Advanced Search
Article Contents

An Extreme Drought over South China in 2020/21 Concurrent with an Unprecedented Warm Northwest Pacific and La Niña


doi: 10.1007/s00376-022-1456-0

  • An extreme drought appeared in South China from October 2020 to March 2021. During that time, sea surface temperatures exhibited an unprecedented warm center over the northwest Pacific (NWP) and a cold center over the tropical eastern Pacific (La Niña). This study demonstrates the combined effects of an exceptionally warm NWP and a moderate La Niña are closely linked to the anomalous drought in South China. The sea surface temperature anomaly in these two regions induced a steeper horizontal geopotential height gradient over South China. As a result, anomalous northeasterly winds prevailed over South China, altering water vapor transport and moisture convergence. A simplified atmospheric general circulation model also verifies the influence of the NWP warm anomaly on South China precipitation. This study points out that the sea surface temperature variation in the NWP was important to the occurrence of extreme drought in South China from October 2020 to March 2021.
    摘要: 2020年10月至2021年3月,华南地区经历了一场极端的季节性连旱事件。在这期间,赤道及北太平洋海表温度呈现出了西北太平洋地区前所未有的暖中心及热带东太平洋(拉尼娜)的冷中心。根据以往研究表明,拉尼娜事件能够通过遥相关引起东亚地区的环流异常,但除了同期中等程度的拉尼娜事件的影响外,西北太平洋地区的异常增温也值得关注。本研究表明,异常增温的西北太平洋和中等程度的拉尼娜现象的协同作用与华南地区的极端干旱密切相关。这两个海区的海表温度异常导致华南地区的水平位势高度场梯度更陡,从而引起华南地区低层上空出现异常的东北风,改变了水汽输送和水汽辐合。简化的大气综合环流模型验证了西北太平洋海温异常对华南地区大气环流的影响。本研究指出,西北太平洋地区海表温度的变化对2020年10月至2021年3月华南地区极端干旱的发生具有重要意义。
  • 加载中
  • Figure 1.  (a) Monthly average of the precipitation anomaly (shading; mm month–1) in South China from October 2020 to March 2021, with the black box indicating the region defined as South China in this study; (b) monthly average precipitation anomaly (blue bars; mm month–1) and the water vapor flux divergence anomaly (red line; 10–5 kg m–2 s–1) in South China for 1979–2020 (October to March); (c) monthly climatological mean precipitation from October to March (green line; mm), mean precipitation minus one standard deviation (red line; mm) and monthly precipitation (blue bars; mm).

    Figure 2.  Composites of water vapor flux (vectors; kg m–1 s–1) and water vapor flux divergence (shading; 10–5 kg m–2 s–1) for (a) the climatological mean and the (b) average and (c) anomaly from October 2020 to March 2021. The water vapor budget for each boundary of South China for (d) the climatological mean and (e) the average and (f) anomaly from October 2020 to March 2021. The numbers in black represent the amount of water vapor (107 kg s–1) transported across each boundary. The arrows in red and blue represent water vapor output and input, respectively, the black box indicating the region defined as South China in this study

    Figure 3.  (a) Climatological mean of UV850 (vectors in m s–1) and geopotential height (shading; m); (b) and (c) similar to (a), but for the average and anomaly from October 2020 to March 2021, respectively; panels (d)–(f) are similar to (a)–(c), but for specific humidity in (d) and (e) (10–2 kg kg–1) and (f) (10–3 kg kg–1), the black box indicating the region defined as South China in this study

    Figure 4.  (a) Composites of UV850 (vectors; m s–1) and geopotential height (shading; m) anomaly from October 2020 to March 2021 in EA-WP; panels (b) and (c) are similar to (a), but for the ENSO-removed and NWP-removed UV850 and geopotential height, respectively, the black box indicating the region defined as South China in this study

    Figure 5.  (a) Linear regression of UV850 (vectors; m s–1 K–1) and geopotential height (shading; m K–1) onto the Niño-3.4 index; panels (b), (c), and (d) are similar to (a), but for late autumn, winter and early spring, respectively. The right panels are similar to the left panels but for the NWP-SST index. In this figure, only values that are significant at the 0.05 level or better are plotted, the black box indicating the region defined as South China in this study

    Figure 5.  (Continued)

    Figure 6.  (a) The average monthly Pacific Ocean sea surface temperature anomalies (shading; K) from October 2020 to March 2021, with black boxes indicating the NWP, NEP, and TEP sea surface temperature regions; (b) Standardized NWP sea surface temperature (red line), NEP sea surface temperature (orange line), geopotential height along mid-latitude East Asia (blue line) and Niño-3.4 (green bar) indices for 1979–2020 (October to March).

    Figure 7.  (a) Difference in 850 hPa geopotential height (m) and wind (m s–1) between the NP_Run and Control_Run simulations; (b), (c), and (d) are similar to (a), but for the TEP_Run, NWP_Run, and NEP_Run, respectively. In this figure, only differences that are significant at the 0.05 significance level or better according to a paired t-test are plotted, the black box indicating the region defined as South China in this study

    Figure 8.  Schematic diagram presenting the air–sea interaction in the western North Pacific, red (blue) shading indicates positive (negative) SST anomalies. Thick green arrows and contours show anomalous horizontal wind and geopotential height at 850 hPa in 2020/21.

  • Ashok, K., M. Shamal, A. K. Sahai, and P. Swapna, 2017: Nonlinearities in the evolutional distinctions between El Niño and La Niña types. J. Geophys. Res.: Oceans, 122, 9649−9662, https://doi.org/10.1002/2017JC013129.
    Ashok, K., S. K. Behera, S. A. Rao, H. Y. Weng, and T. Yamagata, 2007: El Niño Modoki and its possible teleconnection. J. Geophys. Res.: Oceans, 112, C11007, https://doi.org/10.1029/2006JC003798.
    Cai, W. J., and Coauthors, 2015: ENSO and greenhouse warming. Nature Climate Change, 5, 849−859, https://doi.org/10.1038/nclimate2743.
    Chen, J. P., X. Wang, W. Zhou, C. Z. Wang, Q. Xie, G. Li, and S. Chen, 2018: Unusual rainfall in southern China in decaying August during extreme El Niño 2015/16: Role of the western Indian Ocean and north tropical Atlantic SST. J. Climate, 31, 7019−7034, https://doi.org/10.1175/JCLI-D-17-0827.1.
    Chen, M. Y., W. Shi, P. P. Xie, V. B. S. Silva, V. E. Kousky, R. Wayne Higgins, and J. E. Janowiak, 2008: Assessing objective techniques for gauge-based analyses of global daily precipitation. J. Geophys. Res.: Atmos., 113, D04110, https://doi.org/10.1029/2007JD009132.
    Fu, S. M., D. S. Li, J. H. Sun, D. Si, J. Ling, and F. Y. Tian, 2016: A 31-year trend of the hourly precipitation over South China and the underlying mechanisms. Atmospheric Science Letters, 17, 216−222, https://doi.org/10.1002/asl.645.
    Geng, X., W. J. Zhang, Y. Y. Xiang, and F. Jiang, 2021: Dominant spatiotemporal variability of wintertime precipitation days in China and the linkage with large-scale climate drivers. International Journal of Climatology, 41, 3561−3577, https://doi.org/10.1002/joc.7035.
    Gore, M., B. J. Abiodun, and F. Kucharski, 2020: Understanding the influence of ENSO patterns on drought over southern Africa using SPEEDY. Climate Dyn., 54, 307−327, https://doi.org/10.1007/s00382-019-05002-w.
    Hansen, J., M. Sato, R. Ruedy, K. Lo, D. W. Lea, and M. Medina-Elizade, 2006: Global temperature change. Proceedings of the National Academy of Sciences of the United States of America, 103, 14 288−14 293,
    He, C., and T. J. Zhou, 2015: Responses of the western North Pacific subtropical high to global warming under RCP4.5 and RCP8.5 scenarios projected by 33 CMIP5 models: The dominance of tropical Indian Ocean–tropical western Pacific SST gradient. J. Climate, 28, 365−380, https://doi.org/10.1175/JCLI-D-13-00494.1.
    Hersbach, H., and Coauthors, 2020: The ERA5 global reanalysis. Quart. J. Roy. Meteor. Soc., 146, 1999−2049, https://doi.org/10.1002/qj.3803.
    Hu, C. D., D. K. Chen, G. Huang, and S. Yang, 2018: Dipole types of autumn precipitation variability over the subtropical East Asia-western Pacific modulated by shifting ENSO. Geophys. Res. Lett., 45, 9123−9130, https://doi.org/10.1029/2018GL078982.
    Huang, B. Y., and Coauthors, 2017: Extended reconstructed sea surface temperature, version 5 (ERSSTv5): Upgrades, validations, and intercomparisons. J. Climate, 30, 8179−8205, https://doi.org/10.1175/JCLI-D-16-0836.1.
    Jian, Y. T., X. X. Lin, W. Zhou, M. Q. Jian, M. Y. T. Leung, and P. K. Y. Cheung, 2020: Analysis of record-high temperature over southeast coastal China in winter 2018/19: The combined effect of mid- to high-latitude circulation systems and SST forcing over the North Atlantic and tropical western Pacific. J. Climate, 33, 8813−8831, https://doi.org/10.1175/JCLI-D-19-0732.1.
    Jian, Y. T., M. Y. T. Leung, W. Zhou, M. Q. Jian, S. Yang, and X. X. Lin, 2021: Interdecadal shift of the relationship between ENSO and winter synoptic temperature variability over the Asian–Pacific–American region in the 1980s. J. Climate, 34, 5321−5335, https://doi.org/10.1175/JCLI-D-20-0931.1.
    Jiang, T., Q. Yang, Q. Zhang, and K. Li, 2012: Spatial characteristics of drought hazard in Guangdong province. Journal of Lake Sciences, 24, 156−160, https://doi.org/10.18307/2012.0122. (in Chinese with English abstract
    Kucharski, F., F. Molteni, and A. Bracco, 2006: Decadal interactions between the western tropical Pacific and the North Atlantic Oscillation. Climate Dyn., 26, 79−91, https://doi.org/10.1007/s00382-005-0085-5.
    Kucharski, F., F. Molteni, M. P. King, R. Farneti, I.-S. Kang, and L. Feudale, 2013: On the need of intermediate complexity general circulation models: A “SPEEDY” example. Bull. Amer. Meteor. Soc., 94, 25−30, https://doi.org/10.1175/BAMS-D-11-00238.1.
    Lee, S.-K., W. Park, M. O. Baringer, A. L. Gordon, B. Huber, and Y. Y. Liu, 2015: Pacific origin of the abrupt increase in Indian Ocean heat content during the warming hiatus. Nature Geoscience, 8, 445−449, https://doi.org/10.1038/ngeo2438.
    Leung, M. Y.-T., and W. Zhou, 2018: Circumglobal teleconnection and eddy control of variation in summer precipitation over Northwest China. Climate Dyn., 51, 1351−1362, https://doi.org/10.1007/s00382-017-3958-5.
    Leung, M. Y.-T., H. H. N. Cheung, and W. Zhou, 2017: Meridional displacement of the East Asian trough and its response to the ENSO forcing. Climate Dyn., 48, 335−352, https://doi.org/10.1007/s00382-016-3077-8.
    Leung, M. Y.-T., S. Qiu, and W. Zhou, 2018: Modulations of rising motion and moisture on summer precipitation over the middle and lower reaches of the Yangtze River. Climate Dyn., 51, 4259−4269, https://doi.org/10.1007/s00382-018-4247-7.
    Leung, M. Y.-T., W. Zhou, D. X. Wang, P. W. Chan, S. M. Lee, and H. W. Tong, 2020: Remote tropical western Indian Ocean forcing on changes in June precipitation in South China and the Indochina peninsula. J. Climate, 33, 7553−7566, https://doi.org/10.1175/JCLI-D-19-0626.1.
    Li, W. J., H.-C. Ren, J. Q. Zuo, and H.-L. Ren, 2018: Early summer southern China rainfall variability and its oceanic drivers. Climate Dyn., 50, 4691−4705, https://doi.org/10.1007/s00382-017-3898-0.
    Li, X. Z., W. Zhou, and Y. D. Chen, 2015: Assessment of regional drought trend and risk over China: A drought climate division perspective. J. Climate, 28, 7025−7037, https://doi.org/10.1175/JCLI-D-14-00403.1.
    Li, X. Z., Z. P. Wen, D. L. Chen, and Z. S. Chen, 2019b: Decadal transition of the leading mode of interannual moisture circulation over East Asia–Western North Pacific: Bonding to different evolution of ENSO. J. Climate, 32, 289−308, https://doi.org/10.1175/JCLI-D-18-0356.1.
    Li, Y., B. S. Ma, J. Feng, and Y. Lu, 2019a: Influence of the strongest central Pacific El Niño–Southern Oscillation events on the precipitation in eastern China. International Journal of Climatology, 39, 3076−3090, https://doi.org/10.1002/joc.6004.
    Ma, S. M., C. W. Zhu, and J. Liu, 2020: Combined impacts of warm central equatorial Pacific sea surface temperatures and anthropogenic warming on the 2019 severe drought in East China. Adv. Atmos. Sci., 37, 1149−1163, https://doi.org/10.1007/s00376-020-0077-8.
    Mantua, N. J., and S. R. Hare, 2002: The pacific decadal oscillation. Journal of Oceanography, 58, 35−44, https://doi.org/10.1023/A:1015820616384.
    Molteni, F., 2003: Atmospheric simulations using a GCM with simplified physical parametrizations. I: Model climatology and variability in multi-decadal experiments. Climate Dyn., 20, 175−191, https://doi.org/10.1007/s00382-002-0268-2.
    Niu, N., and J. P. Li, 2008: Interannual Variability of autumn precipitation over South China and its relation to atmospheric circulation and SST anomalies. Adv. Atmos. Sci., 25, 117−125, https://doi.org/10.1007/s00376-008-0117-2.
    Qiu, S., W. Zhou, M. Y.-T. Leung, and X. Z. Li, 2017: Regional moisture budget associated with drought/flood events over China. Progress in Earth & Planetary Science, 4, 36, https://doi.org/10.1186/s40645-017-0148-3.
    Rasmusson, E. M., and T. H. Carpenter, 1982: Variations in tropical sea surface temperature and surface wind fields associated with the Southern Oscillation/El Niño. Mon. Wea. Rev., 110, 354−384, https://doi.org/10.1175/1520-0493(1982)110<0354:VITSST>2.0.CO;2.
    Ren, H. L., and F. F. Jin, 2011: Niño indices for two types of ENSO. Geophys. Res. Lett., 38, L04704, https://doi.org/10.1029/2010GL046031.
    Schmitz, J. T., and S. L. Mullen, 1996: Water vapor transport associated with the summertime North American monsoon as depicted by ECMWF analyses. J. Climate, 9, 1621−1634, https://doi.org/10.1175/1520-0442(1996)009<1621:WVTAWT>2.0.CO;2.
    Timmermann, A., and Coauthors, 2018: El Niño–Southern oscillation complexity. Nature, 559, 535−545, https://doi.org/10.1038/s41586-018-0252-6.
    Trenberth, K. E., 1999: Atmospheric moisture recycling: Role of advection and local evaporation. J. Climate, 12, 1368−1381, https://doi.org/10.1175/1520-0442(1999)012<1368:AMRROA>2.0.CO;2.
    Volkov, D. L., S.-K. Lee, F. W. Landerer, and R. Lumpkin, 2017: Decade-long deep-ocean warming detected in the subtropical South Pacific. Geophys. Res. Lett., 44, 927−936, https://doi.org/10.1002/2016GL071661.
    Wang, B., R. G. Wu, and X. H. Fu, 2000: Pacific-East Asian teleconnection: How does ENSO affect East Asian climate? J. Climate, 13, 1517−1536, https://doi.org/10.1175/1520-0442(2000)013<1517:PEATHD>2.0.CO;2.
    Wang, H. J., and H. P. Chen, 2012: Climate control for southeastern China moisture and precipitation: Indian or East Asian monsoon? J. Geophys. Res.: Atmos., 117, D12109, https://doi.org/10.1029/2012JD017734.
    Wang, L., W. Chen, and R. H. Huang, 2008: Interdecadal modulation of PDO on the impact of ENSO on the East Asian winter monsoon. Geophys. Res. Lett., 35, L20702, https://doi.org/10.1029/2008GL035287.
    Wang, L., W. Chen, W. Zhou, and G. Huang, 2015: Teleconnected influence of tropical northwest Pacific sea surface temperature on interannual variability of autumn precipitation in southwest China. Climate Dyn., 45, 2527−2539, https://doi.org/10.1007/s00382-015-2490-8.
    Wang, P., C. Y. Tam, and K. Xu, 2019: El Niño–East Asian monsoon teleconnection and its diversity in CMIP5 models. Climate Dyn., 53, 6417−6435, https://doi.org/10.1007/s00382-019-04938-3.
    Wang, P., C. Y. Tam, N. C. Lau, and K. Xu, 2021: Future impacts of two types of El Niño on East Asian rainfall based on CMIP5 model projections. Climate Dyn., 56, 899−916, https://doi.org/10.1007/s00382-020-05510-0.
    Wang, X. D., and H. L. Liu, 2016: PDO modulation of ENSO effect on tropical cyclone rapid intensification in the western North Pacific. Climate Dyn., 46, 15−28, https://doi.org/10.1007/s00382-015-2563-8.
    Watson, R. T., and Coauthors, 2001: Climate Change 2001: Synthesis Report. Cambridge University Press, 398 pp.
    Xu, K., C. Y. Tam, B. Q. Liu, S. Chen, X. Y. Yang, Z. Q. He, Q. Xie, and W. Q. Wang, 2020: Attenuation of Central Pacific El Niño amplitude by North Pacific sea surface temperature anomalies. J. Climate, 33, 6673−6688, https://doi.org/10.1175/JCLI-D-19-0767.1.
    Yang, S., Z. N. Li, J. Y. Yu, X. M. Hu, W. J. Dong, and S. He, 2018: El Niño–Southern Oscillation and its impact in the changing climate. National Science Review, 5, 840−857, https://doi.org/10.1093/nsr/nwy046.
    Zhang, L. X., Z. M. Chen, and T. J. Zhou, 2021: Human influence on the increasing drought risk over Southeast Asian monsoon region. Geophys. Res. Lett., 48, e2021GL093777, https://doi.org/10.1029/2021GL093777.
    Zhang, Q., C. Y. Xu, T. Jiang, and Y. J. Wu, 2007: Possible influence of ENSO on annual maximum streamflow of the Yangtze River, China. J. Hydrol., 333, 265−274, https://doi.org/10.1016/j.jhydrol.2006.08.010.
    Zhang, R. H., and A. Sumi, 2002: Moisture circulation over East Asia during El Niño episode in northern winter, spring and autumn. J. Meteor. Soc. Japan. Ser. II, 80, 213−227, https://doi.org/10.2151/jmsj.80.213.
    Zhang, W. J., F.-F. Jin, and A. Turner, 2014: Increasing autumn drought over southern China associated with ENSO regime shift. Geophys. Res. Lett., 41, 4020−4026, https://doi.org/10.1002/2014GL060130.
    Zhang, Y., W. Zhou, and M. Y.-T. Leung, 2019a: Phase relationship between summer and winter monsoons over the South China Sea: Indian Ocean and ENSO forcing. Climate Dyn., 52, 5229−5248, https://doi.org/10.1007/s00382-018-4440-8.
    Zhang, Y., W. Zhou, E. C. H. Chow, and M. Y.-T. Leung, 2019b: Delayed impacts of the IOD: Cross-seasonal relationships between the IOD, Tibetan Plateau snow, and summer precipitation over the Yangtze-Huaihe River region. Climate Dyn., 53, 4077−4093, https://doi.org/10.1007/s00382-019-04774-5.
    Zhao, H. K., and C. Z. Wang, 2016: Interdecadal modulation on the relationship between ENSO and typhoon activity during the late season in the western North Pacific. Climate Dyn., 47, 315−328, https://doi.org/10.1007/s00382-015-2837-1.
    Zhou, L.-T., C.-Y. Tam, W. Zhou, and J. C. L. Chan, 2010: Influence of South China Sea SST and the ENSO on winter rainfall over South China. Adv. Atmos. Sci., 27, 832−844, https://doi.org/10.1007/s00376-009-9102-7.
    Zhou, M. Z., G. S. Zhou, X. M. Lv, L. Zhou, and Y. H. Ji, 2019: Global warming from 1.5 to 2°C will lead to increase in precipitation intensity in China. International Journal of Climatology, 39, 2351−2361, https://doi.org/10.1002/joc.5956.
    Zhou, W., W. Chen, and D. X. Wang, 2012: The implications of El Niño-Southern Oscillation signal for South China monsoon climate. Aquatic Ecosystem Health & Management, 15, 14−19, https://doi.org/10.1080/14634988.2012.652050.
  • [1] LI Gang*, LI Chongyin, TAN Yanke, and BAI Tao, 2014: The Interdecadal Changes of South Pacific Sea Surface Temperature in the Mid-1990s and Their Connections with ENSO, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 66-84.  doi: 10.1007/s00376-013-2280-3
    [2] Yang AI, Ning JIANG, Weihong QIAN, Jeremy Cheuk-Hin LEUNG, Yanying CHEN, 2022: Strengthened Regulation of the Onset of the South China Sea Summer Monsoon by the Northwest Indian Ocean Warming in the Past Decade, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 943-952.  doi: 10.1007/s00376-021-1364-8
    [3] Xiaofei WU, Jiangyu MAO, 2019: Decadal Changes in Interannual Dependence of the Bay of Bengal Summer Monsoon Onset on ENSO Modulated by the Pacific Decadal Oscillation, ADVANCES IN ATMOSPHERIC SCIENCES, 36, 1404-1416.  doi: 10.1007/s00376-019-9043-8
    [4] Yawen DUAN, Peili WU, Xiaolong CHEN, Zhuguo MA, 2018: Assessing Global Warming Induced Changes in Summer Rainfall Variability over Eastern China Using the Latest Hadley Centre Climate Model HadGEM3-GC2, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 1077-1093.  doi: 10.1007/s00376-018-7264-x
    [5] KANG Xianbiao, HUANG Ronghui, WANG Zhanggui, ZHANG Rong-Hua, 2014: Sensitivity of ENSO Variability to Pacific Freshwater Flux Adjustment in the Community Earth System Model, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 1009-1021.  doi: 10.1007/s00376-014-3232-2
    [6] Fangxing FAN, Renping LIN, Xianghui FANG, Feng XUE, Fei ZHENG, Jiang ZHU, 2021: Influence of the Eastern Pacific and Central Pacific Types of ENSO on the South Asian Summer Monsoon, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 12-28.  doi: 10.1007/s00376-020-0055-1
    [7] Se-Hwan YANG, LI Chaofan, and LU Riyu, 2014: Predictability of Winter Rainfall in South China as Demonstrated by the Coupled Models of ENSEMBLES, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 779-786.  doi: 10.1007/s00376-013-3172-2
    [8] Ning JIANG, Congwen ZHU, 2021: Seasonal Forecast of South China Sea Summer Monsoon Onset Disturbed by Cold Tongue La Niña in the Past Decade, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 147-155.  doi: 10.1007/s00376-020-0090-y
    [9] ZHENG Fei, ZHANG Rong-Hua, ZHU Jiang, , 2014: Effects of Interannual Salinity Variability on the Barrier Layer in the Western-Central Equatorial Pacific: A Diagnostic Analysis from Argo, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 532-542.  doi: 10.1007/s00376-013-3061-8
    [10] XUE Feng, ZENG Qingcun, HUANG Ronghui, LI Chongyin, LU Riyu, ZHOU Tianjun, 2015: Recent Advances in Monsoon Studies in China, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 206-229.  doi: 10.1007/s00376-014-0015-8
    [11] Kaiming HU, Yingxue LIU, Gang HUANG, Zhuoqi HE, Shang-Min LONG, 2020: Contributions to the Interannual Summer Rainfall Variability in the Mountainous Area of Central China and Their Decadal Changes, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 259-268.  doi: 10.1007/s00376-019-9099-5
    [12] WANG Zhiren, WU Dexing, CHEN Xue'en, QIAO Ran, 2013: ENSO Indices and Analyses, ADVANCES IN ATMOSPHERIC SCIENCES, 30, 1491-1506.  doi: 10.1007/s00376-012-2238-x
    [13] Xinyi XING, Xianghui FANG, Da PANG, Chaopeng JI, 2024: Seasonal Variation of the Sea Surface Temperature Growth Rate of ENSO, ADVANCES IN ATMOSPHERIC SCIENCES, 41, 465-477.  doi: 10.1007/s00376-023-3005-x
    [14] Yuanhai FU, Zhongda LIN, Tao WANG, 2021: Simulated Relationship between Wintertime ENSO and East Asian Summer Rainfall: From CMIP3 to CMIP6, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 221-236.  doi: 10.1007/s00376-020-0147-y
    [15] Yadi LI, Xichen LI, Juan FENG, Yi ZHOU, Wenzhu WANG, Yurong HOU, 2024: Uncertainties of ENSO-related Regional Hadley Circulation Anomalies within Eight Reanalysis Datasets, ADVANCES IN ATMOSPHERIC SCIENCES, 41, 115-140.  doi: 10.1007/s00376-023-3047-0
    [16] ZHOU Lian-Tong, Chi-Yung TAM, ZHOU Wen, Johnny C. L. CHAN, 2010: Influence of South China Sea SST and the ENSO on Winter Rainfall over South China, ADVANCES IN ATMOSPHERIC SCIENCES, 27, 832-844.  doi: 10.1007/s00376--009-9102-7
    [17] ZHAO Haikun, WU Liguang, ZHOU Weican, 2010: Assessing the Influence of the ENSO on Tropical Cyclone Prevailing Tracks in the Western North Pacific, ADVANCES IN ATMOSPHERIC SCIENCES, 27, 1361-1371.  doi: 10.1007/s00376-010-9161-9
    [18] LIN Zhongda, LU Riyu, 2009: The ENSO's Effect on Eastern China Rainfall in the Following Early Summer, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 333-342.  doi: 10.1007/s00376-009-0333-4
    [19] Zhou Tianjun, Yu Rucong, Li Zhaoxin, 2002: ENSO-Dependent and ENSO-Independent Variability over the Mid-Latitude North Pacific: Observation and Air-Sea Coupled Model Simulation, ADVANCES IN ATMOSPHERIC SCIENCES, 19, 1127-1147.  doi: 10.1007/s00376-002-0070-4
    [20] Peng HU, Wen CHEN, Shangfeng CHEN, Lin WANG, Yuyun LIU, 2022: The Weakening Relationship between ENSO and the South China Sea Summer Monsoon Onset in Recent Decades, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 443-455.  doi: 10.1007/s00376-021-1208-6

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 17 December 2021
Manuscript revised: 03 March 2022
Manuscript accepted: 28 March 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

An Extreme Drought over South China in 2020/21 Concurrent with an Unprecedented Warm Northwest Pacific and La Niña

    Corresponding author: Marco Y.-T. LEUNG, liangyt55@mail.sysu.edu.cn
  • 1. School of Marine Sciences, Sun Yat-Sen University, Zhuhai 519082, China
  • 2. Southern Marine Science and Engineering Guangdong Laboratory (Zhuhai), Zhuhai 519082, China
  • 3. Department of Atmospheric and Oceanic Sciences & Institute of Atmospheric Sciences, Fudan University, Shanghai 200433, China
  • 4. Guy Carpenter Asia-Pacific Climate Impact Centre, School of Energy and Environment, City University of Hong Kong, Hong Kong, China

Abstract: An extreme drought appeared in South China from October 2020 to March 2021. During that time, sea surface temperatures exhibited an unprecedented warm center over the northwest Pacific (NWP) and a cold center over the tropical eastern Pacific (La Niña). This study demonstrates the combined effects of an exceptionally warm NWP and a moderate La Niña are closely linked to the anomalous drought in South China. The sea surface temperature anomaly in these two regions induced a steeper horizontal geopotential height gradient over South China. As a result, anomalous northeasterly winds prevailed over South China, altering water vapor transport and moisture convergence. A simplified atmospheric general circulation model also verifies the influence of the NWP warm anomaly on South China precipitation. This study points out that the sea surface temperature variation in the NWP was important to the occurrence of extreme drought in South China from October 2020 to March 2021.

摘要: 2020年10月至2021年3月,华南地区经历了一场极端的季节性连旱事件。在这期间,赤道及北太平洋海表温度呈现出了西北太平洋地区前所未有的暖中心及热带东太平洋(拉尼娜)的冷中心。根据以往研究表明,拉尼娜事件能够通过遥相关引起东亚地区的环流异常,但除了同期中等程度的拉尼娜事件的影响外,西北太平洋地区的异常增温也值得关注。本研究表明,异常增温的西北太平洋和中等程度的拉尼娜现象的协同作用与华南地区的极端干旱密切相关。这两个海区的海表温度异常导致华南地区的水平位势高度场梯度更陡,从而引起华南地区低层上空出现异常的东北风,改变了水汽输送和水汽辐合。简化的大气综合环流模型验证了西北太平洋海温异常对华南地区大气环流的影响。本研究指出,西北太平洋地区海表温度的变化对2020年10月至2021年3月华南地区极端干旱的发生具有重要意义。

    • Precipitation from late autumn to early spring, the period for harvesting and sowing in South China, can strongly influence water resources and agricultural production, with remarkable socioeconomic consequences. In the context of global warming, the frequency of regional autumn droughts has increased over subtropical East Asia in recent decades (Wang et al., 2015). In association with urbanization and population growth, water resources have become a crucial factor in the sustainable development of South China (Jiang et al., 2012). Unfortunately, this region is regularly stricken by droughts (Niu and Li, 2008; Zhang et al., 2021). South China suffered from an extraordinarily long-lasting severe drought from October 2020 to March 2021, which resulted in water scarcity and huge agricultural losses.

      El Niño/Southern Oscillation (ENSO) is the dominant tropical interannual climate phenomenon on interannual timescales (Rasmusson and Carpenter, 1982) and is a major factor modulating precipitation variations over East Asia (Zhou et al. 2012; Li et al., 2015, 2019b; Yang et al., 2018; Wang et al., 2019). In winter, El Niño usually induces more precipitation over South China and the middle to lower reaches of the Yangtze River, consistent with a weaker-than-normal East Asian winter monsoon (Wang et al., 2008; Zhou et al., 2010). The influence of El Niño can persist into the following summer (Zhang et al., 2007). The situation reverses during La Niña events (Qiu et al., 2017). Apart from the canonical El Niño (eastern Pacific El Niño; EP El Niño), the central Pacific (CP) El Niño exerts distinctly different impacts on precipitation over East Asia (Ashok et al., 2007; Li et al., 2019a; Ma et al., 2020; Xu et al., 2020). A significant reduction in winter precipitation in South China is noted along with a CP El Niño. However, for La Niña events, sea surface temperature anomaly patterns are not distinct, in contrast to El Niño diversity (Ren and Jin, 2011; Ashok et al., 2017; Timmermann et al., 2018). Zhang et al. (2014) showed that the impact of La Niña on autumn precipitation over South China has been relatively stable over the past few decades, leading to significant precipitation deficits over South China. In short, the influence of ENSO on South China precipitation varies depending on its phase and type.

      Accompanied by global warming, anomalous circulations and rainfall are expected to persist, amplifying rainfall intensity during mature boreal winter and late summer for both EP and CP El Niño events. Amplification of rainfall related to a CP El Niño seems stronger than that for an EP El Niño, but both are influenced by a moister atmosphere (Wang et al., 2021). Owing to the combined effects of increasing frequency of CP ENSO events concurrent with changes in greenhouse gases (Cai et al., 2015) and the ENSO-like interdecadal climate shift (Zhang et al., 2014; Zhao and Wang, 2016), it is expected that autumn drought frequency over South China will increase (Zhang and Sumi, 2002; Hu et al., 2018). This is mainly attributed to changes in atmospheric circulations and water vapor flux activities (Zhang and Sumi, 2002; Hu et al., 2018).

      According to previous studies, sea surface temperature variation in the different oceans may induce anomalous atmospheric circulation and influence the South China climate (Leung et al., 2017; Chen et al., 2018). In addition, a recent study by Li et al. (2018) demonstrates remote influences from the North Atlantic and the equatorial Indian Ocean on early summer precipitation in South China. Chen et al. (2018) showed that cold sea surface temperature anomalies in the tropical western Indian Ocean and the tropical North Atlantic could induce an anomalous anticyclone in the western North Pacific, leading to a positive precipitation anomaly in South China. In addition, the intensity of the western North Pacific subtropical high, which is controlled by the sea surface temperature gradient between the tropical Indian Ocean and the tropical western Pacific, enhances the intensity and frequency of summer precipitation events in South China (He and Zhou, 2015; Fu et al., 2016). Therefore, the precipitation variation in South China is also linked to sea surface temperature variations in other oceans apart from the tropical Pacific Ocean. On the other hand, a large portion of anomalous heating accompanied by anthropogenic warming is stored in the ocean, leading to significant and inequivalent ocean warming (Lee et al., 2015; Volkov et al., 2017). Hence, this study will further examine the contribution of sea surface temperature anomalies to the extremely low precipitation in South China from late autumn to early spring 2020/21.

      The remainder of this paper is organized as follows. Section 2 outlines the data and methods applied in this study. Subsequently, section 3.1 shows the impact of water vapor flux on precipitation in South China; section 3.2 illustrates the influence of anomalous sea surface temperature based on statistical analysis; section 3.3 elucidates the results of a simplified atmospheric general circulation model. Finally, section 4 discusses and summarizes the results of this study.

    2.   Data and methods
    • To investigate anomalous changes in precipitation over South China, CPC Global Unified Gauge-Based Precipitation data provided by NOAA/OAR/ESRLPSD (https://www.esrl.noaa.gov/psd/) are employed in this study, with a horizontal resolution of 2.5° × 2.5° (Chen et al., 2008). In addition, ERA5 monthly averaged data are utilized to investigate atmospheric variations in conjunction with changes in precipitation (Hersbach et al., 2020), which is obtained from https://cds.climate.copernicus.eu. Different variables with a 0.25° × 0.25° horizontal resolution are used in this study, including geopotential $ \left( \varPhi \right) $, horizontal wind $\left( {\boldsymbol{V}} \right)$, specific humidity $ \left( q \right) $, and the vertically integrated water vapor flux $\left( {\boldsymbol{Q}} \right)$. Horizontal divergence $ \left( {{\text{Div}}} \right) $ is derived from the water vapor flux. For sea surface temperature, data from the NOAA Extended Reconstructed Sea Surface Temperature (SST), version 5 (ERSSTv5; Huang et al., 2017) are used, with a resolution of 2° × 2°, from https://psl.noaa.gov/data/gridded/data.noaa.ersst.v5.html. The study period covers 42 years (1979 to 2020), from late autumn to early summer (October to March).

      Furthermore, the water vapor flux across each boundary is calculated following the method of Schmitz and Mullen (1996):

      where L represents the length of the boundary.

      Linear regression is applied to analyze the influence of sea surface temperature change in the two regions on different atmospheric variables. Linear regression is also employed to remove the ENSO signal and the NWP signal from the variable, with the equations as follows:

      where $ {M}_{X} $ is the linear regression coefficient of specific variable $ Y $ onto the time series X. In the two equations, the Niño-3.4 index (5°N–5°S, 170°–120°W) is the seasonal average from December to February. The specific variable $ Y $ and the NWP time series (SST average over 10°–50°N and 120°E–170°W) are relative to a monthly timescale. In addition, the z-test is employed to examine the significance level of the linear trend and the difference between the two samples.

    3.   Results
    • Figure 1a shows a precipitation anomaly over South China from October 2020 to March 2021. It is noted that the strongest negative anomaly is centered over northeastern Guangdong Province. According to the distribution of the negative precipitation anomaly, South China is hereafter defined as the region from 23.5°N to 27°N and from 112°E to 118°E (dotted box in Fig. 1a). For the climatological mean, the total precipitation from late autumn to early spring in South China is about 468.42 mm, accounting for only 33.36% of the annual precipitation. Further, robustly low precipitation of 232.26 mm occurs over South China from October 2020 to March 2021 (representing the minimum precipitation dating back to 1979), about half of the climatological mean (Fig. 1b). As shown in Fig. 1c, the precipitation anomaly is primarily due to the negative anomaly in early spring, particularly in March 2021. Therefore, the extremely low precipitation over South China supports the occurrence of the extreme drought event from late autumn to early spring. Hence, we will focus mainly on the cause of this extremely low precipitation in the following subsections.

      Figure 1.  (a) Monthly average of the precipitation anomaly (shading; mm month–1) in South China from October 2020 to March 2021, with the black box indicating the region defined as South China in this study; (b) monthly average precipitation anomaly (blue bars; mm month–1) and the water vapor flux divergence anomaly (red line; 10–5 kg m–2 s–1) in South China for 1979–2020 (October to March); (c) monthly climatological mean precipitation from October to March (green line; mm), mean precipitation minus one standard deviation (red line; mm) and monthly precipitation (blue bars; mm).

    • Water vapor flux is a determining factor affecting regional precipitation, as it provides a much greater moisture supply than local evaporation (Trenberth, 1999). Previous studies have indicated that variations in precipitation over East Asia are highly affected by variations in water vapor flux (Wang and Chen, 2012; Leung and Zhou, 2018; Leung et al., 2018). To determine the main causes of the anomalous precipitation in South China, the variation of water vapor divergence is calculated and shown in Fig. 1b. Anomalous low-level water vapor flux divergence is compared with anomalous precipitation, as illustrated in Fig. 1b. A significant negative correlation is apparent between the anomalous water vapor divergence and precipitation, with a Pearson correlation coefficient of –0.89. More importantly, the anomalous water vapor divergence from October 2020 to March 2021 was the greatest since 1979 (Fig. 1b). Therefore, it is necessary to investigate the water vapor flux as it pertains to the anomalous precipitation of this region.

      Figure 2 illustrates the vertically integrated water vapor flux and its horizontal divergence. For the climatological mean, the western boundary of South China is the boundary for water vapor input, followed by the southern boundary. The main output boundary is the eastern boundary (Fig. 2a). However, from October 2020 to March 2021, the western boundary was the only water vapor input boundary, as anomalous water vapor output occurred at the southern boundary (Fig. 2b). In addition, we investigated the contribution of the zonal and meridional components of water vapor flux. There is zonal convergence (–1.07 × 107 kg s–1) and meridional divergence (0.56 × 107 kg s–1) of water vapor from October 2020 to March 2021 over South China (Fig. 2g). Therefore, it is clear that the horizontal divergence of water vapor flux plays an important role in forming extreme drought over South China.

      Figure 2.  Composites of water vapor flux (vectors; kg m–1 s–1) and water vapor flux divergence (shading; 10–5 kg m–2 s–1) for (a) the climatological mean and the (b) average and (c) anomaly from October 2020 to March 2021. The water vapor budget for each boundary of South China for (d) the climatological mean and (e) the average and (f) anomaly from October 2020 to March 2021. The numbers in black represent the amount of water vapor (107 kg s–1) transported across each boundary. The arrows in red and blue represent water vapor output and input, respectively, the black box indicating the region defined as South China in this study

      The contribution of horizontal winds and specific humidity is examined to investigate the underlying causes of the changes in water vapor flux (Fig. 3). For the climatological mean, southwesterly winds in the lower troposphere (850 hPa) transfer moisture from the South China Sea into South China (Fig. 3a). From October 2020 to March 2021, the horizontal wind direction in the lower troposphere notably changes, concurrent with anomalous easterly winds west of Taiwan (Fig. 3b). Figure 3c shows anomalous northeasterly wind flow in South China, which generally parallels the contours of anomalous geopotential height at 850 hPa, which implies the geostrophic balance, $ f\boldsymbol{k}\times \boldsymbol{V}\approx -\nabla {\Phi } $. Hence the change in northeasterly wind flow is associated with the nature of the geopotential height gradient. In addition, the specific humidity from October 2020 to March 2021 is enhanced compared with its climatological mean (Figs. 3df). Therefore, the change in specific humidity is unlikely to have caused the extremely low precipitation.

      Figure 3.  (a) Climatological mean of UV850 (vectors in m s–1) and geopotential height (shading; m); (b) and (c) similar to (a), but for the average and anomaly from October 2020 to March 2021, respectively; panels (d)–(f) are similar to (a)–(c), but for specific humidity in (d) and (e) (10–2 kg kg–1) and (f) (10–3 kg kg–1), the black box indicating the region defined as South China in this study

    • As mentioned above, the anomalous northeasterly winds along with the horizontal gradient of geopotential height at 850 hPa, implies a geostrophic balance between the Coriolis force and pressure gradient force, $f\boldsymbol{k}\times \boldsymbol{V}\approx -\nabla \mathrm{\Phi }$. This means that the relative intensity of the northeasterly wind in this region is related to the anomalous geopotential height in South China from October to March (Fig. 3c).

      To identify the anomalous circulation from October 2020 to March 2021, the geopotential height (Z850) and 850 hPa wind (UV850) anomalies in a larger domain are delineated in Fig. 4a. It is noted that the steeper gradient of geopotential height is attributed to an anomalous positive center along mid-latitude East Asia and the northwest Pacific and a negative center in the South China Sea and surrounding regions.

      Figure 4.  (a) Composites of UV850 (vectors; m s–1) and geopotential height (shading; m) anomaly from October 2020 to March 2021 in EA-WP; panels (b) and (c) are similar to (a), but for the ENSO-removed and NWP-removed UV850 and geopotential height, respectively, the black box indicating the region defined as South China in this study

      Since ENSO is an important feature that can cause variations in atmospheric circulation, it is necessary to clarify the contribution of ENSO to the anomalous geopotential height over South China from October 2020 to March 2021. According to the Niño-3.4 index (Fig. 1b), a moderate La Niña event occurred during this period. To investigate its forcing, the ENSO geopotential height signal is removed by linear regression in Fig. 4b, and linear regression of the geopotential height and wind at 850 hPa onto the Niño-3.4 index is delineated for different periods in Figs. 5ad. The geopotential height and horizontal winds over the northwest Pacific and South China from late autumn to early spring are significantly related to ENSO. This means that the anomalously cold sea surface temperature in Niño-3.4 will be concurrent with the low geopotential height and anomalous northeasterly wind in East Asia and the northwest Pacific. Remarkably, the change in the anomalous wind concurrent with ENSO is more pronounced in autumn to early winter (Figs. 5b-c). According to Wang et al. (2000), ENSO could induce anomalous winds over the western North Pacific that develops rapidly in late fall when a strong cold event matures. The anomalies persist until the following spring or early summer, causing anomalously dry conditions in South China. Hence, La Niña supports a steeper horizontal geopotential height gradient over South China, which exists in tandem with anomalous northeasterly winds.

      Figure 5.  (a) Linear regression of UV850 (vectors; m s–1 K–1) and geopotential height (shading; m K–1) onto the Niño-3.4 index; panels (b), (c), and (d) are similar to (a), but for late autumn, winter and early spring, respectively. The right panels are similar to the left panels but for the NWP-SST index. In this figure, only values that are significant at the 0.05 level or better are plotted, the black box indicating the region defined as South China in this study

      Figure 5.  (Continued)

      Nonetheless, the moderate La Niña from October 2020 to March 2021 only partly explains the extreme drought in South China. Should the ENSO geopotential height signal be removed by linear regression, we find that the negative center of geopotential height in the South China Sea is weakened substantially, and the anomalous northeasterly winds over South China are still controlled by a steeper gradient of geopotential height (Fig. 4b). Notably, the positive geopotential height anomalies along mid-latitude East Asia cannot be explained by La Niña (Figs. 5ad). In addition, the average geopotential height along mid-latitude East Asia (30°–40°N and 108°–120°W) reached a historical peak (Fig. 6b; blue line). Consequently, the positive geopotential anomaly is the other factor that contributed to the steeper gradient of geopotential height over South China from October 2020 to March 2021, which resulted in horizontal wind anomalies.

      Figure 6.  (a) The average monthly Pacific Ocean sea surface temperature anomalies (shading; K) from October 2020 to March 2021, with black boxes indicating the NWP, NEP, and TEP sea surface temperature regions; (b) Standardized NWP sea surface temperature (red line), NEP sea surface temperature (orange line), geopotential height along mid-latitude East Asia (blue line) and Niño-3.4 (green bar) indices for 1979–2020 (October to March).

      Figure 6a shows an anomalously warm sea surface temperature signal found over the northwest Pacific (NWP; 10°–50°N and 120°E–170°W). The time series of the average sea surface temperature over the NWP is indicated by the red line in Fig. 6b. It shows a significant increasing trend and consistent decadal variation. It is also noted that the NWP sea surface temperature from October 2020 to March 2021 is the highest since 1979. Additionally, the regression of geopotential height and wind at 850 hPa onto NWP sea surface temperature is delineated during different periods in Figs. 5eh. A positive signal of geopotential height is found over mid-latitude East Asia, and the anomalous horizontal winds occur along the southern boundary of South China in early spring (Fig. 5e; The anomalous northeasterly wind is noted at the southern boundary of South China). In addition, the change in geopotential height concurrent with warm NWP SST anomaly is more pronounced in early spring, which is different from that of ENSO (Fig. 5h). As shown in Fig. 5, ENSO and NWP SST both play a role, yet with different seasonalities. ENSO contributes to the wind anomalies in autumn to early winter, while the NWP SST plays a more important role in early spring.

      If the NWP geopotential height signal is removed by linear regression, we find that the positive geopotential heights over mid-latitude East Asia are weakened substantially, and the anomalous northeasterly winds over South China also weaken (Fig. 4c). Hence, the warm anomaly in the NWP could potentially lead to steeper geopotential height gradients and anomalous winds over South China from October 2020 to March 2021.

    • A simplified atmospheric general circulation model from the International Center for Theoretical Physics is utilized to clarify the causal relationship between the NWP warm anomaly and the positive height anomaly in East Asia (Molteni, 2003; Kucharski et al., 2006, 2013). This model has been used to investigate the atmospheric response to anomalous sea surface temperature in previous studies (Jian et al., 2020, 2021; Leung et al., 2020; Gore et al., 2020). The model is configured with T30 spectral truncation resolution and eight vertical levels in σ-coordinates. The climatological mean sea surface temperature initially drives the model for 140 years. The first 10 years are used for model spin-up; the remaining years are taken as the control simulation (Control_Run) and the initial condition for the sensitivity simulation. A paired Z-test is employed to examine the difference between the Control_Run and the sensitivity simulations.

      A sensitivity simulation (NP_Run) is driven by the anomalous SST in the North Pacific, whose domain is illustrated in Fig. 6. As presented in Fig. 7a, the NP_Run reveals a significant positive geopotential height difference in midlatitude East Asia and a negative difference in the South China Sea and surrounding regions. The change in the geopotential height gradient is concurrent with significant northeasterly winds in South China. This implies that the anomalous North Pacific SSTs could drive the anomalous atmospheric circulation in East Asia and West Pacific from October 2020 to March 2021.

      Figure 7.  (a) Difference in 850 hPa geopotential height (m) and wind (m s–1) between the NP_Run and Control_Run simulations; (b), (c), and (d) are similar to (a), but for the TEP_Run, NWP_Run, and NEP_Run, respectively. In this figure, only differences that are significant at the 0.05 significance level or better according to a paired t-test are plotted, the black box indicating the region defined as South China in this study

      Furthermore, the other three simulations are carried out to investigate the contribution of different parts of the Pacific to the anomalous northeasterly wind in South China from October 2020 to March 2021. As shown in Figs. 7bd, the anomalous sea surface temperature added to tropical eastern Pacific (TEP; 15°S to 10°N and 165°E to 80°W), NWP (10°N to 50°N and 120°E to 170°W) and Northeast Pacific (NEP; 20°N to 55°N and 170°W to 120°W) from October to March are prescribed to the sensitivity simulations (TEP_Run, NWP_Run, and NEP_Run), respectively.

      As illustrated in the TEP_Run (Fig.7b), the cold anomaly in the tropical eastern Pacific cannot directly induce the northeasterly wind in South China because the model lacks the atmospheric forcing on the ocean, which suggests that the air-sea interaction is important for connecting the tropical eastern Pacific SST anomaly and South China precipitation. In the NWP_Run (Fig. 7c), anomalous northeasterly winds are noted over South China and are induced by the anomalously warm NWP. Hence, the results of numerical simulations emphasize that the NWP could be a crucial factor that transfers the La Niña signal to the anomalous northeasterly winds in South China. This also points out the importance of an unprecedented warm NWP during the extreme drought from October 2020 to March 2021 in South China. In addition, the sea surface temperature average over the NEP reaches a high value (Fig. 6b; orange line). The sensitivity simulation (NEP_Run), driven by the warm anomaly in the Northeast Pacific, is employed to explore its influence on the extreme drought event in South China. As shown in Fig. 7d, a warm NEP SST anomaly exerts a negligible contribution to the atmospheric circulation anomaly in South China.

    4.   Summary and discussion
    • This study suggests a possible physical mechanism to explain the extreme drought from late autumn to early spring over South China from October 2020 to March 2021. The extremely low precipitation is attributed not only to a moderate La Niña, which contributes significantly in autumn to early winter, but also to the NWP warm anomaly that plays a more important role in early spring. The result is summarized in a schematic diagram (Fig. 8). In La Niña, the cooling in the central Pacific could induce a warm anomaly in the NWP SST (Wang et al., 2000). This conjecture supports the unprecedented warm anomaly in the NWP, which causes a positive geopotential height anomaly over mid-latitude East Asia and a negative height anomaly over the South China Sea and surrounding regions. The resulting steeper geopotential height gradient that formed over South China was oriented from the northwest to the southeast and contributed to anomalous northeasterly winds. Therefore, owing to the combined effect of the anomalously warm NWP and the moderate La Niña, water vapor flux input was suppressed from October 2020 to March 2021 over South China.

      Figure 8.  Schematic diagram presenting the air–sea interaction in the western North Pacific, red (blue) shading indicates positive (negative) SST anomalies. Thick green arrows and contours show anomalous horizontal wind and geopotential height at 850 hPa in 2020/21.

      Furthermore, we also confirmed the contribution of the anomalous winds and water vapor flux to the maximum water vapor flux divergence from October 2020 to March 2021 (Figs. 2c and 3c). Previous studies have illustrated the formation of anomalous northeasterly winds due to La Niña (Zhang et al., 2019a, b; Wang et al., 2000). Nonetheless, we emphasize the NWP sea surface temperature’s forcing on the anomalous geopotential height and horizontal wind in the lower troposphere over South China, particularly on decadal timescales. A simplified atmospheric general circulation model also confirmed this result. Therefore, it is important to consider the synergistic effect of an NWP warm anomaly and La Niña on drought intensity and frequency over South China. Apart from an NWP warm anomaly and La Niña, winter droughts in South China could also be induced by the Eurasian teleconnection (Geng et al., 2021), which may have also influenced the extremely low precipitation in South China from October 2020 to March 2021.

      Our study demonstrates a corresponding role of anomalous NWP sea surface temperature in modulating water vapor transport and drought risk over South China. The low-frequency variation in NWP sea surface temperature could be periodically controlled by the Pacific Decadal Oscillation on decadal timescales (Mantua and Hare, 2002; Wang and Liu, 2016), which may lead to decadal variations in precipitation and drought occurrence over South China. In addition, there is significant background warming in the NWP (Watson et al., 2001; Hansen et al., 2006), which may lead to a decrease in precipitation over South China and more extreme drought events from late autumn to early spring in the future. According to the future projections of CMIP5 simulations (Zhou et al., 2019), the precipitation in South China will decrease when global temperatures rise to 1.5°C above pre-industrial levels but will increase with warming above 2°C. These results suggest that the trend in South China precipitation may be altered when global temperature increases by 2°C relative to the pre-industrial period.

      Acknowledgements. This study was jointly supported by the National Natural Science Foundation of China (Grant No. 41805042) and the Science and Technology Program of Guangzhou, China (Grant No. 202102020939).

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return