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:

Episodic deluges in simulated hothouse climates

Abstract

Earth’s distant past and potentially its future include extremely warm ‘hothouse’1 climate states, but little is known about how the atmosphere behaves in such states. One distinguishing characteristic of hothouse climates is that they feature lower-tropospheric radiative heating, rather than cooling, due to the closing of the water vapour infrared window regions2. Previous work has suggested that this could lead to temperature inversions and substantial changes in cloud cover3,4,5,6, but no previous modelling of the hothouse regime has resolved convective-scale turbulent air motions and cloud cover directly, thus leaving many questions about hothouse radiative heating unanswered. Here we conduct simulations that explicitly resolve convection and find that lower-tropospheric radiative heating in hothouse climates causes the hydrologic cycle to shift from a quasi-steady regime to a ‘relaxation oscillator’ regime, in which precipitation occurs in short and intense outbursts separated by multi-day dry spells. The transition to the oscillatory regime is accompanied by strongly enhanced local precipitation fluxes, a substantial increase in cloud cover, and a transiently positive (unstable) climate feedback parameter. Our results indicate that hothouse climates may feature a novel form of ‘temporal’ convective self-organization, with implications for both cloud coverage and erosion processes.

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: Transition to the relaxation oscillator regime due to increased insolation.
Fig. 2: The oscillatory regime is induced by LTRH.
Fig. 3: Mechanism of the oscillatory regime as revealed by high-frequency model output.
Fig. 4: Overview of the relaxation oscillator convective regime.

Similar content being viewed by others

Data availability

Input data files and cloud-resolving model output associated with this work are available in a Zenodo repository at https://doi.org/10.5281/zenodo.5117529.

Code availability

Source code for the stochastic two-layer model, processing cloud-resolving model output, and generating figures is available in a Zenodo repository at https://doi.org/10.5281/zenodo.5117529.

References

  1. Steffen, W. et al. Trajectories of the Earth System in the Anthropocene. Proc. Natl Acad. Sci. USA 115, 8252–8259 (2018).

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  2. Wordsworth, R. D. & Pierrehumbert, R. T. Water loss from terrestrial planets with CO2-rich atmospheres. Astrophys. J. 778, 154 (2013).

  3. Wolf, E. T. & Toon, O. B. The evolution of habitable climates under the brightening Sun. J. Geophys. Res. Atmos. 120, 5775–5794 (2015).

    Article  ADS  Google Scholar 

  4. Kopparapu, R. K. et al. The inner edge of the habitable zone for synchronously rotating planets around low-mass stars using general circulation models. Astrophys. J. 819, 84 (2016).

  5. Popp, M., Schmidt, H. & Marotzke, J. Transition to a Moist Greenhouse with CO2 and solar forcing. Nat. Commun. 7, 10627 (2016).

  6. Wolf, E. T., Haqq-Misra, J. & Toon, O. B. Evaluating climate sensitivity to CO2 across Earth’s history. J. Geophys. Res. Atmos. 123, 11861–11874 (2018).

    Article  ADS  Google Scholar 

  7. Snyder, C. W. Evolution of global temperature over the past two million years. Nature 538, 226–228 (2016).

    Article  CAS  PubMed  ADS  Google Scholar 

  8. Sleep, N. H. The Hadean-Archaean environment. Cold Spring Harb. Perspect. Biol. https://doi.org/10.1101/cshperspect.a002527 (2010).

  9. Charnay, B., Le Hir, G., Fluteau, F., Forget, F. & Catling, D. C. A warm or a cold early Earth? New insights from a 3-D climate-carbon model. Earth Planet. Sci. Lett. 474, 97–109 (2017).

    Article  CAS  ADS  Google Scholar 

  10. Pierrehumbert, R., Abbot, D., Voigt, A. & Koll, D. Climate of the Neoproterozoic. Annu. Rev. Earth Planet. Sci. 39, 417–460 (2011).

    Article  CAS  ADS  Google Scholar 

  11. Goldblatt, C. & Watson, A. J. The runaway greenhouse: implications for future climate change, geoengineering and planetary atmospheres. Philos. Trans. R. Soc. A 370, 4197–4216 (2012).

    Article  CAS  ADS  Google Scholar 

  12. Koll, D. D. B. & Cronin, T. W. Earth’s outgoing longwave radiation linear due to H2O greenhouse effect. Proc. Natl Acad. Sci. USA 115, 10293–10298 (2018).

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  13. Seeley, J. T. & Jeevanjee, N. H2O windows and CO2 radiator fins: a clear-sky explanation for the peak in equilibrium climate sensitivity. Geophys. Res. Lett. 48, e2020GL089609 (2021).

    Article  CAS  ADS  Google Scholar 

  14. Yang, J., Cowan, N. B. & Abbot, D. S. Stabilizing cloud feedback dramatically expands the habitable zone of tidally locked planets. Astrophys. J. Lett. 771, L45 (2013).

  15. Sergeev, D. E. et al. Atmospheric convection plays a key role in the climate of tidally-locked terrestrial exoplanets: insights from high-resolution simulations. Astrophys. J. 894, 84 (2020).

  16. Lefèvre, M., Turbet, M. & Pierrehumbert, R. 3D convection-resolving model of temperate, tidally locked exoplanets. Astrophys. J. 913, 101 (2021).

    Article  ADS  CAS  Google Scholar 

  17. Romps, D. M. The dry-entropy budget of a moist atmosphere. J. Atmos. Sci. 65, 3779–3799 (2008).

    Article  ADS  Google Scholar 

  18. Wang, D. in Wiley Encyclopedia of Electrical and Electronics Engineering (ed. Webster, J. G.) Vol. 18, 396–405 (Wiley, 1999).

  19. Ginoux, J.-M. & Letellier, C. Van der Pol and the history of relaxation oscillations: toward the emergence of a concept. Chaos 22, 023120 (2012).

  20. Pendergrass, A. G. & Hartmann, D. L. The atmospheric energy constraint on global-mean precipitation change. J. Clim. 27, 757–768 (2014).

    Article  ADS  Google Scholar 

  21. Jeevanjee, N. & Romps, D. M. Mean precipitation change from a deepening troposphere. Proc. Natl Acad. Sci. USA 115, 11465–11470 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Torri, G., Kuang, Z. & Tian, Y. Mechanisms for convection triggering by cold pools. Geophys. Res. Lett. 42, 1943–1950 (2015).

    Article  ADS  Google Scholar 

  23. Jeevanjee, N. & Romps, D. M. Effective buoyancy, inertial pressure, and the mechanical generation of boundary layer mass flux by cold pools. J. Atmos. Sci. 72, 3199–3213 (2015).

    Article  ADS  Google Scholar 

  24. Feng, Z. et al. Mechanisms of convective cloud organization by cold pools over tropical warm ocean during the AMIE/DYNAMO field campaign. J. Adv. Model. Earth Syst. 7, 357–381 (2015).

    Article  ADS  Google Scholar 

  25. Torri, G. & Kuang, Z. On cold pool collisions in tropical boundary layers. Geophys. Res. Lett. 46, 399–407 (2019).

    Article  ADS  Google Scholar 

  26. Singh, M. S. & O’Gorman, P. A. Influence of entrainment on the thermal stratification in simulations of radiative-convective equilibrium. Geophys. Res. Lett. 40, 4398–4403 (2013).

    Article  ADS  Google Scholar 

  27. Seeley, J. T. & Romps, D. M. Why does tropical convective available potential energy (CAPE) increase with warming? Geophys. Res. Lett. 42, 10,429–10,437 (2015).

    Article  Google Scholar 

  28. Zhao, M. Uncertainty in model climate sensitivity traced to representations of cumulus precipitation microphysics. J. Clim. 29, 543–560 (2016).

    Article  ADS  Google Scholar 

  29. Popp, M., Schmidt, H. & Marotzke, J. Initiation of a runaway greenhouse in a cloudy column. J. Atmos. Sci. 72, 452–471 (2015).

    Article  ADS  Google Scholar 

  30. Wing, A. A. et al. Clouds and convective self-aggregation in a multimodel ensemble of radiative-convective equilibrium simulations. J. Adv. Model. Earth Syst. 12, e2020MS002138 (2020).

    Article  Google Scholar 

  31. Becker, T. & Wing, A. A. Understanding the extreme spread in climate sensitivity within the radiative-convective equilibrium model intercomparison project. J. Adv. Model. Earth Syst. 12, e2020MS002165 (2020).

  32. Romps, D. M. Climate sensitivity and the direct effect of carbon dioxide in a limited-area cloud-resolving model. J. Clim. 33, 3413–3429 (2020).

    Article  ADS  Google Scholar 

  33. Wing, A. A., Emanuel, K., Holloway, C. E. & Muller, C. Convective self-aggregation in numerical simulations: a review. Surv. Geophys. 38, 1173–1197 (2017).

    Article  ADS  Google Scholar 

  34. Mirollo, R. E. & Strogatz, S. H. Synchronization of pulse-coupled biological oscillators. SIAM J. Appl. Math. 50, 1645–1662 (1990).

    Article  MathSciNet  MATH  Google Scholar 

  35. Pantaleone, J. Synchronization of metronomes. Am. J. Phys. 70, 992–1000 (2002).

    Article  ADS  Google Scholar 

  36. Buck, J. & Buck, E. Biology of synchronous flashing of fireflies. Nature 211, 562–564 (1966).

    Article  ADS  Google Scholar 

  37. Bretherton, C. S. & Smolarkiewicz, P. K. Gravity waves, compensating subsidence, and detrainment around cumulus clouds. J. Atmos. Sci. 46, 740–759 (1989).

  38. Edman, J. P. & Romps, D. M. Beyond the rigid lid: baroclinic modes in a structured atmosphere. J. Atmos. Sci. 74, 3551–3566 (2017).

    Article  ADS  Google Scholar 

  39. Carlson, T. N., Benjamin, S. G. & Forbes, G. S. Elevated mixed layers in the regional severe storm environment: conceptual model and case studies. Mon. Weather Rev. 11, 1453–1473 (1983).

  40. Schultz, D. M., Richardson, Y. P., Markowski, P. M. & Doswell, C. A. Tornadoes in the central United States and the “clash of air masses”. Bull. Am. Meteorol. Soc. 95, 1704–1712 (2014).

    Article  ADS  Google Scholar 

  41. Agard, V. & Emanuel, K. Clausius–Clapeyron scaling of peak CAPE in continental convective storm environments. J. Atmos. Sci. 74, 3043–3054 (2017).

    Article  ADS  Google Scholar 

  42. Arakawa, A. & Schubert, W. H. Interaction of a cumulus cloud ensemble with the large-scale environment, part I. J. Atmos. Sci. 31, 674–701 (1974).

  43. Raymond, D. in The Physics and Parameterization of Moist Atmospheric Convection (ed. Smith, R. K.) 387–397 (Kluwer Academic Publishers, 1997).

  44. Melosh, H. J. in Planetary Surface Processes Ch. 10, 382–433 (Cambridge Univ. Press, 2011).

  45. Villarini, G., Smith, J. A., Baeck, M. L., Marchok, T. & Vecchi, G. A. Characterization of rainfall distribution and flooding associated with U.S. landfalling tropical cyclones: analyses of Hurricanes Frances, Ivan, and Jeanne (2004). J. Geophys. Res. Atmos. 116, D23116 (2011).

  46. Graham, R. J. & Pierrehumbert, R. Thermodynamic and energetic limits on continental silicate weathering strongly impact the climate and habitability of wet, rocky worlds. Astrophys. J. 896, 115 (2020).

    Article  CAS  ADS  Google Scholar 

  47. Dansgaard, W. Stable isotopes in precipitation. Tellus 16, 436–468 (1964).

    Article  ADS  Google Scholar 

  48. Pausata, F. S., Battisti, D. S., Nisancioglu, K. H. & Bitz, C. M. Chinese stalagmite δ18O controlled by changes in the indian monsoon during a simulated heinrich event. Nat. Geosci. 4, 474–480 (2011).

    Article  CAS  ADS  Google Scholar 

  49. Frieling, J. et al. Extreme warmth and heat-stressed plankton in the tropics during the Paleocene-Eocene Thermal Maximum. Sci. Adv. 3, e1600891 (2017).

  50. Wing, A. A. et al. Radiative-Convective Equilibrium Model Intercomparison Project. Geosci. Model Dev. 11, 793–813 (2018).

    Article  CAS  ADS  Google Scholar 

  51. Romps, D. M. Response of tropical precipitation to global warming. J. Atmos. Sci. 68, 123–138 (2011).

    Article  ADS  Google Scholar 

  52. Clough, S. A. et al. Atmospheric radiative transfer modeling: a summary of the AER codes. J. Quant. Spectrosc. Radiat. Transf. 91, 233–244 (2005).

    Article  CAS  ADS  Google Scholar 

  53. Iacono, M. J. et al. Radiative forcing by long-lived greenhouse gases: calculations with the AER radiative transfer models. J. Geophys. Res. Atmos. 113, 2–9 (2008).

    Article  CAS  Google Scholar 

  54. Wordsworth, R. et al. Transient reducing greenhouse warming on early Mars. Geophys. Res. Lett. 44, 665–671 (2017).

    Article  CAS  Google Scholar 

  55. Ding, F. & Wordsworth, R. D. A new line-by-line general circulation model for simulations of diverse planetary atmospheres: initial validation and application to the exoplanet GJ 1132B. Astrophys. J. 878, 117 (2019).

    Article  CAS  ADS  Google Scholar 

  56. Schaefer, L., Wordsworth, R. D., Berta-Thompson, Z. & Sasselov, D. Predictions of the atmospheric composition of GJ 1132b. Astrophys. J. 829, 63 (2016).

    Article  ADS  Google Scholar 

  57. Clough, S. A., Iacono, M. J. & Moncet, J.-L. Line-by-line calculations of atmospheric fluxes and cooling rates: application to water vapor. J. Geophys. Res. 97, 15761 (1992).

    Article  ADS  Google Scholar 

  58. Gordon, I. E. et al. The HITRAN2016 molecular spectroscopic database. J. Quant. Spectrosc. Radiat. Transf. 203, 3–69 (2017).

    Article  CAS  ADS  Google Scholar 

  59. Dudhia, A. The Reference Forward Model (RFM). J. Quant. Spectrosc. Radiat. Transf. 186, 243–253 (2017).

    Article  CAS  ADS  Google Scholar 

  60. Mlawer, E. J. et al. Development and recent evaluation of the MT_CKD model of continuum absorption. Philos. Trans. R. Soc. A 370, 2520–2556 (2012).

    Article  CAS  ADS  Google Scholar 

  61. Claire, M. W. et al. The evolution of solar flux from 0.1 nm to 160 μm: quantitative estimates for planetary studies. Astrophys. J. 757, 95 (2012).

  62. Segura, A. et al. Biosignatures from Earth-like planets around M dwarfs. Astrobiology 5, 706–725 (2005).

    Article  CAS  PubMed  ADS  Google Scholar 

  63. Krueger, S. K., Fu, Q., Liou, K. N. & Chin, H.-N. S. Improvements of an ice-phase mcrophysics parameterization for use in numerical simulations of tropical convection. J. Appl. Meteorol. 34, 281–286 (1995).

  64. Lin, Y.-L., Farley, R. D. & Orville, H. D. Bulk parameterization of the snow field in a cloud model. J. Clim. Appl. Meteorol. 22, 1065–1092 (1983).

  65. Lord, S. J., Willoughby, H. E. & Piotrowicz, J. M. Role of a parameterized ice-phase microphysics in an axisymmetric, nonhydrostatic tropical cyclone model. J. Atmos. Sci. 41, 2836–2848 (1984).

    Article  ADS  Google Scholar 

  66. Seeley, J. T., Jeevanjee, N. & Romps, D. M. FAT or FiTT: are anvil clouds or the tropopause temperature-invariant? Geophys. Res. Lett. 46, 1842–1850 (2019).

    Article  ADS  Google Scholar 

  67. Seeley, J. T., Lutsko, N. J. & Keith, D. W. Designing a radiative antidote to CO2. Geophys. Res. Lett. 48, e2020GL090876 (2021).

    Article  CAS  ADS  Google Scholar 

  68. Romps, D. M. & Kuang, Z. Nature versus nurture in shallow convection. J. Atmos. Sci. 67, 1655–1666 (2010).

    Article  ADS  Google Scholar 

  69. Khairoutdinov, M. F. & Randall, D. A. Cloud resolving modeling of the ARM summer 1997 IOP: model formulation, results, uncertainties, and sensitivities. J. Atmos. Sci. 60, 607–625 (2003).

    Article  ADS  Google Scholar 

  70. Bryan, G. & Fritsch, J. A benchmark simulation for moist nonhydrostatic numerical models. Mon. Weather Rev. 130, 2917–2928 (2002).

    Article  ADS  Google Scholar 

  71. Morrison, H., Curry, J. A. & Khvorostyanov, V. I. A new double-moment microphysics parameterization for application in cloud and climate models. Part I: description. J. Atmos. Sci.62, 1665–1677 (2005).

    Article  ADS  Google Scholar 

  72. Cess, R. D. et al. Intercomparison and interpretation of climate feedback processes in 19 atmospheric general circulation models. J. Geophys. Res. Atmos. 95, 16601–16615 (1990).

    Article  ADS  Google Scholar 

Download references

Acknowledgements

We are grateful to the authors of the cloud-resolving models used in this work: D. Romps, M. Khairoutdinov and G. Bryan. We also thank A. Dudhia for sharing with us the Reference Forward Model. We thank X. Wei for conducting exploratory simulations with SAM. J.T.S. thanks N. Jeevanjee, A. Match, N. Tarshish and Z. Kuang for discussions.

Author information

Authors and Affiliations

Authors

Contributions

J.T.S. and R.D.W. designed the research. J.T.S. performed the simulations, analysed the results and prepared the figures. The manuscript was written jointly by J.T.S. and R.D.W.

Corresponding author

Correspondence to Jacob T. Seeley.

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 Errors in clear-sky RRTM radiative heating rates are corrected by using line-by-line radiative transfer.

Comparison of net (LW+SW; panels ad), longwave (LW; panels eh), and shortwave (SW; panels il) radiative heating rates as computed by RRTM (black) and PCM_LBL (red). The heating rates are computed for moist-adiabatic temperature-pressure profiles with surface temperatures ranging from 305 K to 335 K in 10-K increments (columns, left to right). All columns have a surface pressure of 101325 Pa, 75% tropospheric relative humidity, 400 ppm CO2, and an isothermal stratosphere at 160 K. Note that the discontinuous heating rates calculated by RRTM for the warmer atmospheres (around 20 km altitude) do not appear in the PCM_LBL results.

Extended Data Fig. 2 Top-of-atmosphere radiative fluxes and heating rates from DAM snapshots.

(ac) Outgoing longwave radiation (OLR) from a snapshot from the fixedSST_hires DAM simulation with a surface temperature of 305 K, computed by three different combinations of radiative transfer codes and approximations. Panel (a) is from RRTM alone, panel (b) shows the result of swapping out the clear-sky radiative fluxes from RRTM with those calculated by PCM_LBL, and panel (c) shows the result of swapping out each column’s clear-sky radiative fluxes for those calculated by PCM_LBL for the horizontal-mean column, which is the approach taken for the simulations associated with this work. Panel (d) shows the horizontal-mean longwave radiative heating rates for this snapshot. (eh) As in (ad), but for absorbed shortwave radiation (ASR). (ip) As in (ah), but for a snapshot from the simulation with a surface temperature of 330 K.

Extended Data Fig. 3 Tests of the robustness of the oscillatory transition.

Domain-mean precipitation from two periods of (a) the FCO2 simulation with mean SSTs of 306.1 K and 331.5 K; (b) the fixedSST suite at 305 K and 330 K; (c) the fixedSST_sm suite, which use the simplified microphysics parameterization described in the Methods; (d) fixed-SST simulations with finer horizontal resolution (Δx = 250 m; fixedSST_hires) or on a larger domain (Lx = 512 km; fixedSST_large). (e) The same quantity from simulations conducted with the System for Atmospheric Modeling (SAM)69 at fixed SSTs of 305 and 325 K. (f) As in (e), but for the Cloud Model 1 (CM1)70.

Extended Data Fig. 4 Mean profiles of temperature and cloud fraction.

From the fixedSST simulations, profiles of (a) mean temperature and (b) mean cloud fraction (fraction of grid cells with non-precipitating cloud condensate mass fraction greater than 10−5 kg/kg). In (a), the variability is indicated by the shading, which shows ±2 standard deviations of hourly-mean temperatures at each altitude. In (a), the dashed line shows the mean temperature profile from the simulation without evaporation of precipitating hydrometeors (prevap0) at 330 K.

Extended Data Fig. 5 Sign reversal of the climate feedback parameter indicates transient climate instability.

The feedback parameter λ is defined here as minus the change in net radiative flux at the top-of-atmosphere (TOA) per degree of surface warming (positive downward, so that a negative feedback indicates more radiation escaping to space with warming and hence climate stability, and a positive feedback indicates climate instability; this is often called the “Cess sensitivity”72). We calculated feedbacks using finite differences on a staggered surface temperature grid that interpolates between the surface temperatures of the fixedSST experiment. (a) The solid line shows clear-sky feedbacks calculated for TOA fluxes averaged over the final 100 days of the fixedSST simulations, while the dashed and dot-dashed lines show the feedbacks calculated using the time-mean columns from those simulations with actual or fixed 100% relative humidity profiles, respectively. (b) As in (a), but for the all-sky feedbacks from fixedSST experiments broken down into longwave and shortwave components. The dashed line shows the net all-sky feedback from the final 50 days of the LTRH_off experiment, which does not undergo a steady-to-oscillatory transition and remains stable at all temperatures. (c) Time-mean profiles of relative humidity (RH) in the fixedSST experiments, using temperature within the atmosphere as a vertical coordinate to emphasize the increases in upper-tropospheric relative humidity that occur during the oscillatory transition between 320 and 325 and K. Since the clear-sky climate instability is eliminated by using a fixed relative humidity of 100% (panel a), we attribute the clear-sky climate instability to the increase in upper-tropospheric RH, which lowers spectral emission temperatures and hence OLR.

Extended Data Fig. 6 Spatially-separated subdomains exhibit in-phase pulses of convection.

Timeseries of (a,c) moist static energy in the lowest model level (z = 12.5 m; MSEsurf), and (b,d) precipitation rate, averaged over five different subdomains of the fixedSST_large simulations at 305 K (top row) and 330 K (bottom row). The subdomains (color-coded in panel e) each have an area of 256 km2 and are located an average of 215 km apart from each other.

Extended Data Fig. 7 The steady-to-oscillatory transition in the convection-resolving model and the stochastic two-layer model.

(a) In the convection-resolving model, the radiative heating profile is switched from cool-climate-type to hothouse-type (LTRH_off to LTRH_on) on model day 0 (the transient_SO simulation). (b) In the two-layer model, the inhibition parameter is increased linearly in time between days 0 and 2 and held fixed thereafter.

Extended Data Fig. 8 Probability density functions (PDFs) of 6-hour local rain accumulations.

The precipitation data are from 20-day periods of (a) the fixedSST_large simulations, and (b) the transient_SO simulation in the steady and oscillatory regime. The PDFs are constructed by first dividing the model domains into watershed-sized subdomains (16 × 16 km2 for fixedSST_large, and 12 × 12 km2 for transient_SO). Precipitation is then accumulated in each subdomain for all 6-hour periods during the 20-day intervals, producing the 6-hour local rain accumulations from which the PDFs are constructed. The 99.9th percentile of each of the PDFs is indicated at the top of each plot.

Extended Data Fig. 9 The oscillatory transition occurs more readily for climates instellated by an M-star spectrum.

Comparison of tropospheric radiative heating rates (panels a,b) and timeseries of surface precipitation (panels c,d) in fixed-SST simulations with either the solar instellation spectrum or that of the M-star AD Leonis62. Panel (e) shows the spectral flux for these two stars (normalized to the same total flux), as well as the logarithm of the H2O absorption coefficient at a reference temperature and pressure.

Extended Data Table 1 Summary of key aspects of the suite of DAM simulations conducted for this work

Supplementary information

Supplementary Video 1

The supplementary video is an animation of DAM model output from the fixed-SST simulation at 330 K (from our fixedSST_hires suite; Extended Data Table 1). Each frame in the video consists of six panels showing, from top to bottom and left to right: buoyancy in the near-surface layer, wind speed in the near-surface layer, outgoing solar radiation, temperature anomaly in the near-surface layer, specific humidity anomaly in the near-surface layer, and accumulated rainfall over the preceding 6 h. Anomalies are calculated with respect to the horizontal and time mean. The sampling interval between frames is 15 min, and the animation covers 7 days of model time. The video is also available at https://youtu.be/NALhYFiaeos.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Seeley, J.T., Wordsworth, R.D. Episodic deluges in simulated hothouse climates. Nature 599, 74–79 (2021). https://doi.org/10.1038/s41586-021-03919-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-021-03919-z

This article is cited by

Comments

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

Search

Quick links

Nature Briefing

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