Advanced Search
Article Contents

Interannual Influences of the Surface Potential Vorticity Forcing over the Tibetan Plateau on East Asian Summer Rainfall


doi: 10.1007/s00376-021-1218-4

  • The influences of interannual surface potential vorticity forcing over the Tibetan Plateau (TP) on East Asian summer rainfall (EASR) and upper-level circulation are explored in this study. The results show that the interannual EASR and associated circulations are closely related to the surface potential vorticity negative uniform leading mode (PVNUM) over the TP. When the PVNUM is in the positive phase, more rainfall occurs in the Yangtze River valley, South Korea, Japan, and part of northern China, less rainfall occurs in southern China, and vice versa. A possible mechanism by which PVNUM affects EASR is proposed. Unstable air induced by the positive phase of PVNUM could stimulate significant upward motion and a lower-level anomalous cyclone over the TP. As a result, a dipole heating mode with anomalous cooling over the southwestern TP and anomalous heating over the southeastern TP is generated. Sensitivity experiment results regarding this dipole heating mode indicate that anomalous cooling over the southwestern TP leads to local and northeastern Asian negative height anomalies, while anomalous heating over the southeastern TP leads to local positive height anomalies. These results greatly resemble the realistic circulation pattern associated with EASR. Further analysis indicates that the anomalous water vapor transport associated with this anomalous circulation pattern is responsible for the anomalous EASR. Consequently, changes in surface potential vorticity forcing over the TP can induce changes in EASR.
    摘要: 青藏高原地形强迫对东亚天气气候有重要影响。基于位涡理论,本文研究了青藏高原地表位涡强迫对东亚夏季降水的影响。结果表明:东亚夏季降水的年际变率与青藏高原地表位涡负一致模(PVNUM)密切相关。正(负)位相的PVNUM有利于长江流域、华北部分地区、韩国和日本降水偏多(少),而华南降水偏少(多)。相关机制如下:与PVNUM相关的近地表静力不稳定层结能够激发出高原上空显著的上升运动以及高原南侧低层的气旋式环流。受低层气旋式环流南北风异常的影响,高原东西两侧分别出现非绝热冷却和非绝热加热中心,呈偶极子型加热模态。数值模拟证实,高原西侧的冷极子能够激发出一支沿西风急流传播的罗斯贝波列,导致高原西侧局地和东北亚地区的位势高度负异常;高原东侧的暖极子能够激发其上空的位势高度正异常;在冷暖极子加热的共同激发下,东亚上空相关的环流结构与再分析结果十分一致。进一步分析表明,与环流异常相关的大范围水汽输送异常是导致东亚夏季降水异常的原因。因此,青藏高原地表位涡的强迫能够引起东亚夏季降水的改变。
  • 加载中
  • Figure 1.  Spatial distribution of the first leading mode of variation in boreal summer for (a) surface PV over the TP; (b) precipitation over East Asia; and (c) circulation over East Asia at 200 hPa. (d) Corresponding normalized principal components of the first EOF mode are shown in (a), (b), and (c). The blue line in (a, c) denotes the TP topographic boundary of 2000 m.

    Figure 2.  Spatial distribution of the correlation coefficients between the PVSI and downstream rainfall anomalies. Areas exceeding the 0.05 significance level are highlighted with dots.

    Figure 3.  Spatial distribution of the correlation coefficient between the PVSI and geopotential height anomalies (shading) and circulation anomalies (vector, those passing the 0.05 significance level are shown) at (a) 200 hPa; (b) 500 hPa; and (c) 850 hPa. (d–f) are the same as (a–c), but for PreI. Areas exceeding the 0.05 significance level are highlighted with dots. The blue line denotes the TP topographic boundary of 3000 m.

    Figure 4.  Spatial distribution of the correlation coefficient between the PVSI and (a) surface static stability anomalies and (b) horizontal wind anomalies at the surface (vector, those exceeding the 0.05 significance test are shown) and ω at 500 hPa (shading). Areas exceeding the 0.05 significance level are highlighted with dots. The blue line denotes the TP topographic boundary of 3000 m.

    Figure 5.  Spatial distribution of the (a) correlation coefficient between the PVSI and OLR anomalies and (b) regressed rainfall anomalies against the PVSI (units: mm month−1). Areas exceeding the 0.05 significance level are highlighted with dots. The blue line denotes the TP topographic boundary of 3000 m.

    Figure 6.  Specified ideal diabatic heating (units: K d−1) horizontal distribution in (a) Expt1; (c) Expt2; and (e) Expt3. Specified ideal diabatic heating (units: K d−1) profiles in (b) Expt1 and (d) Expt2. The blue line denotes the TP topographic boundary of 3000 m.

    Figure 7.  Responses of geopotential height (shading, units: gpm) and horizontal wind (vector, units: m s−1) at 200 hPa in (a) Expt1; (b) Expt2; and (c) Expt3. The blue line denotes the TP topographic boundary of 3000 m. The light green shading indicates a westerly jet with a zonal wind speed greater than 20 m s−1 at 200 hPa. The “A” and “C” in Fig. 7c indicate the anomalous anticyclone and cyclone, respectively.

    Figure 8.  Spatial distribution of the correlation coefficient between the PVSI and the divergence of WVF anomalies (shading) and WVF anomalies (vector, those passing the 0.05 significance level are shown) at (a) 500 hPa and (b) 850 hPa. Areas exceeding the 0.05 significance level are highlighted with dots. The blue line denotes the TP topographic boundary of 3000 m.

    Figure 9.  Schematic showing the interannual surface TPPV influence on EASR. (a) Anomalous 200-hPa wind associated with surface TPPV forcing; (b) anomalous surface TPPV; and (c) same as (a) but for rainfall anomalies. The red (blue) circle indicates diabatic heating (cooling). The green (red) vector indicates the lower-level wind (upper-level Rossby wave propagation).

  • Allaart, M. A. F., H. Kelder, and L. C. Heijboer, 1993: On the Relation between Ozone and Potential Vorticity. Geophys. Res. Lett., 20, 811−814, https://doi.org/10.1029/93GL00822.
    Bowley, K. A., J. R. Gyakum, and E. H. Atallah, 2019: A new perspective toward cataloging northern hemisphere rossby wave breaking on the dynamic tropopause. Mon. Wea. Rev., 147, 409−431, https://doi.org/10.1175/MWR-D-18-0131.1.
    Chen, L. X., J. P. Liu, X. J. Zhou, and P. X. Wang, 1999: Impact of uplift of Qinghai-Xizang Plateau and change of land-ocean distribution on climate over Asia. Quaternary Sciences, 314−329. (in Chinese with English abstract)
    Danielsen, E. F., 1968: Stratospheric-tropospheric exchange based on radioactivity, ozone and potential vorticity. J. Atmos. Sci., 25, 502−518, https://doi.org/10.1175/1520-0469(1968)025<0502:STEBOR>2.0.CO;2.
    Ding, Y. H., 2007: The variability of the Asian summer monsoon. J. Meteor. Soc. Japan, 85B, 21−54, https://doi.org/10.2151/jmsj.85B.21.
    Ding, Y. H., and J. C. L. Chan, 2005: The East Asian summer monsoon: An overview. Meteor. Atmos. Phys., 89, 117−142, https://doi.org/10.1007/s00703-005-0125-z.
    Ding, Y. H., P. Liang, Y. J. Liu, and Y. C. Zhang, 2020: Multiscale variability of meiyu and its prediction: A new review. J. Geophys. Res., 125, e2019JD031496, https://doi.org/10.1029/2019JD031496.
    Duan, A. M., Y. M. Liu, and G. X. Wu, 2005: Heating status of the Tibetan Plateau from april to june and rainfall and atmospheric circulation anomaly over East Asia in midsummer. Science in China Series D: Earth Sciences, 48, 250−257, https://doi.org/10.1360/02yd0510.
    Ertel, H., 1942: Ein neuer hydrodynamischer wirbelsatz. Meteorologische Zeitschrift, 59, 271−281.
    Flohn, H., 1957: Large-scale aspects of the “Summer Monsoon” in South and East Asia. J. Meteor. Soc. Japan, 35A, 180−186, https://doi.org/10.2151/jmsj1923.35A.0_180.
    Folkins, I., and C. Appenzeller, 1996: Ozone and potential vorticity at the subtropical tropopause break. J. Geophys. Res., 101, 18787−18792, https://doi.org/10.1029/96JD01711.
    Guan, X. D., J. P. Huang, R. X. Guo, and P. Lin, 2015: The role of dynamically induced variability in the recent warming trend slowdown over the Northern Hemisphere. Scientific Reports, 5, 12669, https://doi.org/10.1038/srep12669.
    Hahn, D. G., and S. Manabe, 1975: The role of mountains in the South Asian monsoon circulation. J. Atmos. Sci., 32, 1515−1541, https://doi.org/10.1175/1520-0469(1975)032<1515:TROMIT>2.0.CO;2.
    Haynes, P. H., and M. E. McIntyre, 1987: On the evolution of vorticity and potential vorticity in the presence of diabatic heating and frictional or other forces. J. Atmos. Sci., 44, 828−841, https://doi.org/10.1175/1520-0469(1987)044<0828:OTEOVA>2.0.CO;2.
    Haynes, P. H., and M. E. McIntyre, 1990: On the conservation and impermeability theorems for potential vorticity. J. Atmos. Sci., 47, 2021−2031, https://doi.org/10.1175/1520-0469(1990)047<2021:OTCAIT>2.0.CO;2.
    Hoskins, B. J., 1991: Towards a PV-θ view of the general circulation. Tellus A, 43, 27−36, https://doi.org/10.3402/tellusa.v43i4.11936.
    Hsu, H.-H., and X. Liu, 2003: Relationship between the Tibetan Plateau heating and East Asian summer monsoon rainfall. Geophys. Res. Lett., 30, 2066, https://doi.org/10.1029/2003GL017909.
    Huang, J. P., W. Chen, Z. P. Wen, G. J. Zhang, Z. X. Li, Z. Y. Zuo, and Q. Y. Zhao, 2019: Review of Chinese atmospheric science research over the Past 70 Years: Climate and climate change. Science China Earth Sciences, 62, 1514−1550, https://doi.org/10.1007/s11430-019-9483-5.
    Huang, R. H., J. L. Chen, and G. Huang, 2007: Characteristics and variations of the East Asian monsoon system and its impacts on climate disasters in China. Adv. Atmos. Sci., 24, 993−1023, https://doi.org/10.1007/s00376-007-0993-x.
    Jiang, D. B., Z. L. Ding, H. Drange, and Y. Q. Gao, 2008: Sensitivity of East Asian climate to the progressive uplift and expansion of the Tibetan Plateau under the Mid-pliocene boundary conditions. Adv. Atmos. Sci., 25, 709−722, https://doi.org/10.1007/s00376-008-0709-x.
    Kitoh, A., 2004: Effects of mountain uplift on East Asian summer climate investigated by a coupled atmosphere-ocean GCM. J. Climate, 17, 783−802, https://doi.org/10.1175/1520-0442(2004)017<0783:EOMUOE>2.0.CO;2.
    Kripalani, R. H., and A. Kulkarni, 1997: Climatic impact of El Niño/La Niña on the Indian Monsoon: A new perspective. Weather, 52, 39−46, https://doi.org/10.1002/j.1477-8696.1997.tb06267.x.
    Kripalani, R. H., and A. Kulkarni, 2001: Monsoon rainfall variations and teleconnections over South and East Asia. International Journal of Climatology, 21, 603−616, https://doi.org/10.1002/joc.625.
    Kubota, H., Y. Kosaka, and S.-P. Xie, 2016: A 117-year long index of the Pacific-Japan pattern with application to interdecadal variability. International Journal of Climatology, 36, 1575−1589, https://doi.org/10.1002/joc.4441.
    Liang, X. Y., Y. M. Liu, and G. X. Wu, 2005: Effect of Tibetan Plateau on the site of onset and intensity of the Asian summer monsoon. Acta Meteorologica Sinica, 63, 799−805, https://doi.org/10.3321/j.issn:0577-6619.2005.05.023. (in Chinese with English abstract
    Liebmann, B., and C. A. Smith, 1996: Description of a complete (interpolated) outgoing longwave radiation dataset. Bull. Amer. Meteor. Soc., 77, 1275−1277.
    Liu, X.-D., 1999: Influences of Qinghai-Xizang (Tibet) Plateau uplift on the atmospheric circulation, global climate and environment changes. Plateau Meteorology, 18, 321−332, https://doi.org/10.3321/j.issn:1000-0534.1999.03.008. (in Chinese with English abstract
    Liu, X. D., and Z.-Y. Yin, 2002: Sensitivity of East Asian monsoon climate to the uplift of the Tibetan Plateau. Palaeogeography, Palaeoclimatology, Palaeoecology, 183, 223−245, https://doi.org/10.1016/S0031-0182(01)00488-6.
    Liu, Y. M., M. M. Lu, H. J. Yang, A. M. Duan, B. He, S. Yang, and G. X. Wu, 2020: Land-atmosphere-ocean coupling associated with the Tibetan Plateau and its climate impacts. National Science Review, 7, 534−552, https://doi.org/10.1093/nsr/nwaa011.
    Lucchesi, R., 2012: File specification for merra products. Gmao Office Note No.1 (Version 2.3), 87 pp. Available from https://gmao.gsfc.nasa.gov/pubs/docs/lucchesi528.pdf.
    Ma, T. T., G. X. Wu, Y. M. Liu, Z. H. Jiang, and J. H. Yu, 2019: Impact of surface potential vorticity density forcing over the Tibetan Plateau on the South China extreme precipitation in January 2008. Part I: Data analysis. Journal of Meteorological Research, 33, 400−415, https://doi.org/10.1007/s13351-019-8604-1.
    Mitchell, T. D., and P. D. Jones, 2005: An improved method of constructing a database of monthly climate observations and associated high-resolution grids. International Journal of Climatology, 25, 693−712, https://doi.org/10.1002/joc.1181.
    Rienecker, M. M., and Coauthors, 2011: MERRA: NASA's modern-era retrospective analysis for research and applications. J. Climate, 24, 3624−3648, https://doi.org/10.1175/JCLI-D-11-00015.1.
    Rossby, C.-G., 1940: Planetary flow patterns in the atmosphere. Quart. J. Roy. Meteor. Soc., 66, 68−87.
    Ryoo, J.-M., Y. Kaspi, D. W. Waugh, G. N. Kiladis, D. E. Waliser, E. J. Fetzer, and J. Kim, 2013: Impact of rossby wave breaking on U. S. west coast winter precipitation during enso events. J. Climate, 26, 6360−6382, https://doi.org/10.1175/JCLI-D-12-00297.1.
    Sandhya, M., S. Sridharan, M. I. Devi, and H. Gadhavi, 2015: Tropical upper tropospheric ozone enhancements due to potential vorticity intrusions over Indian sector. Journal of Atmospheric and Solar-Terrestrial Physics, 132, 147−152, https://doi.org/10.1016/j.jastp.2015.07.014.
    Sheng, C., and Coauthors, 2021: Characteristics of the potential vorticity and its budget in the surface layer over the Tibetan Plateau. International Journal of Climatology, 41, 439−455, https://doi.org/10.1002/joc.6629.
    Wang, B., and Q. Zhang, 2002: Pacific-East Asian teleconnection. Part II: How the philippine sea anomalous anticyclone is established during El Niño development. J. Climate, 15, 3252−3265, https://doi.org/10.1175/1520-0442(2002)015<3252:PEATPI>2.0.CO;2.
    Wang, B., R. G. Wu, and K.-M. Lau, 2001: Interannual variability of the Asian summer monsoon: Contrasts between the indian and the western North Pacific-East Asian monsoons. J. Climate, 14, 4073−4090, https://doi.org/10.1175/1520-0442(2001)014<4073:IVOTAS>2.0.CO;2.
    Watanabe, M., and M. Kimoto, 2000: Atmosphere-ocean thermal coupling in the North Atlantic: A positive feedback. Quart. J. Roy. Meteor. Soc., 126, 3343−3369, https://doi.org/10.1002/qj.49712657017.
    Watanabe, M., and F.-F. Jin, 2003: A moist linear baroclinic model: Coupled dynamical-convective response to El Niño. J. Climate, 16, 1121−1139, https://doi.org/10.1175/1520-0442(2003)16<1121:AMLBMC>2.0.CO;2.
    Watanabe, M., M. Kimoto, T. Nitta, and M. Kachi, 1999: A comparison of decadal climate oscillations in the North Atlantic detected in observations and a coupled GCM. J. Climate, 12, 2920−2940, https://doi.org/10.1175/1520-0442(1999)012<2920:ACODCO>2.0.CO;2.
    Wei, W., R. H. Zhang, M. Wen, X. Y. Rong, and T. Li, 2014: Impact of Indian summer monsoon on the South Asian high and its influence on summer rainfall over China. Climate Dyn., 43, 1257−1269, https://doi.org/10.1007/s00382-013-1938-y.
    Wen, N., Z. Y. Liu, and L. Li, 2019: Direct ENSO impact on East Asian summer precipitation in the developing summer. Climate Dyn., 52, 6799−6815, https://doi.org/10.1007/s00382-018-4545-0.
    Wu, G. X., and Y. P. Cai, 1997: Vertical wind shear and down-sliding slantwise vorticity development. Scientia Atmospheric Sinica, 21, 273−282, https://doi.org/10.3878/j.issn.1006-9895.1997.03.03. (in Chinese with English abstract
    Wu, G. X., and Y. S. Zhang, 1998: Tibetan Plateau forcing and the timing of the monsoon onset over South Asia and the South China Sea. Mon. Wea. Rev., 126, 913−927, https://doi.org/10.1175/1520-0493(1998)126<0913:TPFATT>2.0.CO;2.
    Wu, G. X., Y. P. Cai, and X. J. Tang, 1995: Moist potential vorticity and slantwise vorticity development. Acta Meteorologica Sinica, 53, 387−405, https://doi.org/10.11676/qxxb1995.045. (in Chinese with English abstract
    Wu, G. X., W. J. Li, and H. Guo, 1997: Sensible heat driven air-pump over the Tibetan Plateau and its impacts on the Asian Summer monsoon. Collections on the Memory of Zhao Jiuzhang, D. Z. Ye, Ed., Science Press, 116−126. (in Chinese)
    Wu, G. X., and Coauthors, 2007: The influence of mechanical and thermal forcing by the Tibetan Plateau on Asian Climate. Journal of Hydrometeorology, 8, 770−789, https://doi.org/10.1175/JHM609.1.
    Wu, G. X., Y. M. Liu, B. W. Dong, X. Y. Liang, A. M. Duan, Q. Bao, and J. J. Yu, 2012: Revisiting Asian monsoon formation and change associated with Tibetan Plateau forcing: I. Formation. Climate Dyn., 39, 1169−1181, https://doi.org/10.1007/s00382-012-1334-z.
    Wu, G. X., Y. M. Liu, B. He, Q. Bao, and Z. Q. Wang, 2018: Review of the impact of the Tibetan Plateau sensible heat driven air-pump on the Asian summer monsoon. Chinese Journal of Atmospheric Sciences, 42, 488−504, https://doi.org/10.3878/j.issn.1006-9895.1801.17279.
    Wu, R. G., 2002: A Mid-Latitude Asian circulation anomaly pattern in boreal summer and its connection with the Indian and East Asian summer monsoons. International Journal of Climatology, 22, 1879−1895, https://doi.org/10.1002/joc.845.
    Xie, S.-P., K. M. Hu, J. Hafner, H. Tokinaga, Y. Du, G. Huang, and T. Sampe, 2009: Indian ocean capacitor effect on Indo-Western pacific climate during the summer following El Niño. J. Climate, 22, 730−747, https://doi.org/10.1175/2008JCLI2544.1.
    Xie, S.-P., Y. Kosaka, Y. Du, K. M. Hu, J. S. Chowdary, and G. Huang, 2016: Indo-Western pacific ocean capacitor and coherent climate anomalies in Post-ENSO summer: A review. Adv. Atmos. Sci., 33, 411−432, https://doi.org/10.1007/s00376-015-5192-6.
    Xu, X. D., T. L. Zhao, X. H. Shi, and C. G. Lu, 2015: A study of the role of the Tibetan Plateau's thermal forcing in modulating rainband and moisture transport in Eastern China. Acta Meteorologica Sinica, 73, 20−35, https://doi.org/10.11676/qxxb2014.051. (in Chinese with English abstract
    Yanai, M., C. F. Li, and Z. S. Song, 1992: Seasonal heating of the Tibetan Plateau and its effects on the evolution of the Asian summer monsoon. J. Meteor. Soc. Japan, 70, 319−351, https://doi.org/10.2151/jmsj1965.70.1B_319.
    Yang, J. L., Q. Y. Liu, S.-P. Xie, Z. Y. Liu, and L. X. Wu, 2007: Impact of the Indian ocean Sst basin mode on the Asian summer monsoon. Geophys. Res. Lett., 34, L02708, https://doi.org/10.1029/2006GL028571.
    Ye, D. Z., and Y. X. Gao, 1979: The meteorology (Tibet) Plateau. Science Press, 278 pp. (in Chinese)
    Yeh, T.-C., S. W. Lo, and P. C. Chu, 1957: The wind structure and heat balance in the lower troposphere over Tibetan Plateau and its surrounding. Acta Meteorologica Sinica, 28, 108−121, https://doi.org/10.11676/qxxb1957.010. (in Chinese with English abstract
    Yu, J. H., Y. M. Liu, T. T. Ma, and G. X. Wu, 2019: Impact of surface potential vorticity density forcing over the Tibetan Plateau on the South China extreme precipitation in January 2008. Part Ⅱ: Numerical simulation. Journal of Meteorological Research, 33, 416−432, https://doi.org/10.1007/s13351-019-8606-z.
    Zhang, G. S., J. Y. Mao, G. X. Wu, and Y. M. Liu, 2021: Impact of potential vorticity anomalies around the Eastern Tibetan Plateau on Quasi-Biweekly oscillations of summer rainfall within and South of the Yangtze Basin in 2016. Climate Dyn., 56, 813−835, https://doi.org/10.1007/s00382-020-05505-x.
    Zhao, L. and Y. H. Ding, 2009: Potential vorticity analysis of cold air activities during the East Asian summer monsoon. Chinese Journal of Atmospheric Sciences, 33, 359−374, https://doi.org/10.3878/j.issn.1006-9895.2009.02.14. (in Chinese with English abstract
    Zhao, P., and L. X. Chen, 2001: Climatic features of atmospheric heat source/sink over the Qinghai-Xizang Plateau in 35 years and its relation to rainfall in China. Science in China Series D: Earth Sciences, 44, 858−864, https://doi.org/10.1007/BF02907098.
    Zhou, X. Y., F. Liu, B. Wang, B. Q. Xiang, C. Xing, and H. Wang, 2019: Different responses of East Asian summer rainfall to El Niño decays. Climate Dyn., 53, 1497−1515, https://doi.org/10.1007/s00382-019-04684-6.
    Zuo, Z. Y., R. H. Zhang, and P. Zhao, 2011: The relation of vegetation over the Tibetan Plateau to rainfall in China during the boreal summer. Climate Dyn., 36, 1207−1219, https://doi.org/10.1007/s00382-010-0863-6.
  • [1] MAN Wenmin, and ZHOU Tianjun, 2014: Regional-scale Surface Air Temperature and East Asian Summer Monsoon Changes during the Last Millennium Simulated by the FGOALS-gl Climate System Model, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 765-778.  doi: 10.1007/s00376-013-3123-y
    [2] Ronghui HUANG, Yong LIU, Zhencai DU, Jilong CHEN, Jingliang HUANGFU, 2017: Differences and Links between the East Asian and South Asian Summer Monsoon Systems: Characteristics and Variability, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 1204-1218.  doi: 10.1007/ s00376-017-7008-3
    [3] Xuke LIU, Xiaojing JIA, Min WANG, Qifeng QIAN, 2022: The Impact of Tibetan Plateau Snow Cover on the Summer Temperature in Central Asia, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 1103-1114.  doi: 10.1007/s00376-021-1011-4
    [4] Zhou Yushu, Deng Guo, Gao Shouting, Xu Xiangde, 2002: The Wave Train Characteristics of Teleconnection Caused by the Thermal Anomaly of the Underlying Surface of the Tibetan Plateau. Part Ⅰ: Data Analysis, ADVANCES IN ATMOSPHERIC SCIENCES, 19, 583-593.  doi: 10.1007/s00376-002-0002-3
    [5] Jinghua CHEN, Xiaoqing WU, Chunsong LU, Yan YIN, 2022: Seasonal and Diurnal Variations of Cloud Systems over the Eastern Tibetan Plateau and East China: A Cloud-resolving Model Study, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 1034-1049.  doi: 10.1007/s00376-021-0391-9
    [6] FENG Jinming, WEI Ting, DONG Wenjie, WU Qizhong, and WANG Yongli, 2014: CMIP5/AMIP GCM Simulations of East Asian Summer Monsoon, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 836-850.  doi: 10.1007/s00376-013-3131-y
    [7] SU Tonghua, XUE Feng*, ZHANG He, 2014: Simulating the Intraseasonal Variation of the East Asian Summer Monsoon by IAP AGCM4.0, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 570-580.  doi: 10.1007/s00376-013-3029-8
    [8] Congwen ZHU, Boqi LIU, Kang XU, Ning JIANG, Kai LIU, 2021: Diversity of the Coupling Wheels in the East Asian Summer Monsoon on the Interannual Time Scale: Challenge of Summer Rainfall Forecasting in China, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 546-554.  doi: 10.1007/s00376-020-0199-z
    [9] FENG Juan*, CHEN Wen, 2014: Interference of the East Asian Winter Monsoon in the Impact of ENSO on the East Asian Summer Monsoon in Decaying Phases, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 344-354.  doi: 10.1007/s00376-013-3118-8
    [10] SU Qin, LU Riyu, LI Chaofan, 2014: Large-scale Circulation Anomalies Associated with Interannual Variation in Monthly Rainfall over South China from May to August, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 273-282.  doi: 10.1007/s00376-013-3051-x
    [11] Xue Feng, 2001: Interannual to Interdecadal Variation of East Asian Summer Monsoon and its Association with the Global Atmospheric Circulation and Sea Surface Temperature, ADVANCES IN ATMOSPHERIC SCIENCES, 18, 567-575.  doi: 10.1007/s00376-001-0045-x
    [12] LI Xiaofan, SHEN Xinyong, LIU Jia, 2014: Effects of Doubled Carbon Dioxide on Rainfall Responses to Large-Scale Forcing: A Two-Dimensional Cloud-Resolving Modeling Study, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 525-531.  doi: 10.1007/s00376-013-3030-2
    [13] JIN Liya, WANG Huijun, CHEN Fahu, JIANG Dabang, 2006: A Possible Impact of Cooling over the Tibetan Plateau on the Mid-Holocene East Asian Monsoon Climate, ADVANCES IN ATMOSPHERIC SCIENCES, 23, 543-550.  doi: 10.1007/s00376-006-0543-y
    [14] Huang Ronghui, Wu Bingyi, Sung-Gil Hong, Jai-Ho Oh, 2001: Sensitivity of Numerical Simulations of the East Asian Summer Monsoon Rainfall and Circulation to Different Cumulus Parameterization Schemes, ADVANCES IN ATMOSPHERIC SCIENCES, 18, 23-41.  doi: 10.1007/s00376-001-0002-8
    [15] ZHAI Guoqing, LI Xiaofan, ZHU Peijun, SHEN Hangfeng, ZHANG Yuanzhi, 2014: Surface Rainfall and Cloud Budgets Associated with Mei-yu Torrential Rainfall over Eastern China during June 2011, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 1435-1444.  doi: 10.1007/s00376-014-3256-7
    [16] HAN Jinping, ZHANG Renhe, 2009: The Dipole Mode of the Summer Rainfall over East China during 1958--2001, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 727-735.  doi: 10.1007/s00376-009-9014-6
    [17] Riyu LU, Saadia HINA, Xiaowei HONG, 2020: Upper- and Lower-tropospheric Circulation Anomalies Associated with Interannual Variation of Pakistan Rainfall during Summer, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 1179-1190.  doi: 10.1007/s00376-020-0137-0
    [18] Kai Chi WONG, Senfeng LIU, Andrew G. TURNER, Reinhard K. SCHIEMANN, 2018: Different Asian Monsoon Rainfall Responses to Idealized Orography Sensitivity Experiments in the HadGEM3-GA6 and FGOALS-FAMIL Global Climate Models, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 1049-1062.  doi: 10.1007/s00376-018-7269-5
    [19] Gong-Wang Si, Kuranoshin Kato, Takao Takeda, 1995: The Early Summer Seasonal Change of Large-scale Circulation over East Asia and Its Relation to Change of The Frontal Features and Frontal Rainfall Environment During 1991 Summer, ADVANCES IN ATMOSPHERIC SCIENCES, 12, 151-176.  doi: 10.1007/BF02656829
    [20] Xinyu LI, Riyu LU, 2021: Decadal Change in the Influence of the Western North Pacific Subtropical High on Summer Rainfall over the Yangtze River Basin in the Late 1970s, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 1823-1834.  doi: 10.1007/s00376-021-1051-9

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 07 June 2021
Manuscript revised: 09 August 2021
Manuscript accepted: 27 August 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

Interannual Influences of the Surface Potential Vorticity Forcing over the Tibetan Plateau on East Asian Summer Rainfall

    Corresponding author: Bian HE, heb@lasg.iap.ac.cn
  • 1. State Key Laboratory of Numerical Modeling for Atmospheric Sciences and Geophysical Fluid Dynamics (LASG), Institute of Atmospheric Physics, Chinese Academy of Sciences, Beijing 100029, China
  • 2. College of Earth and Planetary Sciences, University of Chinese Academy of Sciences, Beijing 100049, China

Abstract: The influences of interannual surface potential vorticity forcing over the Tibetan Plateau (TP) on East Asian summer rainfall (EASR) and upper-level circulation are explored in this study. The results show that the interannual EASR and associated circulations are closely related to the surface potential vorticity negative uniform leading mode (PVNUM) over the TP. When the PVNUM is in the positive phase, more rainfall occurs in the Yangtze River valley, South Korea, Japan, and part of northern China, less rainfall occurs in southern China, and vice versa. A possible mechanism by which PVNUM affects EASR is proposed. Unstable air induced by the positive phase of PVNUM could stimulate significant upward motion and a lower-level anomalous cyclone over the TP. As a result, a dipole heating mode with anomalous cooling over the southwestern TP and anomalous heating over the southeastern TP is generated. Sensitivity experiment results regarding this dipole heating mode indicate that anomalous cooling over the southwestern TP leads to local and northeastern Asian negative height anomalies, while anomalous heating over the southeastern TP leads to local positive height anomalies. These results greatly resemble the realistic circulation pattern associated with EASR. Further analysis indicates that the anomalous water vapor transport associated with this anomalous circulation pattern is responsible for the anomalous EASR. Consequently, changes in surface potential vorticity forcing over the TP can induce changes in EASR.

摘要: 青藏高原地形强迫对东亚天气气候有重要影响。基于位涡理论,本文研究了青藏高原地表位涡强迫对东亚夏季降水的影响。结果表明:东亚夏季降水的年际变率与青藏高原地表位涡负一致模(PVNUM)密切相关。正(负)位相的PVNUM有利于长江流域、华北部分地区、韩国和日本降水偏多(少),而华南降水偏少(多)。相关机制如下:与PVNUM相关的近地表静力不稳定层结能够激发出高原上空显著的上升运动以及高原南侧低层的气旋式环流。受低层气旋式环流南北风异常的影响,高原东西两侧分别出现非绝热冷却和非绝热加热中心,呈偶极子型加热模态。数值模拟证实,高原西侧的冷极子能够激发出一支沿西风急流传播的罗斯贝波列,导致高原西侧局地和东北亚地区的位势高度负异常;高原东侧的暖极子能够激发其上空的位势高度正异常;在冷暖极子加热的共同激发下,东亚上空相关的环流结构与再分析结果十分一致。进一步分析表明,与环流异常相关的大范围水汽输送异常是导致东亚夏季降水异常的原因。因此,青藏高原地表位涡的强迫能够引起东亚夏季降水的改变。

    2.   Data, method, and model
    • Monthly mean data on the hybrid σ–p model level obtained from MERRA2 (Rienecker et al., 2011; Lucchesi, 2012) are used to calculate surface PV and surface static stability. Variables include air temperature, zonal and meridional wind, and pressure.

      Additional data used in this study include monthly outgoing longwave radiation (OLR) from the polar-orbiting series of satellites of the National Oceanic and Atmospheric Administration (Liebmann and Smith, 1996), monthly land precipitation from the Climatic Research Unit (Mitchell and Jones, 2005), and monthly specific humidity and zonal, meridional, and vertical wind on pressure level from MERRA2.

      The study period is 1980–2017. The horizontal resolution of all MERRA2 data is 0.625° × 0.5° (longitude × latitude). The horizontal resolutions of the OLR and precipitation data are 2.5° × 2.5° and 0.5° × 0.5°, respectively. Climate mean values calculated over June, July, and August (JJA) are used to represent the boreal summer condition.

    • Surface PV is calculated as follows (Sheng et al., 2021):

      where ${\alpha _{\text{h}}}$ and ${{\boldsymbol\xi} _{\text{ah}}}$ is the specific volume and absolute vorticity in the hybrid σp coordinate system, respectively. The g is gravity, which is 9.8 m s−2, p is pressure, $\theta $ is potential temperature, f is the Coriolis parameter, $ (u,v) $ is horizontal wind, and h indicates that the horizontal difference is carried out at the hybrid σp level. The surface PV is obtained from the two bottom levels at the hybrid σp level. More details about the calculation of surface PV can be found in Sheng et al. (2021).

      The surface static stability is calculated as $\;\; - {{\partial \theta }}/{{\partial p}}$, where $\theta $ and $p$ are obtained from the two bottom levels at the hybrid σp level.

      The horizontal water vapor flux (WVF) is calculated as follows:

      where q is specific humidity.

      Statistical methods, including linear regression, linear correlation, Student’s t-test, empirical orthogonal function (EOF), and multivariate EOF (MVEOF), are used in this study. The linear trends and decadal variation (more than nine years) in the data are removed to highlight the interannual variability.

    • The linear baroclinic model (LBM) (Watanabe et al., 1999; Watanabe and Kimoto, 2000) is employed to investigate the responses of the upper-level circulation to the surface PV forcing over the TP. The model used in this study has 20 σ vertical levels. Each horizontal level is represented by spherical harmonics with a resolution of T42. More details about the model description can be found in Watanabe and Jin (2003). The LBM is a time-varying model that linearizes the basic state based on primitive equations. We take boreal summer climatology as the basic state in this study. Since dissipation terms including biharmonic horizontal diffusion, weak vertical diffusion, Newtonian damping, and Rayleigh friction are adopted, the model response reaches its steady state at approximately 14 days. Therefore, we use the results averaged from the last 15 days in the 30-day integration for analysis.

    3.   Results
    • EOF analysis is applied to analyze the dominant temporal and spatial features of the surface TPPV, EASR, and circulation over East Asia during boreal summer for the 1980–2017 period. Figure 1a shows the first leading mode of the EOF (EOF1) on the summer surface TPPV. The variation in surface PV on the southern slope and in the southeastern corner of the TP has the opposite sign to that in the main TP platform. Although the signs of the surface PV in the two areas are opposite, the southern area is quite small. In general, the distribution of EOF1 on surface PV shows a negative uniform mode (PVNUM). The corresponding normalized first principal component of the EOF pattern (PC1) is defined as the surface PV index (PVSI; red line), as shown in Fig. 1d. The variance percentage explained by the first leading mode is 36%.

      Figure 1.  Spatial distribution of the first leading mode of variation in boreal summer for (a) surface PV over the TP; (b) precipitation over East Asia; and (c) circulation over East Asia at 200 hPa. (d) Corresponding normalized principal components of the first EOF mode are shown in (a), (b), and (c). The blue line in (a, c) denotes the TP topographic boundary of 2000 m.

      Figure 1b shows the EOF1 of EASR. The corresponding normalized PC1 is defined as the precipitation index (PreI; blue dashed line), as shown in Fig. 1d. The maximum variation center of EASR occurs in the Yangtze River valley, South Korea, Japan (YKJ), and southern China. The variation in EASR in YKJ has the opposite sign to that in southern China. There is also a relatively weak dipole variation mode in northern China. The dominant mode of EASR obtained in this study is generally consistent with that obtained by Zuo et al. (2011).

      Using the NCEP-NCAR reanalysis over midlatitude Asia in JJAS (i.e., June, July, August, and September) from 1948 to 1998, Wu (2002) identified an interannual dominant pattern in the upper-level winds called the midlatitude Asian summer (MAS) pattern. The MAS pattern in his study features two anomalous cyclones, with one centered at (37.5°N, 65°E) and the other centered at (42.5°N, 130°E). Following Wu (2002), we applied MVEOF analysis to the 200-hPa wind anomaly to reveal the dominant circulation mode over midlatitude Asia during boreal summer. The domain for the MVEOF analysis is (20°–60°N, 50°–150°E), and the result is not sensitive to the domain. The first leading mode of MVEOF (MVEOF1) is shown in Fig. 1c, and the corresponding normalized PC1 is defined as the midlatitude Asian summer index (MASI; black dashed line) shown in Fig. 1d. Figure 1c shows two anomalous cyclones, one centered northwest of the TP and one centered over northeastern Asia. In between the two large cyclones is an anticyclone over the southeastern corner of the TP. The anticyclone identified in this study is more prominent than that in Wu (2002), which is partly because of the different reanalysis data and partly because of the different periods. From high to low latitudes, the positive MAS generally shows a “CCA” pattern (i.e., cyclone–cyclone–anticyclone pattern). Correspondingly, the negative MAS generally shows an “AAC” pattern (i.e., anticyclone–anticyclone–cyclone pattern).

      To investigate the relationship of surface TPPV forcing with the EASR and upper-level circulation over East Asia, Fig. 1d shows the time series of PVSI, MASI, and PreI. PVSI shows significant correlations with PreI and MASI, with yield correlation coefficients of 0.56 and 0.74 (both passing the 0.05 significance level), respectively. These results indicate that the EASR and the leading mode of circulation (i.e., MAS pattern) over East Asia are closely related to the surface TPPV forcing.

    • Figure 2 shows the spatial pattern of the correlation coefficients between the PVSI and EASR anomalies. The pattern (Fig. 2) is similar to EOF1 of EASR (Fig. 1b). A negative correlation appears in southern China, whereas positive correlations are observed in YKJ. A positive correlation, but with a small area, appears in part of northern China. The correlation coefficients in these regions are significant at the 0.05 significance level. A weak negative correlation is seen in the area south of northern China (approximately 42°N), but this correlation is not significant. These results indicate that the positive phase of PVNUM is associated with more rainfall in YJK and part of northern China and less rainfall in southern China, and the negative phase of PVNUM is related to the opposite rainfall pattern.

      Figure 2.  Spatial distribution of the correlation coefficients between the PVSI and downstream rainfall anomalies. Areas exceeding the 0.05 significance level are highlighted with dots.

      To understand the comprehensive three-dimensional relationship between surface TPPV forcing and EASR, the correlation patterns between the PVSI, PreI, and circulation anomalies at different levels are shown in Fig. 3. Figures 3ac show correlation coefficients between PVSI and circulation at different levels. Corresponding to the positive (negative) PVNUM, at 200 hPa (Fig. 3a), two anomalous cyclones (anticyclones) appear at middle latitudes at approximately 40°N. One center is located northwest of the TP, and the other is located in northeastern China (Fig. 3a). In addition to these two anomalous cyclones (anticyclones), an anomalous anticyclone (cyclone) center appears at the southeastern corner of the TP. This configuration generally shows the MAS pattern (Fig. 1b). Except for the anomalous cyclone (anticyclone) over the northwestern TP at 850 hPa (Fig. 3c), this equivalent barotropic pattern can also be seen at the middle (Fig. 3b) to lower (Fig. 3c) levels. The correlation between PreI and circulation (Figs. 3d3f) closely resembles Figs. 3ac, which further confirms that surface TPPV forcing is significantly related to EASR. In addition, comparing Figs. 3ac and Figs. 3df reveals that the MAS pattern plays an important role in the connection between surface TPPV forcing and EASR.

      Figure 3.  Spatial distribution of the correlation coefficient between the PVSI and geopotential height anomalies (shading) and circulation anomalies (vector, those passing the 0.05 significance level are shown) at (a) 200 hPa; (b) 500 hPa; and (c) 850 hPa. (d–f) are the same as (a–c), but for PreI. Areas exceeding the 0.05 significance level are highlighted with dots. The blue line denotes the TP topographic boundary of 3000 m.

    4.   Possible mechanism of the surface TPPV forcing affecting the EASR
    • To investigate the possible mechanism of surface TPPV forcing affecting EASR, the LBM is employed in this section to examine the response of upper circulation to the anomalous heating related to surface TPPV forcing.

    • Surface PV is closely related to the static stability within the surface layer. Figure 4a shows the correlation coefficient between PVSI and static stability anomalies. The correlation between the PVSI and horizontal wind anomalies at the surface and vertical motion (omega) anomalies at 500 hPa are shown in Fig. 4b. Corresponding to the positive phase of PVNUM, the anomalous static stability over the whole TP is significantly negative (Fig. 4a). This result means that the air within the surface layer over the TP is anomalously statically unstable. As a result, the anomalous omega is significantly negative, accompanied by anomalous upward motion occurring at the main body of the TP (Fig. 4b). Consequently, a significant anomalous cyclonic circulation is generated at the surface south of the TP (Fig. 4b). The negative phase of PVNUM corresponds to the opposite situation of surface static stability and circulation pattern.

      Figure 4.  Spatial distribution of the correlation coefficient between the PVSI and (a) surface static stability anomalies and (b) horizontal wind anomalies at the surface (vector, those exceeding the 0.05 significance test are shown) and ω at 500 hPa (shading). Areas exceeding the 0.05 significance level are highlighted with dots. The blue line denotes the TP topographic boundary of 3000 m.

      The anomalous south and north winds associated with the anomalous circulation to the south of the TP (Fig. 4b) induced by surface TPPV forcing could lead to anomalous heating over the TP. Figure 5 shows the correlation coefficient between PVSI and OLR (Fig. 5a) and regressed rainfall anomalies against PVSI (Fig. 5b) to examine the distribution of anomalous heating. A dipole mode of the OLR anomalies is clearly observed in Fig. 5a. Corresponding to the positive phase of PVNUM, positive OLR anomalies cover the southwestern TP to northern India. The highest negative OLR anomalies mainly occur on the eastern TP, whereas the highest positive OLR anomalies mainly occur on the western TP. The anomalous cyclonic circulation (Fig. 4b), with cool and divergent flow to its western part and wet, warm flow to its eastern part, accounts for this clear dipole mode of the OLR (Fig. 5a). Consistent with the OLR, the rainfall anomalies also show a dipole mode (Fig. 5b) with less rainfall over the southwestern TP and northern India and more rainfall over the eastern TP. The rainfall anomalies indicate that the strongest negative and positive condensational heating anomalies occur at the southwestern TP to northern India and the eastern TP, respectively. Corresponding to the negative phase of PVNUM, the situation is opposite.

      Figure 5.  Spatial distribution of the (a) correlation coefficient between the PVSI and OLR anomalies and (b) regressed rainfall anomalies against the PVSI (units: mm month−1). Areas exceeding the 0.05 significance level are highlighted with dots. The blue line denotes the TP topographic boundary of 3000 m.

    • The response of upper-level circulation to the negative heating anomaly over northern India has been examined in Wei et al. (2014). Because the lower latitude of the negative heating anomaly over northern India (Fig. 5b) is far away from the westerly jet (light green shading in Fig. 7) at high latitudes in the north, the associated atmospheric response regarding the negative heating anomaly over northern India is trapped south of 40°N (Wei et al., 2014; see their Fig. 8), exhibiting a zonal dipole mode with an anomalous cyclone to the west of 80°E and an anomalous anticyclone to the east of 80°E. There are some differences between the circulation caused by the negative heating anomaly over northern India (Wei et al., 2014; see their Fig. 8) and the MAS pattern (Fig. 3) related to EASR. Hence, regarding the negative heating anomalies centered on the southwestern TP and northern India, we focus on the southwestern TP in this study.

      Figure 6.  Specified ideal diabatic heating (units: K d−1) horizontal distribution in (a) Expt1; (c) Expt2; and (e) Expt3. Specified ideal diabatic heating (units: K d−1) profiles in (b) Expt1 and (d) Expt2. The blue line denotes the TP topographic boundary of 3000 m.

      Figure 7.  Responses of geopotential height (shading, units: gpm) and horizontal wind (vector, units: m s−1) at 200 hPa in (a) Expt1; (b) Expt2; and (c) Expt3. The blue line denotes the TP topographic boundary of 3000 m. The light green shading indicates a westerly jet with a zonal wind speed greater than 20 m s−1 at 200 hPa. The “A” and “C” in Fig. 7c indicate the anomalous anticyclone and cyclone, respectively.

      Figure 8.  Spatial distribution of the correlation coefficient between the PVSI and the divergence of WVF anomalies (shading) and WVF anomalies (vector, those passing the 0.05 significance level are shown) at (a) 500 hPa and (b) 850 hPa. Areas exceeding the 0.05 significance level are highlighted with dots. The blue line denotes the TP topographic boundary of 3000 m.

      Based on the diagnostic analysis of heating anomalies induced by surface TPPV forcing, three groups of experiments are conducted. Expt1 is an experiment of positive heating over the southeastern TP (Expt1, Figs. 6a–b). Expt2 is an experiment of negative heating over the southwestern TP (Expt2, Figs. 6cd). Expt3 is an experiment of combined positive and negative heating over the southeastern and southwestern TP, respectively (Expt3, Fig. 6e). According to the reanalysis, the heating centers are at 30°N, 99°E and 32°N, 75°E in Expt1 and Expt2, respectively. The peaks of the ideal heating profiles in Expt1 and Expt2 are 1.0 K d−1 and –1.0 K d−1, respectively. The peaks in the heating profile are at σ= 0.6 and σ= 0.4 in Expt1 and Expt2, respectively. Since the LBM is a linear model, the heating configuration (Fig. 6e) and atmospheric response (Fig. 7c) in Expt3 are linear combinations of Expt1 and Expt2.

      The atmospheric responses are shown in Fig. 7. In Expt1 (Fig. 7a), the positive heating anomaly over the eastern TP triggers a local anomalous anticyclone and high height at 200 hPa. The anomalous high is strengthened locally with slight northward and westward dispersion. In Expt2 (Fig. 7b), the negative heating anomaly over the southwestern TP stimulates a local anomalous cyclone and low height at 200 hPa. Compared with Wei et al. (2014), because the location of the negative heating center over the southwestern TP is farther north than that over northern India, the anomalous cyclone could trigger waves in the westerly jet and lead to an anomalous cyclone downstream over northeastern China. In Expt3 (Fig. 7c), the response of atmospheric circulation with the “CCA” pattern is very similar to the MAS pattern related to EASR (Figs. 3df). These results indicate that the dipole heating mode induced by surface TPPV forcing can lead to a realistic MAS pattern, which is important for the formation of the anomalous EASR.

    • The above results indicate that surface TPPV forcing can lead to the formation of the MAS pattern. In this section, we examine how the MAS pattern affects the EASR anomalies.

      Rainfall anomalies are related to water vapor transport anomalies and their divergence at the lower level. Since atmospheric moisture is mainly concentrated at the middle to lower levels, Fig. 8 shows the correlation between PVSI and WVF and its divergence at 500 hPa (Fig. 8a) and 850 hPa (Fig. 8b). Generally, the anomalous WVF (Fig. 8) shows an equivalent barotropic structure. This equivalent barotropic structure markedly resembles the MAS pattern. Corresponding to the positive phase of PVNUM, the anomalous cyclonic WVF (Fig. 8) over northeastern China embedded in the MAS pattern converges water vapor over South Korea and Japan and, cooperating with the southern anticyclonic WVF, converges the wet, warm southwesterly flow and dry, cold northerly wind to the Yangtze River valley. In addition, an anomalous convergence of WVF (Fig. 8) occurs in part of northern China. Hence, the EASR over YKJ and part of northern China is greater than normal (Fig. 2). Although the WVF transfers water vapor to southern China, the WVF is divergent. Thus, the EASR over southern China is lower than normal (Fig. 2). Corresponding to the negative phase of PVNUM, the situation is opposite. These results indicate that the WVF anomalies related to the MAS pattern are responsible for the EASR anomalies.

    5.   Conclusion and discussion
    • The present study investigates the relationship between the surface PV forcing over the TP and EASR and the associated circulation on the interannual timescale and the possible mechanism. The main conclusions obtained from the results are described as follows.

      The correlation between the time series of the leading mode of the surface PV forcing over the TP (i.e., PVNUM) and EASR and the leading mode of upper-level circulation (i.e., MAS pattern) are as high as 0.56 and 0.74, respectively. Moreover, the circulation related to PVNUM greatly resembles the circulation associated with EASR. These results indicated that the interannual EASR and related upper-level circulation over East Asia are closely linked to the surface PV forcing over the TP, and the MAS pattern plays an important role in the PVNUM affecting EASR. Diagnostic analysis indicates that the positive phase of PVNUM could lead to unstable air within the surface layer over the TP. As a result, anomalous upward motion and cyclonic circulation are generated over the TP. Induced by cyclonic circulation, a dipole heating mode with anomalous cooling over the southwestern TP and anomalous heating over the southeastern TP appeared. Sensitivity experiments prove that the dipole heating mode associated with the surface PV forcing over the TP can trigger the MAS pattern related to EASR anomalies. The MAS pattern converges water vapor to the Yangtze River valley, South Korea, Japan, and part of northern China and diverges water vapor over southern China. Therefore, the EASR over the Yangtze River valley, South Korea, Japan, and part of northern China is greater than normal, and that over southern China is lower. The negative phase of PVNUM is related to the opposite rainfall and circulation pattern. Consequently, the surface PV forcing over the TP exerts a significant influence on EASR by changing the air static stability within the surface layer over the TP and causing the dipole heating mode and the subsequent anomalous water vapor transport related to the MAS pattern.

      The aforementioned major mechanism is briefly shown schematically in Fig. 9. In summary, PVNUM triggers lower-level cyclonic circulation by reducing the surface static stablity (Fig. 9b). The dipole heaing mode induced by this cyclonic circulation could lead to anomalous circulation at the upper level (Fig. 9a). As a result, the anomalous water vapor transport (Fig. 8) embedded in the upper-level anomalous circulation is responsible for EASR anomalies (Fig. 9c).

      Figure 9.  Schematic showing the interannual surface TPPV influence on EASR. (a) Anomalous 200-hPa wind associated with surface TPPV forcing; (b) anomalous surface TPPV; and (c) same as (a) but for rainfall anomalies. The red (blue) circle indicates diabatic heating (cooling). The green (red) vector indicates the lower-level wind (upper-level Rossby wave propagation).

      Previous studies (e.g., Wu, 2002) show that the MAS pattern is related to the Indian summer monsoon. We calculated the concurrent correlation coefficient between the MASI and Indian summer monsoon index (defined by Wang et al., 2001) and found that the correlation coefficient reaches –0.45, passing the 0.05 significance test; however, when the index of the surface TPPV forcing (i.e., PVSI, red line in Fig. 1d) is removed, it is reduced to –0.28, which does not pass the significance test. This result further confirms the important role of surface TPPV forcing on EASR. ENSO is the most prominent interannual signal in climate systems. We also calculated the correlation between the summer PVSI and NINO3.4 index (SST averaged over the region 5°S–5°N, 170°–120°W; https://psl.noaa.gov/data/climateindices/list/) in the preceding winter and concurrent summer. The coefficients are 0.21 and –0.18, respectively, which do not pass the significance test, meaning that the relationship between surface TPPV forcing and EASR is unaffected by ENSO. It should be noted that some studies (e.g., Guan et al., 2015) suggested that North Atlantic Oscillation, Pacific Decadal Oscillation, and Atlantic Multidecadal Oscillation have great impact on surface air temperature, which is related to surface PV. The role of these decadal oscillations in the interdecadal variation of surface TPPV will be investigated in the future.

      The present study only focuses on the impacts of the TP on monsoon rainfall over East Asia. The complicated Asian summer monsoon system, greatly affected by the TP, includes the South Asian summer monsoon and the East Asian summer monsoon. Research on the relationship between the PV anomaly over the TP and monsoon rainfall, as well as wind fields over the Asian region will be conducted in the future. In this study, although the simple model (e.g., LBM) clearly shows the impact of surface TPPV forcing on EASR, it is still necessary to use a fully coupled model to examine the surface TPPV–rainfall–circulation feedback to understand the interactions between the surface TP forcing and EASM system. Furthermore, with the help of the PV budget equation, the relative importance of dynamic PV advection, PV generation due to diabatic heating, and friction effects to the variation of surface PV will be quantificationally answered, which will greatly advance our understanding of the TP’s impacts.

      Acknowledgements. We thank the reviewers for their constructive suggestions and comments. This work is jointly supported by the National Natural Science Foundation of China (Grant Nos. 91837101, 42122035, and 91937302) and the National Key Research and Development Program of China (Grant No. 2018YFC1505706 and 2020YFA0608903).

      Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return