Advanced Search
Article Contents

The Role of Ozone Depletion in the Lack of Cooling in the Antarctic Upper Stratosphere during Austral Winter


doi: 10.1007/s00376-022-2047-9

  • Temperature trends in the upper stratosphere are investigated using satellite measurements from Stratospheric Sounding Unit (SSU) outputs and simulations from chemistry–climate models (CCMs) and the Coupled Model Intercomparison Project Phase 6 (CMIP6). Observational evidence shows a lack of cooling in the Antarctic, in contrast to strong cooling at other latitudes, during austral winter over 1979–97. Analysis of CCM simulations for a longer period of 1961–97 also shows a significant contrast in the upper stratospheric temperature trends between the Antarctic and lower latitudes. Results from two sets of model integrations with fixed ozone-depleting substances (ODSs) and fixed greenhouse gases (GHGs) at their 1960 levels suggest that the ODSs have made a major contribution to the lack of cooling in the Antarctic upper stratosphere. Results from CMIP6 simulations with prescribed GHGs and ozone confirm that changes in the dynamical processes associated with observed ozone depletion are largely responsible for the lack of cooling in the Antarctic upper stratosphere. The lack of cooling is found to be dynamically induced through increased upward wave activity into the upper stratosphere, which is attributed mainly to ODSs forcing. Specifically, the radiative cooling caused by the ozone depletion results in a stronger meridional temperature gradient between middle and high latitudes in the upper stratosphere, allowing more planetary waves propagating upward to warm the Antarctic upper stratosphere. These findings improve our understanding of the chemistry–climate coupling in the southern upper stratosphere.
    摘要: 本文利用平流层探测仪(SSU)的卫星观测数据,化学-气候模式(CCM)资料以及国际耦合模式比较计划第六阶段(CMIP6)模式数据分析了上平流层的温度变化趋势。观测资料表明,在1979-97年南半球冬季,相比于其他纬度地区的显著变冷,南极并无显著变冷趋势。1961-97年长期的CCM模式资料结果同样表明,南极和低纬度地区之间上平流层温度趋势变化存在显著差异。当臭氧消耗物(ODS)和温室气体(GHG)浓度分别固定在1960年水平上时,两组模式积分结果显示,ODS对南极上平流层的冷却缺失具有较大贡献。在CMIP6模式中,固定温室气体和臭氧浓度的模式结果也表明,与臭氧损耗相关的动力过程能够在很大程度上解释南极上平流层的冷却缺失现象。该冷却缺失主要是由ODS强迫更多的行星波进入平流层上层而导致的。具体而言,臭氧损耗引起的辐射冷却使得上平流层中纬度和高纬度之间的经向温度梯度增强,从而导致更多的行星波上传,使得南极上平流层增温。这一发现加深了我们对南半球上平流层化学-气候耦合的认识。
  • 加载中
  • Figure 1.  Latitude profiles of linear trends for monthly zonal mean SSU3 temperature in austral (a) summer, (b) autumn, (c) winter, and (d) spring for 1979–97. Thick black solid lines indicate that the trend at each latitude in the range of 90°–30°S is significantly different from the trend at the equator of 0° at the 95% confidence level. For the specific method of significance test between the two trends, please refer to section 2.3.

    Figure 2.  (a) Same as Fig. 1c. The grey line shows the temperature trends derived from NOAA. The brown, green, dashed orange, and dashed blue lines represent the ensemble-mean trends of CCMs. The brown line represents the ensemble-mean temperature trends of CCMs (Table 1, 12 simulations) with all forcings. The green line is the same as the brown line, but for 34 simulations (Table 3). Dashed orange and dashed blue lines represent the ensemble mean with fixed ODSs and GHGs (Table 1, 12 simulations), respectively. Solid pink and purple lines are from reanalysis data of ERA5 and JRA55 datasets, respectively. (b) is the same as (a), but for 1998–2014. (c) and (d) are similar to (a) and (b), but for the ensemble mean (Table 1, 12 members) of the longer periods of 1961–97 and 1998–2014. (e) and (f) are the same as (c) and (d), but for the ensemble mean of CMIP6 simulations (Table 2, 26 members).

    Figure 3.  Latitude–height cross sections of ensemble-mean linear trends for zonal mean temperature [K (10 yr)–1] in austral winter over (a, c, e) 1961–97 and (b, d, f) 1998–2014. (a) and (b) are for the reference ensemble mean. (c, d) and (e, f) are the same as (a, b), but for the ensemble mean with fixed ODSs and GHGs, respectively. Detailed simulations can be seen in Table 1. The dotted areas indicate that the trends are statistically significant at 95% confidence levels.

    Figure 4.  Same as Fig. 3, but for the ensemble mean of CMIP6 simulations (Table 2, 26 members).

    Figure 5.  Latitude–height cross sections of ensemble-mean linear trends for zonal mean ozone [10−7 mol mol–1 (10 yr)–1] in austral winter for 1961–97 (left panel) and 1998–2014 (right panel). (a) and (b) are for the reference ensemble mean. (c, d) and (e, f) are the same as (a, b), but for the ensemble mean with fixed ODSs and GHGs, respectively. Detailed simulations can be seen in Table 1. The dotted areas indicate that the trends are statistically significant at 95% confidence levels.

    Figure 6.  Latitude–height cross sections of linear trends for observed zonal mean ozone [ppmv (10 yr)–1] in austral winter for (a) 1984–97 and (b) 1998–2014. The dataset is from SWOOSH. The dotted areas indicate that the trends are statistically significant at 95% confidence levels.

    Figure 7.  As in Fig. 5, but for zonal mean wind [m s–1 (10 yr)–1].

    Figure 8.  (a) Same as Fig. 3, but for the average of ERA5 and JRA55. (c) Latitude–height cross sections of linear trends for meridional eddy heat flux [K m s–1 (10 yr)–1] in austral winter for 1979–97. (e) The climatology (black line, K m s–1) and linear trends [red line, K m s–1(10 yr)–1] of vertically integrated meridional eddy heat flux in austral winter for 1979–97. (g) Latitude–height cross sections of linear trends for EP Flux divergence [m s–1 d–1 (10 yr)–1] in austral winter for 1979–97. (b, d, f, h) Same as (a, c, e, g), but for 1998–2014.

    Figure 9.  (a) Latitude–height cross sections of linear trends for refractive index squared (a2n2) for ZWN1 (first row) planetary waves (the average of ERA5 and JRA55) in austral winter for 1979–97. (c) and (e) are the same as (a), but for ZWN2 and ZWN3, respectively. (b, d, f) are the same as (a, c, e), but for 1998–2014.

    Table 1.  Details of the CCMI and CCMVal-2 simulations in which column 3 corresponds to the horizontal resolution, the number of vertical levels, and model lid. For details, please refer to references in column 4. Three types of simulations are employed: the reference simulations and two sensitivity simulations with fixed ODSs and GHGs at their 1960 levels. Specifically, they are REF-B2, SCN-B2b, and SCN-B2c for CCMVal-2 and REF-C2, SEN-C2-fODS, and SEN-C2-fGHG for CCMI.

    Model nameEnsemble
    members
    Domain/Resolution (lat×lon), levels, model lidReference
    CCMVal-2
    CCSRNIES12.8°×2.8°(T42), 34L, 0.01 hPaAkiyoshi et al. (2009),
    Kurokawa et al. (2005)
    CMAM33.75°×3.75°(T31), 71L, 0.06 Pade Grandpré et al. (2000),
    Hitchcock et al. (2009)
    MRI12.8°×2.8°(T42), 68L, 0.01 hPaShibata and Deushi (2008),
    Xiao and Peng (2004)
    SOCOL13.75°×3.75°(T30), 39L, 0.1 hPaSchraner et al. (2008),
    Egorova et al. (2005)
    CCMI
    CCSRNIES-MIROC3.212.8°×2.8°(T42), 34L, 1.2 PaAkiyoshi et al. (2016),
    Sakazaki et al. (2013)
    CMAM12.5°×2.5°(T47), 71L, 0.08 PaScinocca et al. (2008),
    de Grandpré et al. (2000)
    NIWA-UKCA13.75°×2.5°, L60, 84 kmMorgenstern et al. (2009, 2013),
    Zeng et al. (2008, 2010)
    WACCM31.9°×2.5°, L66, 140 kmSolomon et al. (2015),
    Garcia et al. (2017)
    DownLoad: CSV

    Table 2.  Same as Table 1, but for the CMIP6 simulations. Three types of simulations are employed: the historical simulations and two sensitivity simulations with historical stratospheric ozone-only and historical well-mixed GHG-only forcing.

    Model nameEnsemble
    members
    Domain/Resolution (lat×lon), levels, model lidReference
    MIROC631.41° × 1.41°(T85), 19L, 1 hPaTatebe et al. (2019)
    CanESM5102.81° × 2.81°(T42), 19L, 1 PaSwart et al. (2019)
    MRI-ESM2-031.125° × 1.125°(T106), 19L, 1 hPaYukimoto et al. (2019)
    IPSL-CM6A-LR102.5° × 1.26°, 19L, 1 hPaLurton et al. (2020)
    DownLoad: CSV

    Table 3.  Details of the CCMI and CCMVal-2 simulations. Only reference experiments are employed: REF-B2 for CCMVal-2 and REF-C2 for CCMI.

    Model nameEnsemble members Model nameEnsemble members
    CCMVal-2 CCMI
    CCSRNIES1CCSRNIES-MIROC3.21
    CMAM3CMAM1
    MRI2NIWA-UKCA1
    SOCOL3WACCM3
    CAM3.51ACCESS-CCM1
    LMDZrepro1IPSL1
    Niwa-SOCOL1CHASER-MIROC-ESM1
    UMSLIMCAT1CNRM-CM5-31
    WACCM3EMAC-L47MA1
    CNRM-ACM1EMAC-L90MA1
    UMUKCA-METO1GEOSCCM1
    SOCOL31
    HadGEM3-ES1
    MRI-ESM1r11
    DownLoad: CSV
  • Akiyoshi, H., T. Nakamura, T. Miyasaka, M. Shiotani, and M. Suzuki, 2016: A nudged chemistry-climate model simulation of chemical constituent distribution at northern high-latitude stratosphere observed by SMILES and MLS during the 2009/2010 stratospheric sudden warming. J. Geophys. Res.: Atmos., 121(3), 1361−1380, https://doi.org/10.1002/2015JD023334.
    Akiyoshi, H., and Coauthors, 2009: A CCM simulation of the breakup of the Antarctic polar vortex in the years 1980−2004 under the CCMVal scenarios. J. Geophys. Res.: Atmos., 114, D03103, https://doi.org/10.1029/2007JD009261.
    Albers, J. R., and T. R. Nathan, 2013: Ozone loss and recovery and the preconditioning of upward-propagating planetary wave activity. J. Atmos. Sci., 70(12), 3977−3994, https://doi.org/10.1175/JAS-D-12-0259.1.
    Andrews, D. G., and M. E. McIntyre, 1976: Planetary waves in horizontal and vertical shear: The generalized Eliassen-Palm relation and the mean zonal acceleration. J. Atmos. Sci., 33, 2031−2048, https://doi.org/10.1175/1520-0469(1976)033<2031:PWIHAV>2.0.CO;2.
    Andrews, D. G., and M. E. McIntyre, 1978: Generalized Eliassen-Palm and Charney-Drazin theorems for waves on axismmetric mean flows in compressible atmospheres. J. Atmos. Sci., 35, 175−185, https://doi.org/10.1175/1520-0469(1978)035<0175:GEPACD>2.0.CO;2.
    Andrews, D. G., J. R. Holton, and C. B. Leovy, 1987: Middle Atmosphere Dynamics. Academic Press, 489 pp.
    Ball, W. T., and Coauthors, 2016: An upper-branch Brewer–Dobson circulation index for attribution of stratospheric variability and improved ozone and temperature trend analysis. Atmospheric Chemistry and Physics, 16, 15 485−15 500,
    Bell, C. J., L. J. Gray, and J. Kettleborough, 2010: Changes in Northern Hemisphere stratospheric variability under increased CO2 concentrations. Quart. J. Roy. Meteor. Soc., 136(650), 1181−1190, https://doi.org/10.1002/qj.633.
    Boer, G. J., and Coauthors, 2016: The decadal climate prediction project (DCPP) contribution to CMIP6. Geoscientific Model Development, 9(10), 3751−3777, https://doi.org/10.5194/gmd-9-3751-2016.
    Brewer, A. W., 1949: Evidence for a world circulation provided by the measurements of helium and water vapour distribution in the stratosphere. Quart. J. Roy. Meteor. Soc., 75(326), 351−363, https://doi.org/10.1002/qj.49707532603.
    Butchart, N., 2014: The Brewer-Dobson circulation. Rev. Geophys., 52(2), 157−184, https://doi.org/10.1002/2013RG000448.
    Butchart, N., and A. A. Scaife, 2001: Removal of chlorofluorocarbons by increased mass exchange between the stratosphere and troposphere in a changing climate. Nature, 410, 799−802, https://doi.org/10.1038/35071047.
    Charney, J. G., and P. G. Drazin, 1961: Propagation of planetary-scale disturbances from the lower into the upper atmosphere. J. Geophys. Res., 66(1), 83−109, https://doi.org/10.1029/JZ066i001p00083.
    Checa-Garcia, R., K. P. Shine, and M. I. Hegglin, 2016: The contribution of greenhouse gases to the recent slowdown in global-mean temperature trends. Environmental Research Letters, 11, 094018, https://doi.org/10.1088/1748-9326/11/9/094018.
    Davis, S. M., and Coauthors, 2016: The Stratospheric Water and Ozone Satellite Homogenized (SWOOSH) database: A long-term database for climate studies. Earth System Science Data, 8(2), 461−490, https://doi.org/10.5194/essd-8-461-2016.
    de Grandpré, J., S. R. Beagley, V. I. Fomichev, E. Griffioen, J. C. McConnell, A. S. Medvedev, and T. G. Shepherd, 2000: Ozone climatology using interactive chemistry: Results from the Canadian Middle Atmosphere Model. J. Geophys. Res.: Atmos., 105(D21), 26 475−26 491,
    Dobson, G. M. B., 1956: Origin and distribution of the polyatomic molecules in the atmosphere. Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences, 236(1205), 187−193, https://doi.org/10.1098/rspa.1956.0127.
    Ebita, A., and Coauthors, 2011: The japanese 55-year reanalysis “jra-55”: An interim report. SOLA, 7, 149−152, https://doi.org/10.2151/sola.2011-038.
    Egorova, T., E. Rozanov, V. Zubov, E. Manzini, W. Schmutz, and T. Peter, 2005: Chemistry-climate model SOCOL: A validation of the present-day climatology. Atmospheric Chemistry and Physics, 5, 1557−1576, https://doi.org/10.5194/acp-5-1557-2005.
    Eyring, V., and Coauthors, 2005: A strategy for process-oriented validation of coupled chemistry-climate models. Bull. Amer. Meteor. Soc., 86(8), 1117−1134, https://doi.org/10.1175/BAMS-86-8-1117.
    Eyring, V., and Coauthors, 2006: Assessment of temperature, trace species, and ozone in chemistry-climate model simulations of the recent past. J. Geophys. Res.: Atmos., 111(D22), D22308, https://doi.org/10.1029/2006JD007327.
    Eyring, V., and Coauthors, 2008: Overview of the new CCMVal reference and sensitivity simulations in support of upcoming ozone and climate assessments and the planned SPARC CCMVal. SPARC Newsletter, 30, 20−26.
    Eyring, V., and Coauthors, 2010: Multi-model assessment of stratospheric ozone return dates and ozone recovery in CCMVal-2 models. Atmospheric Chemistry and Physics, 10(19), 9451−9472, https://doi.org/10.5194/acp-10-9451-2010.
    Eyring, V., and Coauthors, 2013: Overview of IGAC/SPARC Chemistry-Climate Model Initiative (CCMI) community simulations in support of upcoming ozone and climate assessments. SPARC Newsletter, 40, 48−66.
    Fu, Q., S. Solomon, H. A. Pahlavan, and P. Lin, 2019: Observed changes in Brewer-Dobson circulation for 1980−2018. Environmental Research Letters, 14(11), 114026, https://doi.org/10.1088/1748-9326/ab4de7.
    García-Herrera, R., N. Calvo, R. R. Garcia, and M. A. Giorgetta, 2006: Propagation of ENSO temperature signals into the middle atmosphere: A comparison of two general circulation models and ERA-40 reanalysis data. J. Geophys. Res.: Atmos., 111, D06101, https://doi.org/10.1029/2005JD006061.
    Garcia, R. R., A. K. Smith, D. E. Kinnison, Á. de la Cámara, and D. J. Murphy, 2017: Modification of the gravity wave parameterization in the Whole Atmosphere Community Climate Model: Motivation and results. J. Atmos. Sci., 74(1), 275−291, https://doi.org/10.1175/JAS-D-16-0104.1.
    Gillett, N. P., and D. W. J. Thompson, 2003: Simulation of recent Southern Hemisphere climate change. Science, 302(5643), 273−275, https://doi.org/10.1126/science.1087440.
    Gillett, N. P., and Coauthors, 2011: Attribution of observed changes in stratospheric ozone and temperature. Atmospheric Chemistry and Physics, 11(2), 599−609, https://doi.org/10.5194/acp-11-599-2011.
    Haynes, P. H., M. E. McIntyre, T. G. Shepherd, C. J. Marks, and K. P. Shine, 1991: On the “Downward Control” of extratropical diabatic circulations by eddy-induced mean zonal forces. J. Atmos. Sci., 48(4), 651−678, https://doi.org/10.1175/1520-0469(1991)048<0651:OTCOED>2.0.CO;2.
    Hersbach, H., and D. Dee, 2016: ERA5 reanalysis is in production. ECMWF Newsletter, 147, 5−6.
    Hitchcock, P., T. G. Shepherd, and C. McLandress, 2009: Past and future conditions for polar stratospheric cloud formation simulated by the Canadian Middle Atmosphere Model. Atmospheric Chemistry and Physics, 9, 483−495, https://doi.org/10.5194/acp-9-483-2009.
    Holton, J. R., and C. Mass, 1976: Stratospheric vacillation cycles. J. Atmos. Sci., 33(11), 2218−2225, https://doi.org/10.1175/1520-0469(1976)033<2218:SVC>2.0.CO;2.
    Holton, J. R., P. H. Haynes, M. E. McIntyre, A. R. Douglass, R. B. Rood, and L. Pfister, 1995: Stratosphere-troposphere exchange. Rev. Geophys., 33(4), 403−439, https://doi.org/10.1029/95RG02097.
    Hood, L. L., R. D. Mcpeters, J. P. Mccormack, L. E. Flynn, S. M. Hollandsworth, and J. F. Gleason, 1993: Altitude dependence of stratospheric ozone trends based on nimbus 7 SBUV data. Geophys. Res. Lett., 20(23), 2667−2670, https://doi.org/10.1029/93GL03087.
    Hu, D. Z., W. S. Tian, F. Xie, C. X. Wang, and J. K. Zhang, 2015: Impacts of stratospheric ozone depletion and recovery on wave propagation in the boreal winter stratosphere. J. Geophys. Res.: Atmos., 120(16), 8299−8317, https://doi.org/10.1002/2014JD022855.
    Hu, Y., and Q. Fu, 2009: Stratospheric warming in Southern Hemisphere high latitudes since 1979. Atmospheric Chemistry and Physics, 9(13), 4329−4340, https://doi.org/10.5194/acp-9-4329-2009.
    Hu, Y. Y., 2006: Possible impact of stratospheric polar ozone depletion tropospheric climate. Acta Scientiarum Naturalium Universitatis Pekinensis, 42(5), 561−568, https://doi.org/10.3321/j.issn:0479-8023.2006.05.001. (in Chinese with English abstract
    Hu, Y. Y., and K. K. Tung, 2002: Interannual and decadal variations of planetary wave activity, stratospheric cooling, and northern hemisphere annular mode. J. Climate, 15(13), 1659−1673, https://doi.org/10.1175/1520-0442(2002)015<1659:IADVOP>2.0.CO;2.
    Hu, Y. Y., and K. K. Tung, 2003: Possible ozone-induced long-term changes in planetary wave activity in late winter. J. Climate, 16(18), 3027−3038, https://doi.org/10.1175/1520-0442(2003)016<3027:POLCIP>2.0.CO;2.
    Ivy, D. J., S. Solomon, and H. E. Rieder, 2016: Radiative and dynamical influences on polar stratospheric temperature trends. J. Climate, 29(13), 4927−4938, https://doi.org/10.1175/JCLI-D-15-0503.1.
    Jiang, X., S. J. Eichelberger, D. L. Hartmann, R. Shia, and Y. L. Yung, 2007: Influence of doubled CO2 on ozone via changes in the Brewer-Dobson circulation. J. Atmos. Sci., 64(7), 2751−2755, https://doi.org/10.1175/JAS3969.1.
    Kobayashi, C., and T. Iwasaki, 2016: Brewer-Dobson circulation diagnosed from JRA-55. J. Geophys. Res.: Atmos., 121(4), 1493−1510, https://doi.org/10.1002/2015JD023476.
    Kodera, K., M. E. Hori, S. Yukimoto, and M. Sigmond, 2008: Solar modulation of the Northern Hemisphere winter trends and its implications with increasing CO2. Geophys. Res. Lett., 35(3), L03704, https://doi.org/10.1029/2007GL031958.
    Kurokawa, J. I., H. Akiyoshi, T. Nagashima, H. Masunaga, T. Nakajima, M. Takahashi, and H. Nakane, 2005: Effects of atmospheric sphericity on stratospheric chemistry and dynamics over Antarctica. J. Geophys. Res.: Atmos., 110, D21305, https://doi.org/10.1029/2005JD005798.
    Li, F. N., D. R. Chavas, K. A. Reed, and D. T. Dawson II, 2020: Climatology of severe local storm environments and synoptic-scale features over North America in ERA5 reanalysis and CAM6 simulation. J. Climate, 33(19), 8339−8365, https://doi.org/10.1175/JCLI-D-19-0986.1.
    Lin, P., and Q. Fu, 2013: Changes in various branches of the Brewer-Dobson circulation from an ensemble of chemistry climate models. J. Geophys. Res.: Atmos., 118(1), 73−84, https://doi.org/10.1029/2012JD018813.
    Lurton, T., and Coauthors, 2020: Implementation of the CMIP6 forcing data in the IPSL-CM6A-LR model. Journal of Advances in Modeling Earth Systems, 12, e2019MS001940, https://doi.org/10.1029/2019MS001940.
    Matsuno, T., 1970: Vertical propagation of stationary planetary waves in the winter Northern Hemisphere. J. Atmos. Sci., 27(6), 871−883, https://doi.org/10.1175/1520-0469(1970)027<0871:VPOSPW>2.0.CO;2.
    McLandress, C., A. I. Jonsson, D. A. Plummer, M. C. Reader, J. F. Scinocca, and T. G. Shepherd, 2010: Separating the dynamical effects of climate change and ozone depletion. Part I: Southern Hemisphere stratosphere. J. Climate, 23(18), 5002−5020, https://doi.org/10.1175/2010JCLI3586.1.
    Morgenstern, O., P. Braesicke, F. M. O'Connor, A. C. Bushell, C. E. Johnson, S. M. Osprey, and J. A. Pyle, 2009: Evaluation of the new UKCA climate-composition model - Part 1: The stratosphere. Geoscientific Model Development, 2(1), 43−57, https://doi.org/10.5194/gmd-2-43-2009.
    Morgenstern, O., and Coauthors, 2010: Review of the formulation of present-generation stratospheric chemistry-climate models and associated external forcings. J. Geophys. Res.: Atmos., 115, D00M02, https://doi.org/10.1029/2009JD013728.
    Morgenstern, O., and Coauthors, 2013: Impacts of climate change, ozone recovery, and increasing methane on surface ozone and the tropospheric oxidizing capacity. J. Geophys. Res.: Atmos., 118(2), 1028−1041, https://doi.org/10.1029/2012JD018382.
    Morgenstern, O., and Coauthors, 2017: Review of the global models used within phase 1 of the Chemistry-Climate Model Initiative (CCMI). Geoscientific Model Development, 10(2), 639−671, https://doi.org/10.5194/gmd-10-639-2017.
    Nath, O., and S. Sridharan, 2015: Equatorial middle atmospheric chemical composition changes during sudden stratospheric warming events. Atmospheric Chemistry and Physics, 15(17), 23 969−23 988,
    Nathan, T. R., and E. C. Cordero, 2007: An ozone-modified refractive index for vertically propagating planetary waves. J. Geophys. Res.: Atmos., 112, D02105, https://doi.org/10.1029/2006JD007357.
    Nekola, J. C., and P. S. White, 1999: The distance decay of similarity in biogeography and ecology. Journal of Biogeography, 26(4), 867−878, https://doi.org/10.1046/j.1365-2699.1999.00305.x.
    Newman, P. A., E. R. Nash, and J. E. Rosenfield, 2001: What controls the temperature of the Arctic stratosphere during the spring?. J. Geophys. Res.: Atmos., 106(D17), 19 999−20 010,
    Polvani, L. M., D. W. Waugh, G. J. P. Correa, and S. W. Son, 2011: Stratospheric ozone depletion: The main driver of twentieth-century atmospheric circulation changes in the Southern Hemisphere. J. Climate, 24(3), 795−812, https://doi.org/10.1175/2010JCLI3772.1.
    Polvani, L. M., M. Abalos, R. Garcia, D. Kinnison, and W. J. Randel, 2018: Significant weakening of Brewer-Dobson circulation trends over the 21st century as a consequence of the Montreal Protocol. Geophys. Res. Lett., 45(1), 401−409, https://doi.org/10.1002/2017GL075345.
    Polvani, L. M., and Coauthors, 2019: Large impacts, past and future, of ozone-depleting substances on Brewer-Dobson circulation trends: A multimodel assessment. J. Geophys. Res.: Atmos., 124(13), 6669−6680, https://doi.org/10.1029/2018JD029516.
    Ramaswamy, V., M. D. Schwarzkopf, and W. J. Randel, 1996: Fingerprint of ozone depletion in the spatial and temporal pattern of recent lower-stratospheric cooling. Nature, 382(6592), 616−618, https://doi.org/10.1038/382616a0.
    Ramaswamy, V., and Coauthors, 2001: Stratospheric temperature trends: Observations and model simulations. Rev. Geophys., 39(1), 71−122, https://doi.org/10.1029/1999RG000065.
    Randel, W. J., and F. Wu, 1999: Cooling of the arctic and antarctic polar stratospheres due to ozone depletion. J. Climate, 12(5), 1467−1479, https://doi.org/10.1175/1520-0442(1999)012<1467:COTAAA>2.0.CO;2.
    Randel, W. J., A. K. Smith, F. Wu, C. Z. Zou, and H. F. Qian, 2016: Stratospheric temperature trends over 1979-2015 derived from combined SSU, MLS, and SABER satellite observations. J. Climate, 29(13), 4843−4859, https://doi.org/10.1175/JCLI-D-15-0629.1.
    Rind, D., J. Lerner, J. Perlwitz, C. McLinden, and M. Prather, 2002: Sensitivity of tracer transports and stratospheric ozone to sea surface temperature patterns in the doubled CO2 climate. J. Geophys. Res.: Atmos., 107(D24), 4800, https://doi.org/10.1029/2002JD002483.
    Sakazaki, T., and Coauthors, 2013: Diurnal ozone variations in the stratosphere revealed in observations from the Superconducting Submillimeter-Wave Limb-Emission Sounder (SMILES) on board the International Space Station (ISS). J. Geophys. Res.: Atmos., 118(7), 2991−3006, https://doi.org/10.1002/jgrd.50220.
    Schraner, M., and Coauthors, 2008: Technical Note: Chemistry-climate model SOCOL: Version 2.0 with improved transport and chemistry/microphysics schemes. Atmospheric Chemistry and Physics, 8(19), 5957−5974, https://doi.org/10.5194/acp-8-5957-2008.
    Scinocca, J. F., N. A. McFarlane, M. Lazare, J. Li, and D. Plummer, 2008: Technical Note: The CCCma third generation AGCM and its extension into the middle atmosphere. Atmospheric Chemistry and Physics, 8(23), 7055−7074, https://doi.org/10.5194/acp-8-7055-2008.
    Sekiguchi, Y., 1986: Antarctic ozone change correlated to the stratospheric temperature-field. Geophys. Res. Lett., 13(12), 1202−1205, https://doi.org/10.1029/GL013i012p01202.
    Shibata, K., and M. Deushi, 2008: Long-term variations and trends in the simulation of the middle atmosphere 1980−2004 by the chemistry-climate model of the Meteorological Research Institute. Ann. Geophys., 26(5), 1299−1326, https://doi.org/10.5194/angeo-26-1299-2008.
    Shindell, D. T., and G. A. Schmidt, 2004: Southern Hemisphere climate response to ozone changes and greenhouse gas increases. Geophys. Res. Lett., 31, L18209, https://doi.org/10.1029/2004GL020724.
    Sridharan, S., S. Sathishkumar, and S. Gurubaran, 2012: An unusual reduction in the mesospheric semi-diurnal tidal amplitude over Tirunelveli (8.7°N, 77.8°E) prior to the 2011 minor warming and its relationship with stratospheric ozone. Journal of Atmospheric and Solar-Terrestrial Physics, 89, 27−32, https://doi.org/10.1016/j.jastp.2012.07.012.
    Solomon, S., 1999: Stratospheric ozone depletion: A review of concepts and history. Rev. Geophys., 37(3), 275−316, https://doi.org/10.1029/1999RG900008.
    Solomon, S., D. E. Kinnison, J. Bandoro, and R. Garcia, 2015: Simulations of polar ozone depletion: an update. J. Geophys. Res.: Atmos., 120(15), 7958−7974, https://doi.org/10.1002/2015JD023365.
    Steinitz, O., J. Heller, A. Tsoar, D. Rotem, and R. Kadmon, 2005: Predicting regional patterns of similarity in species composition for conservation planning. Conservation Biology, 19(6), 1978−1988, https://doi.org/10.1111/j.1523-1739.2005.00237.x.
    Steinitz, O., J. Heller, A. Tsoar, D. Rotem, and R. Kadmon, 2006: Environment, dispersal and patterns of species similarity. Journal of Biogeography, 33(6), 1044−1054, https://doi.org/10.1111/j.1365-2699.2006.01473.x.
    Stolarski, R. S., A. R. Douglass, P. A. Newman, S. Pawson, and M. R. Schoeberl, 2010: Relative contribution of greenhouse gases and ozone-depleting substances to temperature trends in the stratosphere: A chemistry-climate model study. J. Climate, 23(1), 28−42, https://doi.org/10.1175/2009JCLI2955.1.
    Swart, N. C., and Coauthors, 2019: The Canadian earth system model version 5 (CanESM5.0.3). Geoscientific Model Development, 12, 4823−4873, https://doi.org/10.5194/gmd-12-4823-2019.
    Tatebe, H., and Coauthors, 2019: Description and basic evaluation of simulated mean state, internal variability, and climate sensitivity in MIROC6. Geoscientific Model Development, 12, 2727−2765, https://doi.org/10.5194/gmd-12-2727-2019.
    Thompson, D. W. J., and S. Solomon, 2002: Interpretation of recent Southern Hemisphere climate change. Science, 296(5569), 895−899, https://doi.org/10.1126/science.1069270.
    Wang, Z., J. K. Zhang, T. Wang, W. H. Feng, Y. H. Hu, and X. R. Xu, 2021: Analysis of the Antarctic ozone hole in November. J. Climate, 34(16), 6513−6529, https://doi.org/10.1175/JCLI-D-20-0906.1.
    Weber, M., S. Dikty, J. P. Burrows, H. Garny, M. Dameris, A. Kubin, J. Abalichin, and U. Langematz, 2011: The Brewer-Dobson circulation and total ozone from seasonal to decadal time scales. Atmospheric Chemistry and Physics, 11(21), 11 221−11 235,
    Xiao, F., and X. D. Peng, 2004: A convexity preserving scheme for conservative advection transport. J. Comput. Phys., 198(2), 389−402, https://doi.org/10.1016/j.jcp.2004.01.013.
    Yukimoto, S., and Coauthors, 2019: The meteorological research institute Earth system model version 2.0, MRI-ESM2.0: Description and basic evaluation of the physical component. J. Meteor. Soc. Japan. Ser. II, 97, 931−965, https://doi.org/10.2151/jmsj.2019-051.
    Zeng, G., J. A. Pyle, and P. J. Young, 2008: Impact of climate change on tropospheric ozone and its global budgets. Atmospheric Chemistry and Physics, 8(2), 369−387, https://doi.org/10.5194/acp-8-369-2008.
    Zeng, G., O. Morgenstern, P. Braesicke, and J. A. Pyle, 2010: Impact of stratospheric ozone recovery on tropospheric ozone and its budget. Geophys. Res. Lett., 37, L09805, https://doi.org/10.1029/2010GL042812.
    Zhang, J. K., W. S. Tian, F. Xie, J. A. Pyle, J. Keeble, and T. Wang, 2020: The influence of zonally asymmetric stratospheric ozone changes on the Arctic polar vortex shift. J. Climate, 33(11), 4641−4658, https://doi.org/10.1175/JCLI-D-19-0647.1.
    Zhang, J. K., and Coauthors, 2022: Responses of Arctic sea ice to stratospheric ozone depletion. Science Bulletin, 67(11), 1182−1190, https://doi.org/10.1016/j.scib.2022.03.015.
    Zou, C. Z., and H. F. Qian, 2016: Stratospheric temperature climate data record from merged SSU and AMSU-A observations. J. Atmos. Oceanic Technol., 33(9), 1967−1984, https://doi.org/10.1175/JTECH-D-16-0018.1.
  • [1] Hyo-Eun JI, Soon-Hwan LEE, Hwa-Woon LEE, 2013: Characteristics of Sea Breeze Front Development with Various Synoptic Conditions and Its Impact on Lower Troposphere Ozone Formation, ADVANCES IN ATMOSPHERIC SCIENCES, 30, 1461-1478.  doi: 10.1007/s00376-013-2256-3
    [2] LI Mingwei, WANG Yuxuan*, and JU Weimin, 2014: Effects of a Remotely Sensed Land Cover Dataset with High Spatial Resolution on the Simulation of Secondary Air Pollutants over China Using the Nested-grid GEOS-Chem Chemical Transport Model, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 179-187.  doi: 10.1007/s00376-013-2290-1
    [3] WANG Mingxing, LIU Qiang, YANG Xin, 2004: A Review of Research on Human Activity Induced Climate Change I. Greenhouse Gases and Aerosols, ADVANCES IN ATMOSPHERIC SCIENCES, 21, 314-321.  doi: 10.1007/BF02915561
    [4] ZHU Renbin, SUN Liguang, YIN Xuebin, LIU Xiaodong, XING Guangxi, 2004: Summertime Surface N2O Concentration Observed on Fildes Peninsula Antarctica: Correlation with Total Atmospheric O3 and Solar Activity, ADVANCES IN ATMOSPHERIC SCIENCES, 21, 204-210.  doi: 10.1007/BF02915706
    [5] XUAN Shouli, ZHANG Qingyun, SUN Shuqing, 2011: Anomalous Midsummer Rainfall in Yangtze River-Huaihe River Valleys and Its Association with the East Asia Westerly Jet, ADVANCES IN ATMOSPHERIC SCIENCES, 28, 387-397.  doi: 10.1007/s00376-010-0111
    [6] Hailiang ZHANG, Yongfu XU, Long JIA, Min XU, 2021: Smog Chamber Study on the Ozone Formation Potential of Acetaldehyde, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 1238-1251.  doi: 10.1007/s00376-021-0407-5
    [7] LIU Yu, LI Weiliang, ZHOU Xiuji, I.S.A.ISAKSEN, J.K.SUNDET, HE Jinhai, 2003: The Possible Influences of the Increasing Anthropogenic Emissions in India on Tropospheric Ozone and OH, ADVANCES IN ATMOSPHERIC SCIENCES, 20, 968-977.  doi: 10.1007/BF02915520
    [8] Liang ZHANG, Bin ZHU, Jinhui GAO, Hanqing KANG, 2017: Impact of Taihu Lake on City Ozone in the Yangtze River Delta, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 226-234.  doi: 10.1007/s00376-016-6099-6
    [9] A.M.Selvam, M.Radhamani, 1994: Signatures of a Universal Spectrum for Nonlinear Variability in Daily Columnar Total Ozone Content, ADVANCES IN ATMOSPHERIC SCIENCES, 11, 335-342.  doi: 10.1007/BF02658153
    [10] XU Jun, ZHANG Yuanhang, WANG Wei, 2006: Numerical Study on the Impacts of Heterogeneous Reactions on Ozone Formation in the Beijing Urban Area, ADVANCES IN ATMOSPHERIC SCIENCES, 23, 605-614.  doi: 10.1007/s00376-006-0605-1
    [11] Junlin AN, Huan LV, Min XUE, Zefeng ZHANG, Bo HU, Junxiu WANG, Bin ZHU, 2021: Analysis of the Effect of Optical Properties of Black Carbon on Ozone in an Urban Environment at the Yangtze River Delta, China, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 1153-1164.  doi: 10.1007/s00376-021-0367-9
    [12] Lan GAO, Xu YUE, Xiaoyan MENG, Li DU, Yadong LEI, Chenguang TIAN, Liang QIU, 2020: Comparison of Ozone and PM2.5 Concentrations over Urban, Suburban, and Background Sites in China, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 1297-1309.  doi: 10.1007/s00376-020-0054-2
    [13] Yawei QU, Tijian WANG, Yanfeng CAI, Shekou WANG, Pulong CHEN, Shu LI, Mengmeng LI, Cheng YUAN, Jing WANG, Shaocai XU, 2018: Influence of Atmospheric Particulate Matter on Ozone in Nanjing, China: Observational Study and Mechanistic Analysis, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 1381-1395.  doi: 10.1007/s00376-018-8027-4
    [14] Junhua YANG, Shichang KANG, Yuling HU, Xintong CHEN, Mukesh RAI, 2022: Influence of South Asian Biomass Burning on Ozone and Aerosol Concentrations Over the Tibetan Plateau, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 1184-1197.  doi: 10.1007/s00376-022-1197-0
    [15] H. Kurtulus OZCAN, Erdem BILGILI, Ulku SAHIN, O. Nuri UCAN, Cuma BAYAT, 2007: Modeling of Trophospheric Ozone Concentrations Using Genetically Trained Multi-Level Cellular Neural Networks, ADVANCES IN ATMOSPHERIC SCIENCES, 24, 907-914.  doi: 10.1007/s00376-007-0907-y
    [16] WANG Feng, AN Junling, LI Ying, TANG Yujia, LIN Jian, QU Yu, CHEN Yong, ZHANG Bing, ZHAI Jing, 2014: Impacts of Uncertainty in AVOC Emissions on the Summer ROx Budget and Ozone Production Rate in the Three Most Rapidly-Developing Economic Growth Regions of China, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 1331-1342.  doi: 10.1007/s00376-014-3251-z
    [17] Zou Han, Ji Chongping, Zhou Libo, Wang Wei, Jian Yongxiao, 2001: ENSO Signal in Total Ozone over Tibet, ADVANCES IN ATMOSPHERIC SCIENCES, 18, 231-238.  doi: 10.1007/s00376-001-0016-2
    [18] Zou Han, Ji Chongping, Zhou Libo, 2000: QBO Signal in Total Ozone over Tibet, ADVANCES IN ATMOSPHERIC SCIENCES, 17, 562-568.  doi: 10.1007/s00376-000-0019-4
    [19] 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
    [20] Zou Han, Zhou Libo, Jian Yongxiao, Liu Yu, 2002: An Observational Study on the Vertical Distribution and Synoptic Variation of Ozone in the Arctic, ADVANCES IN ATMOSPHERIC SCIENCES, 19, 855-862.  doi: 10.1007/s00376-002-0050-8

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 07 June 2022
Manuscript revised: 27 September 2022
Manuscript accepted: 09 October 2022
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 Role of Ozone Depletion in the Lack of Cooling in the Antarctic Upper Stratosphere during Austral Winter

    Corresponding author: Lei WANG, wanglei_ias@fudan.edu.cn
  • 1. Department of Atmospheric and Oceanic Sciences & Institute of Atmospheric Sciences, Fudan University, Shanghai 200438, China
  • 2. Shanghai Qi Zhi Institute, Shanghai 200232, China
  • 3. Big Data Institute for Carbon Emission and Environmental Pollution, Fudan University, Shanghai 200438, China
  • 4. Lamont-Doherty Earth Observatory, Columbia University, Palisades, New York 10964, USA

Abstract: Temperature trends in the upper stratosphere are investigated using satellite measurements from Stratospheric Sounding Unit (SSU) outputs and simulations from chemistry–climate models (CCMs) and the Coupled Model Intercomparison Project Phase 6 (CMIP6). Observational evidence shows a lack of cooling in the Antarctic, in contrast to strong cooling at other latitudes, during austral winter over 1979–97. Analysis of CCM simulations for a longer period of 1961–97 also shows a significant contrast in the upper stratospheric temperature trends between the Antarctic and lower latitudes. Results from two sets of model integrations with fixed ozone-depleting substances (ODSs) and fixed greenhouse gases (GHGs) at their 1960 levels suggest that the ODSs have made a major contribution to the lack of cooling in the Antarctic upper stratosphere. Results from CMIP6 simulations with prescribed GHGs and ozone confirm that changes in the dynamical processes associated with observed ozone depletion are largely responsible for the lack of cooling in the Antarctic upper stratosphere. The lack of cooling is found to be dynamically induced through increased upward wave activity into the upper stratosphere, which is attributed mainly to ODSs forcing. Specifically, the radiative cooling caused by the ozone depletion results in a stronger meridional temperature gradient between middle and high latitudes in the upper stratosphere, allowing more planetary waves propagating upward to warm the Antarctic upper stratosphere. These findings improve our understanding of the chemistry–climate coupling in the southern upper stratosphere.

摘要: 本文利用平流层探测仪(SSU)的卫星观测数据,化学-气候模式(CCM)资料以及国际耦合模式比较计划第六阶段(CMIP6)模式数据分析了上平流层的温度变化趋势。观测资料表明,在1979-97年南半球冬季,相比于其他纬度地区的显著变冷,南极并无显著变冷趋势。1961-97年长期的CCM模式资料结果同样表明,南极和低纬度地区之间上平流层温度趋势变化存在显著差异。当臭氧消耗物(ODS)和温室气体(GHG)浓度分别固定在1960年水平上时,两组模式积分结果显示,ODS对南极上平流层的冷却缺失具有较大贡献。在CMIP6模式中,固定温室气体和臭氧浓度的模式结果也表明,与臭氧损耗相关的动力过程能够在很大程度上解释南极上平流层的冷却缺失现象。该冷却缺失主要是由ODS强迫更多的行星波进入平流层上层而导致的。具体而言,臭氧损耗引起的辐射冷却使得上平流层中纬度和高纬度之间的经向温度梯度增强,从而导致更多的行星波上传,使得南极上平流层增温。这一发现加深了我们对南半球上平流层化学-气候耦合的认识。

    • Observations and model simulations show substantial cooling in the global stratosphere since 1979, especially in the Antarctic due to the strong radiative effect of severe ozone depletion in austral spring and summer between the late 1970s and the late 1990s (Randel and Wu, 1999; Solomon, 1999; Ramaswamy et al., 2001; Thompson and Solomon, 2002; Hu, 2006). However, most studies only focus on the Antarctic stratospheric temperature changes in the lower stratosphere during austral spring (e.g., Sekiguchi, 1986; Randel and Wu, 1999; Gillett and Thompson, 2003; Hu and Fu, 2009; Ivy et al., 2016). Less attention has been paid to the Antarctic upper stratospheric temperature. In contrast to the cooling in the lower stratosphere, some studies recently have shown that the Antarctic upper stratosphere has experienced a lack of cooling since the late 1970s (Randel et al., 2016). Using measurements from the Stratospheric Sounding Unit (SSU), Randel et al. (2016) pointed out that the temperature trends in the Antarctic upper stratosphere are near zero, with significant cooling at other latitudes over 1979–97.

      Ball et al. (2016) pointed out that the upper stratospheric temperature has shown significant responses to changes in the upper deep branch of the Brewer–Dobson circulation (BDC) that lead to adiabatic heating, or cooling. The BDC is mainly driven by the midlatitude upward-propagating planetary and gravity waves through changing westward momentum as they break there (Haynes et al., 1991; Holton et al., 1995; Butchart, 2014). Wave forcing and the mean state of the flow could interact with each other (Charney and Drazin, 1961; Holton and Mass, 1976). Changes in wave forcing could also affect the spatial distribution of ozone by adjusting the strength of the BDC to change the ozone transport (e.g., Brewer, 1949; Dobson, 1956; Rind et al., 2002; García-Herrera et al., 2006; Jiang et al., 2007; Weber et al., 2011). Sridharan et al. (2012) and Nath and Sridharan (2015) showed that the dynamical perturbations at mid-to-high latitudes can directly influence the changes in ozone and temperature. Ozone can also in turn influence the stratospheric circulation (e.g., Hu and Tung, 2003; Nathan and Cordero, 2007; Hu et al., 2015; Zhang et al., 2020, 2022). For example, Hu et al. (2015) pointed out that stratospheric ozone depletion (recovery) leads to weakening (strengthening) of the Arctic polar vortex, and cooling of the Arctic polar vortex was found to be dynamically induced via modulation of planetary wave activity by stratospheric ozone increases during boreal winter. Several modeling studies also found that the BDC becomes stronger (weaker) in response to ozone depletion (recovery) (e.g., McLandress et al., 2010; Polvani et al., 2011, 2018, 2019; Lin and Fu, 2013). The aim of this paper is to gain a better understanding of the coupling between temperature, ozone, and the BDC in the upper stratosphere.

      Ozone-depleting substances (ODSs) and greenhouse gases (GHGs) are considered to be two major factors responsible for global stratospheric temperature changes (Stolarski et al., 2010). Previous studies have suggested a strong decreasing trend in Antarctic stratospheric ozone during the late 20th century and a dramatic increasing trend in global GHGs since the industrial revolution (Randel and Wu, 1999; Thompson and Solomon, 2002; Gillett et al., 2011; Polvani et al., 2011; Checa-Garcia et al., 2016). The ozone depletion and GHGs contribute significantly to the observed cooling in the stratosphere due to their strong radiative cooling effect (Ramaswamy et al., 1996, 2001; Shindell and Schmidt, 2004). Therefore, the lack of cooling in the Antarctic upper stratosphere is likely due to dynamical processes. There is a strong link between ozone and stratospheric temperature, showing negative dynamical feedbacks in the polar upper stratosphere (Albers and Nathan, 2013), and the feedbacks associated with ozone and planetary waves are height dependent. Moreover, planetary waves are also strongly modulated by GHGs, with stronger planetary waves forcing from the troposphere into the stratosphere when GHG concentrations increase (Butchart and Scaife, 2001; Kodera et al., 2008; Bell et al., 2010). Correctly modeling the dynamical response to ODSs and GHGs is essential to understand the processes that drive the temperature changes. Herein, we investigate the role of changes in ODSs and GHGs in the Antarctic upper stratospheric temperature, using simulations from the Chemistry–Climate Model Validation Activity Phase 2 (CCMVal-2) and Chemistry–Climate Model Initiative (CCMI) projects (Eyring et al., 2008, 2013), as well as the Coupled Model Intercomparison Project Phase 6 (CMIP6; Boer et al., 2016).

    2.   Data and methods
    • In the present study, the 1979–2014 monthly mean temperature from the SSU channel 3 (SSU3) on board the National Oceanic and Atmospheric Administration (NOAA) operational satellites is obtained from the NOAA/STAR website (http://www.star.nesdis.noaa.gov/smcd/emb/mscat/) [see Zou and Qian (2016) for more details on this dataset]. The 1961–2014 monthly mean temperature, ozone, and zonal wind are from 12 Chemistry–Climate Model (CCM) simulations (whose output is available for reference and two sensitivity simulations), except that the ozone output is unavailable for the NIWA-UKCA model from the CCMI dataset. The output of six simulations is obtained from the CCMVal-2 (Eyring et al., 2006, 2010; Morgenstern et al., 2010), and the other six simulations are from the CCMI (Eyring et al., 2013; Morgenstern et al., 2017). These models are summarized in Table 1. Three different forcing scenarios are considered here: “REF-B2” for CCMVal-2 and “REF-C2” for CCMI in which the reference models are forced by time-varying concentrations of ODSs and GHGs (All forcings); “SCN-B2b” for CCMVal-2 and “SEN-C2-fODS” for CCMI with fixed ODSs at 1960 levels; “SCN-B2c” for CCMVal-2 and “SEN-C2-fGHG” for CCMI with fixed GHGs at 1960 levels. More details can be found in Morgenstern et al. (2010, 2017). Similar to the above models, three types of simulations from CMIP6 are also employed: historical simulations and two new historical simulations with stratospheric ozone-only and well-mixed GHG-only forcing. Details are shown in Table 2.

      Model nameEnsemble
      members
      Domain/Resolution (lat×lon), levels, model lidReference
      CCMVal-2
      CCSRNIES12.8°×2.8°(T42), 34L, 0.01 hPaAkiyoshi et al. (2009),
      Kurokawa et al. (2005)
      CMAM33.75°×3.75°(T31), 71L, 0.06 Pade Grandpré et al. (2000),
      Hitchcock et al. (2009)
      MRI12.8°×2.8°(T42), 68L, 0.01 hPaShibata and Deushi (2008),
      Xiao and Peng (2004)
      SOCOL13.75°×3.75°(T30), 39L, 0.1 hPaSchraner et al. (2008),
      Egorova et al. (2005)
      CCMI
      CCSRNIES-MIROC3.212.8°×2.8°(T42), 34L, 1.2 PaAkiyoshi et al. (2016),
      Sakazaki et al. (2013)
      CMAM12.5°×2.5°(T47), 71L, 0.08 PaScinocca et al. (2008),
      de Grandpré et al. (2000)
      NIWA-UKCA13.75°×2.5°, L60, 84 kmMorgenstern et al. (2009, 2013),
      Zeng et al. (2008, 2010)
      WACCM31.9°×2.5°, L66, 140 kmSolomon et al. (2015),
      Garcia et al. (2017)

      Table 1.  Details of the CCMI and CCMVal-2 simulations in which column 3 corresponds to the horizontal resolution, the number of vertical levels, and model lid. For details, please refer to references in column 4. Three types of simulations are employed: the reference simulations and two sensitivity simulations with fixed ODSs and GHGs at their 1960 levels. Specifically, they are REF-B2, SCN-B2b, and SCN-B2c for CCMVal-2 and REF-C2, SEN-C2-fODS, and SEN-C2-fGHG for CCMI.

      Model nameEnsemble
      members
      Domain/Resolution (lat×lon), levels, model lidReference
      MIROC631.41° × 1.41°(T85), 19L, 1 hPaTatebe et al. (2019)
      CanESM5102.81° × 2.81°(T42), 19L, 1 PaSwart et al. (2019)
      MRI-ESM2-031.125° × 1.125°(T106), 19L, 1 hPaYukimoto et al. (2019)
      IPSL-CM6A-LR102.5° × 1.26°, 19L, 1 hPaLurton et al. (2020)

      Table 2.  Same as Table 1, but for the CMIP6 simulations. Three types of simulations are employed: the historical simulations and two sensitivity simulations with historical stratospheric ozone-only and historical well-mixed GHG-only forcing.

      Other meteorological datasets, including observed ozone from the Stratospheric Water and Ozone Satellite Homogenized (SWOOSH) datasets (Davis et al., 2016), as well as monthly mean temperature, 6-hourly temperature, and zonal wind from the ECMWF fifth-generation reanalysis, ERA5 (Hersbach and Dee, 2016; Li et al., 2020), and the Japanese 55-year Reanalysis, JRA55 (Ebita et al., 2011; Kobayashi and Iwasaki, 2016), are employed in this paper. The meridional eddy heat flux is obtained from the zonal mean of the product of the temperature and the eddy component of the zonal wind based on 6-hourly datasets. The vertically integrated temperature trends calculated in this study are based on the SSU3 weighting functions. Details of the weighting functions can be found in Randel et al. (2016).

    • The divergence of the Eliassen–Palm (EP) flux is a key diagnostic for wave–mean-flow interaction and can be used to describe the resolved wave breaking (Andrews and McIntyre, 1976, 1978; Andrews et al., 1987). The EP flux is given as (see Andrews et al., 1987):

      where $u$, $v,$ and $w$ are the zonal, meridional, and vertical velocity, respectively. $\theta $ is the potential temperature. $a$ is the radius of the Earth.$ \text{ }{\rho }_{0} $ is the air density. $f$ is the Coriolis parameter. $H$ is the density scale height taken as 7 km. ${P_0}$=1000 hPa, and P is pressure. Subscripts $z$ and $\varphi $ denote derivatives with respect to pressure and latitude, respectively. An overbar represents the zonal mean, and a prime symbol denotes departures from the zonal mean. The divergence of the EP flux calculated in this paper is based on 6-hourly datasets.

      The thermal wind equation indicates that the horizontal temperature gradients are closely associated with vertical wind shear:

      where R is the specific gas constant for dry air and y is northward distance.

      The refractive index (n) of planetary waves developed by Andrews et al. (1987) based on the form of Matsuno (1970) is applied in this study and was expanded by Hu and Tung (2002):

      where

      is the meridional gradient of the zonal mean potential vorticity. Here, $k$, $N$, and $ \Omega $ denote the zonal wave number (ZWN hereafter), buoyancy frequency, and Earth rotation frequency, respectively. Expansion of the third term on the right-hand side of Eq. (7) yields:

    • Nekola and White (1999) proposed a method which can be used to calculate the difference in slope between two datasets [e.g., (x1, y1) and (x2, y2)]. Although it was initially built to calculate the difference in slope between the regression lines of distance decay plots, it could be extended to anything else. First, the y values of the two datasets compared were rescaled to a common mean. Then, the y value for each pair along with the corresponding x was randomly reassigned to one of the two datasets. After this randomization had been carried out, linear regressions were applied to determine the slope of the y function for each of the randomized datasets, and the absolute value of the difference between the two slopes was determined. This procedure was repeated 999 times. The difference between the slopes of the original datasets was also calculated and compared with the distribution of the differences between the slopes of the 999 randomized datasets in order to determine its significance level. Details of this method can be seen in Nekola and White (1999), Steinitz et al. (2005), and Steinitz et al. (2006).

    3.   Results
    • Changes in the Antarctic upper stratospheric temperature are good indicators for the chemistry–climate coupling in the southern upper stratosphere. In the late 20th century, the increased GHGs and decreased stratospheric ozone are considered to be significantly related to the strong cooling in the global stratosphere due to their strong radiative effect. However, using SSU3 observations for 1979–97, Randel et al. (2016) found that the annual mean temperature trends in the upper stratosphere are near zero at southern high latitudes, but there is strong cooling at other latitudes, indicating a significant difference in temperature trends between southern high and low latitudes. To further investigate this problem, the latitudinal distribution of linear trends for zonal mean temperature for each season are examined (Fig. 1). It is found that the difference in temperature trends between southern high and low latitudes is most significant in austral winter (Fig. 1c). However, there is no significant difference in temperature trends between northern high and low latitudes during austral summer (Fig. 1a). Compared with Fig. 1a, in austral autumn and spring, the temperature trends in the Antarctic upper stratosphere are relatively weaker (Figs. 1b, 1d) and may be influenced by lower stratospheric ozone changes and the associated dynamical effects. Therefore, this study focuses on the austral winter temperature trends in the Antarctic upper stratosphere. Using Solar Backscatter Ultraviolet (SBUV) data, Hood et al. (1993) estimated the altitude, latitude, and seasonal dependences of stratospheric ozone trend. Results show that the maximum negative trend occurs in the southern upper stratosphere during austral winter, making it easier to isolate the causality of the temperature trend there. Considering that ozone depletion and increasing GHGs contribute significantly to stratospheric cooling due to their strong radiative cooling effect, why is there a lack of cooling in the Antarctic upper stratosphere during austral winter? The topic of this paper mainly focuses on the lack of cooling in the Antarctic upper stratosphere in austral winter. Temperature trends in other seasons and in the northern hemisphere will be discussed in the final section.

      Figure 1.  Latitude profiles of linear trends for monthly zonal mean SSU3 temperature in austral (a) summer, (b) autumn, (c) winter, and (d) spring for 1979–97. Thick black solid lines indicate that the trend at each latitude in the range of 90°–30°S is significantly different from the trend at the equator of 0° at the 95% confidence level. For the specific method of significance test between the two trends, please refer to section 2.3.

      What is responsible for the lack of cooling in the Antarctic upper stratosphere in austral winter? Zonal mean temperature trends in austral winter derived from three types of simulations are shown in Fig. 2. An obvious difference in temperature trends between southern high and low latitudes can be found in the ensemble mean with all forcings and with fixed GHGs, indicating that the lack of cooling in the upper stratosphere at southern high latitudes may be the result of increased ODSs during 1979–97 (Fig. 2a). Note that 12 chemistry–climate models from the CCMVal-2 and CCMI datasets are analyzed because only these models have all reference and sensitivity simulations available. To test the representativity of these models, we compared the results from the 12 reference simulations with those from all 34 reference simulations obtained from CCMVal-2 and CCMI simulations (Table 3). Results from the 12 members are similar to those from the 34 members, suggesting that the former is a good representation of the latter. In addition, the results from two reanalysis datasets (based on the ERA5 and JRA55) are also in good agreement with the results from satellite observations, with a distinct lack of upper stratospheric cooling in the Antarctic. However, the peak from using the two reanalysis datasets is near 60°S and is more equatorward than the peak from using the satellite observations, which may be attributed to too few observations in the upper Antarctic stratosphere. To address potential sensitivity due to the short time period, results from three types of simulations from CCMs and CMIP6 for a longer period of 1961–2014 are shown in Figs. 2cf, and they are in good agreement with the results for 1979–97. Considering the consistency of the results from CCMs and CMIP6, and our focus on chemical–dynamic processes, an analysis based on the CCMs is given in the following sections.

      Figure 2.  (a) Same as Fig. 1c. The grey line shows the temperature trends derived from NOAA. The brown, green, dashed orange, and dashed blue lines represent the ensemble-mean trends of CCMs. The brown line represents the ensemble-mean temperature trends of CCMs (Table 1, 12 simulations) with all forcings. The green line is the same as the brown line, but for 34 simulations (Table 3). Dashed orange and dashed blue lines represent the ensemble mean with fixed ODSs and GHGs (Table 1, 12 simulations), respectively. Solid pink and purple lines are from reanalysis data of ERA5 and JRA55 datasets, respectively. (b) is the same as (a), but for 1998–2014. (c) and (d) are similar to (a) and (b), but for the ensemble mean (Table 1, 12 members) of the longer periods of 1961–97 and 1998–2014. (e) and (f) are the same as (c) and (d), but for the ensemble mean of CMIP6 simulations (Table 2, 26 members).

      Model nameEnsemble members Model nameEnsemble members
      CCMVal-2 CCMI
      CCSRNIES1CCSRNIES-MIROC3.21
      CMAM3CMAM1
      MRI2NIWA-UKCA1
      SOCOL3WACCM3
      CAM3.51ACCESS-CCM1
      LMDZrepro1IPSL1
      Niwa-SOCOL1CHASER-MIROC-ESM1
      UMSLIMCAT1CNRM-CM5-31
      WACCM3EMAC-L47MA1
      CNRM-ACM1EMAC-L90MA1
      UMUKCA-METO1GEOSCCM1
      SOCOL31
      HadGEM3-ES1
      MRI-ESM1r11

      Table 3.  Details of the CCMI and CCMVal-2 simulations. Only reference experiments are employed: REF-B2 for CCMVal-2 and REF-C2 for CCMI.

      Latitude–height cross sections of ensemble-mean linear trends in the stratospheric temperature are presented in Fig. 3. Significant differences in temperature trends between southern high latitudes and low latitudes can be seen in simulations with all forcings for 1961–97 (Fig. 3a). This may be attributed to the ODSs forcing (Fig. 3e), which produces a significant contrast of temperature trends between the two regions. Furthermore, we quantitatively investigate the temperature trend difference between the southern high (60°–90°S) and low latitudes (0°–30°S). The difference in temperature trends between the two regions exceeds 0.32 K (10 yr)–1 in both the reference and fixed-GHG simulations, with only about –0.03 K (10 yr)–1 in the fixed-ODSs simulations. Moreover, the trend difference between the two sensitivity simulations is significantly different (at the 99% confidence level) in the first period (1961–97). Temperature trends from CMIP6 simulations (Fig. 4) are similar to the results from CCMs (Fig. 3), showing the lack of cooling in the Antarctic upper stratosphere with ozone forcings during 1961–97. These findings suggest that the lack of cooling in the Antarctic upper stratosphere is mainly induced by ozone depletion due to increased ODSs. During 1998–2014, the Antarctic warming may be a statistical artifact due to the short length of this period, and there is no significant warming for a longer period of 1998–2099.

      Figure 3.  Latitude–height cross sections of ensemble-mean linear trends for zonal mean temperature [K (10 yr)–1] in austral winter over (a, c, e) 1961–97 and (b, d, f) 1998–2014. (a) and (b) are for the reference ensemble mean. (c, d) and (e, f) are the same as (a, b), but for the ensemble mean with fixed ODSs and GHGs, respectively. Detailed simulations can be seen in Table 1. The dotted areas indicate that the trends are statistically significant at 95% confidence levels.

      Figure 4.  Same as Fig. 3, but for the ensemble mean of CMIP6 simulations (Table 2, 26 members).

      The above results suggest that the increased ODSs are the main contributor to the lack of cooling in the Antarctic upper stratosphere during 1961–97. However, this is opposite to the direct radiative cooling related to ozone depletion during that period (Figs. 5 and 6). Thus, the lack of cooling in the Antarctic upper stratosphere may be related to dynamical processes with more upward planetary waves into stratosphere indirectly induced by increased ODSs.

      Figure 5.  Latitude–height cross sections of ensemble-mean linear trends for zonal mean ozone [10−7 mol mol–1 (10 yr)–1] in austral winter for 1961–97 (left panel) and 1998–2014 (right panel). (a) and (b) are for the reference ensemble mean. (c, d) and (e, f) are the same as (a, b), but for the ensemble mean with fixed ODSs and GHGs, respectively. Detailed simulations can be seen in Table 1. The dotted areas indicate that the trends are statistically significant at 95% confidence levels.

      Figure 6.  Latitude–height cross sections of linear trends for observed zonal mean ozone [ppmv (10 yr)–1] in austral winter for (a) 1984–97 and (b) 1998–2014. The dataset is from SWOOSH. The dotted areas indicate that the trends are statistically significant at 95% confidence levels.

      Ozone depletion induced by an increase in ODSs can cool the middle and high latitudes via the radiative effect, which would cause a larger meridional temperature gradient between high and low latitudes. By the thermal wind balance, the midlatitude baroclinic zone of the atmosphere is associated with a planetary-scale meridional temperature gradient between the equator and the pole. Zonal wind trends for the three groups of simulations are presented in Fig. 7. For reference and fixed-GHG simulations, significant vertical shear is shown in the midlatitude zonal winds for 1961–97 in Figs. 7a and 7e, consistent with the meridional gradient in the temperature trends displayed in Figs. 3a and 3e and that in the ozone trends shown in Figs. 5a and 5e.

      Figure 7.  As in Fig. 5, but for zonal mean wind [m s–1 (10 yr)–1].

      Results from reanalysis datasets are also examined. Signals of a significant lack of cooling are seen in the Antarctic upper stratosphere during austral winter for 1979–97 (Fig. 8a), consistent with the observations (Fig. 2). The eddy heat flux is a useful proxy for the vertical component of EP flux (Andrews et al., 1987) and can represent wave energy propagating into the stratosphere (Newman et al., 2001; Eyring et al., 2005). Figure 8c shows that the lack of cooling in the Antarctic upper stratosphere is accompanied by a negative trend in the eddy heat flux, indicating more planetary waves propagating into the southern stratosphere. It is in good agreement with the results shown in Fig. 8e. Subsequently, more planetary waves break up in the upper stratosphere at southern high latitudes during austral winter for 1979–97 (Fig. 8g). In contrast, the eddy heat flux trends and EP flux convergence trends afterward did not change significantly.

      Figure 8.  (a) Same as Fig. 3, but for the average of ERA5 and JRA55. (c) Latitude–height cross sections of linear trends for meridional eddy heat flux [K m s–1 (10 yr)–1] in austral winter for 1979–97. (e) The climatology (black line, K m s–1) and linear trends [red line, K m s–1(10 yr)–1] of vertically integrated meridional eddy heat flux in austral winter for 1979–97. (g) Latitude–height cross sections of linear trends for EP Flux divergence [m s–1 d–1 (10 yr)–1] in austral winter for 1979–97. (b, d, f, h) Same as (a, c, e, g), but for 1998–2014.

      Previous studies have found that stratospheric ozone depletion may change the temperature gradient or potential vorticity gradient, which are closely related to the wave refractive index (e.g., Hu et al., 2015; Wang et al., 2021). Figure 9 shows wave refractive index trends during austral winter. As discussed by Matsuno (1970), it is expected that planetary waves of wave number k are able to propagate in regions where $ {n_k}^2 $>0 and are refracted from regions where $ {n_k}^2 $<0, and the larger the $ {n_k}^2 $ in a region, the easier it is for planetary waves to propagate there. Our results show that in austral winter, planetary waves have a great chance to propagate upward from the lower stratosphere to the upper stratosphere at southern middle and high latitudes for zonal wave 1 (ZWN1) during 1979–97 (Fig. 9a). However, the waves are more likely to propagate to the Antarctic middle stratosphere during 1998–2014 (Fig. 9b). Therefore, dynamical warming is more plausible in the Antarctic upper stratosphere during 1979–97. Similar conclusions could be drawn for zonal waves 2 and 3 (ZWN2 and ZWN3). The above conclusions suggest that the lack of cooling in the Antarctic upper stratosphere is mainly attributed to dynamical processes.

      Figure 9.  (a) Latitude–height cross sections of linear trends for refractive index squared (a2n2) for ZWN1 (first row) planetary waves (the average of ERA5 and JRA55) in austral winter for 1979–97. (c) and (e) are the same as (a), but for ZWN2 and ZWN3, respectively. (b, d, f) are the same as (a, c, e), but for 1998–2014.

    4.   Conclusions and discussion
    • Using observations and multi-model ensemble data from CCMs and CMIP6, we evaluated upper stratospheric temperature trends. Observations show that in the late 20th century, Antarctic upper stratospheric temperature trends exhibit distinct seasonal variation with an obvious lack of cooling in austral winter, which can be reproduced by chemistry–climate models.

      ODSs and GHGs are two major contributors to variations in the stratospheric temperature. To further distinguish their contributions, three sets of integrations are investigated in this paper. Our results suggest that the lack of cooling in the Antarctic upper stratosphere is mainly dynamically driven by increased ODSs in the late 20th century. Ozone depletion related to increased ODSs could strengthen the meridional temperature gradient in the middle latitudes and allow more planetary waves into the upper stratosphere, which in turn could weaken zonal winds through the breaking of planetary waves. This would intensify the meridional residual circulation (Fu et al., 2019), resulting in enhanced adiabatic heating in the Antarctic upper stratosphere. These changes suggest that the lack of cooling in the Antarctic upper stratosphere is mainly due to enhanced wave-driven dynamical heating.

      Unlike the Antarctic temperature trends in austral winter, no significant lack of cooling is found in the Arctic upper stratosphere in austral summer. This hemispheric asymmetry is probably due to large internal variability in the northern polar stratosphere, as these models simulate a temperature gradient that is similar to the simulated Antarctic counterpart (now shown) but is missing in observations (Fig. 1a) and reanalysis datasets. In austral autumn and spring, it is difficult to separate the contributions from lower and upper stratospheric ozone for the observed lack of cooling in the polar upper stratosphere (Figs. 1b, 1d). In contrast, in austral winter, it is easier to isolate the causality of the temperature trend there.

      Although the reanalysis data can reproduce the lack of cooling in the Antarctic and its related dynamical mechanism, some issues need to be noted. The quality of the two reanalysis datasets at southern high latitudes is relatively poor. In addition to the latitude of peak warming being a little too far northward compared to the satellite observation, we have also found large residuals in the balance of the thermodynamic equation. This indicates that the quality of the reanalysis data in the upper Antarctic stratosphere needs to be further improved. Improved observations in the upper Antarctic stratosphere are called for in the future.

      Acknowledgements. This project was supported by Grant Nos. 41875047 and 91837206 from the National Natural Science Foundation of China (NSFC) and Grant No. JIH2308007 from Fudan University. We thank NOAA for producing SSU satellite temperature datasets, SWOOSH for the ozone dataset, ECMWF and JMA for meteorological fields from ERA5 and JRA55, the WCRP SPARC Chemistry–Climate Model Validation (CCMVal) Activity, the SPARC/IGAC Chemistry–Climate Model Initiative (CCMI), CMIP6, DAMIP, the participating modelling groups, and the ESGF centres (see details on the CMIP Panel website at http://www.wcrp-climate.org/index.php/wgcm-cmip/about-cmip) for organizing and coordinating the model data analysis activity, as well as the British Atmospheric Data Centre (BADC) for collecting and archiving the model output. We also acknowledge Dr. William RANDEL from NCAR and Dr. Lorenzo POLVANI from Columbia University for providing helpful comments for this study.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return