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.

A large source of cloud condensation nuclei from new particle formation in the tropics

Abstract

Cloud condensation nuclei (CCN) can affect cloud properties and therefore Earth’s radiative balance1,2,3. New particle formation (NPF) from condensable vapours in the free troposphere has been suggested to contribute to CCN, especially in remote, pristine atmospheric regions4, but direct evidence is sparse, and the magnitude of this contribution is uncertain5,6,7. Here we use in situ aircraft measurements of vertical profiles of aerosol size distributions to present a global-scale survey of NPF occurrence. We observe intense NPF at high altitudes in tropical convective regions over both Pacific and Atlantic oceans. Together with the results of chemical-transport models, our findings indicate that NPF persists at all longitudes as a global-scale band in the tropical upper troposphere, covering about 40 per cent of Earth’s surface. Furthermore, we find that this NPF in the tropical upper troposphere is a globally important source of CCN in the lower troposphere, where CCN can affect cloud properties. Our findings suggest that the production of CCN as new particles descend towards the surface is not adequately captured in global models, which tend to underestimate both the magnitude of tropical upper tropospheric NPF and the subsequent growth of new particles to CCN sizes.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: New particle formation and growth to CCN sizes in the tropical convective region.

Image courtesy of K. Bogan, Cooperative Institute for Research in Environmental Sciences, University of Colorado, Boulder.

Fig. 2: Average aerosol properties from ATom 1 and 2.
Fig. 3: The relationship between CS7, temperature and NPF.
Fig. 4: Evidence for particle growth on descent.

Similar content being viewed by others

Data availability

The full ATom data set is publicly available15, as are data specific to this analysis65.

Code availability

Code for the model CESM with the base version of CARMA is available online66 as is code for GEOS-Chem with TOMAS and APM67. Code used to analyse ATom data and model output, and recent modifications to CARMA, GEOS-Chem and CAM5 with APM used here, is available on request.

References

  1. Rosenfeld, D., Sherwood, S., Wood, R. & Donner, L. Climate effects of aerosol-cloud interactions. Science 343, 379–380 (2014).

    ADS  CAS  PubMed  Google Scholar 

  2. Boucher, O. et al. in Climate Change 2013: The Physical Science Basis (eds Stocker, T. F. et al.) Ch. 7 (Cambridge Univ. Press, 2013).

  3. Dusek, U. et al. Size matters more than chemistry for cloud-nucleating ability of aerosol particles. Science 312, 1375–1378 (2006).

    ADS  CAS  PubMed  Google Scholar 

  4. Gordon, H. et al. Causes and importance of new particle formation in the present-day and preindustrial atmospheres. J. Geophys. Res. Atmos. 122, 8739–8760 (2017).

    ADS  Google Scholar 

  5. Merikanto, J., Spracklen, D. V., Mann, G. W., Pickering, S. J. & Carslaw, K. S. Impact of nucleation on global CCN. Atmos. Chem. Phys. 9, 8601–8616 (2009).

    ADS  CAS  Google Scholar 

  6. Pierce, J. R. & Adams, P. J. Uncertainty in global CCN concentrations from uncertain aerosol nucleation and primary emission rates. Atmos. Chem. Phys. 9, 1339–1356 (2009).

    ADS  CAS  Google Scholar 

  7. Clarke, A. D. et al. Free troposphere as a major source of CCN for the equatorial pacific boundary layer: long-range transport and teleconnections. Atmos. Chem. Phys. 13, 7511–7529 (2013).

    ADS  CAS  Google Scholar 

  8. Kulmala, M. et al. Direct observations of atmospheric aerosol nucleation. Science 339, 943–946 (2013).

    ADS  CAS  PubMed  Google Scholar 

  9. Dunne, E. M. et al. Global atmospheric particle formation from CERN CLOUD measurements. Science 354, 1119–1124 (2016).

    ADS  CAS  PubMed  Google Scholar 

  10. Brock, C. A., Hamill, P., Wilson, J. C., Jonsson, H. H. & Chan, K. R. Particle formation in the upper tropical troposphere—a source of nuclei for the stratospheric aerosol. Science 270, 1650–1653 (1995).

    ADS  CAS  Google Scholar 

  11. Clarke, A. D. et al. Nucleation in the equatorial free troposphere: favorable environments during PEM-Tropics. J. Geophys. Res. Atmos. 104, 5735–5744 (1999).

    ADS  CAS  Google Scholar 

  12. Weigel, R. et al. In situ observations of new particle formation in the tropical upper troposphere: the role of clouds and the nucleation mechanism. Atmos. Chem. Phys. 11, 9983–10010 (2011).

    ADS  CAS  Google Scholar 

  13. Clarke, A. D. et al. Particle production in the remote marine atmosphere: cloud outflow and subsidence during ACE 1. J. Geophys. Res. Atmos. 103, 16397–16409 (1998).

    ADS  CAS  Google Scholar 

  14. Clarke, A. D. & Kapustin, V. N. A Pacific aerosol survey. Part I: a decade of data on particle production, transport, evolution, and mixing in the troposphere. J. Atmos. Sci. 59, 363–382 (2002).

    ADS  Google Scholar 

  15. Wofsy, S. C. et al. ATom: merged atmospheric chemistry, trace gases, and aerosols. https://daac.ornl.gov/ATOM/guides/ATom_merge.html (ORNL DAAC, 2018).

    Google Scholar 

  16. Andreae, M. O. et al. Aerosol characteristics and particle production in the upper troposphere over the Amazon Basin. Atmos. Chem. Phys. 18, 921–961 (2018).

    ADS  CAS  Google Scholar 

  17. Ekman, A. M. L., Wang, C., Strom, J. & Krejci, R. Explicit simulation of aerosol physics in a cloud-resolving model: aerosol transport and processing in the free troposphere. J. Atmos. Sci. 63, 682–696 (2006).

    ADS  Google Scholar 

  18. Yue, G. K. & Hamill, P. Homogeneous nucleation rates of H2SO4-H2O aerosol particles in air. J. Aerosol Sci. 10, 609–614 (1979).

    ADS  CAS  Google Scholar 

  19. Gao, R. S. et al. OH in the tropical upper troposphere and its relationships to solar radiation and reactive nitrogen. J. Atmos. Chem. 71, 55–64 (2014).

    CAS  Google Scholar 

  20. Folkins, I. & Martin, R. V. The vertical structure of tropical convection and its impact on the budgets of water vapor and ozone. J. Atmos. Sci. 62, 1560–1573 (2005).

    ADS  Google Scholar 

  21. Fan, J. W. et al. Substantial convection and precipitation enhancements by ultrafine aerosol particles. Science 359, 411–418 (2018).

    ADS  MathSciNet  CAS  PubMed  Google Scholar 

  22. Raes, F. Entrainment of free tropospheric aerosols as a regulating mechanism for cloud condensation nuclei in the remote marine boundary layer. J. Geophys. Res. Atmos. 100, 2893–2903 (1995).

    ADS  CAS  Google Scholar 

  23. Quinn, P. K., Coffman, D. J., Johnson, J. E., Upchurch, L. M. & Bates, T. S. Small fraction of marine cloud condensation nuclei made up of sea spray aerosol. Nat. Geosci. 10, 674–679 (2017).

    ADS  CAS  Google Scholar 

  24. Hegg, D. A., Radke, L. F. & Hobbs, P. V. Particle production associated with marine clouds. J. Geophys. Res. Atmos. 95, 13917–13926 (1990).

    ADS  Google Scholar 

  25. Hoppel, W. A., Frick, G. M., Fitzgerald, J. & Larson, R. E. Marine boundary layer measurements of new particle formation and the effects nonprecipitating clouds have on aerosol-size distribution. J. Geophys. Res. Atmos. 99, 14443–14459 (1994).

    ADS  Google Scholar 

  26. Kazil, J. et al. Modeling chemical and aerosol processes in the transition from closed to open cells during VOCALS-REx. Atmos. Chem. Phys. 11, 7491–7514 (2011).

    ADS  CAS  Google Scholar 

  27. Yu, F. A secondary organic aerosol formation model considering successive oxidation aging and kinetic condensation of organic compounds: global scale implications. Atmos. Chem. Phys. 11, 1083–1099 (2011).

    ADS  CAS  Google Scholar 

  28. Raes, F., VanDingenen, R., Cuevas, E., VanVelthoven, P. F. J. & Prospero, J. M. Observations of aerosols in the free troposphere and marine boundary layer of the subtropical Northeast Atlantic: discussion of processes determining their size distribution. J. Geophys. Res. Atmos. 102, 21315–21328 (1997).

    ADS  CAS  Google Scholar 

  29. Trostl, J. et al. The role of low-volatility organic compounds in initial particle growth in the atmosphere. Nature 533, 527–531 (2016).

    ADS  CAS  PubMed  Google Scholar 

  30. Tsigaridis, K. et al. The AeroCom evaluation and intercomparison of organic aerosol in global models. Atmos. Chem. Phys. 14, 10845–10895 (2014).

    ADS  Google Scholar 

  31. Hoppel, W. A., Frick, G. M. & Larson, R. E. Effect of nonprecipitating clouds on the aerosol size distribution in the marine boundary layer. Geophys. Res. Lett. 13, 125–128 (1986).

    ADS  Google Scholar 

  32. Brock, C. A. et al. Aerosol size distributions during the Atmospheric Tomography Mission (ATom): methods, uncertainties, and data products. Atmos. Meas. Tech. 12, 3081–3099 (2019).

    Google Scholar 

  33. Williamson, C. et al. Fast time response measurements of particle size distributions in the 3-60 nm size range with the nucleation mode aerosol size spectrometer. Atmos. Meas. Tech. 11, 3491–3509 (2018).

    CAS  Google Scholar 

  34. Brock, C. A. et al. Sources of particulate matter in the northeastern United States in summer: 2. Evolution of chemical and microphysical properties. J. Geophys. Res. Atmos. 113, D08301 (2008).

    ADS  Google Scholar 

  35. Brock, C. A. et al. Ultrafine particle size distributions measured in aircraft exhaust plumes. J. Geophys. Res. Atmos. 105, 26555–26567 (2000).

    ADS  CAS  Google Scholar 

  36. Kupc, A., Williamson, C., Wagner, N. L., Richardson, M. & Brock, C. A. Modification, calibration, and performance of the ultra-high sensitivity aerosol spectrometer for particle size distribution and volatility measurements during the Atmospheric Tomography Mission (ATom) airborne campaign. Atmos. Meas. Tech. 11, 369–383 (2018).

    Google Scholar 

  37. Cai, Y., Montague, D. C., Mooiweer-Bryan, W. & Deshler, T. Performance characteristics of the ultra high sensitivity aerosol spectrometer for particles between 55 and 800 nm: laboratory and field studies. J. Aerosol Sci. 39, 759–769 (2008).

    ADS  CAS  Google Scholar 

  38. Murphy, D. M. Time offsets and power spectra of the ER-2 data set from the 1987 airborne antarctic ozone experiment. J. Geophys. Res. Atmos. 94, 16737–16748 (1989).

    ADS  Google Scholar 

  39. Weber, R. J. et al. Spurious aerosol measurements when sampling from aircraft in the vicinity of clouds. J. Geophys. Res. Atmos. 103, 28337–28346 (1998).

    ADS  CAS  Google Scholar 

  40. Seinfeld, J. H. & Pandis, S. N. Atmospheric Chemistry and Physics: From Air Pollution to Climate Change 1st edn (Wiley, 1998).

  41. Lovejoy, E. R., Curtius, J. & Froyd, K. D. Atmospheric ion-induced nucleation of sulfuric acid and water. J. Geophys. Res. Atmos. 109, D08204 (2004).

    ADS  Google Scholar 

  42. Kim, S. et al. PubChem substance and compound databases. Nucleic Acids Res. 44, D1202–D1213 (2016).

    CAS  PubMed  Google Scholar 

  43. Westervelt, D. M. et al. Formation and growth of nucleated particles into cloud condensation nuclei: model-measurement comparison. Atmos. Chem. Phys. 13, 7645–7663 (2013).

    ADS  Google Scholar 

  44. Bowman, K. P. Large-scale isentropic mixing properties of the Antarctic polar vortex from analyzed winds. J. Geophys. Res. Atmos. 98, 23013–23027 (1993).

    ADS  Google Scholar 

  45. NCAR UCAR Research Data Archive/Computational and Information Systems Lab. NCEP GFS 0.25-degree global forecast grids historical archive. https://rda.ucar.edu/datasets/ds084.1/ (2015, updated daily; accessed 28 May 2018).

  46. NASA Langley Cloud and Radiation Research. Satellite imagery and cloud products page www.satcorps.larc.nasa.gov (accessed 28 May 2018). 

  47. Murphy, D. M. et al. Single-particle mass spectrometry of tropospheric aerosol particles. J. Geophys. Res. Atmos. 111, D23S32 (2006).

    ADS  Google Scholar 

  48. DeCarlo, P. F. et al. Field-deployable, high-resolution, time-of-flight aerosol mass spectrometer. Anal. Chem. 78, 8281–8289 (2006).

    CAS  Google Scholar 

  49. Canagaratna, M. R. et al. Chemical and microphysical characterization of ambient aerosols with the aerodyne aerosol mass spectrometer. Mass Spectrom. Rev. 26, 185–222 (2007).

    ADS  CAS  Google Scholar 

  50. Schroder, J. C. et al. Sources and secondary production of organic aerosols in the northeastern United States during WINTER. J. Geophys. Res. Atmos. 123, 7771–7796 (2018).

    ADS  CAS  Google Scholar 

  51. Stith, J. L. et al. An overview of aircraft observations from the Pacific Dust Experiment campaign. J. Geophys. Res. Atmos. 114, D05207 (2009).

    ADS  Google Scholar 

  52. Kodros, J. K., Cucinotta, R., Ridley, D. A., Wiedinmyer, C. & Pierce, J. R. The aerosol radiative effects of uncontrolled combustion of domestic waste. Atmos. Chem. Phys. 16, 6771–6784 (2016).

    ADS  CAS  Google Scholar 

  53. Yu, F. & Luo, G. Simulation of particle size distribution with a global aerosol model: contribution of nucleation to aerosol and CCN number concentrations. Atmos. Chem. Phys. 9, 7691–7710 (2009).

    ADS  CAS  Google Scholar 

  54. Toon, O. B., Turco, R. P., Westphal, D., Malone, R. & Liu, M. S. A multidimensional model for aerosols: description of computational analogs. J. Atmos. Sci. 45, 2123–2143 (1988).

    ADS  Google Scholar 

  55. Yu, P. et al. Evaluations of tropospheric aerosol properties simulated by the community earth system model with a sectional aerosol microphysics scheme. J. Adv. Model. Earth Syst. 7, 865–914 (2015).

    ADS  PubMed  PubMed Central  Google Scholar 

  56. Yu, P. et al. Efficient in-cloud removal of aerosols by deep convection. Geophys. Res. Lett. 46, 1061–1069 (2019).

    ADS  Google Scholar 

  57. Lamarque, J.F. et al. CAM-chem: description and evaluation of interactive atmospheric chemistry in the Community Earth System Model. Geosci. Model Dev. 5, 369–411 (2012).

    ADS  CAS  Google Scholar 

  58. Luo, G. & Yu, F. Simulation of particle formation and number concentration over the Eastern United States with the WRF-Chem + APM model. Atmos. Chem. Phys. 11, 11521–11533 (2011).

    ADS  CAS  Google Scholar 

  59. Granier, C. et al. POET, a database of surface emissions of ozone precursors. http://www.aero.jussieu.fr/projet/ACCENT/POET.php (2005; accessed 23 July 2018).

  60. Keller, C. A. et al. HEMCO v1.0: a versatile, ESMF-compliant component for calculating emissions in atmospheric models. Geosci. Model Dev. 7, 1409–1417 (2014).

    ADS  Google Scholar 

  61. Lamarque, J. F. et al. Historical (1850–2000) gridded anthropogenic and biomass burning emissions of reactive gases and aerosols: methodology and application. Atmos. Chem. Phys. 10, 7017–7039 (2010).

    ADS  CAS  Google Scholar 

  62. Gong, S. L. A parameterization of sea-salt aerosol source function for sub-and super-micron particles. Global Biogeochem. Cycles 17, 1097 (2003).

    ADS  Google Scholar 

  63. Mahowald, N. M. et al. Change in atmospheric mineral aerosols in response to climate: last glacial period, preindustrial, modern, and doubled carbon dioxide climates. J. Geophys. Res. 111, D10202 (2006).

    ADS  Google Scholar 

  64. Kodros, J. K. & Pierce, J. R. Important global and regional differences in cloud-albedo aerosol indirect effect estimates between simulations with and without prognostic aerosol microphysics. J. Geophys. Res. 122, 4003–4018 (2017).

    Google Scholar 

  65. Williamson, C. J. et al. ATom: tropical new particle formation measurements and global-model outputs. https://doi.org/10.3334/ORNLDAAC/1684 (ORNL DAAC, Oak Ridge, TN, USA;2019).

    Book  Google Scholar 

  66. NCAR/UCAR Community Earth System Model CESM1.0 Public Release. http://www.cesm.ucar.edu/models/cesm1.0/ (accessed 19 February 2019).

  67. GEOS-Chem Online User’s Guide. http://acmg.seas.harvard.edu/geos/doc/man/ (accessed 19 February 2019).

  68. Natural Earth Free Vector and Raster Map Data. http://www.naturalearthdata.com (accessed 10 December 2015).

Download references

Acknowledgements

We thank J. Kazil, G. Feingold, T. Goren, D. Fahey and K. Aikin for contributing to this analysis, and the ATom leadership team, science team and crew for contributions to the ATom measurements. We acknowledge support from the US NASA Earth System Science Pathfinder Program under awards NNH15AB12I, NNX15AJ23G, NNX15AH33A and 80NSSC19K0124, and from the US National Oceanic and Atmospheric Administration (NOAA) Health of the Atmosphere and Atmospheric Chemistry, Carbon Cycle, and Climate Programs. A.K. was supported by the Austrian Science Fund’s Erwin Schrodinger Fellowship J-3613. J.R.P., J.K.K., K.R.B. and A.L.H. were supported by the US Department of Energy’s Atmospheric System Research, an Office of Science, Office of Biological and Environmental Research program, under grants DE-SC0019000 and DE-SC0011780; the US National Science Foundation (NSF), Atmospheric Chemistry program, under grant AGS-1559607; and the NOAA, Office of Science, Office of Atmospheric Chemistry, Carbon Cycle, and Climate Program, under cooperative agreement award NA17OAR430001. G.L. and F.Y. acknowledge funding support from NASA under grant NNX17AG35G and from the NSF under grant AGS-1550816. B.W. and M.D. have received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation framework program under grant 640458 (A-LIFE) and from the University of Vienna. P.Y. acknowledges the second Tibetan Plateau Scientific Expedition and Research Program (2019QZKK0604).

Author information

Authors and Affiliations

Authors

Contributions

C.J.W., A.K., C.A.B., M.D., B.W., K.D.F., D.M.M., P.C.-J., B.A.N., J.L.J. and T.B. collected the data, and C.J.W. wrote the manuscript with contributions from C.A.B., A.K., J.R.P., K.D.F., D.M.M., P.C.-J., J.L.J. and E.A.R. C.J.W., A.K. and C.A.B. analysed the size distributions. M.D. and B.W. analysed cloud properties. K.D.F. and D.M.M. analysed single-particle compositions, and P.C.-J., B.A.N. and J.L.J. analysed bulk particle composition. J.K.K., A.L.H., K.R.B. and J.R.P. ran GEOS-Chem-TOMAS and J.R.P. developed methods for understanding relevant tropical dynamics. G.L. and F.Y. ran GEOS-Chem-APM and CAM5-APM, and P.Y. ran CESM-CARMA. E.A.R. ran the ATom back trajectories. D.A. and J.C.W. developed the relative differencing method.

Corresponding author

Correspondence to Christina J. Williamson.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

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

Peer review information Nature thanks Thorsten Hoffmann and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Extended data figures and tables

Extended Data Fig. 1 Location of ATom measurements.

a, ATom 1 (gold) and ATom 2 (blue) measurements by latitude and longitude. b, c, Altitude and latitude of measurements over the Pacific Ocean (b) and Atlantic Ocean (c). TCRs are highlighted in red (for ATom 1) and dark blue (for ATom 2). The world map was made with Natural Earth68

Source Data.

Extended Data Fig. 2 Identifying the tropical convective region.

Shown are average measured relative humidity over water (RHwater; blue); number concentration of particles of 3 nm or more (N3 nm; red); and cloud fraction from reanalysis meteorology (dashed black) at pressures between 200 hPa and 400 hPa. a, b, ATom 1 Pacific (a) and Atlantic (b) transects. c, d, ATom 2 Pacific (c) and Atlantic (d) transects. We take the central peak in relative humidity to be the intertropical convergence zone (ITCZ), and define a tropical convective region (TCR) between the minima on either side of this peak (grey shaded region). These minima correspond to latitudes 2.5° N to 17.5° N for ATom 1 Pacific; 2.5° N to 27.5° N for ATom 1 Atlantic; 27.5° S to 2.5° S for ATom 2 Pacific; and 7.5° S to 22.5° S for ATom 2 Atlantic

Source Data.

Extended Data Fig. 3 Modelled global concentrations of particles larger than 3 nm.

Shown are the monthly mean number concentration of particles bigger than 3 nm in the free troposphere at pressures less than 600 hPa (weighted by grid-box height), modelled for August 2016 (left) and February 2017 (right). a, b, CESM-CARMA; c, d, CAM5-APM; e, f, GEOS-Chem-APM; and g, h, GEOS-Chem-TOMAS. Horizontal black lines mark the TCR defined by the ATom data. GEOS-Chem-TOMAS (g, h) shows higher number concentrations of N3 particles outside the TCR than do the other models. This is partly the effect of the 2009 volcanic emissions, which are included in the emissions database for this model. The world maps are made using Natural Earth68

Source Data.

Extended Data Fig. 4 Condensation and coagulation rates.

a, Relationship between the gas-phase condensation rate onto particles larger than 7 nm, the coagulation rate between 5-nm particles and all other particles, and the total aerosol surface area (2.6–4,800 nm), which show a strongly linear relationship. b, The contribution of each particle-size mode to the average condensation rate in the TCR as a function of pressure. The graph shows that the coagulation sink is not always dominated by particles larger than 60 nm, but that, especially at high altitudes, particles of 12–60 nm (the Aitken mode) can dominate the coagulation sink, and must therefore be considered. Lines show the 50th percentile and shaded areas the 25th to 75th percentile range

Source Data.

Extended Data Fig. 5 TCR back trajectories.

30-day back trajectories were calculated every minute of flight time within the TCRs. ad, Pressures from the time of minimum pressure of the trajectory to the flight track are plotted for all back trajectories from ATom 1 Pacific (a) and Atlantic (b) and ATom 2 Pacific (c) and Atlantic (d) observations. Colours distinguish separate trajectories. The general slope of increasing pressure with time indicates a general descending motion of the air. e, Histogram showing instantaneous descent rates (one point every 3 h) for all trajectories within the indicated pressure bins. The skew at all altitudes towards positive descent rates is evidence of an overall descending motion of the air. The mean descent rate is higher at higher altitudes, and almost 0 at the lowest altitudes, which is to be expected as this is often within the marine boundary layer where the air cannot descend further. f, Average fraction of time that trajectories spent in cloud between the time of minimum pressure and the flight track. In-cloud time is taken to be times when relative humidities are 90% or more (an overestimate). It is binned by the pressure on the flight track (not the pressure of the trajectory itself, as in e). For measurements made at pressures of less than 850 hPa, the air spent less than 5% of its time in cloud on average. For air at pressures of more than 850 hPa, this time increased to around 14%. This shows that most of the particles descend with the air instead of being removed by clouds. g, h, Histograms of the latitudes of trajectories between the minimum pressure and flight track for ATom 1 (g) and ATom 2 (h), coloured by pressure of the point on the flight track. Except for at the lowest altitudes, air parcels entering the flight track mostly remain within the tropics (with histograms peaking around the equator). Peaks shift towards the summer hemisphere with the season, in the same manner as the TCRs

Source Data.

Extended Data Fig. 6 Average size distributions.

ac, Average size distributions in TCRs at pressures between 250 hPa and 300 hPa, 600 hPa and 800 hPa, and 800 hPa and 1,000 hPa respectively. Regions of biomass burning or dust plumes have been excluded. Except for particles larger than 100 nm in CAM5-APM, the models show fewer particles than do the ATom observations. All models except for CAM5-APM also show strong evidence of cloud processing, in the form of the dip in the size distribution around 60–100 nm in b, c. This indication of cloud processing25 is far less pronounced in the ATom data, suggesting that too many particles are being cloud processed in the models. b also shows average size distributions measured between 900 hPa and 1,000 hPa over the contiguous USA and Alaska, as examples of continental size distributions. BL, boundary layer

Source Data.

Extended Data Fig. 7 Chemical composition of particles in the TCRs.

a, Organic and sulfate mass of particles, measured by the AMS (for particles of 50–500 nm) and PALMS (for particles of 150–500 nm), and ambient relative humidity (RHw), with data affected by biomass burning and dust plumes being removed. (The AMS excludes ATom 2 measurements for the Pacific Ocean, where the overall mass was too low to measure sulfate and organic components.) Both methods for measuring organic and sulfate mass have limitations in this regime (the AMS is close to the detection limit, and PALMS cannot measure particles smaller than 150 nm), so perfect agreement is not expected. However, the low sulfate mass at high altitude seems robust, as it is supported by both measurements, and differences between the organic concentrations obtained by PALMS and AMS suggest that organics dominate the composition of smaller particles at high altitudes. b, Size-resolved volume in the TCRs. Between 400 hPa and 800 hPa, the median diameter by volume, and the majority of the aerosol volume, is within the measured 50–500-nm range. The composition results should thus be regarded as informative within this pressure range. dV/dlogDp, log-normalized volume concentration. c, d, Modelled sulfate and organic masses over the same regions, compared with measurements

Source Data.

Extended Data Fig. 8 Details from NMASS data.

a, Identifying instrumental noise in particle measurements, using Fourier transforms of the number concentration of particles larger than 3 nm (N3) as measured in the first NMASS channel for 30 min of data at different altitudes and different total concentrations. b, Concentrations from the first and second channels of NMASS 1 (red solid and blue dotted lines respectively), and the calculated relative difference (black), for an example time period. Missing data occur when the aircraft flew through a cloud and the data were discarded. A relative difference of three standard deviations (3 sigma) is shown with a dashed horizontal line. Any relative difference larger than this is considered statistically significant

Source Data.

Extended Data Fig. 9 Sensitivity studies and radiative effects.

af, Modelled N60 values (ac) and condensation sinks (df) in the GEOS-Chem-TOMAS (a, d), GEOS-Chem-APM (b, e) and CESM-CARMA (c, f) models, showing the base model run (red), the nucleation rates increased by a factor of ten between 28° N and 28° S at pressures of less than 600 hPa (orange), oceanic emissions of DMS tripled (light blue), cloud processing on descent reduced by a factor of ten (dark blue), and nucleation turned off between 28° N and 28° S at pressures of less than 600 hPa (dashed). These modelling results are compared with the ATom 1 Pacific observations (black and grey). gj, Simulated aerosol indirect effect (difference in radiative effect) of tropical upper tropospheric NPF (g, h, GEOS-Chem-TOMAS; i, j, GEOS-Chem-APM), for the base (g, i) and reduced cloud processing (h, j) cases, calculated by turning off nucleation at 28° N to 28° S at pressures below 600 hPa. k, The seasonal cycle of this global aerosol indirect effect from GEOS-Chem-TOMAS, for the base (red) and reduced cloud processing (blue) cases. Maps made using Natural Earth68

Source Data.

Extended Data Table 1 Relevant properties of the models used here

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Williamson, C.J., Kupc, A., Axisa, D. et al. A large source of cloud condensation nuclei from new particle formation in the tropics. Nature 574, 399–403 (2019). https://doi.org/10.1038/s41586-019-1638-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-019-1638-9

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