Advanced Search
Article Contents

The Effect of Super Volcanic Eruptions on Ozone Depletion in a Chemistry-Climate Model


doi: 10.1007/s00376-019-8241-8

  • With the gradual yet unequivocal phasing out of ozone depleting substances (ODSs), the environmental crisis caused by the discovery of an ozone hole over the Antarctic has lessened in severity and a promising recovery of the ozone layer is predicted in this century. However, strong volcanic activity can also cause ozone depletion that might be severe enough to threaten the existence of life on Earth. In this study, a transport model and a coupled chemistry-climate model were used to simulate the impacts of super volcanoes on ozone depletion. The volcanic eruptions in the experiments were the 1991 Mount Pinatubo eruption and a 100× Pinatubo size eruption. The results show that the percentage of global mean total column ozone depletion in the 2050 RCP8.5 100× Pinatubo scenario is approximately 6% compared to two years before the eruption and 6.4% in tropics. An identical simulation, 100× Pinatubo eruption only with natural source ODSs, produces an ozone depletion of 2.5% compared to two years before the eruption, and with 4.4% loss in the tropics. Based on the model results, the reduced ODSs and stratospheric cooling lighten the ozone depletion after super volcanic eruption.
    摘要: 随着平流层中臭氧损耗物质(ODSs)的不断清除,由南极上空臭氧空洞的发现所引发的环境危机正在减轻,并且臭氧层也在逐渐恢复。然而,强火山活动同样会造成严重的臭氧损耗,从而威胁地球上的生命。在本研究中,利用了一个大气输送模式和一个化学气候模式,来模拟超级火山爆发后造成的臭氧损耗。模拟的火山事件为1991年的皮纳图博(Pinatubo)火山以及一个100×Pinatubo级别的火山。结果表明,在2050 RCP8.5 100×Pinatubo情形下,全球平均臭氧损耗和爆发前两年相比大约为6%,赤道地区为6.4%。而一个理想情形,即100×Pinatubo在自然源ODSs全部清除,只剩自然源的背景下爆发后,全球平均臭氧损耗和爆发前两年相比为2.5%,赤道地区为4.4%。根据模式结果,ODSs含量的下降以及平流层降温能够减轻超级火山爆发后造成的臭氧损耗。
  • 加载中
  • Ambrose S. H.,1998: Late Pleistocene human population bottlenecks, volcanic winter, and differentiation of modern humans. Journal of Human Evolution, 34, 623-651, .https://doi.org/10.1006/jhev.1998.0219
    Ansmann A.,I. Mattis, U. Wand inger, F. Wagner, J. Reichardt, and T. Deshler, 1997: Evolution of the Pinatubo aerosol: Raman lidar observations of particle optical depth, effective radius, mass, and surface area over Central Europe at 53.4°N. J. Atmos. Sci., 54, 2630-2641, .https://doi.org/10.1175/1520-0469(1997)054<2630:EOTPAR>2.0.CO;2
    Aquila V.,L. D. Oman, R. Stolarski, A. R. Douglass, and P. A. Newman, 2013: The response of ozone and nitrogen dioxide to the eruption of Mt. Pinatubo at southern and northern midlatitudes. J. Atmos. Sci., 70, 894-900, .https://doi.org/10.1175/jas-d-12-0143.1
    Baldwin, M. P.,Coauthors, 2001: The quasi-biennial oscillation. Rev. Geophys., 39, 179-229, .https://doi.org/10.1029/1999rg000073
    Bekki S.,1995: Oxidation of volcanic SO2: A sink for stratospheric OH and H2O. Geophys. Res. Lett., 22, 913-916, .https://doi.org/10.1029/95gl00534
    Betts A. K.,1986: A new convective adjustment scheme. Part I: Observational and theoretical basis. Quart. J. Roy. Meteor. Soc., 112, 677-691, .https://doi.org/10.1002/qj.49711247307
    Bluth G. J. S.,S. D. Doiron, C. C. Schnetzler, A. J. Krueger, and L. S. Walter, 1992: Global tracking of the SO2 clouds from the June, 1991 Mount-Pinatubo eruptions. Geophys. Res. Lett., 19, 151-154, .https://doi.org/10.1029/91gl02792
    Bobrowski N.,G. Hönninger B. Galle, and U. Platt, 2003: Detection of bromine monoxide in a volcanic plume. Nature, 423, 273-276, .https://doi.org/10.1038/nature01625
    Cadoux A.,B. Scaillet, S. Bekki, C. Oppenheimer, and T. H. Druitt, 2015: Stratospheric ozone destruction by the Bronze-Age Minoan eruption (Santorini Volcano, Greece). Scientific Reports, 5, 12243, .https://doi.org/10.1038/srep12243
    Cagnazzo, C., Coauthors, 2009: Northern winter stratospheric temperature and ozone responses to ENSO inferred from an ensemble of chemistry climate models. Atmospheric Chemistry and Physics, 9, 8935-8948, .https://doi.org/10.5194/acp-9-8935-2009
    Carslaw K. S.,B. P. Luo, S. L. Clegg, T. Peter, P. Brimblecombe, and P. J. Crutzen, 1994: Stratospheric aerosol growth and HNO3 gas-phase depletion from coupled HNO3 and water uptake by liquid particles. Geophys. Res. Lett., 21, 2479-2482, .https://doi.org/10.1029/94GL02799
    Chen W.,T. Li, 2007: Modulation of northern hemisphere wintertime stationary planetary wave activity: East Asian climate relationships by the Quasi-Biennial Oscillation. J. Geophys. Res., 112, D20120, .https://doi.org/10.1029/2007jd008611
    Chen W.,K. Wei, 2009: Interannual variability of the winter stratospheric polar vortex in the northern hemisphere and their relations to QBO and ENSO. Adv. Atmos. Sci., 26(5), 855-863, .https://doi.org/10.1007/s00376-009-8168-6
    Coffey M. T.,1996: Observations of the impact of volcanic activity on stratospheric chemistry. J. Geophys. Res., 101, 6767-6780, .https://doi.org/10.1029/95jd03763
    Crutzen P. J.,F. Arnold, 1986: Nitric acid cloud formation in the cold antarctic stratosphere: A major cause for the springtime "ozone hole". Nature, 324, 651-655, .https://doi.org/10.1038/324651a0
    Dee D. P.,S. Uppala, 2009: Variational bias correction of satellite radiance data in the ERA-Interim reanalysis. Quart. J. Roy. Meteor. Soc., 135, 1830-1841, .https://doi.org/10.1002/qj.493
    Delwiche C. F.,2005: Voyage to the bottom of the tree. Science, 307, 676-677, .https://doi.org/10.1126/science.1105582
    Dvortsov V. L.,S. G. Zvenigorodsky, and S. P. Smyslaev, 1992: On the use of isaksen-luther method of computing photodissociation rates in photochemical models. J. Geophys. Res., 97, 7593-7601, .https://doi.org/10.1029/91JD02861
    Engel A., H. Bönisch, J. Ostermöller, M. P. Chipperfield, S. Dhomse, and P. Jöckel, 2018: A refined method for calculating equivalent effective stratospheric chlorine. Atmospheric Chemistry and Physics, 18, 601-619, .https://doi.org/10.5194/acp-18-601-2018
    Fahey, D. W.,Coauthors, 1993: In situ measurements constraining the role of sulphate aerosols in mid-latitude ozone depletion. Nature, 363, 509- 514.10.1038/363509a098690d7b49672754faf9f72558caa032http%3A%2F%2Fwww.nature.com%2Fdoifinder%2F10.1038%2F363509a0
    Farman J. C.,B. G. Gardiner, and J. D. Shanklin, 1985: Large losses of total ozone in antarctica reveal seasonal CLOx/NOx interaction. Nature, 315, 207-210, .https://doi.org/10.1038/315207a0
    Free M.,J. K. Angell, 2002: Effect of volcanoes on the vertical temperature profile in radiosonde data. J. Geophys. Res., 107, 4101, .https://doi.org/10.1029/2001jd001128
    Free M.,J. Lanzante, 2009: Effect of volcanic eruptions on the vertical temperature profile in radiosonde data and climate models. J. Climate, 22, 2925-2939, .https://doi.org/10.1175/2008jcli2562.1
    Galin V. Y.,S. P. Smyshlyaev, and E. M. Volodin, 2007: Combined chemistry-climate model of the atmosphere. Izvestiya, Atmospheric and Oceanic Physics, 43, 399-412, .https://doi.org/10.1134/s0001433807040020
    Graf H.-F.,D. Zanchettin, C. Timmreck, and M. Bittner, 2014: Observational constraints on the tropospheric and near-surface winter signature of the northern hemisphere stratospheric polar vortex. Climate Dyn., 43, 3245-3266, .https://doi.org/10.1007/s00382-014-2101-0
    Hamilton K.,1993: An examination of observed southern oscillation effects in the northern hemisphere stratosphere. J. Atmos. Sci., 50, 3468-3474, .https://doi.org/10.1175/1520-0469(1993)050<3468:AEOOSO>2.0.CO;2
    Hansen J.,M. Sato, G. Russell, and P. Kharecha, 2013: Climate sensitivity, sea level and atmospheric carbon dioxide. Philosophical Transactions of the Society A: Mathematical, Physical and Engineering Sciences, 371, 20120294, .https://doi.org/10.1098/rsta.2012.0294
    Hanson D. R.,A. R. Ravishankara, 1993: Reaction of ClONO2 with HCl on NAT, NAD, and frozen sulfuric-acid and hydrolysis of N2O5 and ClONO2 on frozen sulfuric-acid. J. Geophys. Res., 98, 22 931-22 936, .https://doi.org/10.1029/93jd01929
    Hines C. O.,1997: Doppler-spread parameterization of gravity-wave momentum deposition in the middle atmosphere. Part 1: Basic formulation. Journal of Atmospheric and Solar-Terrestrial Physics, 59, 371-386, .https://doi.org/10.1016/s1364-6826(96)00079-X
    Hoffmann L.,T. Rößler S. Griessbach, Y. Heng, and O. Stein, 2016: Lagrangian transport simulations of volcanic sulfur dioxide emissions: Impact of meteorological data products. J. Geophys. Res., 121, 4651-4673, .https://doi.org/10.1002/2015jd023749
    Hofmann, D. J.,Coauthors, 1994: Ozone loss in the lower stratosphere over the United-States in 1992-1993: Evidence for heterogeneous chemistry on the pinatubo aerosol. Geophys. Res. Lett., 21, 65-68, .https://doi.org/10.1029/93gl02526
    Holton J. R.,H. C. Tan, 1980: The influence of the equatorial quasi-biennial oscillation on the global circulation at 50 mb. J. Atmos. Sci., 37, 2200-2208, .https://doi.org/10.1175/1520-0469(1980)037<2200:TIOTEQ>2.0.CO;2
    Hunton, D. E.,Coauthors, 2005: In-situ aircraft observations of the 2000 Mt. Hekla volcanic cloud: Composition and chemical evolution in the Arctic lower stratosphere. Journal of Volcanology and Geothermal Research, 145, 23-34, .https://doi.org/10.1016/j.jvolgeores.2005.01.005
    Joshi M. M.,G. S. Jones, 2009: The climatic effects of the direct injection of water vapour into the stratosphere by large volcanic eruptions. Atmospheric Chemistry and Physics, 9, 6109-6118, .https://doi.org/10.5194/acp-9-6109-2009
    Jouzel, J., Coauthors, 2007: Orbital and millennial Antarctic climate variability over the past 800, 000 years.Science, 317, 793-796, .https://doi.org/10.1126/science.1141038
    Klobas J. E.,D. M. Wilmouth, D. K. Weisenstein, J. G. Anderson, and R. J. Salawitch, 2017: Ozone depletion following future volcanic eruptions. Geophys. Res. Lett., 44, 7490-7499, .https://doi.org/10.1002/2017GL073972
    Kremser, S., Coauthors, 2016: Stratospheric aerosol-observations, processes, and impact on climate. Rev. Geophys., 54, 278-335, .https://doi.org/10.1002/2015rg000511
    Kutterolf S.,T. H. Hansteen, K. Appel, A. Freundt, K. Krüger W. Pérez, and H. Wehrmann, 2013: Combined bromine and chlorine release from large explosive volcanic eruptions: A threat to stratospheric ozone? Geology, 41, 707-710, .https://doi.org/10.1130/g34044.1
    Lanzante J. R.,2007: Diagnosis of radiosonde vertical temperature trend profiles: Comparing the influence of data homogenization versus model forcings. J. Climate, 20, 5356-5364, .https://doi.org/10.1175/2007jcli1827.1
    Lean J. L.,D. H. Rind, 2009: How will Earth's surface temperature change in future decades? Geophys. Res. Lett., 36, L15708, .https://doi.org/10.1029/2009gl038932
    LeGrand e, A. N., K. Tsigaridis, S. E. Bauer, 2016: Role of atmospheric chemistry in the climate impacts of stratospheric volcanic injections. Nature Geoscience, 9, 652-655, .https://doi.org/10.1038/ngeo2771
    Legras B.,B. Joseph, and F. Lefèvre, 2003: Vertical diffusivity in the lower stratosphere from Lagrangian back-trajectory reconstructions of ozone profiles. J. Geophys. Res., 108, 4562, .https://doi.org/Artn456210.1029/2002jd003045
    Legras B.,I. Pisso, G. Berthet, and F. Lefèvre, 2005: Variability of the Lagrangian turbulent diffusion in the lower stratosphere. Atmospheric Chemistry and Physics, 5, 1605-1622, .https://doi.org/10.5194/acp-5-1605-2005
    Lisiecki L. E.,M. E. Raymo, 2005: A Pliocene-Pleistocene stack of 57 globally distributed benthic delta $\delta$18O records. Paleoceanography and Paleoclimatology, 20, PA1003, .https://doi.org/10.1029/2004pa001071
    Lurton T.,F. Jégou G. Berthet, J. B. Renard, L. Clarisse, A. Schmidt, C. Brogniez, and T. J. Roberts, 2018: Model simulations of the chemical and aerosol microphysical evolution of the Sarychev Peak 2009 eruption cloud compared to in situ and satellite observations. Atmospheric Chemistry and Physics, 18, 3223-3247, .https://doi.org/10.5194/acp-18-3223-2018
    Marshall, L., Coauthors, 2018: Multi-model comparison of the volcanic sulfate deposition from the 1815 eruption of Mt. Tambora. Atmospheric Chemistry and Physics, 18, 2307-2328, .https://doi.org/10.5194/acp-18-2307-2018
    Mather T. A.,2015: Volcanoes and the environment: Lessons for understanding Earth's past and future from studies of present-day volcanic emissions. Journal of Volcanology and Geothermal Research, 304, 160-179, .https://doi.org/10.1016/j.jvolgeores.2015.08.016
    McCormick M. P.,L. W. Thomason, and C. R. Trepte, 1995: Atmospheric effects of the Mt Pinatubo eruption. Nature, 373, 399-404, .https://doi.org/10.1038/373399a0
    McGee T. J.,P. Newman, M. Gross, U. Singh, S. Godin, A. M. Lacoste, and G. Megie, 1994: Correlation of ozone loss with the presence of volcanic aerosols. Geophys. Res. Lett., 21, 2801-2804, .https://doi.org/10.1029/94gl02350
    Muthers S.,F. Arfeuille, C. C. Raible, and E. Rozanov, 2015: The impacts of volcanic aerosol on stratospheric ozone and the northern hemisphere polar vortex: Separating radiative-dynamical changes from direct effects due to enhanced aerosol heterogeneous chemistry. Atmospheric Chemistry and Physics, 15, 11 461-11 476, .https://doi.org/10.5194/acp-15-11461-2015
    Oppenheimer C.,2002: Limited global change due to the largest known Quaternary eruption, Toba ≈74 kyr BP? Quaternary Science Reviews, 21, 1593-1609, .https://doi.org/10.1016/S0277-3791(01)00154-8
    Oppenheimer, C., Coauthors, 2010: Atmospheric chemistry of an Antarctic volcanic plume. J. Geophys. Res., 115, D04303, .https://doi.org/10.1029/2009jd011910
    Palmer T. N.,G. J. Shutts, and R. Swinbank, 1986: Alleviation of a systematic westerly bias in general circulation and numerical weather prediction models through an orographic gravity wave drag parametrization. Quart. J. Roy. Meteor. Soc., 112, 1001-1039, .https://doi.org/10.1002/qj.49711247406
    Peter T.,1997: Microphysics and heterogeneous chemistry of polar stratospheric clouds. Annu. Rev. Phys. Chem., 48, 785-822, .https://doi.org/10.1146/annurev.physchem.48.1.785
    Petraglia M. D.,R. Korisettar, and J. N. Pal, 2012: The Toba volcanic super-eruption of 74, 000 years ago: Climate change, environments, and evolving humans.Quaternary International, 258, 1-4, .https://doi.org/10.1016/j.quaint.2011.12.001
    Pisso I.,E. Real, K. S. Law, B. Legras, N. Bousserez, J. L. Attié, and H. Schlager, 2009: Estimation of mixing in the troposphere from Lagrangian trace gas reconstructions during long-range pollution plume transport. J. Geophys. Res., 114, 12882, .https://doi.org/10.1029/2008jd011289
    Poberaj C. S.,J. Staehelin, and D. Brunner, 2011: Missing stratospheric ozone decrease at southern hemisphere middle latitudes after Mt. Pinatubo: A dynamical perspective. J. Atmos. Sci., 68, 1922-1945, .https://doi.org/10.1175/JAS-D-10-05004.1
    Rampino M. R.,2002: Supereruptions as a threat to civilizations on earth-like planets. Icarus, 156, 562-569, .https://doi.org/10.1006/icar.2001.6808
    Rampino M. R.,S. Self, 1992: Volcanic winter and accelerated glaciation following the Toba super-eruption. Nature, 359, 50-52, .https://doi.org/10.1038/359050a0
    Rand el, W. J., F. Wu, J. M. Russell III, J. W. Waters, L. Froidevaux, 1995: Ozone and temperature-changes in the stratosphere following the eruption of Mount-Pinatubo. J. Geophys. Res., 100, 16 753-16 764, .https://doi.org/10.1029/95jd01001
    Rand el, W. J., A. K. Smith, F. Wu, C.-Z. Zou, H. F. Qian, 2016: Stratospheric temperature trends over 1979-2015 derived from combined SSU, MLS, and SABER satellite observations. J. Climate, 29, 4843-4859, .https://doi.org/10.1175/JCLI-D-15-0629.1
    Ren R.-C.,M. Cai, C. Y. Xiang, and G. X. Wu, 2012: Observational evidence of the delayed response of stratospheric polar vortex variability to ENSO SST anomalies. Climate Dyn., 38, 1345-1358, .https://doi.org/10.1007/s00382-011-1137-7
    Robock A.,2000: Volcanic eruptions and climate. Rev. Geophys., 38, 191-219, .https://doi.org/10.1029/1998rg000054
    Robock A.,C. M. Ammann, L. Oman, D. Shindell, S. Levis, and G. Stenchikov, 2009: Did the Toba volcanic eruption of ~74 ka B. P. produce widespread glaciation? J. Geophys. Res., 114, D10107, .https://doi.org/10.1029/2008jd011652
    Roscoe H. K.,2001: The risk of large volcanic eruptions and the impact of this risk on future ozone depletion. Natural Hazards, 23, 231-246, .https://doi.org/10.1023/A:1011178016473
    Rosi M., M. Paladio-Melosantos, A. Di Muro, R. Leoni, and T. Bacolcol, 2001: Fall vs flow activity during the 1991 climactic eruption of Pinatubo Volcano (Philippines). Bulletin of Volcanology, 62, 549-566, .https://doi.org/10.1007/s004450000118
    Rößler, T., O. Stein, Y. Heng, P. Baumeister, L. Hoffmann, 2018: Trajectory errors of different numerical integration schemes diagnosed with the MPTRAC advection module driven by ECMWF operational analyses. Geoscientific Model Development, 11, 575-592, .https://doi.org/10.5194/gmd-11-575-2018
    Sand er, S. P.,Coauthors, 2003: Chemical kinetics and photochemical data for use in atmospheric studies. Evaluation Number 14. JPL Publication 02-25, National Aeronautics and Space Administration. [Available online from .http://www.iup.uni-bremen.de/\simbms/lectures/JPL_02-25_rev02.pdf
    Santer, B. D.,Coauthors, 2001: Accounting for the effects of volcanoes and ENSO in comparisons of modeled and observed temperature trends. J. Geophys. Res., 106, 28 033-28 059, .https://doi.org/10.1029/2000jd000189
    Schmidt, A., Coauthors, 2016: Selective environmental stress from sulphur emitted by continental flood basalt eruptions. Nature Geoscience, 9, 77-82, .https://doi.org/10.1038/ngeo2588
    Schoeberl M. R.,D. L. Hartmann, 1991: The dynamics of the stratospheric polar vortex and its relation to springtime ozone depletions. Science, 251, 46-52, .https://doi.org/10.1126/science.251.4989.46
    Smyshlyaev S. P.,V. L. Dvortsov, M. A. Geller, and V. A. Yudin, 1998: A two-dimensional model with input parameters from a general circulation model: Ozone sensitivity to different formulations for the longitudinal temperature variation. J. Geophys. Res., 103, 28 373-28 387, .https://doi.org/10.1029/98JD02354
    Smyshlyaev S. P.,V. Y. Galin, G. Shaariibuu, and M. A. Motsakov. 2010: Modeling the variability of gas and aerosol components in the stratosphere of polar regions. Izvestiya, Atmospheric and Oceanic Physics, 46, 265-280, .http://doi.org/10.1134/s0001433810030011
    Solomon S.,1999: Stratospheric ozone depletion: A review of concepts and history. Rev. Geophys., 37, 275-316, .https://doi.org/10.1029/1999rg900008
    Solomon S.,D. Kinnison, J. Band oro, and R. Garcia, 2015: Simulation of polar ozone depletion: An update. J. Geophys. Res., 120, 7958-7974, .https://doi.org/10.1002/2015JD023365
    Solomon S.,D. J. Ivy, D. Kinnison, M. J. Mills, R. R. Neely III, and A. Schmidt, 2016: Emergence of healing in the Antarctic ozone layer. Science, 353, 269-274, .https://doi.org/10.1126/science.aae0061
    Son S.-W.,Y. Lim, C. Yoo, H. H. Hendon, and J. Kim, 2017: Stratospheric control of the Madden-Julian Oscillation. J. Climate, 30, 1909-1922, .https://doi.org/10.1175/Jcli-D-16-0620.1
    SPARC, 2013: SPARC report on the lifetimes of stratospheric ozone-depleting substances, their replacements, and related species. SPARC Report No. 6, WCRP-15/2013.01831ebb2dfd33d77b95506f9b03781chttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2013AGUFM.A23F0354N
    Stenchikov G.,K. Hamilton, R. J. Stouffer, A. Robock, V. Ramaswamy, B. Santer, and H.-F. Graf, 2006: Arctic Oscillation response to volcanic eruptions in the IPCC AR4 climate models. J. Geophys. Res., 111, D07107, .https://doi.org/10.1029/2005jd006286
    Stohl A.,C. Forster, A. Frank, P. Seibert, and G. Wotawa, 2005: Technical note: The Lagrangian particle dispersion model FLEXPART version 6.2. Atmospheric Chemistry and Physics, 5, 2461-2474, .https://doi.org/10.5194/acp-5-2461-2005
    Tabazadeh A.,R. P. Turco, 1993: Stratospheric chlorine injection by volcanic eruptions: HCI scavenging and implications for ozone. Science, 260, 1082-1086, .https://doi.org/10.1126/science.260.5111.1082
    Tabazadeh A.,K. Drdla, M. R. Schoeberl, P. Hamill, and O. B. Toon, 2002: Arctic "ozone hole" in a cold volcanic stratosphere. Proceedings of the National Academy of Sciences of the United States of America, 99, 2609-2612, .https://doi.org/10.1073/pnas.052518199
    Telford P.,P. Braesicke, O. Morgenstern, and J. Pyle, 2009: Reassessment of causes of ozone column variability following the eruption of Mount Pinatubo using a nudged CCM. Atmospheric Chemistry and Physics, 9, 4251-4260, .https://doi.org/10.5194/acp-9-4251-2009
    Vidal C. M.,N. Métrich J.-C. Komorowski, I. Pratomo, A. Michel, N. Kartadinata, V. Robert, and F. Lavigne, 2016: The 1257 Samalas eruption (Lombok, Indonesia): The single greatest stratospheric gas release of the Common Era. Scientific Reports, 6, 34868, .https://doi.org/10.1038/srep34868
    von Glasow, R., N. Bobrowski, C. Kern, 2009: The effects of volcanic eruptions on atmospheric chemistry. Chemical Geology, 263, 131-142, .https://doi.org/10.1016/j.chemgeo.2008.08.020
    Wei K.,W. Chen, and R. H. Huang, 2007: Association of tropical Pacific sea surface temperatures with the stratospheric Holton-Tan Oscillation in the northern hemisphere winter. Geophys. Res. Lett., 34, L16814, .https://doi.org/10.1029/2007GL030478
    Williams M.,2012: Did the 73 ka Toba super-eruption have an enduring effect? Insights from genetics, prehistoric archaeology, pollen analysis, stable isotope geochemistry, geomorphology, ice cores, and climate models. Quaternary International, 269, 87-93, .https://doi.org/10.1016/j.quaint.2011.03.045
    World Meteorological Organization (WMO), 2007: Scientific assessment of ozone depletion: Global ozone research and monitoring project. Report No. 50, Geneva, Switzerland, 572 pp.
    World Meteorological Organization (WMO), 2014: Scientific Assessment of Ozone Depletion, 2014. World Meteorological Organization, Global Ozone Research And Monitoring Project, Report No. 55, Geneva, Switzerland, 416 pp.62532aecb6f24193caef87e55b56b0a3http%3A%2F%2Fwww.indiaenvironmentportal.org.in%2Freports-documents%2Fscientific-assessment-ozone-depletion-2014
    World Meteorological Organization (WMO), 2018: Scientific Assessment of Ozone Depletion, 2018. World Meteorological Organization, Global Ozone Research and Monitoring Project, Report No. 58, Geneva, Switzerland, 588 pp.695d2bf01defe50667bb45dac4d5b912http%3A%2F%2Fwww.cabdirect.org%2Fabstracts%2F20113082496.html
    Wu S. N.,S. T. Chen, and F. C. Duan, 2012: The relation between the 72 ka BP Event and the toba super-eruption. Advance in Earth Science, 27, 35-41, .https://doi.org/10.11867/j.issn.1001-8166.2012.01.0035
    Wu X.,S. Griessbach, and L. Hoffmann, 2017: Equatorward dispersion of a high-latitude volcanic plume and its relation to the Asian summer monsoon: A case study of the Sarychev eruption in 2009. Atmospheric Chemistry and Physics, 17, 13 439-13 455, .https://doi.org/10.5194/acp-17-13439-2017
    Wu X.,S. Griessbach, and L. Hoffmann, 2018: Long-range transport of volcanic aerosol from the 2010 Merapi tropical eruption to Antarctica. Atmospheric Chemistry and Physics, 18, 15 859-15 877, .https://doi.org/10.5194/acp-18-15859-2018
    Wyser K.,1998: The effective radius in large-scale models: Impact of aerosols and coalescence. Atmospheric Research, 49, 213-234, .https://doi.org/10.1016/S0169-8095(98)00081-7
    Xie F.,J. Li, W. Tian, J. Feng, and Y. Huo, 2012: Signals of El Niño Modoki in the tropical tropopause layer and stratosphere. Atmospheric Chemistry and Physics, 12, 5259-5273, .https://doi.org/10.5194/acp-12-5259-2012
    Xie F.,J. P. Li, W. S. Tian, J. K. Zhang, and C. Sun, 2014: The relative impacts of El Niño Modoki, canonical El Niño, and QBO on tropical ozone changes since the 1980s. Environmental Research Letters, 9, 064020, .https://doi.org/10.1088/1748-9326/9/6/064020
    Xie F.,X. Zhou, J. P. Li, C. Sun, J. Feng, and X. Ma, 2018: The key role of background sea surface temperature over the cold tongue in asymmetric responses of the Arctic stratosphere to El Niño-Southern Oscillation. Environmental Research Letters, 13, 114007, .https://doi.org/10.1088/1748-9326/aae79b
    Yu J.-Y.,H.-Y. Kao, T. Lee, and S. T. Kim, 2011: Subsurface ocean temperature indices for Central-Pacific and Eastern-Pacific types of El Niño and La Niña events. Theor. Appl. Climatol., 103, 337-344, .https://doi.org/10.1007/s00704-010-0307-6
    Zhu, Y. Q.,Coauthors, 2018: Stratospheric aerosols, polar stratospheric clouds, and polar ozone depletion after the Mount Calbuco eruption in 2015. J. Geophys. Res., 123, 12 308-12 331, .https://doi.org/10.1029/2018jd028974
  • [1] Wenshou TIAN, Jinlong HUANG, Jiankai ZHANG, Fei XIE, Wuke WANG, Yifeng PENG, 2023: Role of Stratospheric Processes in Climate Change: Advances and Challenges, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 1379-1400.  doi: 10.1007/s00376-023-2341-1
    [2] TIAN Wenshou, Martyn P. CHIPPERFIELD, LU Daren, 2009: Impact of Increasing Stratospheric Water Vapor on Ozone Depletion and Temperature Change, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 423-437.  doi: 10.1007/s00376-009-0423-3
    [3] ZHANG Yuli, LIU Yi, LIU Chuanxi, V. F. SOFIEVA, 2015: Satellite Measurements of the Madden-Julian Oscillation in Wintertime Stratospheric Ozone over the Tibetan Plateau and East Asia, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 1481-1492.  doi: 10.1007/s00376-015-5005-y
    [4] XIE Fei, LI Jianping, TIAN Wenshou, ZHANG Jiankai, SHU Jianchuan, 2014: The Impacts of Two Types of El Nio on Global Ozone Variations in the Last Three Decades, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 1113-1126.  doi: 10.1007/s00376-013-3166-0
    [5] Kequan ZHANG, Jiakang DUAN, Siyi ZHAO, Jiankai ZHANG, James KEEBLE, Hongwen LIU, 2022: Evaluating the Ozone Valley over the Tibetan Plateau in CMIP6 Models, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 1167-1183.  doi: 10.1007/s00376-021-0442-2
    [6] Fei XIE, Yan XIA, Wuhu FENG, Yingli NIU, 2023: Increasing Surface UV Radiation in the Tropics and Northern Mid-Latitudes due to Ozone Depletion after 2010, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 1833-1843.  doi: 10.1007/s00376-023-2354-9
    [7] Shangrong ZHOU, Fei LIU, 2024: Southern Hemisphere Volcanism Triggered Multi-year La Niñas during the Last Millennium, ADVANCES IN ATMOSPHERIC SCIENCES, 41, 587-592.  doi: 10.1007/s00376-023-3254-8
    [8] LU Daren, YI Fan, XU Jiyao, 2004: Advances in Studies of the Middle and Upper Atmosphere and Their Coupling with the Lower Atmosphere, ADVANCES IN ATMOSPHERIC SCIENCES, 21, 361-368.  doi: 10.1007/BF02915564
    [9] Qian Yongfu, 1993: The Climatic Effects of the Stratospheric Volcanic Ash, ADVANCES IN ATMOSPHERIC SCIENCES, 10, 135-146.  doi: 10.1007/BF02919136
    [10] Jingmei Yang, Jinhuan Qiu, 2009: An Empirical Model for Estimating Stratospheric Ozone Vertical Distributions over China, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 352-358.  doi: 10.1007/s00376-009-0352-1
    [11] V. N. R. Mukku, C. S. Bhosale, 1991: Ozone during Stratospheric Warmings at Uccle, ADVANCES IN ATMOSPHERIC SCIENCES, 8, 251-255.  doi: 10.1007/BF02658099
    [12] Yu FU, Hong LIAO, Yang YANG, 2019: Interannual and Decadal Changes in Tropospheric Ozone in China and the Associated Chemistry-Climate Interactions: A Review, ADVANCES IN ATMOSPHERIC SCIENCES, 36, 975-993.  doi: 10.1007/s00376-019-8216-9
    [13] Yan XIA, Yongyun HU, Jiping LIU, Yi HUANG, Fei XIE, Jintai LIN, 2020: Stratospheric Ozone-induced Cloud Radiative Effects on Antarctic Sea Ice, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 505-514.  doi: 10.1007/s00376-019-8251-6
    [14] CUI Xuedong, GAO Yongqi, SUN Jianqi, 2014: The Response of the East Asian Summer Monsoon to Strong Tropical Volcanic Eruptions, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 1245-1255.  doi: 10.1007/s00376-014-3239-8
    [15] WANG Geli, YAN Jianjun, YANG Peicai, 2012: On the Nonlinear Response of Lower Stratospheric Ozone to External Forces---The Inclusion of BrOx and Radiation Processes, ADVANCES IN ATMOSPHERIC SCIENCES, 29, 855-866.  doi: 10.1007/s00376-012-1177-x
    [16] LIU Yi, LIU Chuanxi, Xuexi TIE, GAO Shouting, 2011: Middle Stratospheric Polar Vortex Ozone Budget during the Warming Arctic Winter, 2002--2003, ADVANCES IN ATMOSPHERIC SCIENCES, 28, 985-996.  doi: 10.1007/s00376-010-0045-9
    [17] WANG Geli, YANG Peicai, 2006: On the Nonlinear Response of Lower Stratospheric Ozone to Nox and ClOx Perturbations for Different CH4 Sources, ADVANCES IN ATMOSPHERIC SCIENCES, 23, 750-757.  doi: 10.1007/s00376-006-0750-6
    [18] LI Lin, LI Chongyin, PAN Jing, TAN Yanke, 2012: On the Differences and Climate Impacts of Early and Late Stratospheric Polar Vortex Breakup, ADVANCES IN ATMOSPHERIC SCIENCES, 29, 1119-1128.  doi: 10.1007/s00376-012-1012-4
    [19] REN Rongcai, WU Guoxiong, Ming CAI, YU Jingjing, 2009: Winter Season Stratospheric Circulation in the SAMIL/LASG General Circulation Model, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 451-464.  doi: 10.1007/s00376-009-0451-z
    [20] REN Rongcai, YANG Yang, 2012: Changes in Winter Stratospheric Circulation in CMIP5 Scenarios Simulated by the Climate System Model FGOALS-s2, ADVANCES IN ATMOSPHERIC SCIENCES, 29, 1374-1389.  doi: 10.1007/s00376-012-1184-y

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 12 November 2018
Manuscript revised: 16 April 2019
Manuscript accepted: 30 April 2019
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

The Effect of Super Volcanic Eruptions on Ozone Depletion in a Chemistry-Climate Model

    Corresponding author: Ke WEI, weike@mail.iap.ac.cn
  • 1. Center for Monsoon System Research, Institute of Atmospheric Physics, Chinese Academy of Sciences, Beijing 100190, China
  • 2. University of Chinese Academy of Sciences, Beijing 100049, China
  • 3. Key Laboratory of Middle Atmosphere and Global Environment Observation, Institute of Atmospheric Physics, Chinese Academy of Sciences, Beijing 100190, China
  • 4. Russian State Hydrometeorological University, Malookhtinskii pr. 98, St. Petersburg 195196, Russia
  • 5. Institute of Numerical Mathematics, Russian Academy of Sciences, ul.Gubkina 8, Moscow 11999, Russia

Abstract: With the gradual yet unequivocal phasing out of ozone depleting substances (ODSs), the environmental crisis caused by the discovery of an ozone hole over the Antarctic has lessened in severity and a promising recovery of the ozone layer is predicted in this century. However, strong volcanic activity can also cause ozone depletion that might be severe enough to threaten the existence of life on Earth. In this study, a transport model and a coupled chemistry-climate model were used to simulate the impacts of super volcanoes on ozone depletion. The volcanic eruptions in the experiments were the 1991 Mount Pinatubo eruption and a 100× Pinatubo size eruption. The results show that the percentage of global mean total column ozone depletion in the 2050 RCP8.5 100× Pinatubo scenario is approximately 6% compared to two years before the eruption and 6.4% in tropics. An identical simulation, 100× Pinatubo eruption only with natural source ODSs, produces an ozone depletion of 2.5% compared to two years before the eruption, and with 4.4% loss in the tropics. Based on the model results, the reduced ODSs and stratospheric cooling lighten the ozone depletion after super volcanic eruption.

摘要: 随着平流层中臭氧损耗物质(ODSs)的不断清除,由南极上空臭氧空洞的发现所引发的环境危机正在减轻,并且臭氧层也在逐渐恢复。然而,强火山活动同样会造成严重的臭氧损耗,从而威胁地球上的生命。在本研究中,利用了一个大气输送模式和一个化学气候模式,来模拟超级火山爆发后造成的臭氧损耗。模拟的火山事件为1991年的皮纳图博(Pinatubo)火山以及一个100×Pinatubo级别的火山。结果表明,在2050 RCP8.5 100×Pinatubo情形下,全球平均臭氧损耗和爆发前两年相比大约为6%,赤道地区为6.4%。而一个理想情形,即100×Pinatubo在自然源ODSs全部清除,只剩自然源的背景下爆发后,全球平均臭氧损耗和爆发前两年相比为2.5%,赤道地区为4.4%。根据模式结果,ODSs含量的下降以及平流层降温能够减轻超级火山爆发后造成的臭氧损耗。

1. Introduction
  • Volcanism has long been considered a strong external force that causes climate and weather variations (Robock, 2000). The most important climatic effect of large volcanic eruptions is caused by their emission of sulfur species into the stratosphere; these sulfur species form sulfur aerosols that change the radiative balance and chemistry in the stratosphere (Coffey, 1996; Robock, 2000; Kremser et al., 2016). The increased stratospheric aerosols reflect more solar radiation, causing decreased shortwave heating in the troposphere and on the surface. Meanwhile, the stratospheric aerosols absorb more solar shortwave radiation and longwave radiation from the surface, leading to a pronounced warming in the tropical stratosphere (Robock, 2000). The adjustment of the vertical and meridional temperature gradients changes the atmospheric circulation and further influences the surface climate (Robock, 2000; LeGrande et al., 2016).

    Stratospheric ozone is critical to life on Earth because it screens harmful ultraviolet radiation from the incident solar beam. When the ozone hole over the Antarctic was discovered in the 1980s (Farman et al., 1985), it caused global concern for its environmental and health effects. Prompt, coordinated, global measures were adopted, such as the Montreal Protocol and its subsequent several amendments, to eliminate the substances that deplete the ozone layer. Meanwhile, more and more research has been conducted to study stratospheric ozone. In addition to anthropogenic chlorofluorocarbons (CFCs), volcanic aerosols were considered a natural external force that could also deplete the ozone layer. Following volcanic eruption, a low ozone concentration has been observed, especially in winter and spring (Bluth et al., 1992; Hofmann et al., 1994; McGee et al., 1994; Randel et al., 1995). The most important impacts of a tropical eruption on stratospheric ozone chemistry can be divided into the following effects: (i) enhanced heterogeneous chemistry from elevated sulfuric acid (H2SO4) aerosol surface area density (SAD); (ii) the temperature and aerosol changes modify the occurrence and types of polar stratospheric clouds (PSCs); (iii) the changes of composition and reaction rate induced by the dynamical perturbations in the stratosphere; (iv) the direct injection of halogen species; and (v) coinjected water.

    The surfaces of sulfate aerosols provide sites for heterogeneous chemical reactions that activate halogen species to destroy ozone. In high-latitude regions, the heating in the tropics enhances the meridional temperature gradient and strengthens the polar vortex, leading to lower temperatures that form more PSC occurrence (Telford et al., 2009; Oppenheimer et al., 2010; Muthers et al., 2015). Furthermore, the presence of H2SO4 in the polar stratosphere within the colder temperatures facilitates the formation of liquid H2SO4 ternary solution particles, which increase the SAD and therefore increases ozone depletion in which there is a high background loading of inorganic chlorine (Cl) (Carslaw et al., 1994). Some volcanic eruptions also release a large load of halogen species [i.e., Cl and bromine (Br)], which directly destroy ozone (Oppenheimer et al., 2010; Kutterolf et al., 2013; Cadoux et al., 2015). However, the role of volcanogenic halogens in stratospheric chemistry remains poorly understood, because it is commonly believed that most halogens produced by volcanic eruptions are removed by hydrometeors in the troposphere and therefore do not enter the stratosphere (Tabazadeh and Turco, 1993). For example, Mount Pinatubo, which erupted in June 1991, occurred simultaneously with Typhoon Yunya, leading to the efficient removal of volcanic halogens over a large region. This eruption did not cause measurable accumulation of stratospheric halogens (McCormick et al., 1995; von Glasow et al., 2009; Kutterolf et al., 2013). Atmospheric water has long been considered an important factor in converting SO2 to sulfate aerosols (Rampino and Self, 1992), and it has been hypothesized to be a limiting factor in volcanic aerosol formation (Bekki, 1995). Water increases the availability of hydroxyl radicals, converting sulfur dioxide (SO2) more quickly into sulfate aerosols and increasing the rate of aerosol growth (LeGrande et al., 2016). Furthermore, SO2 is a greenhouse gas that may temper the negative impact of sulfate aerosols (Schmidt et al., 2016), which makes the balance between SO2 and sulfate aerosols a significant controller of the climate response following an eruption (Joshi and Jones, 2009; LeGrande et al., 2016).

    (Solomon, 1999) noted that the chemical and dynamical effects of volcanic aerosols occur simultaneously, and their effects are not much larger than natural variability. The quasi-biennial oscillation (QBO) and El Niño-Southern Oscillation (ENSO) are the dominant factors that cause the natural variability of ozone. The QBO modulates tropical upwelling and extratropical wave activities, affecting ozone concentrations (Baldwin et al., 2001, Cagnazzo et al., 2009; Xie et al., 2012, Xie et al., 2014). The eruptions of both Mount El Chichón and Mount Pinatubo were followed by the El Niño phase of sea surface temperature (SST). With a positive SST anomaly, El Niño changes circulation patterns, cooling the stratosphere and affecting ozone reaction rates in the tropics. In polar regions, El Niño often features a weaker stratospheric polar vortex (Chen and Wei, 2009; Ren et al., 2012; Graf et al., 2014). When the stratospheric polar vortex is not strong and cold enough, it cannot form many PSCs, which results in fewer sites for heterogeneous reactions and, therefore, less ozone depletion (Solomon, 1999; Solomon et al., 2015).

    The Mount Toba eruption 74 000 years ago was a Volcanic Explosivity Index (VEI)-8 eruption and was approximately 100 times larger than the Mount Pinatubo eruption. After the eruption, the surface temperature of Earth dropped more than 10°C, and a new ice age was triggered (Petraglia et al., 2012; Wu et al., 2012). Some researchers believe that this super volcanic eruption might have killed three-quarters of the plants in the Northern Hemisphere (NH) and that human evolution was at risk (Ambrose, 1998; Rampino, 2002). Though there are still debates about whether this super volcanic eruption caused the subsequent population bottleneck (Oppenheimer, 2002; Delwiche, 2005; Robock et al., 2009; Williams, 2012), all studies have recognized its serious destruction of the global climate.

    As a successful result of the Montreal Protocol, both models and observations have shown that the stratospheric abundances of Cl and Br are decreasing slowly, consistent with the long atmospheric lifetimes of the main ozone-depleting substances (ODSs) (SPARC, 2013; WMO, 2014), and the ozone recovery is quite promising (Solomon et al., 2016). However, some scientists have already raised concerns about the possibility that a major volcanic eruption within the near future may enhance ozone depletion, since there are still significant amounts of ODSs in the stratosphere, which may delay the stratospheric ozone recovery (Roscoe, 2001; Tabazadeh et al., 2002). Recently, (Klobas et al., 2017) also suggested that, in the coming decades, the stratospheric presence of Br supplied by natural, very short-lived compounds makes the ozone layer more susceptible to loss following volcanic eruptions than if this halogen source were not present. Based on the work of other scientists, and with the Mount Pinatubo eruption as a reference, we use model results to examine ozone depletion under different ODS scenarios of a super volcanic eruption in the future. This study may help us to further understand the potential risks of a super volcanic eruption in the ozone recovery period.

2. Model and data
  • Two models were used in this research. One was the highly scalable Massive-Parallel Trajectory Calculations (MPTRAC), in which air parcel trajectories are calculated based on numerical integration using the wind field from global meteorological reanalysis (Hoffmann et al., 2016; Rößler et al., 2018). Diffusion is modeled by uncorrelated Gaussian random displacements of the air parcels with zero mean and standard deviations, $\sigma_x=\sqrt{D_x \Delta t}$ (horizontally) and $\sigma_z=\sqrt{D_z \Delta t}$ (vertically), where Dx and Dz are the horizontal and vertical diffusivities, respectively, and ∆ t is the time step for the trajectory calculations. Depending on the atmospheric conditions, actual values of Dx and Dz may vary by several orders of magnitude (e.g., Legras et al., 2003, Legras et al., 2005; Pisso et al., 2009). In our simulations, we followed the approach of (Stohl et al., 2005) and set Dx and Dz to 50 and 0 m2 s-1 in the troposphere, and 0 and 0.1 m2 s-1 in the stratosphere, respectively. A constant half lifetime of seven days was assumed for SO2 for the stratosphere, and 2.5 days was assumed for the troposphere. More detailed information about the MPTRAC model can be found in (Wu et al., 2018) and (Wu et al., 2017).

    A three-dimensional chemistry-climate model (CCM) developed at the Russian State Hydrometeorological University (RSHMU) was also used in this study. The model includes two important modules: a dynamic module developed at the Institute of Numerical Mathematics of the Russian Academy of Sciences, and a photochemical module developed at the RSHMU. The model resolution is 5°× 4° in longitude and latitude, with 39 sigma levels vertically from the Earth's surface up to 0.003 hPa. The time step is 12 min, and the time integration is conducted by a central-difference scheme combined with a semi-implicit scheme (gravity waves are treated implicitly). In the dynamic module, the equations of atmospheric thermohydrodynamics are solved by the finite-difference method on a C grid. The model also incorporates parametrizations of deep and shallow convection (Betts, 1986) and orographic (Palmer et al., 1986) and nonorographic gravity-wave (Hines, 1997) resistance.

    The chemical module incorporates 74 basic atmospheric gas constituents directly or indirectly influencing the rates of photochemical changes in ozone. The model takes into account reactions of the oxygen, hydrogen, nitrogen, Cl, Br, and sulfur cycles, which makes it possible to treat the influence of chemical processes on the formation and evolution of not only ozone and its related gases but also atmospheric sulfate aerosols. The rates of chemical reactions are based on the JPL-2003 (Sander et al., 2003). The photolysis rates are calculated using the modified Delta-Eddington method (Dvortsov et al., 1992). Oxygen, nitrogen, Cl and Br families are considered and hydrogen species are calculated separately. Halogens (especially Cl and Br) play important roles in the destruction of stratospheric ozone (McCormick et al., 1995; Bobrowski et al., 2003; Cadoux et al., 2015). Recent model and observational studies all suggest significantly high ozone depletion with different halogen yields (Hunton et al., 2005; Cadoux et al., 2015; Klobas et al., 2017; Lurton et al., 2018). Therefore, halogens should be carefully considered after volcanic eruptions. All modeled species are transported independently, with a correction at each time step to conserve mass within families. Boundary emissions are taken from the World Meteorological Organization (WMO)/United Nations Environment Programme (WMO, 2007). Water vapor is specified in the troposphere according to its climatological values. Above the tropopause, water vapor is calculated like all other species, with supersaturation control. The model calculates photochemical production and destruction rates based on the adopted scheme of chemical and photolytic reactions, thus facilitating changes in the photochemical mechanisms. Parameterizations of heterogeneous chemistry on the surfaces of stratospheric sulfate aerosol and PSCs are employed in the model, including Br reactions, which can also play a significant role in stratospheric chemistry. Comparisons were made between the model calculations and observations on the ozone content and temperature by (Galin et al., 2007), and heterogeneous processes on the surface of PSCs were shown to be important for correct simulation of the spatial and temporal distribution of atmospheric ozone. (Smyshlyaev et al., 2010) also conducted model experiments on the evolution of the gas and aerosol compositions of the Arctic and Antarctic atmospheres. A detailed introduction to the parameterization of heterogeneous processes and the model can be found in (Smyshlyaev et al., 1998) and (Galin et al., 2007).

    The variable describing volcanic aerosols in our model is the SAD taken from the CMIP6 Stratospheric Aerosol Data Set for the historical simulation. For the period of 1979-2014, SAD data were based on satellite data from SAGE, SAM, SAGE II, CALIPSO and OSIRIS. Furthermore, CLAES satellite data were used for gap-filling of missing data for the period several months after the Pinatubo eruption. The data are monthly and zonal means averaged in latitudinal bands of five degrees. Data are provided as three-dimensional (time, altitude and latitude) arrays between 90°S and 90°N and from 5 km to 39.5 km at a resolution of 0.5 km. Figure 1 shows the global mean SAD of the CMIP6 Stratospheric Aerosol Data Set. The two eruptions of Mount El Chichón and Mount Pinatubo are clearly reflected in the SAD. For more detailed information about the SAD data, readers are referred to the CMIP6 website (http://www.wcrp-climate.org/wgcm-cmip/wgcm-cmip6).

    Figure 1.  Time-height sections of the global mean SAD (units: cm2 cm-3).

  • The primary dataset used in this study is ERA-Interim, which includes ozone, temperature, wind and SST data (Dee and Uppala, 2009). The ERA-Interim ozone data were used to compare the model's ability to reproduce the variability of ozone after volcanic eruptions. Since ENSO and the QBO have strong influences on ozone, linear regression was applied to remove the ENSO and QBO signals from the reanalysis data (Free and Lanzante, 2009). For the QBO index, the equatorial zonal mean zonal wind at 50 hPa was used, as in previous studies (Holton and Tan, 1980; Hamilton, 1993; Chen and Li, 2007; Wei et al., 2007). For ENSO, the Niño3.4 index was adopted, which is the average SST anomaly in the region bounded by 5°N to 5°S, from 170°W to 120°W. The ENSO index has a four-month lead time, which gives the best correlation for each level (Free and Angell, 2002; Lean and Rind, 2009).

  • Two model runs were conducted using the MPTRAC model: a Pinatubo size run and a 100-times Pinatubo (100× Pinatubo) size run. For the Pinatubo size experiment, 1.7× 1011 kg of SO2 was injected in 100 000 air masses, each containing 170 000 kg of SO2, ranging from 1.5 to 25 km, centered at (15.13°N, 120.35°E) of Pinatubo's location, and distributed homogeneously within a radius of 30 km. For the 100× Pinatubo size experiment, 1.7× 1013 kg of SO2 was injected, ranging from 1.5 to 50 km, with the same distribution as the Pinatubo size experiment, but up to 50 km. The 100× Pinatubo experiment could be used to simulate a VEI-8 super volcanic eruption. The two experiments conducted using the MPTRAC model are summarized in Table 1. Based on these model outputs, the global sulfur aerosol distribution under different SO2 loading scenarios could be analyzed.

    In order to build the connections between the sulfur aerosol distribution and SADs, linear regression was adopted to construct the SADs for the 100× Pinatubo size eruption. The sulfur aerosol distribution at 30 hPa in the Pinatubo size eruption simulated by the MPTRAC model was used to regress the real SADs of the Pinatubo eruption from the CMIP6 datasets (we also used other levels of sulfur aerosol data, and the results did not show large differences). Thus, a linear regression between the SADs and sulfur aerosol was established, and then the SADs of the 100× Pinatubo eruption could be constructed for the simulations of the CCM. Figure 2 shows the constructed SADs for the 100× Pinatubo size eruption.

    Figure 2.  Constructed SADs for the 100× Pinatubo size eruption (units: cm2 cm-3).

    We then performed four model experiments with the CCM. First, a basic run, utilizing all aspects of the model, with prescribed SAD from 1979 to 1999, was used to simulate the Mount Pinatubo eruption in June 1991. Then, to isolate the effects of volcanic aerosols, a fixed run was carried out with the SAD fixed at the 1979 level, which was considered as a baseline. The fixed run also started in 1979 and ended in 1999. As (Klobas et al., 2017) suggested, the response of the stratospheric ozone layer to enhanced aerosol loading is both a function of equivalent effective stratospheric chlorine (EESC) and stratospheric temperature. Meanwhile, model studies show that EESC is likely to decline rapidly (Engel et al., 2018) and the stratospheric temperature is likely to cool (Randel et al., 2016) in the near future. Thus, in order to simulate the stratospheric ozone layer response to a super volcanic eruption, two more model runs were performed with the Community Earth System Model (CESM) SST output of RCP8.5 2040-60 future simulations and future ODSs suggested by WMO (2007). The CESM SST output of RCP8.5 2040-60 was used to reproduce a colder stratosphere, and the 2050 100× Pinatubo experiment was set to erupt in June 2050 with approximately half the ODSs compared to the 1990s level. The background 100× Pinatubo experiment was incorporated with all anthropogenic ODSs (such as CFCs, CH3CCl3, CHClF2, C2H3FCl2, C2H3F2Cl, C2HF3Cl2, CF2ClBr and CF3Br) having disappeared and the remaining ODSs produced by natural processes only [mainly CH3Br (~7 ppbv) and CH3Cl (~483 ppbv)]. All the CCM model runs are listed in Table 2.

    Two different methods were adopted to analyze the effects of volcanic aerosols on stratospheric ozone: Method A used two years before the eruption as the baseline and determined two years after the eruption minus the baseline as the effects of volcanic aerosols (Free and Lanzante, 2009). This method was applied to both the model results and the reanalysis data. (Stenchikov et al., 2006) used a longer base period (six years for El Chichón and Mount Pinatubo), and (Lanzante, 2007) used two to four years, while (Santer et al., 2001) used just four to twelve months before the eruption as the baseline. We obtained similar results if other time slices were adopted, such as one or four years before and after the eruption (results not shown). Method B used the fixed run as the reference and determined the basic run results minus the fixed run results. Since the fixed run did not contain any SAD variations, it represented normal conditions with no volcanic aerosols. Method B only applied to model runs.

3. Model results of ozone depletion after volcanic eruptions
  • Laboratory studies have shown that the nitrogen oxides N2O5 and ClONO2 react rapidly on the surface of H2SO4 solutions (Hanson and Ravishankara, 1993). These reactions decrease NOx (= NO + NO2) and increase nitric acid (HNO3) within the reactive nitrogen reservoir (NOy). Reduced NOx decreases inactive Cl reservoir species, ClONO2 and HCl, thereby increasing active chlorine, ClO, and its associated ozone loss cycles (Fahey et al., 1993). A schematic diagram of NOx and ClO is depicted in Fig. 3. Meanwhile, observations have also revealed large changes in NOx/NOy and ClO/Cly partitioning, after Mt. Pinatubo erupted, associated with the increase in SADs at midlatitudes (Fahey et al., 1993; Solomon, 1999). Figure 4 shows the modeled changes in chemical partitioning of NOx/NOy and ClO/Cly at 50 hPa midlatitudes. The decreasing of NOx/NOy and increasing of ClO/Cly are clearly simulated after the Mt. El Chichón and Pinatubo eruptions. The injected sulfur aerosols provide sites for heterogeneous chemical reactions of N2O5 and ClONO2, which consume NO and NO2, produce active chlorine, ClO and Cl, and finally decrease ozone.

    Figure 3.  Schematic diagram of NOx, ClO and heterogeneous reactions after strong volcanic eruptions.

    Figure 4.  Modeled changes in chemical partitioning of (NO+NO2)/NOy and ClO/Cly at 50 hPa midlatitudes (60°S-60°N). The triangles indicate the April 1982 El Chichón and June 1991 Pinatubo eruptions.

    Figure 5 shows the vertical ozone mixing ratio of the ERA-Interim data and model simulations after the Mount Pinatubo eruption in different regions. When a strong volcano erupts, large loads of sulfur aerosols are injected into the stratosphere; the prevailing Brewer-Dobson Circulation then transports the volcanic aerosols around the globe, causing worldwide ozone depletion through heterogeneous reactions (Robock, 2000). For the global mean (Fig. 5a), both the model and ERA-Interim data show negative ozone anomalies after the volcanic eruption between 50 hPa and 10 hPa, as expected, but some differences still exist. First, the ENSO and QBO signals have a strong impact on global ozone content, and distinct differences could be found between the ENSO-and-QBO-removed reanalysis data (purple lines) and the ENSO-and-QBO-unremoved data (blue lines). Table 3 shows the percentage change of ozone at 20 hPa in four different regions after the Mount Pinatubo eruption. ENSO and the QBO make opposite contributions to ozone variations. Thus, to analyze the effect of volcanic aerosols on ozone, ENSO and the QBO must be carefully treated. Although linear signals of ENSO and the QBO have been removed from the reanalysis data, the interaction of atmospheric circulation is quite nonlinear. Nonlinear signals of ENSO and the QBO may still remain in the data and are mixed with the volcanic influences. Meanwhile, the two methods of identifying volcanic effects also have inconsistencies. The ozone content of Method B (basic run minus fixed run) decreases less than that of Method A (after minus before). As mentioned above, Method B determines the basic-run-minus-fixed-run results, and the chemical effect of the eruption on ozone is caused by the changes in SAD; while Method A uses two years before the eruption as the baseline and determines two years after the eruption minus the baseline as the effects of volcanic aerosols. Thus, Method B only demonstrates the chemical influences of volcanic aerosols (Telford et al., 2009). However, the model is coupled with chemistry, radiation and dynamics; there must be interactions between these modules, and some dynamic and radiative signals must remain in the chemistry part of the model. Therefore, Method B mainly contains the chemical effect of volcanic aerosols, but some other factors also exist.

    Figure 5.  Comparison of the vertical ozone mixing ratio between the model runs and ERA-Interim reanalysis data after the Mount Pinatubo eruption in terms of the (a) global mean, (b) tropical mean (30°S-30°N), (c) north polar mean (60°-90°N), and (d) south polar mean (60°-90°S). Units: ppm (10-6 kg kg-1).

    As Mt. Pinatubo (15°N) erupted in the tropics, the direct injection of sulfur aerosols caused ozone depletion simultaneously and significantly. The tropical mean (Fig. 5b) shows that the ENSO-and-QBO-unremoved reanalysis data (blue line) depict more ozone depletion than the ENSO-and-QBO-removed data (purple line), which means that the impact of ENSO and the QBO in the tropics is completely contrary to that in the north polar regions (Fig. 5c). The difference between them is also quite large, indicating that ENSO and the QBO have a very strong or even controlling impact on the ozone content at low latitudes, despite these two volcanoes both erupting in equatorial regions.

    In the north polar regions, besides the volcanic aerosols, the PSCs in the NH winter also decrease ozone (Schoeberl and Hartmann, 1991; Peter, 1997; Zhu et al., 2018). ENSO and the QBO also play important roles in ozone natural variability (Baldwin et al., 2001; Yu et al., 2011; Son et al., 2017; Xie et al., 2018), and the winter following Pinatubo's eruption featured an El Niño cycle. Usually, the stratospheric polar vortex is weaker than normal during ENSO years (Chen and Wei, 2009; Ren et al., 2012; Graf et al., 2014), which is unfavorable for the formation of PSCs and thus leads to fewer heterogeneous reactions and less ozone depletion (Solomon, 1999; Solomon et al., 2015). Therefore, when the ENSO and QBO signals were removed from the reanalysis data, the ozone depletion was less severe than that of the data without those signals removed (Fig. 5c).

    The same as in the NH, the abundant PSCs and volcanic aerosols both decrease ozone in the south polar winters. However, the stratospheric vortex in the Southern Hemisphere (SH) is usually colder and more isolated than in the NH, leading to the formation of more PSCs, and severer ozone depletion (Crutzen and Arnold, 1986; Schoeberl and Hartmann, 1991; Zhu et al., 2018). For the south polar regions (Fig. 5d), however, the ERA-Interim data and model simulation show opposite results: the ozone decreases in the model simulation but increases in the ERA-Interim data. (Poberaj et al., 2011) suggested that the increased ozone in the SH was caused by the enhanced ozone transport in late 1991 and early 1992, which compensated for the aerosol-induced chemical ozone depletion. This enhanced transport was related to the enhanced wave activities of the SH stratosphere. (Aquila et al., 2013) used a set of CCM simulations to show that aerosol-induced longwave radiation heating in the lower stratosphere would increase tropical upwelling, leading to increased Brewer-Dobson circulation (BDC), which transported more ozone to the SH. These two mechanisms likely acted together to explain the ozone increase in the SH, despite the eruption of Mount Pinatubo. Meanwhile, it is worth noting that the CCM may overestimate ozone loss due to heterogeneous processes in the polar regions, as (Galin et al., 2007) suggested. On the one hand, the CCM failed to represent the anomalous transport of ozone from the tropics to the south polar regions, while on the other hand it tends to overestimate ozone loss in the polar regions. These two reasons lead to the large inconsistencies between the simulations and reanalysis data in Fig. 5d.

4. Ozone depletion caused by a super eruption
  • As the Mount Pinatubo eruption led to severe ozone depletion in the stratosphere, it is natural to wonder about the effects caused by super volcanoes, which might be much stronger than the 1991 Pinatubo eruption. The true effect of such eruptions on the global scale might be very complicated and uncertain, and the volcanic products should be more complex than simple sulfate aerosols. For instance, the 1257 Samalas eruption released a very large load of gases into the atmosphere, including approximately 158±12 Tg of SO2, 227±18 Tg of Cl and 1.3±0.3 Tg of Br. These emissions represent the greatest volcanogenic gas injection of the Common Era (Vidal et al., 2016). The Mount Toba eruption 74 000 years ago was a VEI-8 level event and approximately 100 times larger than the Mount Pinatubo eruption. The Toba catastrophe theory has illustrated its destructive power on global climate (Rampino and Self, 1992; Petraglia et al., 2012; Williams, 2012). However, from the perspective of stratospheric ozone, we considered sulfate aerosols as a major point to simply study the ozone changes after a super volcanic eruption. This could give us some references and precautions if such an eruption truly happens.

    Figure 6 shows the vertical ozone depletion of the 2050 100× Pinatubo and background 100× Pinatubo scenarios in terms of the global mean, tropical mean, north polar mean, and south polar mean. The anomalies were calculated by the two-year mean after the eruption minus that before the eruption. The ODS level in the 2050 100× Pinatubo scenarios is approximately half of the 1990s level, and in the background 100× Pinatubo scenario is only naturally produced. For the global mean ozone depletion (Fig. 6a), the 2050 100× Pinatubo scenario (blue line) shows strong ozone depletion under half the ODSs of the 1990s level. The ozone depletion peaks at 5 hPa (approximately 1.3 ppm), which is higher than that of the Pinatubo eruption (Fig. 5b) at 20 hPa. This is due to the SO2 maximum height set in the model simulations; the injected SO2 reached 50 km (approximately 1 hPa) in the 100× Pinatubo scenarios. The higher volcanic aerosol caused higher ozone depletion. The background 100× Pinatubo scenario (red line) also shows strong ozone depletion, but less than the 2050 100× Pinatubo scenario. The depletion also peaks at 5 hPa, at approximately 0.9 ppm. As the anthropogenic ODSs all disappear from the stratosphere, the naturally produced ODSs also cause serious ozone depletion when a VEI-8 level volcano erupts. For the tropical mean ozone depletion (Fig. 6b), it shows worse ozone reduction than the global mean (Fig. 6a). In the 2050 100× Pinatubo scenario, the worst ozone depletion reaches approximately 1.4 ppm at 5 hPa and 1 ppm in the background 100× Pinatubo scenario.

    Figure 6.  Ozone depletion of the 2050 100× Pinatubo and background 100× Pinatubo scenarios in terms of the (a) global mean, (b) tropical mean (30°S-30°N), (c) north polar mean (60°-90°N), and (d) south polar mean (60°-90°S). The anomalies were calculated by Method A: two-year mean after the eruption minus that before the eruption. Units: ppm (10-6 kg kg-1).

    For the north polar mean (Fig. 6c), the 2050 100× Pinatubo scenario (blue line) shows a worst ozone depletion of approximately 1.3 ppm at 3 hPa, while the background 100× Pinatubo scenario (red line) shows a worst ozone depletion of approximately 1 ppm at 3 hPa. However, for the south polar mean (Fig. 6d), the ozone depletion is not as serious as in other regions. In the 2050 100× Pinatubo scenario (blue line), the worst ozone depletion is at 30 hPa, at approximately 0.8 ppm, which is less than that of the north polar region and tropics. In the background 100× Pinatubo scenario (red line), a VEI-8 volcano does not cause serious ozone depletion in the south polar regions when all anthropogenic ODSs are removed. Ozone only shows a slight depletion at 30 hPa and above; the worst reduction is around 3 hPa, at approximately 0.2 ppm.

    The most recent ozone assessment report (WMO, 2018) suggested that, outside the polar regions, upper-stratospheric ozone has increased by 1%-3% (10 yr)-1 since 2000, while total column ozone declined across most of the globe during the 1990s and early 2000s by approximately 2.5%, averaged over 60°S to 60°N (WMO, 2014). Figure 7 shows total column ozone depletion as a function of time in the 2050 100× Pinatubo and background 100× Pinatubo scenarios of the model runs in terms of the global mean, tropical mean, north polar mean, and south polar mean. The ozone anomalies were calculated by subtracting two years before the eruption. For the 2050 100× Pinatubo scenario, the global mean ozone shows a strong decline (Fig. 7a), with the worst global mean depletion being about 6% compared to two years before the eruption, and the ozone has fallen to approximately 289 DU. For the background 100× Pinatubo scenario (red line), the ozone content also shows depletion, but less than the 2050 100× Pinatubo scenario (blue line). The worst global mean depletion is about 2.5% compared to two years before the eruption. For the tropical mean (Fig. 7b), the 2050 100× Pinatubo scenario shows a worst depletion of approximately 6.4%, and the background 100× Pinatubo scenario is approximately 4.4%, both worse than the global mean ozone depletion. For the north polar mean (Fig. 7c) and the south polar mean (Fig. 7d), the ozone depletion depicts large variations, especially in the north polar regions. After a strong volcanic eruption in the tropics, large loads of sulfur aerosols are injected into the tropical stratosphere, and then transported to the poles by BDC (Robock, 2000). Besides the volcanic aerosols that destroy polar ozone, PSCs forming in winter also decrease ozone. Sulfur aerosols and PSCs together cause ozone depletion in winter with ODSs. However, when winter passes, the cold and strong polar vortex that forms PSCs breaks down, along with PSCs, and thus the rate of ozone decline slows down, causing large ozone variations in polar regions.

    Figure 7.  Global mean total column ozone depletion as a function of time in the 2050 100× Pinatubo and background 100× Pinatubo scenarios of the model runs in terms of the (a) global mean, (b) tropical mean (30°S-30°N), (c) north polar mean (60°-90°N), and (d) south polar mean (60°-90°S). The ozone anomalies were calculated by subtracting two years before the eruption. Units: DU.

5. Conclusions and discussion
  • A transport model and a coupled CCM were used to study the effect of volcanic aerosols on stratospheric ozone. The MPTRAC model was used to simulate the aerosol particle distribution of a Pinatubo eruption and 100× Pinatubo size eruption. Linear regression was adopted to build the correlation between the volcanic aerosol distribution and the SADs. The CCM was then used to simulate the ozone depletion after the Pinatubo size and 100× Pinatubo size eruption scenarios. The Pinatubo scenario was analyzed to verify the ozone depletion simulated by the CCM and ERA-Interim datasets. The impacts of volcanic eruptions on ozone may not be larger than its natural variability (Solomon, 1999), so the method of (Free and Lanzante, 2009) was used to remove the ENSO and QBO signals from the reanalysis data. Two different methods were adopted to calculate the impacts of volcanic aerosols on ozone. Furthermore, two more model runs were conducted with CESM RCP8.5 SST from 2040 to 2060 and future ODSs of WMO (2007) to simulate 2050 100× Pinatubo and background 100× Pinatubo scenarios, respectively. The ODS level in the 2050 100× Pinatubo experiment was approximately half that of the 1990s, while in the background 100× Pinatubo scenario the anthropogenic ODSs all disappeared and only natural products remained.

    For the reanalysis data, the ENSO and QBO signals should be carefully treated because the data with removed and unremoved signals have distinct differences. The two methods used to analyze the volcanic effects indicated that Method A (mean of two years after the eruption minus two years before) reflects more comprehensive effects on ozone, including the chemistry, radiation and dynamics, while Method B (basic run minus fixed run) mainly shows the chemical effect.

    The 2050 100× Pinatubo scenario showed that the worst global mean total column ozone depletion is approximately 6% compared to the two years before the eruption, and the tropical mean ozone depletion is approximately 6.4%, while the ozone depletion of the background 100× Pinatubo scenario is approximately 2.5% in terms of the global mean and 4.4% in terms of the tropical mean. The decreased ODSs and stratospheric cooling indeed lighten the ozone depletion after a super volcanic eruption.

    However, it should be pointed out that, even if the aerosol transport has been explicitly calculated with the MPTRAC model, as in our research, the aerosol particle growth processes, which will significantly affect the SAD (LeGrande et al., 2016; Marshall et al., 2018) and aerosol effective radius (Ansmann et al., 1997; Wyser, 1998) relative to the Pinatubo eruption, were not included in the simulations. This needs to be improved in future studies.

    As discussed, the total impacts of volcanic eruptions on stratospheric ozone should be more complicated than the single effect of volcanic aerosols, although the enhanced SAD caused by volcanic sulfate is the most prominent effect. The Mount Pinatubo eruption occurred simultaneously with Typhoon Yunya, probably leading to efficient removal of volcanic halogens over a large region, and thus that eruption caused no measurable injections of stratospheric halogens (Rosi et al., 2001). Therefore, the impacts of the Mount Pinatubo eruption on ozone were mostly equal to the effect of volcanic aerosols. For the other effects, we mentioned the composition and reaction rate changes induced by dynamic perturbations in the stratosphere; the direct injection of halogen species and coinjected water might have important impacts on ozone, according to the different types of volcanic eruptions. The abundant halogens, which are directly injected by volcanic eruptions, may be an uncertain factor involved in the impacts of volcanic aerosols on ozone depletion (Cadoux et al., 2015; LeGrande et al., 2016). Water also plays an important role in stratospheric ozone (Joshi and Jones, 2009; LeGrande et al., 2016), but a large load of water injection is strongly dependent on the location of the volcanic eruption; specifically, whether it is in the vicinity of a large body of water (Joshi and Jones, 2009).

    The volcanically injected halogens have the largest uncertainties for the ozone depletion caused by strong volcanic eruptions. Although the 1991 Pinatubo eruption resulted in stratospheric ozone loss, it was due to heterogeneous chemistry on volcanic sulfate aerosols involving Cl of anthropogenic rather than volcanogenic origin (Cadoux et al., 2015). Until now, it is a common recognition that the volcanically produced halogens indeed play an important role in decreasing ozone, and many researchers have suggested to take volcanic halogens into account when simulating ozone depletion caused by volcanic eruptions in the future or in the past (Cadoux et al., 2015, Mather, 2015). However, there are still few observations or references able to estimate the quantities of volcanic halogens that enter the stratosphere. As (Cadoux et al., 2015) suggested, even if only 2% of these halogens reach the stratosphere, it will result in strong global ozone depletion. Models should be sensitive to the volcanic halogens that enter the stratosphere: the more volcanic halogens in the stratosphere, the greater the level of ozone depletion. Thus, there are still uncertainties in simulating the effects of strong volcanic eruptions on ozone depletion, especially volcanic halogens.

    For paleoclimatology and paleobiology, however, the surface temperature features strong centennial and decadal oscillation (Lisiecki and Raymo, 2005; Jouzel et al., 2007; Hansen et al., 2013), which also makes the stratospheric temperature quite uncertain during the Toba eruption. Furthermore, due to the absence of anthropogenic CFCs, the sources of ODSs also include volcanic injection, which makes the ozone depletion caused by the ancient Toba eruption more uncertain.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return