acad. Sci. USA 116, 8190–8199 (2019).

    Article  Google Scholar 

  42. Markwick, P. J. & Valdes, P. J. Palaeo-digital elevation models for use as boundary conditions in coupled ocean–atmosphere GCM experiments: a Maastrichtian (late Cretaceous) example. Palaeogeogr. Palaeoclimatol. Palaeoecol. 213, 37–63 (2004).

    Article  Google Scholar 

  43. Upchurch, G. R. et al. Latitudinal temperature gradients and high-latitude temperatures during the latest Cretaceous: congruence of geologic data and climate models. Geology 43, 683–686 (2015).

    Article  Google Scholar 

  44. O’Brien, C. L. et al. Cretaceous sea-surface temperature evolution: constraints from TEX86 and planktonic foraminiferal oxygen isotopes. Earth Sci. Rev. 172, 224–247 (2017).

    Article  Google Scholar 

  45. Niezgodzki, I. et al. Late Cretaceous climate simulations with different CO2 levels and subarctic gateway configurations: a model–data comparison. Paleoceanography 32, 980–998 (2017).

    Article  Google Scholar 

  46. Bralower, T. et al. Grain size of Cretaceous–Paleogene boundary sediments from Chicxulub to the open ocean: implications for interpretation of the mass extinction event. Geology 38, 199–202 (2010).

    Article  Google Scholar 

  47. Vellekoop, J. et al. Type-Maastrichtian gastropod faunas show rapid ecosystem recovery following the Cretaceous–Palaeogene boundary catastrophe. Palaeontology 63, 349–367 (2020).

    Article  Google Scholar 

  48. Donovan, M. P. et al. Rapid recovery of Patagonian plant–insect associations after the end-Cretaceous extinction. Nat. Ecol. Evol. 1, 0012 (2016).

    Article  Google Scholar 

  49. Belza, J. Petrography and geochemistry of ejecta material from the K-Pg boundary Chicxulub crater (Yucatan, Mexico). PhD thesis, Vrije Universiteit Brussel (2015).

  50. Pierazzo, E. & Artemieva, N. Local and global environmental effects of impacts on Earth. Elements 8, 55–60 (2012).

    Article  Google Scholar 

  51. Toonen, W. H. J. et al. Lower Rhine historical flood magnitudes of the last 450 years reproduced from grain-size measurements of flood deposits using end member modelling. Catena 130, 69–81 (2015).

    Article  Google Scholar 

  52. Konert, M. & Vandenberghe, J. E. F. 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  Google Scholar 

  53. Mahowald, N. et al. The size distribution of desert dust aerosols and its impact on the Earth system. Aeolian Res. 15, 53–71 (2014).

    Article  Google Scholar 

  54. Haberle, R. M. et al. Documentation of the NASA/Ames Legacy Mars Global Climate Model: simulations of the present seasonal water cycle. Icarus 333, 130–164 (2019).

    Article  Google Scholar 

  55. Li, J. et al. Accounting for dust aerosol size distribution in radiative transfer. J. Geophys. Res. Atmos. 120, 6537–6550 (2015).

    Article  Google Scholar 

  56. Leschonski, K. Representation and evaluation of particle size analysis data. Part. Part. Syst. Charact. 1, 89–95 (1984).

    Article  Google Scholar 

  57. Richardson, M. I., Toigo, A. D. & Newman, C. E. PlanetWRF: a general purpose, local to global numerical model for planetary atmospheric and climate dynamics. J. Geophys. Res. Planets 112, E09001 (2007).

    Article  Google Scholar 

  58. Newman, C. E. et al. Simulating Titan’s methane cycle with the TitanWRF general circulation model. Icarus 267, 106–134 (2016).

    Article  Google Scholar 

  59. Lee, C. et al. The sensitivity of solsticial pauses to atmospheric ice and dust in the MarsWRF general circulation model. Icarus 311, 23–34 (2018).

    Article  Google Scholar 

  60. Temel, O. et al. Large eddy simulations of the Martian convective boundary layer: towards developing a new planetary boundary layer scheme. Atmos. Res. 250, 105381 (2021).

    Article  Google Scholar 

  61. Senel, C. B. et al. Interannual, seasonal and regional variations in the Martian convective boundary layer derived from GCM simulations with a semi-interactive dust transport model. J. Geophys. Res. Planets 126, e2021JE006965 (2021).

    Article  Google Scholar 

  62. Skamarock, W. C. et al. A Description of the Advanced Research WRF Version 3 (No. NCAR/TN-475 + STR) (University Corporation for Atmospheric Research. 2008).

  63. Chou, M.-D. & Suarez, M. J. A Solar Radiation Parameterization for Atmospheric Studies Technical Report 104606, Vol. 15 (NASA, 1999).

  64. Chou, M.-D. et al. A Thermal Infrared Radiation Parameterization for Atmospheric Studies Technical Report 104606, Vol. 19 (NASA, 2001).

  65. Feichter, J. et al. Simulation of the tropospheric sulfur cycle in a global climate model. Atmos. Environ. 30, 1693–1707 (1996).

    Article  Google Scholar 

  66. Feng, Q., Cui, S. & Zhao, W. Effect of particle shape on dust shortwave direct radiative forcing calculations based on MODIS observations for a case study. Adv. Atmos. Sci. 32, 1266–1276 (2015).

    Article  Google Scholar 

  67. Dufresne, J.-L. et al. Longwave scattering effects of mineral aerosols. J. Atmos. Sci. 59, 1959–1966 (2002).

    Article  Google Scholar 

  68. Hess, M., Koepke, P. & Schult, I. Optical properties of aerosols and clouds: the software package OPAC. Bull. Am. Meteorol. Soc. 79, 831–844 (1998).

    Article  Google Scholar 

  69. Binkowski, F. S. & Shankar, U. The regional particulate matter model: 1. Model description and preliminary results. J. Geophys. Res. Atmos. 100, 26191–26209 (1995).

    Article  Google Scholar 

  70. Zhang, L., Gong, S., Padro, J. & Barrie, L. A size-segregated particle dry deposition scheme for an atmospheric aerosol module. Atmos. Environ. 35, 549–560 (2001).

    Article  Google Scholar 

  71. Zhang, J. & Shao, Y. A new parameterization of particle dry deposition over rough surfaces. Atmos. Chem. Phys. 14, 12429–12440 (2014).

    Article  Google Scholar 

  72. Emerson, E. W. et al. Revisiting particle dry deposition and its role in radiative effect estimates. Proc. Natl Acad. Sci. USA 117, 26076–26082 (2020).

    Article  Google Scholar 

  73. Xu, Y. & Carmichael, G. R. Modeling the dry deposition velocity of sulfur dioxide and sulfate in Asia. J. Appl. Meteorol. Climatol. 37, 1084–1099 (1998).

    Article  Google Scholar 

  74. Feichter, J., Brost, R. A. & Heimann, M. Three-dimensional modeling of the concentration and deposition of 210Pb aerosols. J. Geophys. Res. Atmos. 96, 22447–22460 (1991).

    Article  Google Scholar 

  75. Seinfeld, J. & Pandis, S. N. Atmospheric Chemistry and Physics: From Air Pollution to Climate Change (John Wiley & Sons, 2016).

  76. Tsarpalis, K. et al. The implementation of a mineral dust wet deposition scheme in the GOCART-AFWA module of the WRF model. Remote Sens. 10, 1595 (2018).

    Article  Google Scholar 

  77. Shao, Y. Simplification of a dust emission scheme and comparison with data. J. Geophys. Res. Atmos. 109, D10202 (2004).

    Article  Google Scholar 

  78. Shao, Y. et al. Parameterization of size-resolved dust emission and validation with measurements. J. Geophys. Res. Atmos. 116, D08203 (2011).

    Article  Google Scholar 

  79. Liu, X. et al. Toward a minimal representation of aerosols in climate models: description and evaluation in the Community Atmosphere Model CAM5. Geosci. Model Dev. 5, 709–739 (2012).

    Article  Google Scholar 

  80. Dudhia, J. A multilayer soil temperature model for MM5. In Sixth PSU/NCAR Mesoscale Model Users’ Workshop. 49–50 (1996).

  81. Chen, S.-H. & Sun, W.-Y. A one-dimensional time dependent cloud model. J. Meteorol. Soc. Jpn. 2 80, 99–118 (2002).

    Article  Google Scholar 

  82. Tiedtke, M. A comprehensive mass flux scheme for cumulus parameterization in large-scale models. Mon. Weather Rev. 117, 1779–1800 (1989).

    Article  Google Scholar 

  83. Zhang, C., Wang, Y. & Hamilton, K. Improved representation of boundary layer clouds over the southeast Pacific in ARW-WRF using a modified Tiedtke cumulus parameterization scheme. Mon. Weather Rev. 139, 3489–3513 (2011).

    Article  Google Scholar 

  84. Senel, C. B. et al. A new planetary boundary layer scheme based on LES: application to the XPIA campaign. J. Adv. Model. Earth Syst. 11, 2655–2679 (2019).

    Article  Google Scholar 

  85. Jimenez, P. A. et al. A revised scheme for the WRF surface layer formulation. Mon. Weather Rev. 140, 898–918 (2012).

    Article  Google Scholar 

  86. Pollard, R. T., Rhines, P. B. & Thompson, R. The deepening of the wind-mixed layer. Geophys. Fluid Dyn. 4, 381–404 (1973).

    Article  Google Scholar 

  87. Davis, C. et al. Prediction of landfalling hurricanes with the advanced hurricane WRF model. Mon. Weather Rev. 136, 1990–2005 (2008).

    Article  Google Scholar 

  88. Pierrehumbert, R. & Gaidos, E. Hydrogen greenhouse planets beyond the habitable zone. Astrophys. J. Lett. 734, L13 (2011).

    Article  Google Scholar 

  89. Su, W., Charlock, T. P., Rose, F. G. & Rutan, D. Photosynthetically active radiation from clouds and the Earth’s Radiant Energy System (CERES) products. J. Geophys. Res. Biogeosci. 112, G02022 (2007).

    Article  Google Scholar 

  90. García-Rodríguez, A. et al. Modelling photosynthetic active radiation (PAR) through meteorological indices under all sky conditions. Agric. For. Meteorol. 310, 108627 (2021).

    Article  Google Scholar 

  91. Hatzianastassiou, N. et al. Ten year radiation budget of the earth: 1984–93. Int. J. Climatol. 24, 1785–1802 (2004).

    Article  Google Scholar 

  92. Khoder, M. I. Atmospheric conversion of sulfur dioxide to particulate sulfate and nitrogen dioxide to particulate nitrate and gaseous nitric acid in an urban area. Chemosphere 49, 675–684 (2002).

    Article  Google Scholar 

  93. Loftus, K., Wordsworth, R. D. & Morley, C. V. Sulfate aerosol hazes and SO2 gas as constraints on rocky exoplanets’ surface liquid water. Astrophys. J. 887, 231 (2019).

    Article  Google Scholar 

  94. Gulick, S. et al. Importance of pre-impact crustal structure for the asymmetry of the Chicxulub impact crater. Nat. Geosci. 1, 131–135 (2008).

    Article  Google Scholar 

  95. Bekki, S. Oxidation of volcanic SO2: a sink for stratospheric OH and H2O. Geophys. Res. Lett. 22, 913–916 (1995).

    Article  Google Scholar 

  96. Ohno, S. et al. Production of sulphate-rich vapour during the Chicxulub impact and implications for ocean acidification. Nat. Geosci. 7, 279–282 (2014).

    Article  Google Scholar 

Download references

Acknowledgements

This research is supported by the Belgian Federal Science Policy (BELSPO) through the Chicxulub BRAIN-be (Belgian Research Action through Interdisciplinary Networks) project (to P.C. & Ö.K.) and FED-tWIN project Prf-2020-038 (to J.V.), as well as the Research Foundation-Flanders (FWO; project G0A6517N, grant 12AM624N to C.B.S., grant 11E6621N to P.K., 12Z6621N to J.V., 12ZZL20N to O.T.). S.G. and P.C. acknowledge support of the VUB strategic programme. Ö.K. acknowledges the support of BELSPO through the ESA/PRODEX programme. M. Hagen and U. van Buuren (VU Amsterdam) are thanked for their assistance during the laser-diffraction particle-size analyses.

Author information

Authors and Affiliations

Authors

Contributions

C.B.S. and P.K. led the writing of the paper. C.B.S., P.K., O.T., J.V., S.G., P.C. and Ö.K. built the conceptualization of study and wrote the original text. C.B.S., P.K., O.T., J.V., S.G., R.D., M.A.P., P.C. and Ö.K. commented on and edited the original and revised manuscripts. C.B.S., O.T. and Ö.K. developed the general circulation model, implemented microphysics and radiation models and performed palaeoclimate simulations and post-processing of the results. P.K. and R.D. collected sediment samples during fieldwork at the Tanis K-Pg site in August 2017. P.K. carried out laser-diffraction grain-size analyses, with lab supervision of M.A.P. P.K. created Figs. 1 and 2 and Extended Data Fig. 1. C.B.S. created all other figures. All authors approved the final draft of the manuscript.

Corresponding author

Correspondence to Cem Berk Senel.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Geoscience thanks Teruyuki Maruoka, Julia Brugger and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Primary Handling Editor: Tamara Goldin, in collaboration with the Nature Geoscience team.

Additional information

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

Extended data

Extended Data Fig. 1 Geological context of the Chicxulub impact ejecta stratigraphy at the Tanis K-Pg boundary site.

a The inset shows a paleogeographic reconstruction and relief map for the latest Cretaceous42 as used in this modeling study with locations of Chicxulub and Tanis indicated. The base map is based on the latest Cretaceous paleogeographic data42. b Stratigraphy of the Tanis K-Pg boundary event deposit highlighting the lithological units (adapted from41; based on sections X-2741-A and X-2761) together with data on grain-size classes (clay, silt, and sand fractions), median grain-size values (in µm) and different types of impact ejecta found within this deposit. HCF = Hell Creek Formation (Upper Cretaceous). FUF = Fort Union Formation (Paleocene). c Representative grain-size distribution curves throughout the section, the colors match the stratigraphic units in b. The uppermost K-Pg boundary claystone (unit P1), indicated with a bold orange line, corresponds to final phases of atmospheric fallout of silicate dust injected by the impact and is used in present GCM simulations.

Extended Data Fig. 2 Mass and number density spectra of the grain-size dataset.

a Mass density spectrum of Tanis sediment sample X-2761-8A displayed by filled orange circles, which corresponds to the uppermost K-Pg claystone interval just below a Paleocene lignite and yields a median grain-size of 2.88 µm. The measured mass density spectrum was fitted by a trimodal lognormal size distribution as depicted by the blue solid line. This fitted model curve is the sum of lognormal size distributions comprising 3 modes. Model parameters are as follows: w0 = 0.002, Dpg,0 = 0.18 μm, lnσg,0 = 0.3087 (mode 1, green solid line), w0 = 0.125, Dpg,0 = 2.6 μm, lnσg,0 = 1.1193 (mode 2, cyan solid line), w0 = 0.01, Dpg,0 = 30 μm, lnσg,0 = 0.7284 (mode 3, magenta solid line), b Converted grain-size distribution into number density spectrum displayed by filled cyan circles. The converted spectrum is fitted by a lognormal size distribution (blue solid line), which is the input parameter for our GCM study. Converted model median grain-size corresponds to 0.125 µm with a logarithmic standard deviation of 0.446.

Extended Data Fig. 3 Number density spectrum of soot along with the particle size from K-Pg boundary layer.

It indicates a median diameter of 0.22 μm (Toon et al., 201639; Wolbach et al., 198520). Concerning the aerosol life cycle and processes, coagulation is one of the crucial microphysical mechanisms that might pose as important for the transport of aerosols in the atmosphere. Nanometric particles below 0.1 μm, that is, 0.015 μm < Dp < 0.052 μm79, whose range is referred to as the Aitken-mode (within the cyan dashed lines), are formed by two processes: (i) condensational growth on existing aerosol particles, and (ii) coagulation due to the random particle collisions. These nanometric particles can further grow into larger particles or chains, resulting in the so-called accumulation-mode (0.056 μm < Dp < 0.26 μm)79 (within the red dashed lines) where the coagulation can occur especially at high particle concentrations following the K-Pg impact. The median diameter of soot (0.22 μm20,39) in our simulations are prominently larger than the Aitken-mode interval, while lying within the range of accumulation-mode.

Extended Data Fig. 4 Latest Cretaceous surface temperatures from our GCM simulations, one year before the impact, in comparison with proxy observations43.

Proxy temperature data are presented as mean values +/- standard error of mean (SEM), displayed by black circles and horizontal error bars. Here, the proxy data consists of N = 66 samples at different latitudes. Green solid and dashed lines display the zonal mean of land temperatures during the boreal summer and winter seasons, from GCM simulations. Blue solid (boreal summer) and dashed (boreal winter) lines indicate the zonal mean of ocean temperatures. Both green (land) and blue (ocean) shaded areas show the region between the mean boreal summer and winter profiles. The black solid line refers to the GCM-based annual average of land and ocean surface temperatures at each latitude.

Extended Data Fig. 5 Global surface temperature reconstructions using the combined fine-grained ejecta scenario.

Results are displayed on a latest Cretaceous paleogeographic map (Extended Data Fig. 1a) and shown for different time snapshots. a Latest Cretaceous, 1 year before impact (annual mean). b Latest Cretaceous, 1 week before impact. c Impact winter, 1 month after impact. d 6 months after impact. e 2 years after impact. f 10 years after impact (annual-mean). Base maps are based on the latest Cretaceous paleogeographic data42.

Extended Data Fig. 6 Impact-generated global surface net radiative responses.

The temporal evolution from ~2 years of the latest Cretaceous towards 25 years of post-impact conditions, for the individual silicate dust, sulfur, soot, and combined scenarios. a Global-average surface net shortwave radiation flux. b Global-average surface net longwave radiation flux. Here in x-axis, the year of 0 refers to the start of the year where the impact event occurs. The solid purple dashed line denotes the moment of Chicxulub impact, that is, boreal spring season38. Our paleoclimate simulations indicate that the drastic changes in surface net shortwave/longwave radiation stabilize to pre-impact levels within the first 3 years after impact. Accordingly, this timescale, in which large radiative anomalies emerged, determines the timescale of the initial extreme cold (Fig. 4a).

Extended Data Fig. 7 Global PAR flux reconstructions in the latest Cretaceous.

Land-Ocean PAR flux a 1 day before impact (boreal spring) and b 6 months before impact (austral summer), displayed on a latest Cretaceous paleogeographic map (Extended Data Fig. 1a). The range of the green and purple colorbar represents the photosynthetically high and low radiative flux, varying between 0-160 W/m2. Base maps are based on the latest Cretaceous paleogeographic data42.

Extended Data Fig. 8 PAR flux reconstructions following the Chicxulub impact.

Land and ocean PAR flux from 1 day (post-impact state, instantaneous) to 1 and 2 weeks after impact for a silicate dust; b sulfur; and c soot scenarios, displayed on a latest Cretaceous paleogeographic map (Extended Data Fig. 1a). The range of the green-white colorbar denotes the photosynthetically high and low radiative flux. Base maps are based on the latest Cretaceous paleogeographic data42.

Extended Data Fig. 9 Effect of silicate dust particle size on the global column-integrated fine-grained ejecta mass.

Orange line refers to the present study, using the Tanis K-Pg silicate dust. Gray and black lines show GCM results using particle size constraints reported in previous modeling studies18,39. The gray line displays the response of nanometric sized particles, indicating a deposition rate with an atmospheric lifetime of ~7 years (as a lower threshold). The black line refers to the type 2 spherules39, representing microkrystites of 250 µm in diameter prone to very swift gravitational settling within a few days after impact. Cyan dashed line displays the shocked ejected quartz grains (mean diameter of 50 µm) defined as clastic debris27. We use the same amount of ejecta release in each GCM simulation, in the order of 2×1018 g as an upper limit. The optical properties of nanoparticles and type 2 spherules are the same as in the Tanis K-Pg silicate dust simulation, as we compare the microphysical response to the changes in particle size. Here in x-axis, the year of 0 refers to the start of the year where the impact event occurs. The purple dashed line denotes the moment of Chicxulub impact, that is, boreal spring season38. Regarding nanoparticles, those nanometric sized particles (median diameter of 20 nm) would grow into larger particles in atmosphere due to the coagulation. Such larger aggregates would have lower deposition rates on land and ocean (Fig. 3), hence higher atmospheric lifetimes. To illustrate, the deposition rate of nanoparticles (gray dashed line) would have occasionally shifted rightward through the response of silicate dust (orange dashed line) depending on the rate of coagulation. Therefore, the present simulation of nanoparticles, excluding coagulation, would serve as the minimum threshold for the atmospheric lifetime (t ~ 7 years). The inclusion of coagulation mechanism forming larger aggregates would lead to lower deposition rates on land and ocean (Fig. 3) for some fraction of nanoparticles, thus relatively high atmospheric lifetime of more than 7 years. Nevertheless, we do not expect nanoparticles to have an atmospheric lifetime and PAR response as substantial as single soot or micrometer-sized silicate dust.

Extended Data Fig. 10 Global-average surface temperature.

It is same as Fig. 4a, yet the time evolution is shown from 15 years before the Chicxulub impact instead of 2 years, for the individual silicate dust, sulfur, soot, and combined scenarios. The first 15 years correspond to the model initial spin-up simulation of 15-years, in which the latest Cretaceous conditions stabilized.

Supplementary information

Supplementary Information

Supplementary Table 1.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) 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

Senel, C.B., Kaskes, P., Temel, O. et al. Chicxulub impact winter sustained by fine silicate dust. Nat. Geosci. 16, 1033–1040 (2023). https://doi.org/10.1038/s41561-023-01290-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41561-023-01290-4

This article is cited by

Search

Quick links

Nature Briefing

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

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