Advanced Search
Article Contents

Interdecadal Variations of the March Atmospheric Heat Source over the Southeast Asian Low-Latitude Highlands


doi: 10.1007/s00376-023-2146-2

  • Based on the fifth-generation reanalysis dataset from the European Centre for Medium-Range Weather Forecasts for 1979–2019, we investigated the effects of the circumglobal teleconnection (CGT) on the interdecadal variation of the March atmospheric heat source (AHS) over the Southeast Asian low-latitude highlands (SEALLH). The dominant mode of the March AHS over the SEALLH features a monopole structure with an 8–11-year period. Decadal variations in the AHS make an important contribution to the 11-year low-pass filtered component of the AHS index, whichexplains 54.3% of the total variance. The CGT shows a clear interdecadal variation, which explains 59.3% of the total variance. The March AHS over the SEALLH is significantly related to the CGT on interdecadal timescales. When the CGT is optimally excited by a significant cyclonic vorticity source near northern Africa (i.e., in its positive phase), the SEALLH is dominated by anomalous southerly winds and ascending motions on the east of the anomalous cyclone. The enhanced advection and upward transfer result in a high-enthalpy air mass that converges into and condenses over the SEALLH, leading to a larger-than-average March AHS over this region. The key physical processes revealed by this diagnostic analysis are supported by numerical experiments.
    摘要: 本文利用1979–2019年欧洲中期天气预报中心提供的ERA5资料,研究了初春环球遥相关(CGT)对东南亚低纬高原(SEALLH)大气热源年代际变化的影响。研究发现初春SEALLH 大气热源显著的8–11年变化主要呈现全场一致型的分布。11年低通滤波的初春大气热源指数年代际分量的方差贡献率高达54.3%。11年低通滤波的初春CGT的方差贡献率为59.3%,也呈现显著的年代际变化。在年代际时间尺度上,初春SEALLH 大气热源与CGT存在显著的相关关系。当北非出现异常气旋式罗斯贝波源时,在北非至东亚上空激发出异常“反气旋、气旋、反气旋、气旋、反气旋”的CGT波列。异常的偏南风使得高湿焓气团输送到SEALLH并发生异常辐合上升,进而导致初春SEALLH 大气热源偏强。数值模拟结果验证了观测事实所揭示的关键物理过程。
  • 加载中
  • Figure 1.  Location and terrain (units: m) of the SEALLH (18°–30°N, 96°–108°E). Rivers are indicated by the thick blue lines.

    Figure 2.  (a) Annual mean AHS (units: W m–2) and (b) long-term monthly means of the AHS (units: W m–2) averaged for the SEALLH and TP, along with their standard deviations. In (a), the region bounded by the green (red) solid line indicates the SEALLH (TP).

    Figure 3.  (a) First mode of the EOF analysis for the March AHS (units: W m–2) of the SEALLH region (18°–30°N, 96°–108°E) for 1979–2019. (b) Time series of the first EOF mode (PC1; thin line) and its interdecadal component (thick line with in-filled colors). (c) Explained variances (units: %) of the first five EOF modes, the standard error bars result from the North et al. (1982) rule; (d) Power spectral density of PC1. In (a), the thick green line denotes the domain of the SEALLH. In (d), the red line shows the 95% confidence level of the corresponding red spectrum.

    Figure 4.  Same as in Fig. 3, but for the 300-hPa non-divergent meridional wind (units: m s–1) of the region (0°–45°N, 0°–120°E).

    Figure 5.  Composite differences of the March (a) 200-hPa horizontal winds (vectors; units: m s–1), and (b) 15°–45°N averaged relative vorticity (contour intervals: 10–6 s–1) between the positive phase of the CGT-ID and negative phase of the CGT-ID. In (a), the solid green line indicates the SEALLH. Light-to-dark shading indicates the 90%, 95%, and 99% confidence levels based on a two-tailed Student’s t-test.

    Figure 6.  Regression of the (a) AHS (contour intervals: 20 W m–2), (b) column-integrated local tendency (contour intervals: 0.4 W m–2), (c) column-integrated advection (contour intervals: 10 W m–2), and (d) column-integrated vertical transfer (contour intervals: 5 W m–2) onto the normalized CGT-ID in March during the period 1979–2019. Light-to-dark shading indicates the 90%, 95%, and 99% confidence levels based on a two-tailed Student’s t-test. The solid green line indicates the SEALLH.

    Figure 7.  Regression of the (a) 700-hPa and (b) 200-hPa horizontal winds (vectors; units: m s−1), (c) vertical velocity (contour intervals: 4×10–3 Pa s–1) along line ABCDEFG shown in (b), and (d) column-integrated moist enthalpy flux (vectors; units: J m–1 s–1) and its divergence (contour intervals: 2×102 J m–2 s–1) onto the normalized CGT-ID in March during the period 1979–2019. In (b), the letters A (40°N, 30°W), B (35°N, 5°E), C (25°N, 20°E), D (15°N, 55°E), E (15°N, 90°E), F (23°N, 100°E) and G (26°N, 125°E) stand for the seven nodes on the CGT path (blue line). In (d), vectors exceeding the 95% confidence level are shown, and zero contours are omitted. Light-to-dark shading indicates the 90%, 95%, and 99% confidence levels based on the two-tailed Student’s t-test. The solid green line indicates the SEALLH.

    Figure 8.  Regression of the 200-hPa (a) RWS (contour intervals: 2.5 × 10–11 s–2) and (b) wave-activity flux (m2 s−2) onto the normalized CGT-ID in March during 1979–2019. (c) Time series of the RWS index (thin line) and its interdecadal component (thick line with infilled colors). (d) Power spectral density of the RWS index. In (a), the light-to-dark shading indicates the 90%, 95%, and 99% confidence levels based on the two-tailed Student’s t-test. In (a) and (b), the solid purple line indicates the 40 m s–1 contour of climatological zonal wind. The solid green line indicates the SEALLH. In (d), the red line shows the 95% confidence level of the corresponding red spectrum.

    Figure 9.  Composite differences of the March (a) 200-hPa horizontal winds (vectors; units: m s–1), (b) 15°–45°N averaged relative vorticity (contour intervals: 10–6 s–1), and (c) column-integrated moist enthalpy flux (vectors; units: J m–1 s–1) and its divergence (contour intervals: 2×102 J m–2 s–1) between SE1 and SE2. In (a), the solid green line indicates the SEALLH. Light-to-dark shading indicates the 90%, 95%, and 99% confidence levels based on a two-tailed Student’s t-test.

    Figure 10.  A schematic diagram of the effects of the CGT on atmospheric heat sources in the SEALLH. The letters A and C denote an anomalous anticyclone and cyclone, respectively.

  • Branstator, G., 2002: Circumglobal teleconnections, the jet stream waveguide, and the North Atlantic oscillation. J. Climate, 15, 1893−1910, https://doi.org/10.1175/1520-0442(2002)015<1893:CTTJSW>2.0.CO;2.
    Bretherton, C. S., M. Widmann, V. P. Dymnikov, J. M. Wallace, and I. Bladé, 1999: The effective number of spatial degrees of freedom of a time-varying field. J. Climate, 12, 1990−2009, https://doi.org/10.1175/1520-0442(1999)012<1990:TENOSD>2.0.CO;2.
    Cao, J., J. M. Hu, and Y. Tao, 2012: An index for the interface between the Indian summer monsoon and the East Asian summer monsoon. J. Geophys. Res.: Atmos., 117, D18108, https://doi.org/10.1029/2012JD017841.
    Cao, J., S. Gui, Q. Su, and Y. L. Yang, 2016: The variability of the Indian-East Asian summer monsoon interface in relation to the spring seesaw mode between the Indian Ocean and the central-western Pacific. J. Climate, 29, 5027−5040, https://doi.org/10.1175/JCLI-D-15-0839.1.
    Chang, C. P., Z. Wang, J. McBride, and C. H. Liu, 2005: Annual cycle of Southeast Asia-Maritime continent rainfall and the asymmetric monsoon transition. J. Climate, 18, 287−301, https://doi.org/10.1175/JCLI-3257.1.
    Chen, G. S., and R. H. Huang, 2012: Excitation mechanisms of the teleconnection patterns affecting the July precipitation in northwest China. J. Climate, 25, 7834−7851, https://doi.org/10.1175/JCLI-D-11-00684.1.
    Chow, K. C., Y. M. Liu, J. C. L. Chan, and Y. H. Ding, 2006: Effects of surface heating over Indochina and India landmasses on the summer monsoon over south China. International Journal of Climatology, 26, 1339−1359, https://doi.org/10.1002/joc.1310.
    Enomoto, T., 2004: Interannual variability of the Bonin high associated with the propagation of Rossby waves along the Asian jet. J. Meteor. Soc. Japan, 82, 1019−1034, https://doi.org/10.2151/jmsj.2004.1019.
    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.
    Giorgetta, M. A., and Coauthors, 2013: Climate and carbon cycle changes from 1850 to 2100 in MPI-ESM simulations for the coupled model intercomparison project phase 5. Journal of Advances in Modeling Earth Systems, 5, 572−597, https://doi.org/10.1002/jame.20038.
    Hersbach, H., and Coauthors, 2019: Global reanalysis: Goodbye ERA-Interim, hello ERA5. ECMWF Newsletter, 159, 17−24, https://doi.org/10.21957/vf291hehd7.
    Hoffmann, L., and Coauthors, 2019: From ERA-Interim to ERA5: The considerable impact of ECMWF's next-generation reanalysis on Lagrangian transport simulations. Atmospheric Chemistry and Physics, 19, 3097−3124, https://doi.org/10.5194/acp-19-3097-2019.
    Holmes, J. A., E. R. Cook, and B. Yang, 2009: Climate change over the past 2000 years in Western China. Quaternary International, 194, 91−107, https://doi.org/10.1016/j.quaint.2007.10.013.
    Hoskins, B. J., and T. Ambrizzi, 1993: Rossby wave propagation on a realistic longitudinally varying flow. J. Atmos. Sci., 50, 1661−1671, https://doi.org/10.1175/1520-0469(1993)050<1661:RWPOAR>2.0.CO;2.
    Hsu, H. H., and S. H. Lin, 1992: Global teleconnections in the 250-mb streamfunction field during the northern hemisphere winter. Mon. Wea. Rev., 120, 1169−1190, https://doi.org/10.1175/1520-0493(1992)120<1169:GTITMS>2.0.CO;2.
    Huang, G., Y. Liu, and R. H. Huang, 2011: The interannual variability of summer rainfall in the arid and semiarid regions of northern China and its association with the Northern Hemisphere circumglobal teleconnection. Adv. Atmos. Sci., 28, 257−268, https://doi.org/10.1007/s00376-010-9225-x.
    Lee, M. H., S. Lee, H. J. Song, and C. H. Ho, 2017: The recent increase in the occurrence of a boreal summer teleconnection and its relationship with temperature extremes. J. Climate, 30, 7493−7504, https://doi.org/10.1175/JCLI-D-16-0094.1.
    Li, C. F., and M. Yanai, 1996: The onset and interannual variability of the Asian summer monsoon in relation to land-sea thermal contrast. J. Climate, 9, 358−375, https://doi.org/10.1175/1520-0442(1996)009<0358:TOAIVO>2.0.CO;2.
    Li, Y. N., S. Yang, Y. Deng, and B. Zheng, 2020: Signals of spring thermal contrast related to the interannual variations in the onset of the South China Sea summer monsoon. J. Climate, 33, 27−38, https://doi.org/10.1175/JCLI-D-19-0174.1.
    Liu, X. F., Q. Li, J. H. He, and P. Wang, 2010: Effects of the thermal contrast between Indo-China Peninsula and South China Sea on SCS monsoon onset. Acta Meteorologica Sinica, 67, 100−107, https://doi.org/10.11676/qxxb2009.011. (in Chinese with English abstract
    Mantua, N. J., S. R. Hare, Y. Zhang, J. M. Wallace, and R. C. Francis, 1997: A Pacific interdecadal climate oscillation with impacts on salmon production. Bull. Amer. Meteor. Soc., 78, 1069−1079, https://doi.org/10.1175/1520-0477(1997)078,1069:APICOW.2.0.CO;2.
    Meng, L., D. Long, S. M. Quiring, and Y. J. Shen, 2014: Statistical analysis of the relationship between spring soil moisture and summer precipitation in East China. International Journal of Climatology, 34, 1511−1523, https://doi.org/10.1002/joc.3780.
    Mo, K., and E. M. Rasmusson, 1993: The 200-mb climatological vorticity budget during 1986-1989 as revealed by NMC analyses. J. Climate, 6, 577−594, https://doi.org/10.1175/1520-0442(1993)006<0577:TMCVBD>2.0.CO;2.
    Neelin, J. D., and H. Su, 2005: Moist teleconnection mechanisms for the tropical South American and Atlantic sector. J. Climate, 18, 3928−3950, https://doi.org/10.1175/JCLI3517.1.
    Neelin, J. D., I. M. Held, and K. H. Cook, 1987: Evaporation-wind feedback and low-frequency variability in the tropical atmosphere. J. Atmos. Sci., 44, 2341−2348, https://doi.org/10.1175/1520-0469(1987)044<2341:EWFALF>2.0.CO;2.
    North, G. R., T. L. Bell, R. F. Cahalan, and F. J. Moeng, 1982: Sampling errors in the estimation of empirical orthogonal functions. Mon. Wea. Rev., 110, 699−706, https://doi.org/10.1175/1520-0493(1982)110<0699:SEITEO>2.0.CO;2.
    Qin, J., and J. H. Ju, 1997: Weather and Climate in Low Latitudes Plateau. China Meteorology Press, 210 pp. (in Chinese)
    Sardeshmukh, P. D., and B. J. Hoskins, 1985: Vorticity balances in the tropics during the 1982-83 El Niño-Southern oscillation event. Quart. J. Roy. Meteor. Soc., 111, 261−278, https://doi.org/10.1256/smsqj.46801.
    Sardeshmukh, P. D., and B. J. Hoskins, 1988: The generation of global rotational flow by steady idealized tropical divergence. J. Atmos. Sci., 45, 1228−1251, https://doi.org/10.1175/1520-0469(1988)045<1228:TGOGRF>2.0.CO;2.
    Schlesinger, M. E., and N. Ramankutty, 1994: An oscillation in the global climate system of period 65-70 years. Nature, 367, 723−726, https://doi.org/10.1038/367723a0.
    Son, J. H., K. H. Seo, and B. Wang, 2019: Dynamical control of the Tibetan Plateau on the East Asian summer monsoon. Geophys. Res. Lett., 46, 7672−7679, https://doi.org/10.1029/2019GL083104.
    Son, J. H., K. H. Seo, and B. Wang, 2020: How does the Tibetan Plateau dynamically affect downstream monsoon precipitation? Geophys. Res. Lett., 47, e2020GL090543, https://doi.org/10.1029/2020GL090543.
    Stevens, B., and Coauthors, 2013: Atmospheric component of the MPI-M Earth system model: ECHAM6. Journal of Advances in Modeling Earth Systems, 5, 146−172, https://doi.org/10.1002/jame.20015.
    Takaya, K., and H. Nakamura, 1997: A formulation of a wave-activity flux for stationary Rossby waves on a zonally varying basic flow. Geophys. Res. Lett., 24, 2985−2988, https://doi.org/10.1029/97GL03094.
    Takaya, K., and H. Nakamura, 2001: A formulation of a phase-independent wave-activity flux for stationary and migratory quasigeostrophic eddies on a zonally varying basic flow. J. Atmos. Sci., 58, 608−627, https://doi.org/10.1175/1520-0469(2001)058<0608:AFOAPI>2.0.CO;2.
    Tao, Y., J. Cao, G. D. Lan, and Q. Su, 2016: The zonal movement of the Indian-East Asian summer monsoon interface in relation to the land-sea thermal contrast anomaly over East Asia. Climate Dyn., 46, 2759−2771, https://doi.org/10.1007/s00382-015-2729-4.
    Wang, H., B. Wang, F. Huang, Q. G. Ding, and J. Y. Lee, 2012: Interdecadal change of the boreal summer circumglobal teleconnection (1958−2010). Geophys. Res. Lett., 39, L12704, https://doi.org/10.1029/2012GL052371.
    Wang, Y. M., B. Wu, and T. J. Zhou, 2022: Maintenance of western north Pacific anomalous anticyclone in boreal summer by wind-induced moist enthalpy advection mechanism. J. Climate, 35, 4499−4511, https://doi.org/10.1175/JCLI-D-21-0708.1.
    Watanabe, M., 2004: Asian jet waveguide and a downstream extension of the North Atlantic Oscillation. J. Climate, 17, 4674−4691, https://doi.org/10.1175/JCLI-3228.1.
    Wu, B., J. S. Lin, and T. J. Zhou, 2016a: Interdecadal circumglobal teleconnection pattern during boreal summer. Atmos. Sci. Lett., 17, 446−452, https://doi.org/10.1002/asl.677.
    Wu, B., T. J. Zhou, and T. M. Li, 2016b: Impacts of the Pacific-Japan and circumglobal teleconnection patterns on the interdecadal variability of the East Asian summer monsoon. J. Climate, 29, 3253−3271, https://doi.org/10.1175/JCLI-D-15-0105.1.
    Wu, B., T. J. Zhou, C. Li, W. A. Müller, and J. S. Lin, 2019: Improved decadal prediction of Northern-Hemisphere summer land temperature. Climate Dyn., 53, 1357−1369, https://doi.org/10.1007/s00382-019-04658-8.
    Xie, M. E., and Y. Liu, 1998: Climatic features primary study of global low latitude plateau regions. Yunnan Geographic Environment Research, 10, 25−33. (in Chinese with English abstract)
    Xie, S. P., H. M. Xu, N. H. Saji, Y. Q. Wang, and W. T. Liu, 2006: Role of narrow mountains in large-scale organization of Asian monsoon convection. J. Climate, 19, 3420−3429, https://doi.org/10.1175/JCLI3777.1.
    Xu, H. M., J. H. He, M. Wen, and M. Dong, 2002: A numerical study of effects of the Indo-China Peninsula on the establishment and maintenance of the South China Sea summer monsoon. Chinese Journal of Atmospheric Sciences, 26, 330−342, https://doi.org/10.3878/j.issn.1006-9895.2002.03.04. (in Chinese with English abstract
    Yanai, M., and C. F. Li, 1994: Mechanism of heating and the boundary layer over the Tibetan Plateau. Mon. Wea. Rev., 122, 305−323, https://doi.org/10.1175/1520-0493(1994)122<0305:MOHATB>2.0.CO;2.
    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. Q., H. S. Chen, Y. D. Song, S. G. Zhu, B. T. Zhou, and J. Zhang, 2021: Atmospheric circumglobal teleconnection triggered by spring land thermal anomalies over West Asia and its possible impacts on early summer climate over northern China. J. Climate, 34, 5999−6021, https://doi.org/10.1175/JCLI-D-20-0911.1.
    Yasui, S., and M. Watanabe, 2010: Forcing processes of the summertime circumglobal teleconnection pattern in a dry AGCM. J. Climate, 23, 2093−2114, https://doi.org/10.1175/2009JCLI3323.1.
    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. (in Chinese with English abstract)
    Yuan, J. C., W. H. Li, and Y. Deng, 2015: Amplified subtropical stationary waves in boreal summer and their implications for regional water extremes. Environmental Research Letters, 10, 104009, https://doi.org/10.1088/1748-9326/10/10/104009.
    Zhu, Z. W., and T. M. Li, 2017: Empirical prediction of the onset dates of South China Sea summer monsoon. Climate Dyn., 48, 1633−1645, https://doi.org/10.1007/s00382-016-3164-x.
    Zhuang, M. R., A. M. Duan, R. Y. Lu, P. X. Li, and J. L. Yao, 2022: Relative impacts of the orography and land–sea contrast over the Indochina peninsula on the Asian Summer Monsoon between early and late summer. J. Climate, 35, 3037−3055, https://doi.org/10.1175/JCLI-D-21-0576.1.
  • [1] Guoxiong WU, Bian HE, Anmin DUAN, Yimin LIU, Wei YU, 2017: Formation and Variation of the Atmospheric Heat Source over the Tibetan Plateau and Its Climate Effects, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 1169-1184.  doi: 10.1007/s00376-017-7014-5
    [2] Yizhe HAN, Dabang JIANG, Dong SI, Yaoming MA, Weiqiang MA, 2024: Time-lagged Effects of the Spring Atmospheric Heat Source over the Tibetan Plateau on Summer Precipitation in Northeast China during 1961–2020: Role of Soil Moisture, ADVANCES IN ATMOSPHERIC SCIENCES.  doi: 10.1007/s00376-023-2363-8
    [3] HUANG Gang, LIU Yong, HUANG Ronghui, 2011: The Interannual Variability of Summer Rainfall in the Arid and Semiarid Regions of Northern China and Its Association with the Northern Hemisphere Circumglobal Teleconnection, ADVANCES IN ATMOSPHERIC SCIENCES, 28, 257-268.  doi: 10.1007/s00376-010-9225-x
    [4] LIU Jing, ZHAI Panmao, 2014: Changes in Climate Regionalization Indices in China during 1961-2010, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 374-384.  doi: 10.1007/s00376-013-3017-z
    [5] JU Jianhua, Lü Junmei, CAO Jie, REN Juzhang, 2005: Possible Impacts of the Arctic Oscillation on the Interdecadal Variation of Summer Monsoon Rainfall in East Asia, ADVANCES IN ATMOSPHERIC SCIENCES, 22, 39-48.  doi: 10.1007/BF02930868
    [6] BIAN Jianchun, YANG Peicai, 2005: Interdecadal Variations of Phase Delays Between Two Ni(n)o Indices at Different Time Scales, ADVANCES IN ATMOSPHERIC SCIENCES, 22, 122-125.  doi: 10.1007/BF02930875
    [7] Xiaowei HONG, Riyu LU, Shuanglin LI, 2018: Asymmetric Relationship between the Meridional Displacement of the Asian Westerly Jet and the Silk Road Pattern, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 389-397.  doi: 10.1007/s00376-017-6320-2
    [8] Ja-Yeon MOON, Kyung-Ja HA, 2003: Association between Tropical Convection and Boreal Wintertime Extratropical Circulation in 1982/83 and 1988/89, ADVANCES IN ATMOSPHERIC SCIENCES, 20, 593-603.  doi: 10.1007/BF02915502
    [9] Leying ZHANG, Haiming XU, Ning SHI, Jiechun DENG, 2017: Responses of the East Asian Jet Stream to the North Pacific Subtropical Front in Spring, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 144-156.  doi: 10.1007/s00376-016-6026-x
    [10] ZHU Yimin, YANG Xiuqun, 2003: Joint Propagating Patterns of SST and SLP Anomalies in the North Pacific on Bidecadal and Pentadecadal Timescales, ADVANCES IN ATMOSPHERIC SCIENCES, 20, 694-710.  doi: 10.1007/BF02915396
    [11] WANG Huijun, SUN Jianqi, 2009: Variability of Northeast China River Break-up Date, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 701-706.  doi: 10.1007/s00376-009-9035-1
    [12] LI Lun, ZHANG Renhe, WEN Min, 2011: Diagnostic Analysis of the Evolution Mechanism for a Vortex over the Tibetan Plateau in June 2008, ADVANCES IN ATMOSPHERIC SCIENCES, 28, 797-808.  doi: 10.1007/s00376-010-0027-y
    [13] WU Bingyi, YANG Kun, ZHANG Renhe, 2009: Eurasian Snow Cover Variability and Its Association with Summer Rainfall in China, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 31-44.  doi: 10.1007/s00376-009-0031-2
    [14] WANG Lin, CHEN Wen, 2010: How Well do Existing Indices Measure the Strength of the East Asian Winter Monsoon?, ADVANCES IN ATMOSPHERIC SCIENCES, 27, 855-870.  doi: 10.1007/s00376-009-9094-3
    [15] Chuandong ZHU, Rongcai REN, Guoxiong WU, 2018: Varying Rossby Wave Trains from the Developing to Decaying Period of the Upper Atmospheric Heat Source over the Tibetan Plateau in Boreal Summer, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 1114-1128.  doi: 10.1007/s00376-017-7231-y
    [16] Lu Keli, Zhu Yongchun, 1994: Seasonal Variation of Stationary and Low-Frequency Rossby Wave Rays, ADVANCES IN ATMOSPHERIC SCIENCES, 11, 427-435.  doi: 10.1007/BF02658163
    [17] WEI Na, GONG Yuanfa, HE Jinhai, 2009: Structural Variation of Atmospheric Heat Source over the Qinghai-Xizang Plateau and its Influence on Precipitation in Northwest China the Qinghai-Xizang Plateau and Its Influence on Precipitation in Northwest China, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 1027-1041.  doi: 10.1007/s00376-009-7207-7
    [18] Fang Juan, Wu Rongsheng, 1998: Influences of Vorticity Source and Momentum Source on Atmospheric Circulation, ADVANCES IN ATMOSPHERIC SCIENCES, 15, 41-46.  doi: 10.1007/s00376-998-0016-6
    [19] Xu Xiangde, 1991: The Effect of Spatial Structure Character of Heat Source on the Ray Path and the Evolution of Wave Energy of Meridional Wave Train, ADVANCES IN ATMOSPHERIC SCIENCES, 8, 87-98.  doi: 10.1007/BF02657367
    [20] Zhao Ping, Chen Longxun, 2001: Interannual Variability of Atmospheric Heat Source/Sink over the Qinghai-Xizang (Tibetan) Plateau and its Relation to Circulation, ADVANCES IN ATMOSPHERIC SCIENCES, 18, 106-116.  doi: 10.1007/s00376-001-0007-3

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 21 May 2022
Manuscript revised: 27 January 2023
Manuscript accepted: 22 February 2023
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

Interdecadal Variations of the March Atmospheric Heat Source over the Southeast Asian Low-Latitude Highlands

    Corresponding author: Jie CAO, caoj@ynu.edu.cn
  • Yunnan Key Laboratory of Meteorological Disasters and Climate Resources in the Greater Mekong Subregion, Yunnan University, Kunming 650091, China

Abstract: Based on the fifth-generation reanalysis dataset from the European Centre for Medium-Range Weather Forecasts for 1979–2019, we investigated the effects of the circumglobal teleconnection (CGT) on the interdecadal variation of the March atmospheric heat source (AHS) over the Southeast Asian low-latitude highlands (SEALLH). The dominant mode of the March AHS over the SEALLH features a monopole structure with an 8–11-year period. Decadal variations in the AHS make an important contribution to the 11-year low-pass filtered component of the AHS index, whichexplains 54.3% of the total variance. The CGT shows a clear interdecadal variation, which explains 59.3% of the total variance. The March AHS over the SEALLH is significantly related to the CGT on interdecadal timescales. When the CGT is optimally excited by a significant cyclonic vorticity source near northern Africa (i.e., in its positive phase), the SEALLH is dominated by anomalous southerly winds and ascending motions on the east of the anomalous cyclone. The enhanced advection and upward transfer result in a high-enthalpy air mass that converges into and condenses over the SEALLH, leading to a larger-than-average March AHS over this region. The key physical processes revealed by this diagnostic analysis are supported by numerical experiments.

摘要: 本文利用1979–2019年欧洲中期天气预报中心提供的ERA5资料,研究了初春环球遥相关(CGT)对东南亚低纬高原(SEALLH)大气热源年代际变化的影响。研究发现初春SEALLH 大气热源显著的8–11年变化主要呈现全场一致型的分布。11年低通滤波的初春大气热源指数年代际分量的方差贡献率高达54.3%。11年低通滤波的初春CGT的方差贡献率为59.3%,也呈现显著的年代际变化。在年代际时间尺度上,初春SEALLH 大气热源与CGT存在显著的相关关系。当北非出现异常气旋式罗斯贝波源时,在北非至东亚上空激发出异常“反气旋、气旋、反气旋、气旋、反气旋”的CGT波列。异常的偏南风使得高湿焓气团输送到SEALLH并发生异常辐合上升,进而导致初春SEALLH 大气热源偏强。数值模拟结果验证了观测事实所揭示的关键物理过程。

    • Low-latitude highlands are located between 30°S and 30°N and have an average elevation of at least 1000 m. Ten low-latitude highlands, including the Southeast Asian low-latitude highlands (SEALLH; 18°–30°N, 96°–108°E), meet these criteria worldwide (Qin and Ju, 1997; Xie and Liu, 1998). The SEALLH mainly includes the Yun–Gui Plateau and its southern extension in northeastern Myanmar, northern Laos, and northern Vietnam (Fig. 1). The SEALLH has its highest and lowest elevations in its northwestern and southeastern parts, respectively. The SEALLH and its meridional mountains and rivers cover an area of approximately 9 × 105 km2, about one-third of the size of the Tibetan Plateau (TP). Its complex landforms and distinctive geographical position result in water vapor from the tropical Indian Ocean and tropical western Pacific mixing over the SEALLH and entering the East Asian monsoon region (Holmes et al., 2009; Cao et al., 2012, 2016; Tao et al., 2016). In this way, at least to some extent, the SEALLH modulates the weather and climate over East Asia (Chang et al., 2005; Xie et al., 2006; Meng et al., 2014).

      Figure 1.  Location and terrain (units: m) of the SEALLH (18°–30°N, 96°–108°E). Rivers are indicated by the thick blue lines.

      In comparison with the numerous studies of the atmospheric heat source (AHS) over the TP and its effects (Son et al., 2019, 2020), since the pioneering studies of Yeh et al. (1957) and Flohn (1957), there have been relatively few studies of the AHS over the SEALLH adjacent to the TP. The SEALLH and southern margin of the TP are characterized by a large average annual heat flux of >200 W m–2, equivalent to that of tropical convection (Fig. 2a). The AHS in the SEALLH exhibits a clear annual cycle, with larger variability from November to April. Given that the SEALLH generally transitions from a heat sink to a heat source in March, one month earlier than that of the TP, it is one of the earliest AHS to develop in the East Asian summer monsoon region (Fig. 2b). Previous studies show that the land–sea thermal contrast between the region south of the SEALLH and its surrounding oceans in spring can be a precursory signal of the South China Sea summer monsoon onset (Xu et al., 2002; Chow et al., 2006; Li et al., 2020; Zhuang et al., 2022). The earlier the atmospheric heat source over the region south of the SEALLH becomes persistently greater than that compared to the South China Sea, the earlier the summer monsoon establishes itself over the South China Sea (Liu et al., 2010; Zhu and Li, 2017). The AHS variability of the SEALLH in its transition month may provide possible theoretical guidance for understanding the establishment of the summer monsoon. However, the variability of the March AHS over the SEALLH and what causes these variations remain unclear.

      Figure 2.  (a) Annual mean AHS (units: W m–2) and (b) long-term monthly means of the AHS (units: W m–2) averaged for the SEALLH and TP, along with their standard deviations. In (a), the region bounded by the green (red) solid line indicates the SEALLH (TP).

      The circumglobal teleconnection (CGT) with a wavenumber five structure in the upper-tropospheric meridional wind and streamfunction fields is a hemispheric-scale, low-frequency pattern that propagates along with the waveguide for Rossby waves in the boreal winter (Hsu and Lin, 1992; Hoskins and Ambrizzi, 1993; Branstator, 2002). The CGT is an important factor that affects the temperature, precipitation, and extreme weather in East Asia (Huang et al., 2011; Yuan et al., 2015; Wu et al., 2016b). Recently, some studies have suggested that the CGT undergoes significant interdecadal variations (Wang et al., 2012; Wu et al., 2016a, b, 2019). Therefore, an important unresolved issue is whether the CGT modulates the March AHS over the SEALLH, along with what key physical processes link the CGT and March AHS. This study was undertaken to address these two issues.

    2.   Data and Methods
    • The datasets used in this analysis include the fifth-generation reanalysis dataset from the European Centre for Medium-Range Weather Forecasts (ERA5; Hersbach et al., 2019; Hoffmann et al., 2019) covering 1979–2019. The variables in the ERA5 dataset include air temperature, specific humidity, u- and v-components of wind, and surface pressure. The horizontal resolution in the EAR5 dataset is 0.25° × 0.25°. The ERA5 data has 37 pressure levels ranging from 1000 to 1 hPa.

    • We adopted the atmospheric general circulation model ECHAM (version 6.3; ECHAM6) to confirm the excitation mechanism of the wave train by the Rossby wave source near northern Africa. The ECHAM6 was developed by the Max Planck Institute for Meteorology in Germany (Giorgetta et al., 2013; Stevens et al., 2013). Here, ECHAM6 was adopted with T63L47 resolution, corresponding to a horizontal resolution of about 1.875° and 47 vertical levels. The time step was set to 450 s following the default settings.

    • The statistical methods used in this paper include empirical orthogonal function (EOF) analysis without any rotation and regression analysis. The interdecadal component of a variable is extracted using a Lanczos low-pass filter with an 11-year cutoff period. The confidence levels of the linear regression and correlation are evaluated with a two-tailed Student’s t-test and using the effective degrees of freedom ($ {N}_{\mathrm{d}\mathrm{o}\mathrm{f}} $), as defined by Bretherton et al. (1999):

      where $ N $ represents the sample size, and $ {r}_{1} $ and $ {r}_{2} $ are the lag-1 autocorrelations of the time series. The trends of all data are removed to avoid possible influences of global warming on the results.

    • The AHS is estimated using the hourly ERA5 data and the thermodynamic equations developed by (Yanai et al., 1992; Yanai and Li, 1994; Li and Yanai, 1996):

      where $ \boldsymbol{V} $ is horizontal velocity (m s–1), $ \theta $ is potential temperature (K), $ \omega $ is vertical velocity at the isobaric surface (Pa s–1), ${\boldsymbol{\nabla}}$ is the isobaric gradient operator, $ t $ is time (s), $ p $ is pressure (hPa), $ {C}_{p} $ (1004 J kg–1 K–1) is the specific heat at a constant pressure of dry air, $ {p}_{0} $ = 1000 hPa, $ \kappa =R/{C}_{p} $, and $ R $ is the gas constant of dry air (287.05 J kg–1 K–1). In Eq. (2), $ \partial \theta /\partial t $, $ \boldsymbol{V}\cdot \nabla \theta $, $ \omega (\partial \theta /\partial p) $ reflect the contributions of the local tendency, the advection, and the vertical transfer to the apparent heat source. The column-integrated apparent heat source is calculated using Eq. (3):

      In Eq. (3), $ {p}_{\mathrm{t}} $ = 100 hPa, $ {p}_{\mathrm{s}} $ is the surface pressure (hPa), and $ g $ is the gravitational acceleration (m s–2).

    • The enthalpy ($ {h}_{m} $; J kg–1) of moist air mass is calculated by the method of Neelin et al. (1987) and Neelin and Su (2005):

      where $ {L}_{v} $ (2.5×106 J kg–1) is the latent heat of vaporization for typical atmospheric temperatures, $ T $ is the ambient temperature (K), and $ q $ is the mixing ratio (kg kg–1), i.e., higher moist enthalpy means higher temperature and humidity.

      The column-integrated moist enthalpy flux is defined as:

      where $ \boldsymbol{V}=(u,v) $ is the horizontal wind (m s–1), $ ⟨\boldsymbol{V}{h}_{m}⟩ $ and ${\boldsymbol{\nabla}} \cdot ⟨\boldsymbol{V}{h}_{m}⟩$ denote the horizontal enthalpy transport (J m–1 s–1) and its divergence (J m–2 s–1), respectively (Wang et al., 2022).

    • The Rossby wave source (RWS; units: s–1) is calculated following Sardeshmukh and Hoskins (1988):

      where $ {\boldsymbol{V}}_{\chi } $ represents the horizontally divergent wind velocity (m s–1) and is defined as the gradient of velocity potential $ \chi $ (m2 s–2). $ \zeta $ is the relative vorticity (s–1), and $ f $ is the Coriolis parameter. The RWS includes contributions from both the vortex stretching term and the advection of absolute vorticity by the divergent part of the flow (Sardeshmukh and Hoskins, 1985; Mo and Rasmusson, 1993).

      The anomalous RWS is computed following Watanabe (2004):

      where the overbar and prime denote the climatological mean and monthly perturbation subtracted from the climatological mean, respectively.

    • The three-dimensional wave-activity flux was derived by Takaya and Nakamura (1997, 2001) based on the Plumb wave flux, making it more suitable for complex background airflow. The background field takes the average monthly climatological field for many years, which can better describe the energy dispersion characteristics of the stationary Rossby wave and the propagation anomaly of the Rossby wave. The three-dimensional expression of the wave-activity flux is as follows:

      Here, $\boldsymbol{U}=(U,V)$ is a horizontal basic flow velocity (m s–1), $ p $ is pressure normalized by 1000 hPa, $ {\psi }^{'} $ denotes the perturbation geostrophic streamfunction, $ ({u}^{'},{v}^{'}) $ is perturbation geostrophic wind velocity, $ {H}_{0} $ is the constant scale height, $ {f}_{0} $ is the Coriolis force constant at the central latitude of 43° (${f}_{0}=9.87\times {10}^{-5}\;{\mathrm{s}}^{-1}$), $ {N}^{2} $ is the Brunt-Väisälä frequency.

    3.   Interdecadal links between the March AHS in the SEALLH and CGT
    • Figure 3 shows the spatial pattern of the first empirical orthogonal function (EOF1; Fig. 3a) mode of the March AHS over the SEALLH and its normalized principal component time-series (PC1; i.e., the AHS index; Fig. 3b). EOF1 is significantly separated from the remaining EOF modes (Fig. 3c), based on the criteria of North et al. (1982), and has a monopole structure that explains 68.4% of the total variance of the March AHS in the SEALLH (Fig. 3a). The positive phase of the EOF1 features a larger than average AHS over the entire SEALLH. The power spectra of the normalized AHS index time-series (Fig. 3b) show a peak with an 8–11-year period (Fig. 3d). Given that the peak passes a significance test at the 95% confidence level, a Lanczos low-pass filter with an 11-year cutoff period is applied to the AHS index to extract its interdecadal component (AHS-ID; Fig. 3b). The AHS-ID explains 54.3% of the total variance of the AHS index, and is one of the most important components contributing to the variability of the March AHS in the SEALLH.

      Figure 3.  (a) First mode of the EOF analysis for the March AHS (units: W m–2) of the SEALLH region (18°–30°N, 96°–108°E) for 1979–2019. (b) Time series of the first EOF mode (PC1; thin line) and its interdecadal component (thick line with in-filled colors). (c) Explained variances (units: %) of the first five EOF modes, the standard error bars result from the North et al. (1982) rule; (d) Power spectral density of PC1. In (a), the thick green line denotes the domain of the SEALLH. In (d), the red line shows the 95% confidence level of the corresponding red spectrum.

      Following Branstator (2002), the EOF1 of the March 300-hPa non-divergent meridional wind is featured by a clear wave train-like structure with two negative centers over the Mediterranean and western India, and two positive centers over Somalia and the SEALLH in the upper troposphere (Fig. 4a). We subsequently refer to the phase shown in Fig. 4a as the positive phase of the CGT. The EOF1 is significantly separated from the remaining EOF modes and explains 27.0% of the total variance (Fig. 4c). The PC1 time series of the March 300-hPa non-divergent meridional wind (i.e., the CGT index, Fig. 4b) exhibits a significant interdecadal variation with a 10-year periodicity (Fig. 4d). Consequently, a Lanczos low-pass filter with an 11-year cutoff period is also applied to the CGT index to extract its interdecadal component (CGT-ID; Fig. 4b). The CGT-ID can explain 59.3% of the total variance of the CGT index, indicating that interdecadal variability is one of the most important components in the variability of the March CGT.

      To delineate the structures and variations of the CGT on interdecadal time-scales, we further calculate the difference of associated variables using the CGT-ID averaged by 16 negative-phase years subtracted from the CGT-ID averaged by 15 positive-phase years (Fig. 5). Figure 5a shows a clear wave train-like structure in the upper troposphere, with three anticyclonic centers over the Mediterranean, northwestern Indian subcontinent, and northwestern Pacific, and two anomalous cyclonic centers over northeastern Africa and the SEALLH. The wave train–like pattern presents an equivalent barotropic structure, with the maximum amplitude near the upper troposphere (Fig. 5b). This difference shares the same structure as the CGT in boreal winter on the interannual timescale (Branstator, 2002). These results, confirming the result shown in Fig. 4d, indicate that the March CGT undergoes interdecadal variations.

      Figure 4.  Same as in Fig. 3, but for the 300-hPa non-divergent meridional wind (units: m s–1) of the region (0°–45°N, 0°–120°E).

      Figure 5.  Composite differences of the March (a) 200-hPa horizontal winds (vectors; units: m s–1), and (b) 15°–45°N averaged relative vorticity (contour intervals: 10–6 s–1) between the positive phase of the CGT-ID and negative phase of the CGT-ID. In (a), the solid green line indicates the SEALLH. Light-to-dark shading indicates the 90%, 95%, and 99% confidence levels based on a two-tailed Student’s t-test.

      The correlation coefficient between the AHS-ID and CGT-ID is 0.96, which exceeds the 99% confidence level even after decreasing the effective degree of freedom to 12 (Bretherton et al., 1999). This result suggests that the CGT is probably the driver of the March AHS variability over the SEALLH on an interdecadal timescale. To examine the pattern of the March AHS in the SEALLH and its link with the CGT, the March AHS data were regressed onto the CGT-ID for 1979–2019. It is clear that a wave train-like structure starts over Western Europe and passes through northeastern Africa and the Arabian Peninsula before finally reaching the SEALLH (Fig. 6a). However, the anomalous central value of the AHS decreases to some extent. If we further divide the March AHS in the SEALLH into the column-integrated local tendency (CILT), column-integrated advection (CIA), and column-integrated vertical transfer (CIVT), these parameters can also be regressed onto the CGT-ID for 1979–2019. The regression of CILT onto the CGT-ID mostly differs from the CGT in terms of both its spatial structure and anomalous centers, indicating the CILT is not significantly related to the CGT on an interdecadal timescale (Fig. 6b). The CIA regressed onto the CGT-ID, being most like the CGT (Figs. 4a, 5), shows a positive CIA center with a maximum > 40 W m–2 over the SEALLH (Fig. 6c). Figure 6d also resembles the CGT (Figs. 4a, 5). An anomalous center appears over the SEALLH with a maximum of > 25 W m–2. These results indicate that the CGT may affect the anomalous AHS by the advection and vertical transfer on an interdecadal timescale.

      Figure 6.  Regression of the (a) AHS (contour intervals: 20 W m–2), (b) column-integrated local tendency (contour intervals: 0.4 W m–2), (c) column-integrated advection (contour intervals: 10 W m–2), and (d) column-integrated vertical transfer (contour intervals: 5 W m–2) onto the normalized CGT-ID in March during the period 1979–2019. Light-to-dark shading indicates the 90%, 95%, and 99% confidence levels based on a two-tailed Student’s t-test. The solid green line indicates the SEALLH.

    4.   What physical processes link the March AHS in the SEALLH with the CGT on an interdecadal timescale?
    • Figure 7a shows the 700-hPa horizontal wind anomalies regressed onto the CGT-ID. There are three anomalous anticyclonic centers over the Mediterranean, northwestern Indian subcontinent, and northwestern Pacific, and two anomalous cyclonic centers over northeastern Africa and the SEALLH. The 200-hPa horizontal wind anomalies regressed onto the CGT-ID (Fig. 7b) exhibit a similar pattern to Fig. 7a. The intensities of these anomalous anticyclones and anomalous cyclones gradually increase from the lower to upper troposphere, but their positions remain almost unchanged. Such a vertical distribution indicates that the interdecadal CGT is also a zonally oriented equivalent barotropic wave train resembling Fig. 5. Because lines ABCDEFG (blue lines in Fig. 7b) are mainly located in three anomalous anticyclonic centers and two anomalous cyclonic centers and traverse the SEALLH, the cross-section of anomalous vertical velocity is drawn along this line to illustrate the vertical transfer (Fig. 7c). During the positive phase of the CGT, the significant ascending branches are mainly located around point B, between points C and D, and around point F, and the significant descending branches occur between point C, and between points D and E. The SEALLH is located in one of the significant ascending branches (around point F). To reveal more clearly how the anomalous AHS is induced, the column-integrated moist enthalpy flux and its divergence are further calculated (Fig. 7d). It is clear that the air mass with higher moist enthalpy converges in the SEALLH (Figs. 7a, b, d). Stronger than normal southerly winds are located on the eastern flank of the anomalous cyclone in the lower–upper troposphere, and an air mass of higher moist enthalpy converges and condenses over the SEALLH (Fig. 7d), which results in a larger than average AHS over this region (Fig. 6a).

      Figure 7.  Regression of the (a) 700-hPa and (b) 200-hPa horizontal winds (vectors; units: m s−1), (c) vertical velocity (contour intervals: 4×10–3 Pa s–1) along line ABCDEFG shown in (b), and (d) column-integrated moist enthalpy flux (vectors; units: J m–1 s–1) and its divergence (contour intervals: 2×102 J m–2 s–1) onto the normalized CGT-ID in March during the period 1979–2019. In (b), the letters A (40°N, 30°W), B (35°N, 5°E), C (25°N, 20°E), D (15°N, 55°E), E (15°N, 90°E), F (23°N, 100°E) and G (26°N, 125°E) stand for the seven nodes on the CGT path (blue line). In (d), vectors exceeding the 95% confidence level are shown, and zero contours are omitted. Light-to-dark shading indicates the 90%, 95%, and 99% confidence levels based on the two-tailed Student’s t-test. The solid green line indicates the SEALLH.

      The 200-hPa RWS is calculated to further determine the source of the CGT on an interdecadal timescale. The wave train structure is primarily excited by an anomalous RWS in the upper troposphere over northern Africa (Fig. 8a; i.e., the Asian jet entrance indicated by the 40 m s−1 contour of the climatological zonal wind) and propagates eastward to East Asia through the Arabian Peninsula, India, and the SEALLH. This is consistent with the disturbances near the core of the westerly jet that propagate eastward to the downstream area in a Rossby wave pattern because the jet acts as a waveguide (Branstator, 2002; Watanabe, 2004). Figure 8b shows that the 200-hPa wave-activity flux regressed onto the CGT-ID diverges from the positive RWS anomalies over northern Africa, crosses the negative RWS anomalies over the Arabian Peninsula, propagates eastward to the positive RWS anomalies over India, continues downstream and has negative RWS anomalies over the SEALLH, and then disperses in Japan. To address whether there is a decadal periodicity in the RWS over northern Africa, we calculate the 200-hPa RWS averaged over the rectangle in Fig. 8a (20°–35°N, 5°W–30°E) as the RWS index (Fig. 8c). The RWS index has a strong interdecadal signal with a peak value around the 10-year band (Fig. 8d). The correlation coefficient between the CGT-ID and the interdecadal component of the March RWS index is 0.83, which exceeds the 99% confidence level even after decreasing the effective degree of freedom to 11. These results confirm that the CGT affects the interdecadal variations of the March AHS over the SEALLH, which is mainly caused by the Rossby wave originating from northern Africa that propagates along the Asian jet stream in March.

      Figure 8.  Regression of the 200-hPa (a) RWS (contour intervals: 2.5 × 10–11 s–2) and (b) wave-activity flux (m2 s−2) onto the normalized CGT-ID in March during 1979–2019. (c) Time series of the RWS index (thin line) and its interdecadal component (thick line with infilled colors). (d) Power spectral density of the RWS index. In (a), the light-to-dark shading indicates the 90%, 95%, and 99% confidence levels based on the two-tailed Student’s t-test. In (a) and (b), the solid purple line indicates the 40 m s–1 contour of climatological zonal wind. The solid green line indicates the SEALLH. In (d), the red line shows the 95% confidence level of the corresponding red spectrum.

    • Here, two sensitive experiments are conducted to confirm the possible mechanism of the CGT wave train modulated by the RWS over northern Africa. Sensitive experiment 1 (SE1) is forced by the March RWS anomalies over northern Africa (20°–35°N, 5°W–30°E) regressed onto the CGT-ID in March (rectangle in Fig. 8a). Sensitive experiment 2 (SE2) is forced by the same RWS but with opposite sign. In SE1 and SE2, the climatology RWS is used in other regions and other months except for northern Africa (20°–35°N, 5°W–30°E) and March. Other boundary conditions keep their climatological values. Each experiment is integrated for 20 years, and the first 5 years are discarded as the spin-up period. Analysis of the modeling results was performed over the last 15 years based on the differences between SE1 and SE2 (SE1 minus SE2).

      We analyzed simulated differences in the 200-hPa horizontal winds, 15°–45°N averaged relative vorticity, column-integrated moist enthalpy flux, and its divergence in March. The simulated differences of the 200-hPa horizontal winds almost share the same pattern as the diagnostic results for the research domain (Figs. 5a and 9a), although the centers of the anomalous anticyclones around the Mediterranean and near northern Africa somewhat differ from the observational analysis. Two anomalous anticyclonic centers over the northwestern Indian subcontinent and northwestern Pacific, and one anomalous cyclonic center over the SEALLH, perfectly match the observational results. The wave train–like structure also presents an equivalent barotropic structure, with the maximum amplitude near the upper troposphere (Fig. 9b). The relative vorticity in the regions west of North Africa shows an opposite distribution to Fig. 5b, suggesting that the cyclonic RWS anomalies over northern Africa only propagate downstream along the westerly jet. These results confirm the major role of the RWS over northern Africa in driving the phase of the CGT on an interdecadal timescale. The anomalous southerly column-integrated moist enthalpy flux is well reproduced in the SEALLH (Fig. 9c). The higher moist enthalpy air mass converges over the SEALLH, leading to an increase in the AHS in the SEALLH. The simulated results obtained above confirm that the RWS over northern Africa forcing the CGT plays a significant role in regulating the interdecadal variation in AHS over the SEALLH.

      Figure 9.  Composite differences of the March (a) 200-hPa horizontal winds (vectors; units: m s–1), (b) 15°–45°N averaged relative vorticity (contour intervals: 10–6 s–1), and (c) column-integrated moist enthalpy flux (vectors; units: J m–1 s–1) and its divergence (contour intervals: 2×102 J m–2 s–1) between SE1 and SE2. In (a), the solid green line indicates the SEALLH. Light-to-dark shading indicates the 90%, 95%, and 99% confidence levels based on a two-tailed Student’s t-test.

    5.   Summary and discussion
    • This study has investigated the interdecadal variations of the March AHS over the SEALLH and its possible causes based on the fifth-generation reanalysis dataset from the European Centre for Medium-Range Weather Forecasts for the period 1979–2019. The positive EOF1 phase of the March AHS over the SEALLH is characterized by a monopole structure with a larger-than-average AHS over the entire SEALLH. The interdecadal component of the AHS index explains 54.3% of the total variance, highlighting its importance in controlling the variability of the March AHS over the SEALLH. The March CGT is characterized by an equivalent barotropic wave train along the Asian westerly jet stream in the upper troposphere, which also has a significant interdecadal variation that explains 59.3% of the total variance.

      The interdecadal variation of the March AHS over the SEALLH is significantly related to the CGT on the same timescale. When the CGT is optimally excited by the significant vorticity source near northern Africa in its positive phase, the SEALLH is dominated by anomalous southerly winds and ascending motions. The enhanced southerly winds and upward motions located on the region east of the anomalous cyclone, further transport a higher moist enthalpy air mass into the SEALLH, which converges and condenses, ultimately resulting in a larger than average March AHS over this region. When the CGT is in its negative phase, the opposite condition can be observed to a large degree. The simulated results confirm the impact of the RWS over northern Africa on the interdecadal variation of the AHS over the SEALLH through its phase modulation of the CGT. This key physical process links the March AHS over the SEALLH with the CGT on an interdecadal timescale, as is schematically shown in Fig. 10.

      Figure 10.  A schematic diagram of the effects of the CGT on atmospheric heat sources in the SEALLH. The letters A and C denote an anomalous anticyclone and cyclone, respectively.

      The Atlantic Multidecadal Oscillation (AMO; Schlesinger and Ramankutty, 1994) and Pacific Decadal Oscillation (PDO; Mantua et al., 1997) are two significant signals on the interdecadal timescale. The correlation coefficient between the AHS-ID and the interdecadal component of the March AMO index is 0.36. The correlation coefficient between the AHS-ID and the interdecadal component of the March PDO index is –0.26. These two correlation coefficients are insignificant, even at the 90% confidence level. Apparently, the AMO and PDO are not the main factors that directly influence the interdecadal variations of the AHS over the SEALLH on interdecadal timescales.

      Previous studies have pointed out that the AHS and surface heat flux over the anomalous centers of the CGT contribute somewhat to the formation and variability of the CGT (Enomoto, 2004; Yasui and Watanabe, 2010; Chen and Huang, 2012; Lee et al., 2017; Yang et al., 2021). In this study (Figs. 8a, b), we found a significant relationship between the RWS over northern Africa and the CGT-ID. All results imply that an interaction exists between the CGT and AHS located around the anomalous centers of the CGT. The relative contributions of the CGT and associated AHS to the interaction is beyond the scope of this study. In subsequent studies, we intend to clarify the physical processes involved in the interaction between the RWS in northern Africa and CGT through numerical experiments.

      Acknowledgements. This work was supported by the National Natural Science Foundation of China (Grant No. 42030603), the Natural Science Foundation of Yunnan Province (2019FY003006), and the Postgraduate Research and Innovation foundation of Yunnan University (2021Z017).

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return