Advanced Search
Article Contents

The Significant Role of Radiosonde-measured Cloud-base Height in the Estimation of Cloud Radiative Forcing

Fund Project:

We acknowledge the support from the National Key R&D Program of China under Grants Nos. 2017YFC1501401 and 2017YFC0212803, the National Natural Science Foundation under Grant No. 41771399, and the Chinese Academy of Meteorological Sciences under Grant No. 2018Y014. Also, we are grateful to NASA for granting access to the level-3 cloud products from MODIS/Aqua. Last but not least, special thanks go to the National Meteorological Information Center, China Meteorological Administration for generously providing the radiosonde datasets used here (https://data.cma.cn/en)


doi: 10.1007/s00376-021-0431-5

  • The satellite-based quantification of cloud radiative forcing remains poorly understood, due largely to the limitation or uncertainties in characterizing cloud-base height (CBH). Here, we use the CBH data from radiosonde measurements over China in combination with the collocated cloud-top height (CTH) and cloud properties from MODIS/Aqua to quantify the impact of CBH on shortwave cloud radiative forcing (SWCRF). The climatological mean SWCRF at the surface (SWCRFSUR), at the top of the atmosphere (SWCRFTOA), and in the atmosphere (SWCRFATM) are estimated to be −97.14, −84.35, and 12.79 W m−2, respectively for the summers spanning 2010 to 2018 over China. To illustrate the role of the cloud base, we assume four scenarios according to vertical profile patterns of cloud optical depth (COD). Using the CTH and cloud properties from MODIS alone results in large uncertainties for the estimation of SWCRFATM, compared with those under scenarios that consider the CBH. Furthermore, the biases of the CERES estimation of SWCRFATM tend to increase in the presence of thick clouds with low CBH. Additionally, the discrepancy of SWCRFATM relative to that calculated without consideration of CBH varies according to the vertical profile of COD. When a uniform COD vertical profile is assumed, the largest SWCRF discrepancies occur during the early morning or late afternoon. By comparison, the two-point COD vertical distribution assumption has the largest uncertainties occurring at noon when the solar irradiation peaks. These findings justify the urgent need to consider the cloud vertical structures when calculating the SWCRF which is otherwise neglected.
    摘要: 现有基于卫星观测的云辐射强迫定量估算精度仍十分有限,这在很大程度上归因于云底高度(CBH)表征的局限性或不确定性。本文利用中国区域L波段秒级探空数据计算得到的CBH数据,及MODIS / Aqua的云特性和云顶高度(CTH)产品数据,量化了CBH对短波云辐射强迫(SWCRF)的影响。结果表明:中国2010年至2018年夏季云的短波辐射强迫在地表(SWCRFSUR)、大气顶部(SWCRFTOA)和大气中(SWCRFATM)的平均值分别为–97.14、–84.35和12.79 W m−2。为阐明云底作用,我们在利用辐射传输模式估算云短波辐射强迫时,根据云光学厚度(COD)垂直分布廓线做了四种情景假设。与考虑CBH的方案相比,仅使用MODIS CTH和云特性的方案会导致估计的SWCRFATM产生较大的不确定性。除此之外,在云底高度较低且云层较厚的情况下,云与地球辐射能量系统(CERES)的SWCRFATM偏差呈现增加的趋势。另外,相对于不考虑CBH的情景,估算得到的SWCRFATM偏差会随COD的垂直廓线而变化。当假定COD在垂直方向呈均匀分布时,最大的SWCRF差异出现在清晨或午后。相比之下,COD在垂直方向呈两点分布的假设具有最大的不确定性,该不确定性发生在一天中太阳辐射达到峰值的正午。鉴于云底高度在云短波辐射强迫估算中的重要作用,因此在计算SWCRF时须考虑云的垂直结构特征。
  • 加载中
  • Figure 1.  The normalized cloud optical depth (COD; ${\tau }_{{\rm{c}}}$) shown as a function of normalized cloud height (km) for three different vertical profile Scenarios of COD: (a) Scenario 1, (b) Scenario 2, and (c) Scenario 3, based on the assumption of the true number of the cloud layer is equal to 10 and the height of cloud is uniform.

    Figure 2.  Spatial distributions of (a) cloud-top height (CTH) from MODIS/Aqua, (b) cloud-base height (CBH) from radiosonde measurements, and (c) cloud thickness (CT) from MODIS/Aqua over China in summer (June−August) for the period from 2010 to 2018. The blue bars on the right-hand side represent the frequency of cloud height in different latitudes.

    Figure 3.  The frequency distribution of (a) CTH (red bars), (b) CBH (blue bars), and (c) cloud thickness (green bars) over China in summer for the period from 2010 to 2018. The data in the histogram are pooled into 1.0 km bins. The red, blue, and green curves depict the kernel density estimate (KDE) of the probability distribution function (PDF) for CTH, CBH, and cloud thickness, respectively.

    Figure 4.  Evolution, as a function of latitude and percentile, for (a) cloud-base height, (b) cloud-top height, and (c) cloud thickness over China in the summers during the period from 2010 to 2018. The 10th−90th percentiles in 10 percentile increments are shown, with the 50th and 90 th percentiles highlighted in blue and red, respectively.

    Figure 5.  Scatterplots of (a) SWCRFSUR, (b) SWCRFTOA, and (c) SWCRFATM using the CBH from radiosonde measurement, MODIS-observed CTH, and ancillary cloud parameters under Scenario 2 versus that from the CERES product in China. The regression lines of estimated SWCRFs against CERES-derived SWCRFs are given as well. Note that all samples are limited to the period in summer in 2013.

    Figure 6.  The difference between the CERES-derived SWCRFATM and the simulated SWCRFATM (ΔSWCRFATM) shown as a function of (a) cloud thickness, (b) cloud-base height, and (c) cloud-top height. The red points represent those samples with negative SWCRFATM from CERES.

    Figure 7.  Spatial distributions of (a) SWCRFSUR, (b) SWCRFTOA, and (c) SWCRFATM under Scenario 2 in China in summer for the period from 2010 to 2018. The blue bars on the right-hand side represent their corresponding frequency of SWCRF in various latitudes.

    Figure 8.  The (a) SWCRFSUR, (b) SWCRFTOA, and (c) SWCRFATM shown as a function of CBH under four vertical configuration scenarios of cloud by assuming the vertical profile of COD: Scenario 0 (black), Scenario 1 (red), Scenario 2 (green), and Scenario 3 (blue).

    Figure 9.  Comparison of the nationwide mean summertime values for (a) SWCRFSUR, (b) SWCRFTOA, and (c) SWCRFATM in China during the period from 2010 to 2018 under four vertical configuration scenarios of cloud by assuming the vertical profile of COD: Scenario 0 (black), Scenario 1 (red), Scenario 2 (green), and Scenario 3 (blue).

    Figure 10.  SWCRFSUR (a−c), SWCRFTOA (d−f), and SWCRFATM (g−i) shown as a function of the increase in CBH (Δhb) under four vertical configuration scenarios of cloud by assuming the vertical profile of the COD: Scenario 0 (black), Scenario 1 (red), Scenario 2 (green), and Scenario 3 (blue).

    Table 1.  Summary for the vertical profile scenarios of cloud optical depth (COD) assumed in the calculation of cloud optical depth, which is an important input for the calculation of cloud radiative forcing.

    ScenariosVertical profile pattern of CODNumber of cloud layers (N)
    Scenario 0Uniform1
    Scenario 1two-point2
    Scenario 2Uniformtrue cloud layer
    Scenario 3Normaltrue cloud layer
    DownLoad: CSV
  • Allan, R. P., C. L. Liu, N. G. Loeb, M. D. Palmer, M. Roberts, D. Smith, and P.-L. Vidale, 2014: Changes in global net radiative imbalance 1985−2012. Geophys. Res. Lett., 41, 5588−5597, https://doi.org/10.1002/2014GL060962.
    Clough, S. A., M. W. Shephard, E. J. Mlawer, J. S. Delamere, M. J. Iacono, K. CadyPereira, S. Boukabara, and P. D. Brown, 2005: Atmospheric radiative transfer modeling: A summary of the AER codes. J. Quant. Spectrosc. Radiat. Transf., 91, 233−244, https://doi.org/10.1016/j.jqsrt.2004.05.058.
    Chen, D. D., J. P. Guo, H. Q. Wang, J. Li, M. Min, W. H. Zhao, and D. Yao, 2018: The cloud top distribution and diurnal variation of clouds over East Asia: Preliminary results from advanced Himawari imager. J. Geophys. Res., 123, 3724−3739, https://doi.org/10.1002/2017JD028044.
    Chen, D. D., and Coauthors, 2019a: Mesoscale convective systems in the asian monsoon region from Advanced Himawari imager: Algorithms and preliminary results. J. Geophys. Res., 124, 2210−2234, https://doi.org/10.1029/2018JD029707.
    Chen, T. M., J. P. Guo, Z. Q. Li, C. F. Zhao, H. Liu, M. Cribb, F. Wang, and J. He, 2016: A CloudSat perspective on the cloud climatology and its association with aerosol perturbations in the vertical over Eastern China. J. Atmos. Sci., 73, 3599−3616, https://doi.org/10.1175/JAS-D-15-0309.1.
    Chen, Y. L., and Y. F. Fu, 2018: Tropical echo-top height for precipitating clouds observed by multiple active instruments aboard satellites. Atmos. Res., 199, 54−61, https://doi.org/10.1016/j.atmosres.2017.08.008.
    Chen, Y. L., K. Z. Chong, and Y. F. Fu, 2019b: Impacts of distribution patterns of cloud optical depth on the calculation of radiative forcing. Atmos. Res., 218, 70−77, https://doi.org/10.1016/j.atmosres.2018.11.007.
    Costa-Surós, M., J. Calbó, J. A. González, and J. Martin-Vide, 2013: Behavior of cloud base height from ceilometer measurements. Atmospheric Research, 127, 64−76, https://doi.org/10.1016/j.atmosres.2013.02.005.
    Dai, A. G., T. R. Karl, B. M. Sun, and K. E. Trenberth, 2006: Recent trends in cloudiness over the United States: A tale of monitoring inadequacies. Bull. Amer. Meteor. Soc., 87, 597−606, https://doi.org/10.1175/BAMS-87-5-597.
    George, G., C. Sarangi, S. N. Tripathi, T. Chakraborty, and A. Turner, 2018: Vertical structure and radiative forcing of monsoon clouds over Kanpur during the 2016 INCOMPASS field campaign. J. Geophys. Res., 123, 2152−2174, https://doi.org/10.1002/2017JD027759.
    Guo, J. P., and Coauthors, 2016: The climatology of planetary boundary layer height in China derived from radiosonde and reanalysis data. Atmospheric Chemistry and Physics, 16, 13 309−13 319, https://doi.org/10.5194/acp-16-13309-2016.
    Guo, J. P., and Coauthors, 2019: Shift in the temporal trend of boundary layer height in China using long‐term (1979−2016) radiosonde data. Geophys. Res. Lett., 46(11), 6080−6089, https://doi.org/10.1029/2019GL082666.
    Henderson, D. S., T. L’Ecuyer, G. Stephens, P. Partain, and M. Sekiguchi, 2013: A multisensor perspective on the radiative impacts of clouds and aerosols. J. Appl. Meteorol. Climatol., 52, 853−871, https://doi.org/10.1175/JAMC-D-12-025.1.
    Kato, S., and Coauthors, 2011: Improvements of top-of-atmosphere and surface irradiance computations with CALIPSO-, CloudSat-, and MODIS-derived cloud and aerosol properties. J. Geophys. Res., 116, D19209, https://doi.org/10.1029/2011JD016050.
    Li, Z. Q., and Coauthors, 2019: East Asian study of tropospheric aerosols and their impact on regional clouds, precipitation, and climate (EAST‐AIRCPC). J. Geophys. Res., 124, 13 026−13 054, https://doi.org/10.1029/2019JD030758.
    Loeb, N. G., S. Kato, W. Y. Su, T. Wong, F. G. Rose, D. R. Doelling, J. R. Norris, and X. L. Huang, 2012: Advances in understanding top-of-atmosphere radiation variability from satellite observations. Surveys in Geophysics, 33, 359−385, https://doi.org/10.1007/s10712-012-9175-1.
    Lou, M. Y., and Coauthors, 2019: On the relationship between aerosol and boundary layer height in Summer in China under different thermodynamic conditions. Earth and Space Science, 6(5), 887−901, https://doi.org/10.1029/2019EA000620.
    Lü, Q. Y., J. M. Li, T. H. Wang, and J. P. Huang, 2015: Cloud radiative forcing induced by layered clouds and associated impact on the atmospheric heating rate. Journal of Meteorological Research, 29, 779−792, https://doi.org/10.1007/s13351-015-5078-7.
    Martucci, G., C. Milroy, and C. D. O’Dowd, 2010: Detection of cloud-base height using jenoptik CHM15K and vaisala CL31 ceilometers. J. Atmos. Oceanic Technol., 27, 305−318, https://doi.org/10.1175/2009JTECHA1326.1.
    Meloni, D., A. di Sarra, T. Di Iorio, and G. Fiocco, 2005: Influence of the vertical profile of Saharan dust on the visible direct radiative forcing. Journal of Quantitative Spectroscopy and Radiative Transfer, 93, 397−413, https://doi.org/10.1016/j.jqsrt.2004.08.035.
    Menzel, W. P., and Coauthors, 2008: MODIS global cloud-top pressure and amount estimation: Algorithm description and results. J. Appl. Meteorol. Climatol., 47, 1175−1198, https://doi.org/10.1175/2007JAMC1705.1.
    Min, M., and Z. B. Zhang, 2014: On the influence of cloud fraction diurnal cycle and sub-grid cloud optical thickness variability on all-sky direct aerosol radiative forcing. Journal of Quantitative Spectroscopy and Radiative Transfer, 142, 25−36, https://doi.org/10.1016/j.jqsrt.2014.03.014.
    Min, M., J. Li, F. Wang, Z. J. Liu, and W. P. Menzel, 2020: Retrieval of cloud top properties from advanced geostationary satellite imager measurements based on machine learning algorithms. Remote Sens. Environ., 239, 111616, https://doi.org/10.1016/j.rse.2019.111616.
    Mishra, S., and A. Datta-Gupta, 2018: Applied Statistical Modeling and Data Analytics: A Practical Guide for the Petroleum Geosciences. Elsevier Inc.
    Mitchell, D. L., and W. Finnegan, 2009: Modification of cirrus clouds to reduce global warming. Environmental Research Letters, 4, 045102, https://doi.org/10.1088/1748-9326/4/4/045102.
    Platnick, S., M. D. King, S. A. Ackerman, W. P. Menzel, B. A. Baum, J. C. Riedi, and R. A. Frey, 2003: The MODIS cloud products: Algorithms and examples from Terra. IEEE Trans. Geosci. Remote Sens., 41, 459−473, https://doi.org/10.1109/TGRS.2002.808301.
    Poore, K. D., J. H. Wang, and W. B. Rossow, 1995: Cloud layer thicknesses from a combination of surface and upper-air observations. J. Climate, 8, 550−568, https://doi.org/10.1175/1520-0442(1995)008<0550:CLTFAC>2.0.CO;2.
    Ramanathan, V., R. D. Cess, E. F. Harrison, P. Minnis, B. R. Barkstrom, E. Ahmad, and D. Hartmann, 1989: Cloud-radiative forcing and climate: Results from the Earth radiation budget experiment. Science, 243, 57−63, https://doi.org/10.1126/science.243.4887.57.
    Remer, L. A., and Coauthors, 2005: The MODIS aerosol algorithm, products, and validation. J. Atmos. Sci., 62, 947−973, https://doi.org/10.1175/JAS3385.1.
    Salomonson, V. V., W. L. Barnes, P. W. Maymon, H. E. Montgomery, and H. Ostrow, 1989: MODIS: Advanced facility instrument for studies of the Earth as a system. IEEE Trans. Geosci. Remote Sens., 27, 145−153, https://doi.org/10.1109/36.20292.
    Schaaf, C. B., and Coauthors, 2002: First operational BRDF, albedo nadir reflectance products from MODIS. Remote Sens. Environ., 83, 135−148, https://doi.org/10.1016/S0034-4257(02)00091-3.
    Sharma, S., R. Vaishnav, M. V. Shukla, P. Kumar, P. Kumar, P. K. Thapliyal, S. Lal, and Y. B. Acharya, 2016: Evaluation of cloud base height measurements from Ceilometer CL31 and MODIS satellite over Ahmedabad, India. Atmospheric Measurement Techniques, 9, 711−719, https://doi.org/10.5194/amt-9-711-2016.
    Shupe, M. D., and J. M. Intrieri, 2004: Cloud radiative forcing of the Arctic surface: The influence of cloud properties, surface albedo, and solar zenith angle. Journal of climate, 17, 616−628, https://doi.org/10.1175/1520-0442(2004)017<0616:CRFOTA>2.0.CO;2.
    Slingo, A., and J. M. Slingo, 1988: The response of a general circulation model to cloud longwave radiative forcing. I: Introduction and initial experiments. Quart. J. Roy. Meteor. Soc., 114, 1027−1062, https://doi.org/10.1002/qj.49711448209.
    Stephens, G. L., and Coauthors, 2012: An update on Earth's energy balance in light of the latest global observations. Nature Geoscience, 5, 691−696, https://doi.org/10.1038/NGEO1580.
    Thampi, B. V., and R. Roca, 2014: Investigation of negative cloud radiative forcing over the Indian subcontinent and adjacent oceans during the summer monsoon season. Atmospheric Chemistry and Physics, 14, 6739−6758, https://doi.org/10.5194/acp-14-6739-2014.
    Trenberth, K. E., J. T. Fasullo, and M. A. Balmaseda, 2014: Earth's energy imbalance. J. Climate, 27, 3129−3144, https://doi.org/10.1175/JCLI-D-13-00294.1.
    Verlinden, K. L., and S. P. de Szoek, 2018: Simulating radiative fluxes through southeastern pacific stratocumulus clouds during VOCALS-REx. J. Atmos. Oceanic Technol., 35, 821−836, https://doi.org/10.1175/JTECH-D-17-0169.1.
    Viúdez-Mora, A., M. Costa-Surós, J. Calbó, and J. A. González, 2015: Modeling atmospheric longwave radiation at the surface during overcast skies: The role of cloud base height. J. Geophys. Res., 120, 199−214, https://doi.org/10.1002/2014JD022310.
    Wang, B., and L. Ho, 2002: Rainy season of the Asian-Pacific summer monsoon. Journal of Climate, 15(4), 386−398.
    Wang, F., Y. H. Ding, and Y. Xu, 2007: Cloud and radiation processes simulated by a coupled atmosphere-ocean model. Acta Meteorologica Sinica, 21, 397−408.
    Wang, F., and Coauthors, 2015: Multi-sensor quantification of aerosol-induced variability in warm clouds over eastern China. Atmos. Environ., 113, 1−9, https://doi.org/10.1016/j.atmosenv.2015.04.063.
    Wang, W.-C., W. Gong, W.-S. Kau, C.-T. Chen, H.-H. Hsu, and C.-H. Tu, 2004: Characteristics of cloud radiation forcing over east China. J. Climate, 17, 845−853, https://doi.org/10.1175/1520-0442(2004)017<0845:COCRFO>2.0.CO;2.
    Wang, X. J., and J. R. Key, 2003: Recent trends in Arctic surface, cloud, and radiation properties from space. Science, 299(5613), 1725−1728, https://doi.org/10.1126/science.1078065.
    Wielicki, B. A., B. R. Barkstrom, E. F. Harrison, R. B. Lee III, G. L. Smith, and J. E. Cooper, 1996: Clouds and the Earth’s radiant energy system (CERES): An Earth observing system experiment. Bull. Amer. Meteor. Soc., 77, 853−868, https://doi.org/10.1175/1520-0477(1996)077<0853:CATERE>2.0.CO;2.
    Winker, D. M., M. A. Vaughan, A. Omar, Y. X. Hu, K. A. Powell, Z. Y. Liu, W. H. Hunt, and S. A. Young, 2009: Overview of the CALIPSO mission and CALIOP data processing algorithms. J. Atmos. Oceanic Technol., 26, 2310−2323, https://doi.org/10.1175/2009JTECHA1281.1.
    Xu, Z. Q., H. S. Chen, J. P. Guo, and W. C. Zhang, 2021: Contrasting effect of soil moisture on the daytime boundary layer under different thermodynamic conditions in summer over China. Geophys. Res. Lett., 48, e2020GL090989, https://doi.org/10.1029/2020GL090989.
    Yin, J. F., D. H. Wang, H. B. Xu, and G. Q. Zhai, 2015: An investigation into the three-dimensional cloud structure over East Asia from the CALIPSO-GOCCP Data. Science China Earth Sciences, 58, 2236−2248, https://doi.org/10.1007/s11430-015-5205-4.
    Yu, R. C., Y. Q. Yu, and M. H. Zhang, 2001: Comparing cloud radiative properties between the eastern China and the Indian monsoon region. Adv. Atmos. Sci., 18, 1090−1102, https://doi.org/10.1007/s00376-001-0025-1.
    Zhang, J. Q., H. B. Chen, Z. Q. Li, X. H. Fan, L. Peng, Y. Yu, and M. Cribb, 2010: Analysis of cloud layer structure in Shouxian, China using RS92 radiosonde aided by 95 GHz cloud radar. J. Geophys. Res., 115, D00K30, https://doi.org/10.1029/2010JD014030.
    Zhang, L., X. Q. Dong, A. Kennedy, B. K. Xi, and Z. Q. Li, 2017: Evaluation of NASA GISS post-CMIP5 single column model simulated clouds and precipitation using ARM Southern Great Plains observations. Adv. Atmos. Sci., 34, 306−320, https://doi.org/10.1007/s00376-016-5254-4.
    Zhang, Y., and Coauthors, 2018: Climatology of cloud-base height from long-term radiosonde measurements in China. Adv. Atmos. Sci., 35, 158−168, https://doi.org/10.1007/s00376-017-7096-0.
    Zhang, Y., Q. Zhou, S. S. Lv, S. Z. Jia, F. Tao, D. D. Chen, and J. P. Guo, 2019: Elucidating cloud vertical structures based on three-year Ka-band cloud radar observations from Beijing, China. Atmospheric Research, 222, 88−99, https://doi.org/10.1016/j.atmosres.2019.02.007.
    Zhou, C., M. D. Zelinka, and S. A. Klein, 2016: Impact of decadal cloud variations on the Earth’s energy budget. Nature Geoscience, 9, 871−874, https://doi.org/10.1038/ngeo2828.
    Zhou, Q., and Coauthors, 2019: Cloud-base and cloud-top heights determined from a ground-based cloud radar in Beijing, China. Atmos. Environ., 201, 381−390, https://doi.org/10.1016/j.atmosenv.2019.01.012.
  • [1] Yong ZHANG, Lejian ZHANG, Jianping GUO, Jinming FENG, Lijuan CAO, Yang WANG, Qing ZHOU, Liangxu LI, Bai LI, Hui XU, Lin LIU, Ning AN, Huan LIU, 2018: Climatology of Cloud-base Height from Long-term Radiosonde Measurements in China, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 158-168.  doi: 10.1007/s00376-017-7096-0
    [2] TANG Yanbing, GAN Jingjing, ZHAO Lu, GAO Kun, 2006: On the Climatology of Persistent Heavy Rainfall Events in China, ADVANCES IN ATMOSPHERIC SCIENCES, 23, 678-692.  doi: 10.1007/s00376-006-0678-x
    [3] GUO Xueliang, FU Danhong, LI Xingyu, HU Zhaoxia, LEI Henchi, XIAO Hui, HONG Yanchao, 2015: Advances in Cloud Physics and Weather Modification in China, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 230-249.  doi: 10.1007/s00376-014-0006-9
    [4] Dabang JIANG, Dan HU, Zhiping TIAN, Xianmei LANG, 2020: Differences between CMIP6 and CMIP5 Models in Simulating Climate over China and the East Asian Monsoon, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 1102-1118.  doi: 10.1007/s00376-020-2034-y
    [5] Jinqiang ZHANG, Hongbin CHEN, Xiang'ao XIA, Wei-Chyung WANG, 2016: Dynamic and Thermodynamic Features of Low and Middle Clouds Derived from Atmospheric Radiation Measurement Program Mobile Facility Radiosonde Data at Shouxian, China, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 21-33.  doi: 10.1007/s00376-015-5032-8
    [6] Xiaoyan WU, Jinyuan XIN, Wenyu ZHANG, Chongshui GONG, Yining MA, Yongjing MA, Tianxue WEN, Zirui LIU, Shili TIAN, Yuesi WANG, Fangkun WU, 2020: Optical, Radiative and Chemical Characteristics of Aerosol in Changsha City, Central China, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 1310-1322.  doi: 10.1007/s00376-020-0076-9
    [7] Boru MAI, Xuejiao DENG, Zhanqing LI, Jianjun LIU, Xiang'ao XIA, Huizheng CHE, Xia LIU, Fei LI, Yu ZOU, Maureen CRIBB, 2018: Aerosol Optical Properties and Radiative Impacts in the Pearl River Delta Region of China during the Dry Season, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 195-208.  doi: 10.1007/s00376-017-7092-4
    [8] Su-Bin OH, Yeon-Hee KIM, Ki-Hoon KIM, Chun-Ho CHO, Eunha LIM, 2016: Verification and Correction of Cloud Base and Top Height Retrievals from Ka-band Cloud Radar in Boseong, Korea, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 73-84.  doi: 10.1007/s00376-015-5058-y
    [9] JIANG Dabang, YU Ge, ZHAO Ping, CHEN Xing, LIU Jian, LIU Xiaodong, WANG Shaowu, ZHANG Zhongshi, YU Yongqiang, LI Yuefeng, JIN Liya, XU Ying, JU Lixia, ZHOU Tianjun, YAN Xiaodong, 2015: Paleoclimate Modeling in China: A Review, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 250-275.  doi: 10.1007/s00376-014-0002-0
    [10] YANG Shili, FENG Jinming, DONG Wenjie, CHOU Jieming, 2014: Analyses of Extreme Climate Events over China Based on CMIP5 Historical and Future Simulations, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 1209-1220.  doi: 10.1007/s00376-014-3119-2
    [11] REN Guoyu, DING Yihui, ZHAO Zongci, ZHENG Jingyun, WU Tongwen, TANG Guoli, XU Ying, 2012: Recent Progress in Studies of Climate Change in China, ADVANCES IN ATMOSPHERIC SCIENCES, 29, 958-977.  doi: 10.1007/s00376-012-1200-2
    [12] CHE Huizheng, ZHANG Xiaoye, Stephane ALFRARO, Bernadette CHATENET, Laurent GOMES, ZHAO Jianqi, 2009: Aerosol Optical Properties and Its Radiative Forcing over Yulin, China in 2001 and 2002, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 564-576.  doi: 10.1007/s00376-009-0564-4
    [13] Rui YANG, Lingkun RAN, Yuli ZHANG, Yi LIU, 2019: Analysis and Simulation of the Stratospheric Quasi-zero Wind Layer over Korla, Xinjiang Province, China, ADVANCES IN ATMOSPHERIC SCIENCES, 36, 1143-1155.  doi: 10.1007/s00376-019-9045-6
    [14] Ho Nam CHEUNG, ZHOU Wen, Hing Yim MOK, Man Chi WU, Yaping SHAO, 2013: Revisiting the Climatology of Atmospheric Blocking in the Northern Hemisphere, ADVANCES IN ATMOSPHERIC SCIENCES, 30, 397-410.  doi: 10.1007/s00376-012-2006-y
    [15] K.-M. Lau, Song Yang, 1997: Climatology and Interannual Variability of the Southeast Asian Summer Monsoon, ADVANCES IN ATMOSPHERIC SCIENCES, 14, 141-162.  doi: 10.1007/s00376-997-0016-y
    [16] Yufang TIAN, Daren LÜ, 2017: Comparison of Beijing MST Radar and Radiosonde Horizontal Wind Measurements, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 39-53.  doi: 10.1007/s00376-016-6129-4
    [17] LIAO Hong, CHANG Wenyuan, YANG Yang, 2015: Climatic Effects of Air Pollutants over China: A Review, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 115-139.  doi: 10.1007/s00376-014-0013-x
    [18] SUN Zhian, WANG Xiaoyun, ZENG Xianning, 2006: Radiative Forcing of SO2 and NOx: A Case Study in Beijing, ADVANCES IN ATMOSPHERIC SCIENCES, 23, 317-322.  doi: 10.1007/s00376-006-0317-6
    [19] WANG Zhili, ZHANG Hua, SHEN Xueshun, 2011: Radiative Forcing and Climate Response Due to Black Carbon in Snow and Ice, ADVANCES IN ATMOSPHERIC SCIENCES, 28, 1336-1344.  doi: 10.1007/s00376-011-0117-5
    [20] MA Jianzhong, GUO Xueliang, ZHAO Chunsheng, ZHANG Yijun, HU Zhijin, 2007: Recent Progress in Cloud Physics Research in China, ADVANCES IN ATMOSPHERIC SCIENCES, 24, 1121-1137.  doi: 10.1007/s00376-007-1121-7

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 16 December 2020
Manuscript revised: 15 February 2021
Manuscript accepted: 09 March 2021
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 Significant Role of Radiosonde-measured Cloud-base Height in the Estimation of Cloud Radiative Forcing

    Corresponding author: Jianping GUO, jpguocams@gmail.com
  • 1. State Key Laboratory of Severe Weather, Chinese Academy of Meteorological Sciences, Beijing 100081, China
  • 2. College of Earth Sciences, University of Chinese Academy of Sciences, Beijing 100049, China
  • 3. Meteorological Observation Center, China Meteorological Administration, Beijing 100081, China

Abstract: The satellite-based quantification of cloud radiative forcing remains poorly understood, due largely to the limitation or uncertainties in characterizing cloud-base height (CBH). Here, we use the CBH data from radiosonde measurements over China in combination with the collocated cloud-top height (CTH) and cloud properties from MODIS/Aqua to quantify the impact of CBH on shortwave cloud radiative forcing (SWCRF). The climatological mean SWCRF at the surface (SWCRFSUR), at the top of the atmosphere (SWCRFTOA), and in the atmosphere (SWCRFATM) are estimated to be −97.14, −84.35, and 12.79 W m−2, respectively for the summers spanning 2010 to 2018 over China. To illustrate the role of the cloud base, we assume four scenarios according to vertical profile patterns of cloud optical depth (COD). Using the CTH and cloud properties from MODIS alone results in large uncertainties for the estimation of SWCRFATM, compared with those under scenarios that consider the CBH. Furthermore, the biases of the CERES estimation of SWCRFATM tend to increase in the presence of thick clouds with low CBH. Additionally, the discrepancy of SWCRFATM relative to that calculated without consideration of CBH varies according to the vertical profile of COD. When a uniform COD vertical profile is assumed, the largest SWCRF discrepancies occur during the early morning or late afternoon. By comparison, the two-point COD vertical distribution assumption has the largest uncertainties occurring at noon when the solar irradiation peaks. These findings justify the urgent need to consider the cloud vertical structures when calculating the SWCRF which is otherwise neglected.

摘要: 现有基于卫星观测的云辐射强迫定量估算精度仍十分有限,这在很大程度上归因于云底高度(CBH)表征的局限性或不确定性。本文利用中国区域L波段秒级探空数据计算得到的CBH数据,及MODIS / Aqua的云特性和云顶高度(CTH)产品数据,量化了CBH对短波云辐射强迫(SWCRF)的影响。结果表明:中国2010年至2018年夏季云的短波辐射强迫在地表(SWCRFSUR)、大气顶部(SWCRFTOA)和大气中(SWCRFATM)的平均值分别为–97.14、–84.35和12.79 W m−2。为阐明云底作用,我们在利用辐射传输模式估算云短波辐射强迫时,根据云光学厚度(COD)垂直分布廓线做了四种情景假设。与考虑CBH的方案相比,仅使用MODIS CTH和云特性的方案会导致估计的SWCRFATM产生较大的不确定性。除此之外,在云底高度较低且云层较厚的情况下,云与地球辐射能量系统(CERES)的SWCRFATM偏差呈现增加的趋势。另外,相对于不考虑CBH的情景,估算得到的SWCRFATM偏差会随COD的垂直廓线而变化。当假定COD在垂直方向呈均匀分布时,最大的SWCRF差异出现在清晨或午后。相比之下,COD在垂直方向呈两点分布的假设具有最大的不确定性,该不确定性发生在一天中太阳辐射达到峰值的正午。鉴于云底高度在云短波辐射强迫估算中的重要作用,因此在计算SWCRF时须考虑云的垂直结构特征。

    • Clouds cover approximately two-thirds of Earth's surface area and play a significant role in both the Earth’s radiative balance and hydrological cycle (e.g., Ramanathan et al., 1989; Stephens et al., 2012; Chen et al., 2016; Zhou et al., 2016; Min et al., 2020). It is well established that clouds exert a cooling effect by reflecting incoming shortwave radiation back to space, and also a warming effect by trapping outgoing long-wave radiation (Slingo and Slingo, 1988; Wang and Key, 2003; Min and Zhang, 2014). The net magnitude of this cloud-induced perturbation in the radiative balance at the surface (SUR), at the top of the atmosphere (TOA), and in the atmosphere (ATM) is typically measured in terms of cloud radiative forcing (CRF, Shupe and Intrieri, 2004). The CRF in the ATM, which represents the energy absorbed in the cloud layer, is shown to significantly affect the atmospheric circulation and the water cycle, thereby altering global weather and climate (Wang et al., 2007; Henderson et al., 2013; Lü et al., 2015). However, large uncertainties exist regarding the estimation of the CRF (especially the CRF in ATM) in current climate and weather models.

      The magnitudes of the CRF are highly dependent on the cloud vertical structures, including cloud-top height (CTH), cloud-base height (CBH), cloud thickness (CT), and the vertical overlapping pattern of cloud layers (Ramanathan et al., 1989; Mitchell and Finnegan, 2009; Li et al., 2019). The impact of cloud vertical structures on the CRF is often investigated through the role of the CTH in the estimation of the CRF at the TOA (Loeb et al., 2012; Chen and Fu, 2018). However, recent studies suggest a strong dependence of the CRF on the CBH, which considerably affects the land-atmosphere energy exchanges. For example, Kato et al. (2011) indicated that the underestimation of the CBH results in an increase of the global annual mean of instantaneous surface downward long-wave irradiance. Viúdez-Mora et al. (2015) quantified the average CRF at the surface using the CBH determined from ceilometer measurements and demonstrated that the CRF decreased with CBH at a rate of 4−5 W m−2 km−1 with the average CRF observed to be in the range of 50−80 W m−2. George et al. (2018) investigated the relationships of the CBH with the CRF at the surface and TOA using soundings and found that the shortwave CRF depended on the total depth of cloud layers, whereas the longwave CRF depended on the height of cloud layers. However, there are few previous studies on the CBH impact on the estimates of the CRF in the atmosphere, and thus the role of the CBH upon the CRF estimates remains largely unknown, particularly in China.

      One of the major obstacles in quantifying the impact of the CBH on the CRF lies in the difficulties in determining the CBH on a large scale. On the one hand, ground-based instruments such as lidars, ceilometers, and millimeter wavelength cloud radars (e.g., Martucci et al., 2010; Costa-Surós et al., 2013), gain competitive edges in determining CBHs for their relatively higher accuracy and stronger temporally continuous sampling capabilities (e.g., Sharma et al., 2016; Zhou et al., 2019). However, these instruments are generally deployed at very few locations (Dai et al., 2006; Zhang et al., 2019), thereby having poor spatial representation. On the other hand, the CBH cannot be directly retrieved from passive satellite observations such as the moderate resolution imaging spectroradiometer (MODIS) onboard both Terra and Aqua. Active sensors such as the Cloud-Aerosol Lidar with Orthogonal Polarization (CALIOP) onboard the CALIPSO and the millimeter-wavelength cloud profiling radar (CPR) onboard the CloudSat can resolve the vertical distribution of cloud layers, including the CBH and CTH (Winker et al., 2009; Wang et al., 2015). However, in terms of the accuracy of the CBH detected from the aforementioned active sensors, large uncertainties exist due to the strong attenuation of radar or laser beams, especially in the presence of deep convective clouds (e.g., Zhang et al., 2017; Zhou et al., 2019).

      The operational radiosonde networks have higher detection accuracy of cloud layers, making it possible to determine CBH over a large spatial domain, albeit with poor temporal coverage (Poore et al., 1995; Zhang et al., 2010). Therefore, the present study aims to investigate the impact of the CBH on the estimated shortwave CRF (SWCRF) by using a high-resolution dataset from a radiosonde network in China that spanned nine years (2010−2018). As a first step, the climatology of SWCRF at the surface (SWCRFSUR), at the top of the atmosphere (SWCRFTOA), and in the atmosphere (SWCRFATM) is compiled, by combining the CBH from the radiosonde measurements and the CTH and cloud properties from MODIS for the summers spanning 2010 to 2018. The accuracy of the SWCRF climatology is established through comparisons with the Clouds and the Earth's Radiant Energy System (CERES) product. Then, the SWCRFs calculated under four scenarios which vary the vertical profile of cloud optical depth (COD) are compared to deduce the role of the CBH on the estimation of the SWCRF. In addition, we will test the sensitivity of both the CBH and solar zenith angle on the SWCRF.

    2.   Data and Methods
    • Radiosonde data provide in situ measurements of atmospheric environmental variables, including temperature, pressure, RH, wind speed, and wind direction, which is recognized to be able to derive the overlapping pattern of cloud in the vertical direction. The China radiosonde network (CRN) consists of 120 radiosonde sites across China, in which the high-resolution (5−8 m) sounding balloon is launched twice per day at 0800 LST (LST = UTC + 8) and 2000 LST (Guo et al., 2016, 2019). Based on these fine-resolution radiosonde measurements, the CBHs are calculated for the summer (June−August) during the period from 2010 to 2018 using the improved relative humidity (RH) threshold method (Zhang et al., 2018). Coincidently, up to two additional soundings are launched occasionally at 1400 LST and 0200 LST in summer during certain intensive observing periods at selected stations (Guo et al., 2016; Lou et al., 2019). Given the spatial inhomogeneity, the CBHs from all sites are interpolated onto a regular 5° × 5° grid, following the method by Zhang et al. (2018). The significant diurnal variation of summertime clouds in China (Chen et al., 2018) justifies the use of the 1400 LST soundings to derive CBHs unless otherwise noted. This is expected to better match with the CTH products from MODIS onboard Aqua, one of the polar-orbiting satellites.

    • The MODIS instrument, an important payload of both Terra and Aqua satellites, can observe the Earth with a total swath width of 2300 km, with a revisit cycle of one to two days, depending on the location (Remer et al., 2005). The MODIS has 36 spectral bands ranging from 0.4 to 14.2 μm (Salomonson et al., 1989), providing global coverage of cloud optical and physical parameters at both granule and grid levels (Platnick et al., 2003). As one of important cloud optical properties, cloud optical depth (τ) can affects the shortwave CRF by reflecting the attenuation of radiation intensity by cloud in the transmission path (Chen et al., 2019b). As the global gridded product, the level-3 cloud products of MODIS are made publicly available on a grid size of 1° × 1° at daily, 8-day, and monthly timescales (Menzel et al., 2008).

      Therefore, the MODIS level-3 daily cloud products onboard Aqua (MYD08_D3) are used in this study to characterize the optical and physical properties of cloud tops. In order to be collocated with the CBHs derived from the 1400 LST radiosonde data, the MODIS cloud products, including the CTH (ht), the daily averaged liquid and ice COD (τc), liquid and ice cloud effective radius (rc), as well as cloud fraction (ƒc), are resampled to 5° latitude× 5°longitude grid by simply averaging out the original MODIS products.

    • The MODIS albedo product (MCD43) could provide both white-sky albedos and black-sky albedos for spectral and broad bands. The MCD43 is a daily product, obtained by the surface reflectance observations built up over a 16-day period (Schaaf et al., 2002). As a climate modeling grid product, MCD43C3 provides albedo d ata at a spatial resolution of 0.05° latitude× 0.05° longitude grid. In the present study, a climatology of daily-averaged surface albedo (α) data was built using nine years (2010−18) of MCD43C3 black-sky visible albedo, in which only those α values with a quality flag less than or equal to two were used.

    • The Clouds and the Earth's Radiant Energy System (CERES) instrument directly measures global radiance at the TOA. By combining the MODIS and GEOS-derived cloud properties, CERES could also provide the radiance at the surface (Wielicki et al., 1996). The high accuracy and robustness of the CERES radiance products in revealing Earth’s energy imbalance have been adequately verified (Loeb et al., 2012; Allan et al., 2014; Trenberth et al., 2014). As such, the CERES Synoptic TOA and surface fluxes and clouds (SYN) products under the clear-sky and all-sky conditions with a spatial resolution of 1° × 1° are used here as “ground-truth” to validate the SWCRF estimated by the radiative transfer model. Prior to its application in the following analysis, we first calculate the CERES SWCRFTOA and SWCRFSUR using the clear-sky and all-sky incoming and outgoing SW flux at the TOA and surface, respectively. Then, the CERES-derived SWCRF are resampled from a 1° × 1° to 5° × 5° latitude/longitude grid.

    • The shortwave rapid radiative transfer model (RRTM_SW) has been used to calculate both atmospheric radiative fluxes and heating rates for the spectral wavenumber of 820−50 000 cm−1 (Clough et al., 2005). By utilizing the correlated-k approach, the fluxes provided by RRTM SW are accurate, computationally efficient, and fast. Due to its high accuracy and efficiency, the RRTM_SW has been widely applied to those studies associated with atmospheric radiative transfer and general circulation models (GCMs) (e.g., Thampi and Roca, 2014; Verlinden and de Szoek, 2018).

    • Here, the SWCRF (ΔF; W m−2) is defined as the difference of net solar flux between an environment with clouds and one without clouds present at the surface, at the TOA, or in the atmosphere, which can be expressed as:

      where i represents the surface, TOA, and the ATM, respectively, $ {F_{{\rm{net}},i}} $ represents the net flux with clouds, and $ {F'_{{\rm{net}},i}}$ represents the net flux without clouds.

      $ {F_{{\rm{net}},i}} $ can be calculated as the difference between the downwelling flux with cloud ($ {F}_{\downarrow,i} $) and upwelling flux with cloud ($ {F}_{\uparrow,i} $), which is formulated as:

      Similarly, $ {F'_{{\rm{net}},i}} $ denotes the difference between the downwelling flux without cloud ($ {F'_{ \downarrow,i}} $) and the upwelling flux without cloud ($ {F'_{ \uparrow,{\rm{i}}}} $), which can be formulated as follows:

      Combining Eqs. (1)−(3), we obtain the following expression:

      Since the downward flux at the top of the atmosphere is independent of the presence of clouds (i.e.,$ {F_{ \downarrow,{\rm{TOA}}}} = {F'_{ \downarrow,{\rm{TOA}}}} $), the SWCRF at TOA (ΔFTOA) can be derived as:

      Meanwhile, the SWCRF in the atmosphere (ΔFATM) is defined as:

    • Normally, the vertical distribution of COD is assumed to be uniform in the current climate model. Here, we define four scenarios regarding the vertical distribution of COD, including a two-point distribution (Scenario 1), a uniform distribution (Scenario 0 & Scenario 2), and a normal distribution (Scenario 3), which are described as follows:

      (1)Two-point distribution

      The two-point distribution assumes the total COD is concentrated in only two cloud layers. Thus, the COD for each cloud layer is 1/2 of the total COD. In addition, we define the two heights of the cloud layer to be 1/4 and 3/4 of the total cloud height. The distribution can be formulated as follows:

      where the ${\tau _{\rm{t}}}$ and ${\tau _i}$ stands for the total COD and the COD of the ith cloud layer, respectively. ht and hb represent the cloud-top height and cloud-base height, respectively, and hi stands for the height of the ith cloud layer.

      (2)Uniform distribution

      The uniform distribution assumes that the total COD is uniformly distributed across the cloud layers in the vertical, and the COD of the ith cloud layer is expressed as follows:

      where the hi,t, and hi,b stand for the top and base height for the ith cloud layer, respectively, and N represents the number of the cloud layer.

      (3)Normal distribution

      The symmetrical distribution of many natural processes and phenomena is the normal distribution. In this study, we assume the vertical distribution of COD is a normal distribution. For the given interval [μ ± 3σ], the probability of the actual COD falling within that interval is nearly 1.0 (Mishra and Datta-Gupta, 2018). Thus, the COD of the ith cloud layer can be described as follows:

    • To calculate the upward flux reaching the top-of-atmosphere ($ {F'_{ \uparrow,{\rm{TOA}}}} $ and $ {F_{ \uparrow,{\rm{TOA}}}} $) and the downward flux reaching the surface ($ {F'_{ \downarrow,{\rm{SUR}}}} $ and $ {F_{ \downarrow,{\rm{SUR}}}} $), we employ the RRTM_SW model. A tropical standard atmosphere model is chosen for the model set up, with the vertical atmospheric profiles of pressure and temperature set to their default values. The model differentiates between ice and liquid clouds. As key input parameters of the RRTM_SW model, MODIS-derived cloud physical parameters, such as cloud fraction, liquid droplet effective radius, and cloud effective radius of ice, have been directly incorporated into the model simulation. The cloud water path is calculated using the COD from MODIS. The vertical profiles of the COD are calculated by the CBH, CTH, and the total COD, using the method described in section 2.2.2. The profiles of cloud fraction and cloud effective radius are assumed to be invariant with the height h and set equal to fc and rc, respectively. For the surface albedo, the climatological mean values derived from MODIS have been utilized. It is noteworthy that aerosol properties are not included in the input file for the RRTM_SW model since the radiative effect induced by aerosol-cloud interaction is more complicated and beyond the scope of this study.

      Besides the CBH, it is well known that the SWCRF is influenced by the vertical profile of the COD. In order to investigate the impact of CBH on the SWCRF, we assume two main scenarios: (1) the SWCRF is calculated from only the CTH with the CBH unknown (Scenario 0), and (2) the SWCRF is calculated from both the CTH and CBH (Scenarios 1−3), the details of which are summarized in Table 1. Schematically shown in Fig. 1 is an example of the vertical profile of the COD as a function of the normalized total COD.

      ScenariosVertical profile pattern of CODNumber of cloud layers (N)
      Scenario 0Uniform1
      Scenario 1two-point2
      Scenario 2Uniformtrue cloud layer
      Scenario 3Normaltrue cloud layer

      Table 1.  Summary for the vertical profile scenarios of cloud optical depth (COD) assumed in the calculation of cloud optical depth, which is an important input for the calculation of cloud radiative forcing.

      Figure 1.  The normalized cloud optical depth (COD; ${\tau }_{{\rm{c}}}$) shown as a function of normalized cloud height (km) for three different vertical profile Scenarios of COD: (a) Scenario 1, (b) Scenario 2, and (c) Scenario 3, based on the assumption of the true number of the cloud layer is equal to 10 and the height of cloud is uniform.

    • We perform a sensitivity analysis to better understand the impact of CBH on the SWCRF. This consisted of a series of experiments on the CRF estimation, that have been conducted by running the RRTM_SW with varying CBH values, under different cloud vertical profile patterns of COD. In an ideal experiment, the liquid effective particle radius is set to be uniform at 15 μm, whereas the solar zenith angles (SZA) are set to 0°, 30°, and 60°, respectively. The CBH values are set to increase from 0.5 km upwards until 5.0 km at 0.5 km intervals, and the increase in CBH each time is defined as Δhb. The total COD is constant, and the vertical profile of COD is calculated using the cloud-top height (ht) and the changed cloud-base height (hb + Δhb).

    3.   Results and discussion
    • Figure 2 presents the geographic distributions of summertime cloud properties in China within 5° latitude × 5° longitude grid boxes during the period from 2010 to 2018, including the CTH derived from the MODIS/Aqua, the CBH derived from the radiosonde measurements, and the cloud thickness (CT). The mean CTH, CBH, and CT correspond to 7.23 km, 3.56 km, and 4.71 km above ground level (AGL), respectively. Overall, the CTH exhibits a noticeable geographical variability in summer across China. For instance, the cloud top tends to develop much higher over the Tibetan Plateau (TP) and South China Sea (SCS), where the South Asian Monsoon and mesoscale convective systems most frequently occur (Wang and Ho, 2002; Chen et al., 2019a). In the latitude zone between 30°N and 45°N, western China has higher CTHs than eastern China, which is likely due to the terrain difference between eastern and western China. By comparison, the summertime CBHs exhibit a “northwest-high-southeast-low” pattern: CBHs as high as 4.0 km AGL or higher are found in Northwest China versus much smaller CBHs (less than 3.0 km AGL) in Southeast China. The lowest CBH tends to occur in the coastal region of East China, and the northern part of the South China Sea, which is lower than 2.0 km AGL. This could be attributed to a lower lifting condensation level due to higher dew points near the coast. Another pronounced signal is evidenced by the decreasing trend of CT with increasing latitude, indicating that the CT spatial pattern exhibits a strong south-to-north gradient, with greater CT in southern China.

      Figure 2.  Spatial distributions of (a) cloud-top height (CTH) from MODIS/Aqua, (b) cloud-base height (CBH) from radiosonde measurements, and (c) cloud thickness (CT) from MODIS/Aqua over China in summer (June−August) for the period from 2010 to 2018. The blue bars on the right-hand side represent the frequency of cloud height in different latitudes.

      Figure 3 shows the spatial distributions of height-resolved occurrence frequency for the CTH, CBH, and CT, respectively. The probability density function (PDF) of CTH peaks at around 6−8 km AGL and features a heavy tail. Specifically, most of the CTHs (~80%) are mainly located above 5 km, indicating the dominance of middle- and high-level clouds in China. By comparison, the tail shifts towards low CBHs, and the CBHs falling within 0−3 km are responsible for approximately 65% of all cases. Interestingly, the PDF of cloud thickness exhibits a roughly normal distribution, with more than half of the CTs falling within 2−5 km. Overall, the frequency distributions of CTH and CBH revealed here are broadly consistent with the previous remote sensing studies (e.g., Yin et al., 2015; Chen et al., 2018). Note that the PDFs of CTH, CBH, and CT given in 0.1 km vertical range bins are also shown in Fig. 3. The peak in the PDF of CBH is generally located at 1−2 km, and then gradually decreases with height above 2 km, indicating that the bases of cloud layers are lower than 2 km AGL under most circumstances.

      Figure 3.  The frequency distribution of (a) CTH (red bars), (b) CBH (blue bars), and (c) cloud thickness (green bars) over China in summer for the period from 2010 to 2018. The data in the histogram are pooled into 1.0 km bins. The red, blue, and green curves depict the kernel density estimate (KDE) of the probability distribution function (PDF) for CTH, CBH, and cloud thickness, respectively.

      Figure 4 illustrates the percentile analysis for CBH, CTH, and CT as a function of latitude. It is clear that CBHs tend to increase with latitude, reaching the peaks in the latitude range between 37.5°N and 42.5°N. Coincidently, the planetary boundary layer height (PBLH) exhibits an interesting spatial pattern of “northwest high southeast low” in China (i.e., Zhang et al., 2018; Guo et al., 2019; Xu et al., 2021). Therefore, the dependence of CBH on latitude resembles the spatial pattern of PBLH in China, to some extent. This could be attributed to the spatial discrepancy of soil moisture (Xu et al., 2021) since the dominant factor driving the development of the boundary layer comes from sensible heat flux that is closely associated with soil moisture. In particular, as the latitude increases, the gap of CBH shifts towards larger values particularly when the CBH is below the 50th percentile, suggesting a patchier CBH distribution at high altitude regions in China (under lower cloud base circumstances). As for the linear meridional trends in CBH percentiles, the lower percentiles of CBH increase slowly, in contrast to the higher percentiles of CBH. The trends of high CBHs (e.g., 80th and 90th percentile), on average, are much larger than those of low CBHs (e.g., 50th percentile and below). Regarding the CTHs, low percentiles of CTHs do not show any pronounced latitudinal dependence, while high percentiles of CTHs tend to decrease with increasing latitudes. The average trend of high CTHs (90th percentile) is far less than twice as large as the median CTHs (50th percentile). This is also the case for CT, since the high percentiles of CTs decrease much more rapidly with the increasing latitude compared to the median percentile of CT.

      Figure 4.  Evolution, as a function of latitude and percentile, for (a) cloud-base height, (b) cloud-top height, and (c) cloud thickness over China in the summers during the period from 2010 to 2018. The 10th−90th percentiles in 10 percentile increments are shown, with the 50th and 90 th percentiles highlighted in blue and red, respectively.

    • Using the radiosonde-derived CBH and the MODIS-derived CTH as the two main input parameters, the SWCRF climatology is simulated in summer over China. In order to confirm the reliability of this climatology, the simulated SWCRF is compared with its counterpart from CERES in Fig. 5. In general, the SWCRFSUR and SWCRFTOA are negative, whereas the SWCRFATM is positive. This means that clouds tend to exhibit a cooling effect at the SUR and TOA, but exhibit a warming effect in the ATM during summer over China. A high correlation (R = 0.8) is found between the simulated SWCRFTOA and CERES observation. Similarly, the SWCRFSUR agrees well with CERES (R = 0.79). These robust relationships indicate that the simulated SWCRF climatology is consistent with the CERES measurements at the TOA and the surface.

      Figure 5.  Scatterplots of (a) SWCRFSUR, (b) SWCRFTOA, and (c) SWCRFATM using the CBH from radiosonde measurement, MODIS-observed CTH, and ancillary cloud parameters under Scenario 2 versus that from the CERES product in China. The regression lines of estimated SWCRFs against CERES-derived SWCRFs are given as well. Note that all samples are limited to the period in summer in 2013.

      However, there are large differences between the simulated and the CERES-derived SWCRFATM. We hypothesize that the probable reason for this could be the inability of the satellite observation of CERES to capture the cloud bottom. To confirm our hypothesis, we conducted a correlation analysis concerning the absolute difference between the CERES-derived and the simulated SWCRFATM (ΔSWCRFATM) versus cloud height. Figure 6a shows that the biases which exist in the CERES-derived SWCRFATM positively correlate with cloud thickness, while it negatively correlates with cloud bottom (Fig. 6b). This indicates that the biases of CERES SWCRFATM tend to increase when thick clouds with low base height are present. Also, the negative SWCRFATM from the CERES product (red points in Fig. 6) further demonstrates that large biases from satellites often occur for low and thick clouds. The reason for this could be that the satellite observation could not provide an accurate estimation of cloud bottom, especially in the presence of thick clouds with low bottoms.

      Figure 6.  The difference between the CERES-derived SWCRFATM and the simulated SWCRFATM (ΔSWCRFATM) shown as a function of (a) cloud thickness, (b) cloud-base height, and (c) cloud-top height. The red points represent those samples with negative SWCRFATM from CERES.

    • Figure 7 presents the spatial distribution of summertime SWCRFSUR, SWCRFTOA, and SWCRFATM in China during the period from 2010 to 2018. The mean values of SWCRFSUR, SWCRFTOA and SWCRFATM are −97.14 W m−2, −84.35 W m−2, and 12.79 W m−2, respectively. This result is consistent with previous studies. For example, Yu et al. (2001) estimated SWCRFTOA values as large as −85.0 W m−2 using the observations from the Earth Radiation Budget Experiment (ERBE) experiment during summertime in eastern China. Similarly, Wang et al. (2004) obtained the mean SWCRFTOA over the Yangtze River−Huaihe River valley at about −110.0, −100.0, and −70.0 W m−2 in June, July, and August, respectively. In particular, relatively small SWCRFSUR is observed in Northwest China (Fig. 7a), which may be explained by the thin cloud layer caused by the high CBH and low CTH common to this region (Fig. 2). By comparison, the SWCRFSUR in central and eastern China is high, which is attributed to the large cloud thickness in this region. This result demonstrates that the SWCRFSUR largely depends on the cloud thickness but has no relation with the cloud base. Our results are consistent with previous findings (e.g., Zhang et al., 2017; George et al., 2018), which suggests that cloud depth contributes significantly to the estimate of SWCRFSUR.

      Figure 7.  Spatial distributions of (a) SWCRFSUR, (b) SWCRFTOA, and (c) SWCRFATM under Scenario 2 in China in summer for the period from 2010 to 2018. The blue bars on the right-hand side represent their corresponding frequency of SWCRF in various latitudes.

      The spatial distribution of SWCRFTOA (Fig. 7b), to some extent, is roughly similar to that of the SWCRFSUR. This indicates that SWCRFTOA is also influenced by cloud depth but also has no relation with the cloud base. A probable reason for this could be that SWCRFTOA is highly dependent on cloud optical depth, which is positively correlated with cloud depth.

      Also noteworthy is that the spatial distribution of SWCRFATM and CBH exhibits an obvious negative correlation. As shown in Fig. 7c, the SWCRFATM is low in northwest China where high clouds always occur. In contrast, the SWCRFATM in southeast China is high, which is attributed to the low cloud base in this region. A possible reason may be that low clouds tend to increase the multiple scattering and absorption of solar radiation between the clouds and the surface, which enhances the solar radiation absorbed by the cloud, thus, resulting in a marked increase in the SWCRFATM. This result demonstrates that the height of the cloud base significantly contributes to the estimate of the SWCRFATM. By comparison, the frequencies of the SWCRFATM as a function of latitude are consistent with the height of cloud base, which increases first and then decreases as the latitude increases. This further highlights the importance of the cloud base in modulating SWCRFATM.

    • To further examine the relationship between the CBH and SWCRF, the SWCRFSUR, SWCRFTOA, and SWCRFATM are calculated by differentiating between the various vertical distribution patterns of COD, which has already been clarified in Table 1 and Fig. 1. Figure 8 shows the comparison of the CBH and SWCRF at the surface, top-of-atmosphere, and in the atmosphere. As expected, the SWCRFATM negatively correlated with cloud base, while the SWCRFSUR and SWCRFTOA have no relation to cloud base. Under Scenarios 2 and 3, the correlation of SWCRFATM with the height of cloud base is strong, with the R values as high as 0.76. However, under Scenario 0, the relationship between SWCRFATM estimates and cloud base is weak, since it does not consider the cloud base at all.

      Figure 8.  The (a) SWCRFSUR, (b) SWCRFTOA, and (c) SWCRFATM shown as a function of CBH under four vertical configuration scenarios of cloud by assuming the vertical profile of COD: Scenario 0 (black), Scenario 1 (red), Scenario 2 (green), and Scenario 3 (blue).

    • In this section, we will perform a quantitative analysis to deduce how the SWCRFATM is dependent on cloud base. The SWCRF estimates, under different vertical profile scenarios of COD, are shown in Fig. 9. Overall, the estimates of SWCRFSUR do not differ with the vertical distribution of COD (Fig. 9a). On average, the relative deviations of the SWCRFSUR with (Scenarios 1−3) and without (Scenario 0) considering cloud bottom is very small (only 0.32%), which indicates that the impact of cloud base on the SWCRFSUR can be ignored. This result is consistent with previous studies pointing out that the SWCRF at the surface is independent of the cloud bottom height (e.g., George et al., 2018).

      Figure 9.  Comparison of the nationwide mean summertime values for (a) SWCRFSUR, (b) SWCRFTOA, and (c) SWCRFATM in China during the period from 2010 to 2018 under four vertical configuration scenarios of cloud by assuming the vertical profile of COD: Scenario 0 (black), Scenario 1 (red), Scenario 2 (green), and Scenario 3 (blue).

      Similar to the SWCRFSUR, the error in the SWCRFTOA estimates considering no cloud bottom is not as obvious (Fig. 9b). However, it is inferred that the SWCRFTOA tends to be overestimated if the cloud base is not considered in the estimation of cloud forcing, especially under Scenario 1. The relative deviations of the SWCRFTOA under Scenario 1 from that under Scenario 0 is nearly 6.1%, which demonstrates that an accurate cloud base is more essential to the SWCRFTOA than to the SWCRFSUR.

      In stark contrast to the SWCRFSUR and the SWCRFTOA, the magnitudes of estimated SWCRFATM under different profile scenarios of COD differ greatly (Fig. 9c). Given the lack of consideration of an accurately described cloud bottom, the SWCRFATM is expected to be underestimated. The possible reason is that the true cloud bottom would be overestimated if the cloud base is not considered in the calculation of the SWCRFATM, which would reduce the multiple scattering and absorption of solar radiation between the surface and cloud layer, ultimately resulting in an underestimate of the SWCRFATM.

      The absolute (relative) discrepancy of the SWCRFATM between Scenario 0 and Scenarios 1−3 reaches up to 9.61 (57%), 5.43 (42%), and 4.82 (40%) W m-2, suggesting that the influence of cloud base on the SWCRFATM estimate under Scenario 1 is much larger compared to that under Scenario 2 and 3. This could be likely associated with the multiple scattering and absorption of solar radiation. With the assumption of a two-point COD vertical distribution, the multiple scattering and absorption of solar radiation become strongest, which allows more solar radiation to be absorbed by the clouds, resulting in a commensurate increase of the SWCRFATM. Overall, the SWCRFATM estimates without using cloud base as input are considerably lower than those that use the cloud base, indicating the importance of using an in situ observed cloud base for an accurate calculation of the SWCRFATM.

    • Provided that the solar zenith angle (SZA) can greatly modulate the total solar radiation that is available to interact with cloud layers, this subsection will elaborate on the sensitivity of the SWCRF to CBH and SZA in an attempt to further quantify how the SWCRF changes with the cloud base. As shown in Figs. 10a-c, the SWCRFSUR estimates are nearly unaffected by increases in the height of cloud base, irrespective of SZA and the vertical profile of COD. This indicates that the calculation of SWCRFSUR does not change appreciably with the cloud base and/or the SZA, which has been well documented in previous studies. For example, Meloni et al. (2005) reported on the modest dependency of the shortwave radiative forcing on the vertical profile of the cloud. The reasons behind this may be that the total COD and other cloud properties do not change much, and thus the capability for the cloud to scatter and absorb solar radiation is kept nearly constant.

      Figure 10.  SWCRFSUR (a−c), SWCRFTOA (d−f), and SWCRFATM (g−i) shown as a function of the increase in CBH (Δhb) under four vertical configuration scenarios of cloud by assuming the vertical profile of the COD: Scenario 0 (black), Scenario 1 (red), Scenario 2 (green), and Scenario 3 (blue).

      The sensitivity of the SWCRFTOA to CBH and the SZA exhibits similar behavior as those for the SWCRFSUR (Figs. 10d-f). The sensitivity of the SWCRFTOA to both cloud base and the SZA is weak. As the cloud base increases, the magnitude of SWCRFTOA tends to increase. This may be related to the expectation that a higher cloud base corresponds to a larger COD for the top of the cloud, which enhances the scattering of the solar radiation, and thus results in an increased SWCRF at the top of the atmosphere.

      In sharp contrast to the SWCRFSUR and SWCRFTOA, the calculated SWCRFATM is found to be sensitive to both cloud base and SZA, and the slope or sensitivity becomes much stronger (Figs. 10g-i). Interestingly, the magnitude of the SWCRFATM tends to decrease with a commensurate increase of cloud base in a monotonic scaling way.

      Also noteworthy is that the scaling rate of the SWCRFATM with cloud base differs greatly according to the scenario. The sensitivity under Scenario 1 and Scenario 3 is much larger than Scenario 2. More striking is that the SZA not only changes the slope, but also the absolute magnitude of SWCRFATM. For instance, the magnitude of SWCRFATM under Scenario 1 is found to decline from 17−23 W m−2 at SZA of 0° (Fig. 9g) to −7−2 W m−2 at SZA of 60° (Fig. 9i) as the Δhb increases from 0.5 km to 5.0 km, which indicates that the SWCRFATM decreases along with the increase of SZA. Besides, this suggests that the largest discrepancy of SWCRFATM under Scenario 1 preferentially occurs under small SZA conditions, which generally corresponds to late morning, noon, or early afternoon, which is roughly consistent with the results mentioned in section 3.5. By comparison, the largest discrepancy of the SWCRFATM under Scenario 2 preferentially occurs under large SZA conditions, which generally corresponds to early morning and late afternoon. We also notice that the increases in SZA are closely associated with a decreased SWCRFATM that is estimated with CBH as input, further illustrating the important role of cloud base.

    4.   Summary and conclusions
    • In this present study, a long-term record of the summertime SWCRF at the surface, at the top of the atmosphere, and in the atmosphere during the period from 2010 to 2018 has been generated using the CBH estimated by radiosonde measurements and the CTH and cloud properties derived from MODIS/Aqua product. To our best knowledge, this work is the first to take the in situ measured cloud base into account when calculating the SWCRF. Overall, the CBHs show significant spatial variability that is characterized by a “northwest-high-southeast-low” pattern. Compared with the CBH, the CTH has a different spatial pattern. Consequently, the SWCRFATM follows the spatial pattern of CBH, which suggests that the cloud bottom is the main factor in modulating the shortwave cloud forcing in the atmosphere.

      Overall, there are good correlations between the simulated and the CERES-derived SWCRFs. However, an obvious disagreement occurs between the simulated and CERES-derived SWCRFATM especially for thick clouds with low cloud bottoms. Also, a series of experiments have been done by assuming two kinds of vertical profiles of COD, Scenario 0 and Scenarios 1−3. The former assumption features only CTH and cloud properties from MODIS as input, and the latter combines both CTH and cloud properties and CBH.

      Comparison analyses further suggest that a large discrepancy exists in SWCRFATM between Scenario 0 and Scenarios 1−3. Meanwhile, Scenario 1 produces the largest uncertainties of the SWCRFATM. By comparison, no dramatic difference is observed for the estimate of the SWCRFSUR and SWCRFTOA, under the different scenarios concerning the vertical profiles of the COD. Given the nature of the COD assumed as a two-point distribution under Scenario 1, it is inferred that adequate consideration of the cloud base, would lead to much more realistic estimates of cloud forcing in the atmosphere. Furthermore, the sensitivity analyses also corroborate the findings mentioned above, Scenario 1 produces much more realistic cloud forcing compared with those without consideration of cloud layers in the vertical. As for Scenario 2, the influence of the cloud base on the cloud forcing is largest under large SZA conditions. Therefore, we argue that the impact of CBH depends upon an accurate estimate of the SWCRFATM, especially under Scenario 1, at the time period from late morning through early afternoon when the SZA is relatively small, or under Scenario 2 for the early morning and late afternoon.

      Nevertheless, we have to emphasize that our study makes relatively simple assumptions for the vertical profiles of the COD, which inevitably impairs our ability to accurately estimate the cloud forcings. Our understanding and accurate computation of cloud forcing will be greatly improved with the inclusion of additional instrumentation that can resolve the vertical structure of clouds, which will result in more realistic forcing caused by the changing clouds under global warming. Besides, MODIS/Aqua only provides a single snapshot of the clouds, which makes it impossible to characterize the diurnal cycles of cloud properties (e.g., the cloud fraction, COD, and cloud effective radius) and ultimately cloud radiative effect. Improvement in this regard is worthy of separate analysis and should be based on deriving cloud properties from geostationary satellites, like Himawari-8 and Fengyun-4.

      Acknowledgments. We acknowledge the support from the National Key R&D Program of China under Grants Nos. 2017YFC1501401 and 2017YFC0212803, the National Natural Science Foundation under Grant No. 41771399, and the Chinese Academy of Meteorological Sciences under Grant No. 2018Y014. Also, we are grateful to NASA for granting access to the level-3 cloud products from MODIS/Aqua. Last but not least, special thanks go to the National Meteorological Information Center, China Meteorological Administration for generously providing the radiosonde datasets used here (https://data.cma.cn/en).

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return