Advanced Search
Article Contents

Evolution of the Mid-tropospheric Vortex during the Formation of Super Typhoon Megi (2010)

Fund Project:

This work was supported in part by the National Key Research and Development Program of China (Grant No. 2017YFC1501601) and the National Natural Science Foundation of China (Grant No. 41875067). Computations were performed at the Texas Advanced Computing Center and at the High Performance Computing Center of Nanjing University


doi: 10.1007/s00376-020-9178-7

  • As a follow-up of a previously published article on the contribution of tropical waves, this study explores the evolution of the mid-tropospheric mesoscale cyclonic vortex (MV) during the formation of Typhoon Megi (2010) with a successful cloud-resolving simulation. It is found that the formation and intensification of the MV were related to the deep convection and subsequent stratiform precipitation, while the weakening of the MV was related to the shallow convection. Both the upward transport of vorticity related to the deep convection and the horizontal convergence associated with the stratiform precipitation contributed to the formation and intensification of the MV. Even though the latter was dominant, the former could not be ignored, especially in the early stage of the MV. The MV played dual roles in the formation of Megi. On the one hand, the formation and intensification of MV were primarily associated with the stratiform precipitation, which induced the low-level divergence inhibiting the spin-up of the near-surface cyclonic circulation. On the other hand, the coupled low-level cold core under the MV benefited the accumulation of the convective available potential energy (CAPE), which was favorable for the convective activity. A sensitivity experiment with the evaporative cooling turned off indicated that the development of the MV retarded the genesis process of Megi.
    摘要: 利用高精度数值模拟资料,本文分析了超强台风“鲇鱼”形成过程中中层涡旋的发展和演变。研究结果表明,中层涡旋的形成和发展与深对流和随后的层云有关,而其减弱则与浅对流的发展有关。层云产生的中层辐合和深对流引起的垂直方向上涡度输送都对中层涡旋的形成有着贡献。其中,前者起主要作用,但后者的作用也是不可忽略的,尤其是在中层涡旋发展的早期阶段。进一步的分析表明,中层涡旋的形成与发展对“鲇鱼”的形成具有双重影响。一方面,中层涡旋的形成伴随着明显的低层辐散,导致低层气旋性环流减弱;另一方面,与中层涡旋耦合的低层冷心加大了大气的不稳定,有利于对流的爆发。敏感性试验进一步表明,中层涡旋的发展对“鲇鱼”的形成过程有抑制作用。
  • 加载中
  • Figure 1.  (a) GFS-derived (black circles) and model-derived (red circles) track of the disturbance, overlaid by the simulated 1000 hPa wind (vectors), and sea level pressure (blue lines) at 0000 UTC 13 October. The values of the contours are 1006.5, 1007, 1007.5, 1008, 1009, 1010, and 1011 hPa. (b) Area-mean vorticity (units: 10−5 s−1) at 850 hPa within 200-km radius of the disturbance center. The vertical dashed line denotes the genesis time (0000 UTC 13 October, from the JMA). (c) Area-mean vertical wind shear (units: m s−1) between 850 hPa and 200 hPa within a 20° × 20° area centered at the disturbance center.

    Figure 2.  (a) Time−longitude section of GMS/MTSAT-derived blackbody temperature (shading; units: °C) and GFS-analysis-derived 850-hPa winds averaged between 8°N and 16°N (vectors) in the period from 0600 UTC 8 to 0600 UTC 13 October. (b) As in (a) but for the summation of positive vorticity (units: 10−5 s−1) between 8° and 16°N derived from the GFS analysis data at 850 hPa. (c) As in (b) but for 600 hPa. (d−f) As in (a−c) except for (d) the model-derived column-maximum reflectivity (units: dBZ) and 850-hPa winds, (e) 850-hPa and (f) 600-hPa vorticity (units: 10−5 s−1).

    Figure 3.  The re-gridded relative vorticity at 1000 hPa (red contours; units: 10−5 s−1) and 600 hPa (blue contours; units: 10−5 s−1), the 1000−600-hPa-averaged vertical velocity (shading; units: m s−1), as well as the 1000-hPa winds (vectors) in the region centered at pre-Megi’s center with side length of ~1125 km. The contour intervals are 2 × 10−5 s−1 and dashed contours designate negative values. The dashed circles center at pre-Megi’s center with radius of 2°. The vector in the corner of each panel denotes the vertical wind shear between 200 and 850 hPa. The number below the vector represents the magnitude of the shear.

    Figure 4.  (a) Time−height diagrams of model-derived frequency of convection top height (shading) and mean vertical velocity (dashed contours are negative) averaged over the center area of pre-Megi’s disturbance. (b) As in (a) but for the vorticity (shading; units: 10−5 s−1) and divergence (dashed contours are negative). (c) As in (a) but for the temperature perturbation (shading; units: K) relative to the whole domain average, relative humidity (thick contours) and equivalent potential temperature (thin contours). The values of contours are −3, −2, −1, 1, 2, 3, 4, 6, 8, and 10 cm s−1 in (a); −3.8, −2.8, −1.8, −0.8, 0, 0.8, 1.8, 3.8, 5.8, 7.8, 9.8, and 11.8 × 10−5 s−1 in (b); and 10, 20, 30, 40, 50, 60, 70, 75, 80, 84, 88, 90, and 95% in (c). The black dashed, blue, red, and purple thick contours denote the values of 60%, 84% 90%, and 95% in (c), respectively, and the values of dashed (solid) thin contours start from 340 K (344 K) with intervals of 1 K (2 K).

    Figure 5.  (a) Results of the vorticity budget (units: 10−9 s−2) and vorticity (thick black solid line for 600 hPa; dashed line for 950 hPa; ordinate on the right; units: 10−5 s−1) averaged within a 200-km radius from the disturbance center at 600 hPa. (b) Vertical advection, tilting, and divergence of the non-advective vorticity flux (units: 10−9 s−2).

    Figure 6.  (a) Time series of 1000-hPa mean equivalent potential temperature (black; units: K), potential temperature (blue; units: K) and specific humidity (red; units: g kg−1) averaged over the region with 600-hPa re-gridded vorticity greater than 2 × 10−5 s−1 and radius less than 4° during the period from 0600 UTC 8 to 0000 UTC 11 October. (c) As in (a) but for the potential temperature tendency (black; units: 10−4 K s−1), and its components induced by vertical advection (blue), horizontal advection (green) of potential temperature, diabatic heating (red), as well as boundary-layer processes and turbulent diffusion and air−sea interaction (orange). (e) As in (c) except for the specific humidity tendency (units: 10−4 g kg−1 s−1), and its components induced by vertical advection (blue), horizontal advection (green) of specific humidity, phase transition of vapor (red), as well as boundary-layer processes and turbulent diffusion air−sea interaction (orange). (b, d, f) As in (a, c, e) but for the variables at 850 hPa.

    Figure 7.  Skew-T log-P vertical sounding averaged over the center area of pre-Megi’s disturbance. The temperature and dewpoint temperature sounding are denoted by the thick solid and cyan solid lines, respectively. The red dashed line delineates the lifted parcel’s ascent path.

    Figure 8.  Evolution of CAPE (black line; units: J kg−1) and CIN (red line with ordinate on the right; units: J kg−1) averaged over a 200-km radius of the disturbance center.

    Figure 9.  (a−c) $ {\theta }' $ (shading; units: K) at 1000 hPa, vertical wind shear (vectors) between 850 hPa and 200 hPa, relative vorticity (white lines; units: 10−5 s−1) at 600 hPa, and vertical velocity at 850 hPa (red lines; units: m s−1) in the region centered at pre-Megi’s center with side length of ~450 km at 2300 UTC 9 October, 0300 UTC 10 October, and 0600 UTC 10 October, respectively. The dashed circles center at pre-Megi’s center with the radius of 200 km. The values of vorticity and vertical velocity are 2, 4, 8, 12, and 16 × 10−5 s−1, and 0.3 and 0.4 m s−1, respectively, in (a−c). The black box in (a−c) outlines the cross section of (d−f). (d−f) Vertical cross section showing the $ {\theta }' $ (shading; units: K), wind vectors (vectors; units: m s−1), and vorticity (white lines; units: 10−5 s−1). Vertical velocity has been scaled by a factor of 10. The contour intervals are 20 × 10−5 s−1 starting from zero and negative values are ignored. The locations of the cross section appear as black boxes in (a−c).

    Figure 10.  As in Fig. 4 but for NoEvp.

  • Bister, M., and K. A. Emanuel, 1997: The genesis of Hurricane Guillermo: TEXMEX analyses and a modeling study. Mon. Wea. Rev., 125, 2662−2682, https://doi.org/10.1175/1520-0493(1997)125<2662:TGOHGT>2.0.CO;2.
    Bosart, L. R., and F. Sanders, 1981: The Johnstown flood of July 1977: A long-lived convective system. J. Atmos. Sci., 38, 1616−1642, https://doi.org/10.1175/1520-0469(1981)038<1616:TJFOJA>2.0.CO;2.
    Davis, C. A., 2015: The formation of moist vortices and tropical cyclones in idealized simulations. J. Atmos. Sci., 72, 3499−3516, https://doi.org/10.1175/JAS-D-15-0027.1.
    Davis, C. A., and D. A. Ahijevych, 2012: Mesoscale structural evolution of three tropical weather systems observed during PREDICT. J. Atmos. Sci., 69, 1284−1305, https://doi.org/10.1175/JAS-D-11-0225.1.
    Dudhia, J., 1989: Numerical study of convection observed during the Winter Monsoon Experiment using a mesoscale two-dimensional model. J. Atmos. Sci., 46, 3077−3107, https://doi.org/10.1175/1520-0469(1989)046<3077:NSOCOD>2.0.CO;2.
    Dunkerton, T. J., M. T. Montgomery, and Z. Wang, 2009: Tropical cyclogenesis in a tropical wave critical layer: Easterly waves. Atmospheric Chemistry and Physics, 9, 5587−5646, https://doi.org/10.5194/acp-9-5587-2009.
    Fang, J., and F. Q. Zhang, 2016: Contribution of tropical waves to the formation of Supertyphoon Megi (2010). J. Atmos. Sci., 73, 4387−4405, https://doi.org/10.1175/JAS-D-15-0179.1.
    Ge, X. Y., T. M. Li, and S. T. Peng, 2013: Tropical cyclone genesis efficiency: Mid-level versus bottom vortex. Journal of Tropical Meteorology, 19, 197−213, https://doi.org/10.16555/j.1006-8775.2013.03.001.
    Gray, W. M., 1998: The formation of tropical cyclones. Meteorol. Atmos. Phys., 67, 37−69, https://doi.org/10.1007/BF01277501.
    Harr, P. A., R. L. Elsberry, and J. C. L. Chan, 1996: Transformation of a large monsoon depression to a tropical storm during TCM-93. Mon. Wea. Rev., 124, 2625−2643, https://doi.org/10.1175/1520-0493(1996)124<2625:TOALMD>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.
    Hendricks, E. A., M. T. Montgomery, and C. A. Davis, 2004: The role of “vortical” hot towers in the formation of tropical cyclone Diana (1984). J. Atmos. Sci., 61, 1209−1232, https://doi.org/10.1175/1520-0469(2004)061<1209:TROVHT>2.0.CO;2.
    Hong, S. Y., J. Dudhia, and S. H. Chen, 2004: A revised approach to ice microphysical processes for the bulk parameterization of clouds and precipitation. Mon. Wea. Rev., 132, 103−120, https://doi.org/10.1175/1520-0493(2004)132<0103:ARATIM>2.0.CO;2.
    Houze, R. A., Jr., 1993: Cloud Dynamics. Academic, San Diego, Calif., 573 pp.
    Kieu, C. Q., N. M. Truong, H. T. Mai, and T. Ngo-Duc, 2012: Sensitivity of the track and intensity forecasts of Typhoon Megi (2010) to satellite-derived atmospheric motion vectors with the ensemble Kalman filter. J. Atmos. Sci., 29, 1794−1810, https://doi.org/10.1175/JTECH-D-12-00020.1.
    Kilroy, G., R. K. Smith, and M. T. Montgomery, 2017: A unified view of tropical cyclogenesis and intensification. Quart. J. Roy. Meteorol. Soc., 143, 450−462, https://doi.org/10.1002/qj.2934.
    Komaromi, W. A., 2013: An investigation of composite dropsonde profiles for developing and nondeveloping tropical waves during the 2010 PREDICT field campaign. J. Atmos. Sci., 70, 542−558, https://doi.org/10.1175/JAS-D-12-052.1.
    Mapes, B. E., and R. A. Houze Jr., 1995: Diabatic divergence profiles in western Pacific mesoscale convective systems. J. Atmos. Sci., 52, 1807−1828, https://doi.org/10.1175/1520-0469(1995)052<1807:DDPIWP>2.0.CO;2.
    Mlawer, E. J., S. J. Taubman, P. D. Brown, M. J. Iacono, and S. A. Clough, 1997: Radiative transfer for inhomogeneous atmospheres: RRTM, a validated correlated-k model for the longwave. J. Geophys. Res., 102, 16 663−16 682, https://doi.org/10.1029/97JD00237.
    Montgomery, M. T., M. E. Nicholls, T. A. Cram, and A. B. Saunders, 2006: A vortical hot tower route to tropical cyclogenesis. J. Atmos. Sci., 63, 355−386, https://doi.org/10.1175/JAS3604.1.
    Montgomery, M. T., Z. Wang, and T. J. Dunkerton, 2010: Coarse, intermediate and high resolution numerical simulations of the transition of a tropical wave critical layer to a tropical storm. Atmospheric Chemistry and Physics, 10, 10 803−10 827, https://doi.org/10.5194/acp-10-10803-2010.
    Noh, Y., W. G. Cheon, S. Y. Hong, and S. Raasch, 2003: Improvement of the K-profile model for the planetary boundary layer based on large eddy simulation data. Bound.-Layer Meteorol., 107, 401−427, https://doi.org/10.1023/A:1022146015946.
    Nolan, D. S., 2007: What is the trigger for tropical cyclogenesis? Aust Meteorol. Mag., 56, 241−266.
    Qian, C. H., F. Q. Zhang, B. W. Green, J. Zhang, and X. Q. Zhou, 2013: Probabilistic evaluation of the dynamics and prediction of Supertyphoon Megi (2010). Wea. Forecasting, 28, 1562−1577, https://doi.org/10.1175/WAF-D-12-00121.1.
    Raymond, D. J., S. L. Sessions, and C. López Carrillo, 2011: Thermodynamics of tropical cyclogenesis in the northwest Pacific. J. Geophys. Res., 116, D18101, https://doi.org/10.1029/2011JD015624.
    Raymond, D. J., S. Gjorgjievska, S. L. Sessions, and Ž. Fuchs, 2014: Tropical cyclogenesis and mid-level vorticity. Australian Meteorological and Oceanographic Journal, 64, 11−25, https://doi.org/10.22499/2.6401.003.
    Ritchie, E. A., and G. J. Holland, 1997: Scale interactions during the formation of Typhoon Irving. Mon. Wea. Rev., 125, 1377−1396, https://doi.org/10.1175/1520-0493(1997)125<1377:SIDTFO>2.0.CO;2.
    Shi, W. L., J. F. Fei, X. G. Huang, X. P. Cheng, J. L. Ding, and Y. Q. He, 2014: A numerical study on the combined effect of midlatitude and low-latitude systems on the abrupt track deflection of Typhoon Megi (2010). Mon. Wea. Rev., 142, 2483−2501, https://doi.org/10.1175/MWR-D-13-00283.1.
    Simpson, J., E. Ritchie, G. J. Holland, J. Halverson, and S. Stewart, 1997: Mesoscale interactions in tropical cyclone genesis. Mon. Wea. Rev., 125, 2643−2661, https://doi.org/10.1175/1520-0493(1997)125<2643:MIITCG>2.0.CO;2.
    Skamarock, W. C., J. B. Klemp, J. Dudhia, D. O. Gill, D. M. Barker, W. Wang, and J. G. Powers, 2005: A Description of the Advanced Research WRF Version 2. NCAR Technical Note NCAR/TN-468+STR, 88 pp.
    Tory, K. J., M. T. Montgomery, 2008: Tropical cyclone formation: A synopsis of the internal dynamics. Extended abstracts. 28th Conf. on Hurricanes and Tropical Meteorology, Orlando, FL, Amer. Meteor. Soc., 10A.1. [Available online at http://ams.confex.com/ams/28Hurricanes/techprogram/paper_138062.htm.]
    Wang, H., and Y. Q. Wang, 2014: A numerical study of Typhoon Megi (2010). Part I: Rapid intensification. Mon. Wea. Rev., 142, 29−48, https://doi.org/10.1175/MWR-D-13-00070.1.
    Wang, Y. Q., and H. Wang, 2013: The inner-core size increase of Typhoon Megi (2010) during its rapid intensification phase. Tropical Cyclone Research and Review, 2, 65−80, https://doi.org/10.6057/2013TCRR02.01.
    Wang, Z., 2012: Thermodynamic aspects of tropical cyclone formation. J. Atmos. Sci., 69, 2433−2451, https://doi.org/10.1175/JAS-D-11-0298.1.
    Wang, Z., 2014: Role of cumulus congestus in tropical cyclone formation in a high-resolution numerical model simulation. J. Atmos. Sci., 71, 1681−1700, https://doi.org/10.1175/JAS-D-13-0257.1.
    Wang, Z., M. T. Montgomery, and T. J. Dunkerton, 2010a: Genesis of Pre-Hurricane Felix (2007). Part I: The role of the easterly wave critical layer. J. Atmos. Sci., 67, 1711−1729, https://doi.org/10.1175/2009JAS3420.1.
    Wang, Z., M. T. Montgomery, and T. J. Dunkerton, 2010b: Genesis of Pre-Hurricane Felix (2007). Part II: Warm core formation, precipitation evolution, and predictability. J. Atmos. Sci., 67, 1730−1744, https://doi.org/10.1175/2010JAS3435.1.
    Wu, S., and J. Fang, 2019: The Evolution and Role of Midtropospheric Cyclonic Vortex in the Formation of Super Typhoon Nepartak (2016). J. Geophys. Res. Atmos., 124, 9277−9298, https://doi.org/10.1029/2019JD030631.
    Zawislak, J., and E. J. Zipser, 2014: Analysis of the thermodynamic properties of developing and nondeveloping tropical disturbances using a comprehensive dropsonde dataset. Mon. Wea. Rev., 142, 1250−1264, https://doi.org/10.1175/MWR-D-13-00253.1.
    Zehr, R. M., 1992: Tropical cyclogenesis in the western North Pacific. PhD dissertation, Colorado State University, 181 pp.
    Zipser, E. J., 1977: Mesoscale and convective-scale downdrafts as distinct components of squall-line structure. Mon. Wea. Rev., 105, 1568−1589, https://doi.org/10.1175/1520-0493(1977)105<1568:MACDAD>2.0.CO;2.
  • [1] Jieshun ZHU, Entcho DEMIROV, Ying ZHANG, and Ania POLOMSKA-HARLICK, 2014: Model Simulations of Mesoscale Eddies and Deep Convection in the Labrador Sea, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 743-754.  doi: 10.1007/s00376-013-3107-y
    [2] HUANG Wenyu, WANG Bin*, LI Lijuan, DONG Li, LIN Pengfei, YU Yongqiang, ZHOU Tianjun, LIU Li, XU Shiming, XIA Kun, PU Ye, WANG Lu, LIU Mimi, SHEN Si, HU Ning, WANG Yong, SUN Wenqi, and DONG Fang, 2014: Variability of Atlantic Meridional Overturning Circulation in FGOALS-g2, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 95-109.  doi: 10.1007/s00376-013-2155-7
    [3] Haikun ZHAO, Chunzai WANG, Ryuji YOSHIDA, 2016: Modulation of Tropical Cyclogenesis in the Western North Pacific by the Quasi-Biweekly Oscillation, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 1361-1375.  doi: 10.1007/s00376-016-5267-z
    [4] Yongguang ZHENG, Yanduo GONG, Jiong CHEN, Fuyou TIAN, 2019: Warm-Season Diurnal Variations of Total, Stratiform, Convective, and Extreme Hourly Precipitation over Central and Eastern China, ADVANCES IN ATMOSPHERIC SCIENCES, 36, 143-159.  doi: 10.1007/s00376-018-7307-3
    [5] Yafei YAN, Yimin LIU, 2019: Vertical Structures of Convective and Stratiform Clouds in Boreal Summer over the Tibetan Plateau and Its Neighboring Regions, ADVANCES IN ATMOSPHERIC SCIENCES, 36, 1089-1102.  doi: 10.1007/s00376-019-8229-4
    [6] Shuang LUO, Yunfei FU, Shengnan ZHOU, Xiaofeng WANG, Dongyong WANG, 2020: Analysis of the Relationship between the Cloud Water Path and Precipitation Intensity of Mature Typhoons in the Northwest Pacific Ocean, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 359-376.  doi: 10.1007/s00376-020-9204-9
    [7] Yan SUN, Fan WANG, De-Zheng SUN, 2016: Weak ENSO Asymmetry Due to Weak Nonlinear Air-Sea Interaction in CMIP5 Climate Models, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 352-364.  doi: 10.1007/s00376-015-5018-6
    [8] Lijuan WANG, Aiguo DAI, Shuaihong GUO, Jing GE, 2017: Establishment of the South Asian High over the Indo-China Peninsula During Late Spring to Summer, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 169-180.  doi: 10.1007/s00376-016-6061-7
    [9] Xiaofan LI, SHEN Xinyong, 2012: Temporal and Spatial Scale Dependence of Precipitation Analysis over the Tropical Deep Convective Regime, ADVANCES IN ATMOSPHERIC SCIENCES, 29, 1390-1394.  doi: 10.1007/s00376-012-1269-7
    [10] Tuanjie HOU, Hengchi LEI, Youjiang HE, Jiefan YANG, Zhen ZHAO, Zhaoxia HU, 2021: Aircraft Measurements of the Microphysical Properties of Stratiform Clouds with Embedded Convection, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 966-982.  doi: 10.1007/s00376-021-0287-8
    [11] Tuanjie HOU, Baojun CHEN, Hengchi LEI, Lei WEI, Youjiang HE, Qiujuan FENG, 2023: Evaluation of the Predicted Particle Properties (P3) Microphysics Scheme in Simulations of Stratiform Clouds with Embedded Convection, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 1859-1876.  doi: 10.1007/s00376-023-2178-7
    [12] Quanliang CHEN, Guolu GAO, Yang LI, Hongke CAI, Xin ZHOU, Zhenglin WANG, 2019: Main Detrainment Height of Deep Convection Systems over the Tibetan Plateau and Its Southern Slope, ADVANCES IN ATMOSPHERIC SCIENCES, 36, 1078-1088.  doi: 10.1007/s00376-019-9003-3
    [13] Yongjie HUANG, Yaping WANG, Xiaopeng CUI, 2019: Differences between Convective and Stratiform Precipitation Budgets in a Torrential Rainfall Event, ADVANCES IN ATMOSPHERIC SCIENCES, 36, 495-509.  doi: 10.1007/s00376-019-8159-1
    [14] Jiang HUANGFU, Zhiqun HU, Jiafeng ZHENG, Lirong WANG, Yongjie ZHU, 2024: Study on Quantitative Precipitation Estimation by Polarimetric Radar Using Deep Learning, ADVANCES IN ATMOSPHERIC SCIENCES.  doi: 10.1007/s00376-023-3039-0
    [15] ZHU Congwen, Tetsuo NAKAZAWA, LI Jianping, 2003: Modulation of Twin Tropical Cyclogenesis by the MJO Westerly Wind Burst during the Onset Period of 1997/98 ENSO, ADVANCES IN ATMOSPHERIC SCIENCES, 20, 882-898.  doi: 10.1007/BF02915512
    [16] Xi CAO, Shangfeng CHEN, Guanghua CHEN, Renguang WU, 2016: Intensified Impact of Northern Tropical Atlantic SST on Tropical Cyclogenesis Frequency over the Western North Pacific after the Late 1980s, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 919-930.  doi: 10.1007/s00376-016-5206-z
    [17] Lu YANG, Jianfang FEI, Xiaogang HUANG, Xiaoping CHENG, Xiangrong YANG, Juli DING, Wenli SHI, 2016: Asymmetric Distribution of Convection in Tropical Cyclones over the Western North Pacific Ocean, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 1306-1321.  doi: 10.1007/s00376-016-5277-x
    [18] Jae-Young BYON, Gyu-Ho LIM, 2005: Diurnal Variation of Tropical Convection during TOGA COARE IOP, ADVANCES IN ATMOSPHERIC SCIENCES, 22, 685-702.  doi: 10.1007/BF02918712
    [19] Tuanjie HOU, Hengchi LEI, Zhaoxia HU, Jiefan YANG, Xingyu LI, 2020: Simulations of Microphysics and Precipitation in a Stratiform Cloud Case over Northern China: Comparison of Two Microphysics Schemes, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 117-129.  doi: 10.1007/s00376-019-8257-0
    [20] Hong WANG, Hengchi LEI, Jiefan YANG, 2017: Microphysical Processes of a Stratiform Precipitation Event over Eastern China: Analysis Using Micro Rain Radar data, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 1472-1482.  doi: 10.1007/s00376-017-7005-6

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 12 September 2019
Manuscript revised: 17 April 2020
Manuscript accepted: 22 April 2020
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

Evolution of the Mid-tropospheric Vortex during the Formation of Super Typhoon Megi (2010)

    Corresponding author: Juan FANG, fangjuan@nju.edu.cn
  • 1. Key Laboratory of Mesoscale Severe Weather (MOE), School of the Atmospheric Sciences, Nanjing University, 163 Xianlin Road, Nanjing 210046, China
  • 2. Yubei Meteorological Office of Chongqing, Chongqing 401147, China
  • 3. Chongqing Institute of Meteorological Sciences, Chongqing 401147, China

Abstract: As a follow-up of a previously published article on the contribution of tropical waves, this study explores the evolution of the mid-tropospheric mesoscale cyclonic vortex (MV) during the formation of Typhoon Megi (2010) with a successful cloud-resolving simulation. It is found that the formation and intensification of the MV were related to the deep convection and subsequent stratiform precipitation, while the weakening of the MV was related to the shallow convection. Both the upward transport of vorticity related to the deep convection and the horizontal convergence associated with the stratiform precipitation contributed to the formation and intensification of the MV. Even though the latter was dominant, the former could not be ignored, especially in the early stage of the MV. The MV played dual roles in the formation of Megi. On the one hand, the formation and intensification of MV were primarily associated with the stratiform precipitation, which induced the low-level divergence inhibiting the spin-up of the near-surface cyclonic circulation. On the other hand, the coupled low-level cold core under the MV benefited the accumulation of the convective available potential energy (CAPE), which was favorable for the convective activity. A sensitivity experiment with the evaporative cooling turned off indicated that the development of the MV retarded the genesis process of Megi.

摘要: 利用高精度数值模拟资料,本文分析了超强台风“鲇鱼”形成过程中中层涡旋的发展和演变。研究结果表明,中层涡旋的形成和发展与深对流和随后的层云有关,而其减弱则与浅对流的发展有关。层云产生的中层辐合和深对流引起的垂直方向上涡度输送都对中层涡旋的形成有着贡献。其中,前者起主要作用,但后者的作用也是不可忽略的,尤其是在中层涡旋发展的早期阶段。进一步的分析表明,中层涡旋的形成与发展对“鲇鱼”的形成具有双重影响。一方面,中层涡旋的形成伴随着明显的低层辐散,导致低层气旋性环流减弱;另一方面,与中层涡旋耦合的低层冷心加大了大气的不稳定,有利于对流的爆发。敏感性试验进一步表明,中层涡旋的发展对“鲇鱼”的形成过程有抑制作用。

1.   Introduction
  • Due to its multiscale nature and the lack of conventional observations over the remote ocean, tropical cyclogenesis remains a topic of great challenge in the fields of tropical meteorology. Via advanced observational diagnoses and high-resolution numerical simulations, an increasing number of mesoscale and small-scale features associated with tropical cyclogenesis have been identified in the past several decades. Among them, the mid-tropospheric mesoscale cyclonic vortex (MV) has been widely accepted as a precursor disturbance for the formation of a tropical cyclone (TC) and its role in tropical cyclogenesis has been discussed in detail.

    With aircraft observing data, Harr et al. (1996) proposed that the structure of a mesoscale convective system (MCS) is beneficial to the transformation of the monsoon depression to a TC by amplifying the vorticity of the monsoon depression. Bister and Emanuel (1997) and Ritchie and Holland (1997) investigated the formation of Hurricane Guillermo (1991) and Typhoon Irving (1992), respectively, and both found that a midlevel mesocyclone in the stratiform precipitation region preceded the development of the near-surface cyclone. Simpson et al. (1997) argued that the formation of TC Oliver might have been induced by the interactions between MCSs and associated mesoscale vortices at the mid-levels. Based on a large amount of statistical observational data of the early stage of TC formation, Zehr (1992) and Gray (1998) suggested that tropical cyclogenesis is a two-stage process, with the MV establishing in the first stage. Recent observational work has continued to indicate that TC formation is preceded by a moist, mid-level mesoscale vortex (Komaromi, 2013; Zawislak and Zipser, 2014). Idealized numerical simulations confirm the observational studies. In idealized numerical simulations with a pre-existing, weak tropical disturbance in a quiescent atmosphere, the genesis of a TC only occurs if the vortex in the mid-troposphere has intensified (Nolan, 2007; Ge et al., 2013). With a cloud model initialized by a radiative−convective equilibrium state perturbed by random signals with amplitude of 0.25 K, the simulation by Davis (2015) exhibited the development of rotating, coherent mid-tropospheric structures under the condition without large-scale forcing and an initial vortex, revealing that a TC formed about 6.5 days later after the emergence of the mid-level vortex.

    As a significant precursor of TC formation, the roles of an MV in tropical cyclogenesis have been discussed in detail in the literature. Simpson et al. (1997) and Ritchie and Holland (1997) found that a strengthened MV with increased spatial scale can result from the concentration of mid-level vortices, which can lead to the growth of the near-surface cyclonic circulation and thus favor the intensification of a tropical disturbance. Bister and Emanuel (1997) hypothesized that the spin-up of the surface vortex can happen under the effect of the downward transport of mid-level cyclonic vorticity by the mesoscale downdraft within the stratiform precipitation. Gray (1998) argued that the convectively initiated mid-level cyclonic vortex can extend downwards to the surface in the form of a much smaller and more concentrated cyclonic vortex, which can be the embryo of the future storm.

    Different from the above-mentioned works focusing on the downward extension of the mid-level vorticity, there are many other studies in the literature that argue the MV is closely linked to vortical hot towers (VHTs), which are critical to the intensification of the surface vortex in tropical cyclogenesis (e.g., Hendricks et al., 2004; Montgomery et al., 2006). Montgomery et al. (2006) suggested that the positive vorticity-rich environment provided by an MV favors VHT development. Dunkerton et al. (2009) argued that, because of the approximately closed Lagrangian recirculation, the mid-level cyclonic circulation associated with the easterly wave can protect the TC embryo inside from the hostile environment, such as dry air intrusion, strain/shear deformation and so on, thus providing a critical point for VHT concentration. This perspective was advocated by Wang et al. (2010a, b) and Montgomery et al. (2010), who showed that the region occupied by an MV is characterized by strong rotation as well as weak deformation and vertical wind shear, favoring further VHT aggregation and persistent convection. In addition, Wang (2012) indicated that the closed cyclonic region of the MV in the central area of pre-Hurricane Felix (2007) was beneficial for VHT aggregation and tropical cyclogenesis because of the abundant moisture and short incubation time scale.

    With a focus on the thermodynamic structure of the MV, Raymond et al. (2011, 2014) argued that the cold anomaly below the MV favors the outbreak of surface-based convection and then the spin-up of TCs in that the downdrafts induced by evaporative cooling are inhibited. However, Davis (2015) argued that, during TC genesis, the downdrafts intensify/weaken along with the strengthening/weakening of the upward motion, and the bottom-heavy mass flux profile within an MV can primarily be ascribed to the intense updrafts rather than the weakened downdrafts in the lower troposphere. In addition, the simulation by Davis (2015) highlighted that the enhanced low−mid-level stratification corresponding to the remarkable mid-level cyclonic vorticity was unfavorable for deep convective bursts. His work suggested that, prior to genesis, strong and large surface cold pools induced by strengthened downdrafts would interplay with the vertical wind shear in organizing the low-level updrafts that contribute to the strengthening of the surface cyclonic circulation. It is worth mentioning that, although the MV has been widely accepted as a precursor in tropical cyclogenesis, the idealized experiments performed by Kilroy et al. (2017), in which the ice microphysical processes were absent, indicated that the mid-level vortex may not be necessary for the occurrence of TC formation.

    Although the relationships between MVs and TC formation have been discussed in detail since Bosart and Sanders (1981), the evolution of the MV in the pre-genesis stage of a TC has not been discussed comprehensively owing to the lack of continuous observations. With the aim to further understand the tropical cyclogenesis associated with the MV, the current study examines the formation of super Typhoon Megi (2010) on the basis of a cloud-resolving simulation. Similar to Davis (2015), this work focuses on the MV evolution in TC genesis, but for a real-world TC. Megi was the strongest typhoon over the globe in 2010, with a minimum sea-level pressure of 895 hPa and maximum wind of 72 m s−1 (best-track data of the Japan Meteorological Agency; JMA). It originated from several positive vorticity anomalies (PVAs) initially scattered in the western tropical Pacific Ocean and was designated as a tropical depression by the JMA at 0000 UTC 13 October. During Megi’s lifetime, rapid intensification, sudden track changes, and rapid weakening occurred sequentially, which induced large prediction errors in the operational forecasting. As a result, Megi drew the attention of many studies (e.g., Kieu et al., 2012; Qian et al., 2013; Shi et al., 2014; Wang and Wang, 2013, 2014). The current study focuses on its formation, which has not been fully explored in the literature.

    The rest of this paper is organized as follows: The numerical simulation and verification are briefly introduced in section 2. Section 3 presents the model-derived evolution of the MV and associated convection during Megi’s formation. Section 4 focuses on the influence of the MV on the formation of Megi. Section 5 gives some concluding remarks.

2.   Numerical simulation and verification
  • The data for the analysis of the MV evolution in the formation of Megi were derived from the control experiment in Fang and Zhang (2016). A description of the experimental design can be found in their paper. Here, some key aspects of the setup for the numerical simulation are given for reference. The simulation utilized the Weather Research and Forecasting (WRF) model, version 3.5.1 (Skamarock et al., 2005). Three, two-way nested domains were employed, with horizontal grid spacings of 13.5, 4.5 and 1.5 km, containing 481 × 361, 811 × 601 and 811 × 691 grid points, respectively. All three domains had 35 levels, with the top level at 10 hPa. The physics packages used included the WRF single-moment 6-class microphysics scheme (with graupel) (Hong et al., 2004), the Yonsei University boundary layer scheme (Noh et al., 2003), the Rapid Radiative Transfer Model for longwave radiation (Mlawer et al., 1997), and the Dudhia shortwave radiation scheme (Dudhia, 1989). The model was initialized at 0600 UTC 8 October with the NCEP’s Global Forecast System (GFS) analysis fields at a resolution of 0.5°. To better simulate the large-scale circulation around the pre-Megi disturbance, spectral nudging was applied to the outermost domain to nudge the disturbances with a scale larger than 810 km towards the GFS analysis data during the five-day simulation. Spectral nudging was also used in the inner domains in the first 24 h of integration with the aim to reduce the impact of a cold start on the simulation. In this work, we mainly focus on the mesoscale processes derived from the innermost domain in the time period starting from 24 h, during which the spectral nudging was turned off.

    The model-simulated and GFS-derived track and area-mean vorticity of the pre-Megi disturbance, as well as the environmental vertical wind shear between 850 and 200 hPa following the disturbance, are shown in Fig. 1. The observational data indicate that the formation of Megi occurred in the environment with moderate vertical wind shear (~7.5 m s−1; Fig. 1c). At the beginning, the pre-Megi disturbance moved westwards, and then took a northwest route and a southwest route sequentially (Fig. 1a), while the vorticity experienced fluctuating variations so that the disturbance developed into a TC at 0000 UTC 13 October 2010 according to the JMA best track (Fig. 1b). In general, the above-mentioned characteristics of the environment, track and intensity of the pre-Megi disturbance were properly described by the numerical model. The model-simulated vertical wind shear was larger than that derived from GFS, but the difference was less than ~2 m s−1. The deviation of the model-simulated track from that derived from GFS was smaller than 150 km. The model-derived vorticity was a bit larger than that derived from GFS. This difference can be partially ascribed to the fact the horizontal resolution of the model data was higher than the observation. In addition to the environment, track and intensity, the key mesoscale processes in the formation of Megi, such as the first episode of vigorous deep convection starting from 0600 UTC 8 October, the vorticity aggregation after 0600 UTC 9 October, and convection reinvigoration after 0000 UTC 10 October, were also represented pretty well in the numerical simulation (Fig. 2). We acknowledge that the spectral nudging may have a role in the model performance, but it was only used in the first 24 h in the innermost domain. Overall, the good performance of the numerical model in the simulation of Megi’s formation provided four-dimensional and dynamically consistent data for our investigation of the formation of Megi.

    Figure 1.  (a) GFS-derived (black circles) and model-derived (red circles) track of the disturbance, overlaid by the simulated 1000 hPa wind (vectors), and sea level pressure (blue lines) at 0000 UTC 13 October. The values of the contours are 1006.5, 1007, 1007.5, 1008, 1009, 1010, and 1011 hPa. (b) Area-mean vorticity (units: 10−5 s−1) at 850 hPa within 200-km radius of the disturbance center. The vertical dashed line denotes the genesis time (0000 UTC 13 October, from the JMA). (c) Area-mean vertical wind shear (units: m s−1) between 850 hPa and 200 hPa within a 20° × 20° area centered at the disturbance center.

    Figure 2.  (a) Time−longitude section of GMS/MTSAT-derived blackbody temperature (shading; units: °C) and GFS-analysis-derived 850-hPa winds averaged between 8°N and 16°N (vectors) in the period from 0600 UTC 8 to 0600 UTC 13 October. (b) As in (a) but for the summation of positive vorticity (units: 10−5 s−1) between 8° and 16°N derived from the GFS analysis data at 850 hPa. (c) As in (b) but for 600 hPa. (d−f) As in (a−c) except for (d) the model-derived column-maximum reflectivity (units: dBZ) and 850-hPa winds, (e) 850-hPa and (f) 600-hPa vorticity (units: 10−5 s−1).

3.   Evolution of MV during the formation of Megi
  • In consistent with Gray (1998), Figs. 2a and d show that the pre-Megi disturbance experienced two episodes of deep convection separated by an intermediate one-day inactive period of little convection. The first episode of massive intense convection began from the model start to around 0600 UTC 9 October, which was associated with the passage of a tropical Kelvin wave (Fang and Zhang, 2016), while the second episode of deep convection started from about 0600 UTC 10 October and continued towards the end. Figures 2b-c and 2e-f further indicate that the distinct increase of lower-tropospheric vorticity mainly occurred in the second episode of deep convection, prior to which, i.e., in the interval between the first and second deep convection episode, the mid-tropospheric vorticity enhanced significantly. According to the variations of the convection and low- and mid-level vorticity described above, and for the convenience of ensuing discussions, the genesis process of Megi is roughly partitioned into three phases (with 0600 UTC 9 October and 0600 UTC 10 October subjectively chosen as the dividing points): (1) massive convective burst associated with the tropical Kelvin wave; (2) decaying convection and mid-level vorticity enhancement; and (3) deep convection reinvigoration, low-level vortex intensification, and formation of Megi. Such a divide is consistent with that in Fang and Zhang (2016).

    Figure 3 displays the convection and low- and mid-tropospheric vorticity in the pre-Megi disturbance during the formation of Megi with time intervals of 12 h. At the beginning of the model integration (0600 UTC 8 October; Fig. 3a), a small PVA and two PVAs were presented in the lower and middle troposphere, respectively. After a short time of model spin-up, there was a convection burst in the pre-Megi disturbance and the low-level PVA enlarged while one more PVA appeared in the middle troposphere (stage 1; Fig. 3b). In the following 12 h, the low-level PVA shrank a little and its vorticity increased considerably, while the mid-level PVAs merged into a mesoscale vorticity anomaly (Fig. 3c). Afterwards, the deep convection and the low-level vorticity decayed gradually in the pre-Megi disturbance, while the mid-level mesoscale vorticity anomaly kept strengthening and developed into a well-defined mid-level mesoscale vortex, i.e., the MV, at about 1800 UTC 9 October (stage 2; Fig. 3d). As the convection became active again in the pre-Megi disturbance after 0600 UTC 10 October (stage 3; Fig. 3e), the low-level vorticity began to enhance significantly, while the MV weakened slightly (Figs. 3e and f). Meanwhile, the MV began to shift eastwards relative to the low-level vorticity under the impact of the westerly shear. The westerly shear also exerted some influences on the convection, as they tended to occur more frequently in the downshear region (Figs. 3e-g). In the following period of Megi’s genesis process (Figs. 3f-l), both the low-level PVA and MV persistently occupied the center area of the pre-Megi disturbance and exhibited fluctuating intensification. The fluctuation of the intensity of the MV can also be detected in Fig. 2c. After 0000 UTC 13 October, i.e., Megi’s genesis time, the low-level mesoscale vortex and MV began to couple together to form a monopole structure, though there still existed little displacement between the low-level vortex and MV due to the vertical wind shear (Fig. 3l).

    Figure 3.  The re-gridded relative vorticity at 1000 hPa (red contours; units: 10−5 s−1) and 600 hPa (blue contours; units: 10−5 s−1), the 1000−600-hPa-averaged vertical velocity (shading; units: m s−1), as well as the 1000-hPa winds (vectors) in the region centered at pre-Megi’s center with side length of ~1125 km. The contour intervals are 2 × 10−5 s−1 and dashed contours designate negative values. The dashed circles center at pre-Megi’s center with radius of 2°. The vector in the corner of each panel denotes the vertical wind shear between 200 and 850 hPa. The number below the vector represents the magnitude of the shear.

    Figures 3e-l also show that the convection in the center area of the pre-Megi disturbance was particularly active at about 1800 UTC 10, 18 UTC 11, 18 UTC 12 and 18 UTC 13 October, while comparatively weak at about 0600 UTC 12 and 0600 UTC 13 October. Such a quasi-periodic oscillation of the convection can also be detected in Figs. 2a and d, and is more apparent in the time−height diagrams of model-derived frequency of convection top height and the mean vertical velocity averaged over the center area of the pre-Megi disturbance as shown in Fig. 4a, which shows that the convection, especially the deep convection, was characterized by diurnal variations in the pre-Megi disturbance after 0600 UTC 10 October. Such a characteristic has also been found in the formation of Hurricanes Karl (2010) and Matthew (2010) (Davis and Ahijevych, 2012), as well as TC Fay (2008) (Wang, 2014).

    Figure 4.  (a) Time−height diagrams of model-derived frequency of convection top height (shading) and mean vertical velocity (dashed contours are negative) averaged over the center area of pre-Megi’s disturbance. (b) As in (a) but for the vorticity (shading; units: 10−5 s−1) and divergence (dashed contours are negative). (c) As in (a) but for the temperature perturbation (shading; units: K) relative to the whole domain average, relative humidity (thick contours) and equivalent potential temperature (thin contours). The values of contours are −3, −2, −1, 1, 2, 3, 4, 6, 8, and 10 cm s−1 in (a); −3.8, −2.8, −1.8, −0.8, 0, 0.8, 1.8, 3.8, 5.8, 7.8, 9.8, and 11.8 × 10−5 s−1 in (b); and 10, 20, 30, 40, 50, 60, 70, 75, 80, 84, 88, 90, and 95% in (c). The black dashed, blue, red, and purple thick contours denote the values of 60%, 84% 90%, and 95% in (c), respectively, and the values of dashed (solid) thin contours start from 340 K (344 K) with intervals of 1 K (2 K).

    It is worth mentioning that the fluctuating growth of the MV mentioned before, which is more prominent in Fig. 4b, is different from that derived from the idealized simulation initialized by a radiative−convective equilibrium state perturbed by random signals, in which the MV strengthened persistently (Davis, 2015). Comparing Fig. 4a to Fig. 4b, the evolution of the MV was closely related to the convective activity in the pre-Megi disturbance, in that the maximum vorticity of the MV usually appeared after the deep convection. The evolution of the MV and its relationship with the convection in the pre-Megi disturbance are discussed in detail in the following section.

  • To understand the processes responsible for the evolution of the MV during the formation of Megi, the flux form of vorticity budget analysis was conducted at 600 hPa following Haynes and McIntyre (1987) and Wang et al. (2010a). The vorticity equation is written as

    where $ \omega $ is the vertical velocity in the pressure coordinate, p denotes pressure, k is the unit vector in the vertical direction, and $\eta $ is the vertical component of absolute vorticity, i.e.,

    where u, v is the eastward and northward component of velocity, x, y is the distance in the eastward and northward direction, and ${{{V}}'}$ denotes disturbance-relative flow, i.e.,

    where V and C is the earth-relative flow and the velocity of the disturbance. The term on the left-hand side of Eq. (1) is the tendency of the absolute vertical vorticity. The first and second terms on the right-hand side of Eq. (1) are referred to as the advective vorticity flux and the non-advective vorticity flux, respectively, following Tory and Montgomery (2008). The former is the divergence of the vorticity flux and consists of the horizontal advection of absolute vorticity and the stretching effect. The latter combines the vertical vorticity advection and the tilting effect. R in Eq. (1) represents the horizontal components of subgrid-scale terms. Since R generally yields a small contribution to the net vorticity tendency, it is ignored in the calculation, as done in Montgomery et al. (2006).

    Figure 5 displays the area-mean 600-hPa vorticity and its tendencies derived by applying Eq. (1) to the domain centered at the vorticity centroid of the MV with the radius of 200 km. During the spin-up period of the model, both the advective vorticity flux and non-advective flux terms in Eq. (1) were negative, and thus the vorticity in MV decreased a little bit. As the deep convection became active in the pre-Megi disturbance after 1800 UTC 8 October, the vorticity of MV began to increase, which was primarily induced by the convergence of horizontal vorticity flux (Figs. 4a and 5a). Although the vorticity tendency induced by the vertical vorticity advection was mostly offset by the tilting effect, it caused the non-advective flux term to be positive and contributed to the development of the MV (Fig. 5b), which indicates that the deep convection can also enhance the mid-level vorticity by transporting the low-level vorticity upwards.

    Figure 5.  (a) Results of the vorticity budget (units: 10−9 s−2) and vorticity (thick black solid line for 600 hPa; dashed line for 950 hPa; ordinate on the right; units: 10−5 s−1) averaged within a 200-km radius from the disturbance center at 600 hPa. (b) Vertical advection, tilting, and divergence of the non-advective vorticity flux (units: 10−9 s−2).

    After about 0600 UTC 9 October, i.e., in stage 2, the deep convection decayed in the pre-Megi disturbance, while the stratiform precipitation developed gradually (Figs. 3c and 4a). Under the effect of the upper-level outflow of the deep convection and the moderate to strong vertical wind shear shown in Fig. 1c, the ice particles moved away from the deep convection region, and then descended and began to melt, inducing the stratiform precipitation (Houze, 1993). The evaporative cooling associated with the precipitation provided the negative buoyancy, leading to the evident downdrafts in the lower troposphere. Accordingly, marked subsidence emerged in the lower troposphere while the updraft shrank to the mid−upper levels in the pre-Megi disturbance (Fig. 4a). Corresponding to the downdraft, a cold core formed in the low levels (Fig. 4c). In addition, divergence gradually replaced convergence in the lower troposphere and evident convergence appeared in the mid-levels (Fig. 4b). These phenomena are usually observed in the stratiform precipitation region (Mapes and Houze, 1995), indicating that the stratiform precipitation became dominant in the pre-Megi disturbance in stage 2.

    Corresponding to the stratiform precipitation development, the vertical vorticity advection, and accordingly the non-advective flux term, weakened, while the horizontal convergence of vorticity flux enhanced considerably. Therefore, the MV strengthened rapidly after about 0600 UTC 9 October and reached maximum intensity at around 1800 UTC 9 October (Fig. 5a).

    From the above description, we can see that the formation and development of the MV primarily resulted from the marked mid-level convergence associated with the stratiform precipitation in the pre-Megi disturbance. In addition, the intense convection also transported the low-level vorticity upwards to facilitate the enhancement of the mid-level vorticity. It is worth noting that the variation of the mid-level vorticity was not synchronized to the low-level vorticity in the pre-Megi disturbance. Figures 4b and 5a indicate that, as the mid-level vorticity increased rapidly in the period from ~0600−1800 UTC 9 October, the low-level vorticity decreased considerably. At 1800 UTC 9 October, when the MV reached maximum intensity, the low-level vorticity was at its minimum and anticyclonic circulation appeared in the near-surface levels (Figs. 3d, 4b and 5a).

  • As the mid-level vorticity increased from about 1800 UTC 8 to 1800 UTC 9 October, the low-level thermodynamic characteristics below it also varied considerably (Figs. 4c and 6). From Fig. 4c, we can see that a distinct cold core developed in the lower troposphere below the MV at ~0000 UTC 9 October, about 6 h after the deep convection burst in the pre-Megi disturbance. In the framework of balanced dynamics, the cold core was coupled with the MV to satisfy the thermal wind relationship. However, the cold core weakened after ~0600 UTC 9 October and a warm core emerged in the lower troposphere below the MV. At the same time, the lower troposphere dried considerably below the MV (Fig. 4c and 6b). At around 1800 UTC 9 October when the mid-level vorticity reached the maximum, a warm and dry core appeared below the MV (Fig. 4c). To understand the thermodynamic variations below the MV, the potential temperature and moisture budgets were calculated using the following equations:

    Figure 6.  (a) Time series of 1000-hPa mean equivalent potential temperature (black; units: K), potential temperature (blue; units: K) and specific humidity (red; units: g kg−1) averaged over the region with 600-hPa re-gridded vorticity greater than 2 × 10−5 s−1 and radius less than 4° during the period from 0600 UTC 8 to 0000 UTC 11 October. (c) As in (a) but for the potential temperature tendency (black; units: 10−4 K s−1), and its components induced by vertical advection (blue), horizontal advection (green) of potential temperature, diabatic heating (red), as well as boundary-layer processes and turbulent diffusion and air−sea interaction (orange). (e) As in (c) except for the specific humidity tendency (units: 10−4 g kg−1 s−1), and its components induced by vertical advection (blue), horizontal advection (green) of specific humidity, phase transition of vapor (red), as well as boundary-layer processes and turbulent diffusion air−sea interaction (orange). (b, d, f) As in (a, c, e) but for the variables at 850 hPa.

    where $ \theta $ and $ {q}_{\mathrm{v}} $ denote the potential temperature and water vapor mixing ratio, V and C are same as those in Eq. (3), w is the vertical velocity, $ {\dot{Q}}_{1} $ ($ {\dot{Q}}_{\mathrm{v}1} $) is the heating rate (moistening rate) related to diabatic heating (water phase transitions), and $ {\dot{Q}}_{2} $ ($ {\dot{Q}}_{\mathrm{v}2} $) is the heating rate (moistening rate) associated with the other physical processes except microphysics. $ {\dot{Q}}_{1} $ ($ {\dot{Q}}_{\mathrm{v}1} $) and $ {\dot{Q}}_{2} $ ($ {\dot{Q}}_{\mathrm{v}2} $) are derived from the model outputs directly. The local tendency term was computed from the model output with a time interval of 1 h. From Figs. 6c and d, we can see that the evaporative cooling made an important contribution to the decrease in potential temperature and the occurrence of negative potential temperature anomalies before 0600 UTC 9 October. After 0900 UTC 9 October, the near-surface cold advection and evaporative cooling reduced rapidly, and hence the potential temperature under the MV at 1000 hPa increased under the impact of surface enthalpy fluxes (Figs. 6a and c). Different from that at 1000 hPa, the increase in potential temperature at 850 hPa was primarily attributable to the adiabatic warming related to the remarkable descent dominant in the region below the MV (Figs. 5a-d and Fig. 6d). Along with the downward motion, the relatively dry air at higher levels was brought to the lower troposphere and contributed to the significant moisture reduction in the lower troposphere, especially at about 850 hPa, below the MV (Figs. 6e and f). As a result, the mean sounding in the pre-Megi disturbance manifested as an onion shape (Fig. 7a), indicating that the warm and dry core developed beneath the MV (Fig. 4c), which is usually related to the downdrafts in the sub-saturated environment as suggested by Zipser (1977).

    Figure 7.  Skew-T log-P vertical sounding averaged over the center area of pre-Megi’s disturbance. The temperature and dewpoint temperature sounding are denoted by the thick solid and cyan solid lines, respectively. The red dashed line delineates the lifted parcel’s ascent path.

    After the first episode of deep convection and subsequent stratiform precipitation, a well-defined MV developed in the pre-Megi disturbance at around 1800 UTC 9 October. However, the MV was characterized by near-surface anticyclonic vorticity (Fig. 3d and Fig. 4b) and an onion-shaped sounding (Fig. 7a), both of which are unfavorable for deep moist convection and low-level vorticity enhancement. But with the dissipation of the stratiform precipitation, the shallow dry convection soon developed in the pre-Megi’s central area (Fig. 4a), corresponding to the mid-level divergence and thus the weakening of the MV (Figs. 4b and 5a). Due to the vertical advection effect associated with the convection, the potential temperature decreased and the moisture increased in the lower troposphere (Figs. 6b and d). After ~1800 UTC 9 October, the warm and dry core began to transform into a cold anomaly between 900 and 600 hPa below the weakening MV (Figs. 4c and 6b). The cold anomaly situated above the near-surface warm anomaly induced by the sea-surface fluxes enhanced the convective available potential energy (CAPE) (Figs. 4c, 6c, 7b and 8). Meanwhile, the convection inhibition (CIN) decreased due to the low-level humidification (Figs. 7 and 8). All these processes promoted the outbreak of the moist convection after ~0000 UTC 10 October.

    Figure 8.  Evolution of CAPE (black line; units: J kg−1) and CIN (red line with ordinate on the right; units: J kg−1) averaged over a 200-km radius of the disturbance center.

    Figure 9 illustrates the development and organization of the convection under the favorable condition in the period from 2300 UTC 9 October to 0600 UTC 10 October. At 2300 UTC 9 October, the MV was coupled with a negative potential temperature anomaly ($ {\theta }' $) in the lower troposphere with the minimum at about 850−750 hPa (Fig. 9d). Meanwhile, positive $ {\theta }' $ appeared in the near-surface under the effect of surface heat fluxes (Fig. 6c and Figs. 9a and d). Such stratification destabilized the atmosphere and thus favored the convective activity (Fig. 8). As a result, moist convection burst in the pre-Megi’s central area (Fig. 9b). The associated evaporative cooling led to the formation of the near-surface cold pool. From Fig. 9e, one can see the evident forced uplifts induced by the combined effect of the cold pool and the local vertical wind shear, which helped trigger the convection. This is in accordance with Davis (2015) and Wu and Fang (2019). Along with the convection outbreak, the near-surface cold pool enlarged and more convection burst on its downshear side (Figs. 9c and f). From the above discussion, we can see that the cold core related to the MV benefited the increase in CAPE, while the near-surface cold pool caused by the convection and the vertical wind shear played an important role in triggering and organizing the convection. As the convection continuously transported moisture upwards (Fig. 6f), the mid-troposphere was considerably humidified (Figs. 6b and 7c), which facilitated the massive deep intense convection burst after ~1200 UTC 10 October (Figs. 4a and c).

    Figure 9.  (a−c) $ {\theta }' $ (shading; units: K) at 1000 hPa, vertical wind shear (vectors) between 850 hPa and 200 hPa, relative vorticity (white lines; units: 10−5 s−1) at 600 hPa, and vertical velocity at 850 hPa (red lines; units: m s−1) in the region centered at pre-Megi’s center with side length of ~450 km at 2300 UTC 9 October, 0300 UTC 10 October, and 0600 UTC 10 October, respectively. The dashed circles center at pre-Megi’s center with the radius of 200 km. The values of vorticity and vertical velocity are 2, 4, 8, 12, and 16 × 10−5 s−1, and 0.3 and 0.4 m s−1, respectively, in (a−c). The black box in (a−c) outlines the cross section of (d−f). (d−f) Vertical cross section showing the $ {\theta }' $ (shading; units: K), wind vectors (vectors; units: m s−1), and vorticity (white lines; units: 10−5 s−1). Vertical velocity has been scaled by a factor of 10. The contour intervals are 20 × 10−5 s−1 starting from zero and negative values are ignored. The locations of the cross section appear as black boxes in (a−c).

  • From 0600 UTC 9 October, the deep convection in the pre-Megi disturbance was characterized by diurnal variations (Fig. 4a). Between the intervals of the deep convection, the vertical distribution of the vertical velocity, horizontal divergence, vertical vorticity, relative humidity and potential temperature perturbation shown in Fig. 4 indicate that stratiform precipitation occurred, albeit less evident than that in the period from 0600 to 1800 UTC 9 October. Correspondingly, the MV demonstrated fluctuating growth after 0600 UTC 10 October (Figs. 4b and 5). From 1200 UTC 10 to ~0000 UTC 11 October, the deep intense convection and stratiform precipitation occurred sequentially in the pre-Megi disturbance, and the MV intensified. In the following 12 h, the mid-level divergence of vorticity fluxes related to the shallow convection caused the MV to weaken. As the deep convection burst again after ~1200 UTC 11 October, the MV re-intensified. Due to the relatively weak shallow convection at around 0000 UTC 12 October, the MV did not exhibit evident weakening. After ~0000 UTC 12 October, as massive deep intense convection appeared almost at the same time as the shallow moist convection, and the amount of deep intense convection was much greater than that in the previous two episodes of convection (Fig. 4a), the advective vorticity flux term dominated the positive tendency of the mid-level vorticity while the contribution from the non-advective vorticity flux was almost close to zero and negligible. As a result, the mid-level vorticity increased sharply in terms of magnitude towards genesis. It is worth mentioning that such an increase was primarily attributable to the horizontal convergence of the vorticity fluxes, and the contribution of the non-advective flux term was trivial because the vertical vorticity advection was nearly offset by the tilting effect.

4.   The role of the MV in the formation of Megi
  • Figures 4b and 5a show that the formation of the MV in stage 2 was accompanied by the weakening of the near-surface cyclonic circulation, and the rapid intensification of the MV in stage 3 usually corresponded to the decrease in the low-level vorticity in the pre-Megi disturbance. Such a relationship between the MV and the low-level vorticity means that the development of the MV was probably unfavorable for the spin-up of the low-level cyclonic circulation and thus the formation of Megi. However, as the MV formed, it did benefit the accumulation of the CAPE and favor the convective activity. Thus, it seems that the MV played dual roles in the formation of Megi. To examine which role dominated Megi’s genesis, a sensitivity experiment was carried out in this study.

    The sensitivity experiment shared the same configuration as the control experiment described in section 2 (CNTRL) except that the evaporative cooling below 600 hPa was disabled in the period from 0000 UTC to 1800 UTC 9 October (NoEvp). Figure 10a displays the time−height diagrams of model-derived frequency of convection top height and the mean vertical velocity averaged over the center area of the pre-Megi disturbance derived from the sensitivity experiment. Due to the lack of evaporative cooling below 600 hPa, the strong subsidence did not appear after 0600 UTC 9 October. Accordingly, the mid-level convergence and thus the MV were weaker in NoEvp than that in CNTRL at ~1200 UTC 9 October (Figs. 4b and 10b). After 1200 UTC 9 October, both the MVs in NoEvp and CNTRL intensified, but the former was mainly associated with the deep convection (Figs. 10a and b), while the latter was related to the stratiform precipitation (Figs. 4a and b). Although the area-mean vorticity in NoEvp also attained a maximum at 1800 UTC 9 October, it was much smaller than that in CNTRL. However, the low-level vorticity did not experience a significant decrease in NoEvp in the period from 0600 to 1800 UTC 9 October. After 1800 UTC 9 October, when the evaporative cooling was turned on, the stratiform precipitation began to occur in the pre-Megi disturbance (Figs. 10a-c). Due to the low-level divergence associated with the stratiform precipitation, the low-level positive vorticity decreased. However, the vorticity below 600 hPa was still much larger in NoEvp than that in CNTRL at 1200 UTC 10 October. Accordingly, the deep convection was more active in NoEvp than that in CNTRL after 1200 UTC 10 October (Fig. 4a). With the strong stretching effects of the convection, the vorticity associated with the disturbance in NoEvp developed much more quickly than that in CNTRL (Figs. 10b and 4b). In addition, the more active deep convection caused the warm core in the upper levels to become more distinct in NoEvp than that in CNTRL (Figs. 4c and 10c), which also facilitated the intensification of the low-level cyclonic circulation.

    Figure 10.  As in Fig. 4 but for NoEvp.

    The above analysis implies that the disturbance in NoEvp may develop much more quickly than that in CNTRL. This is further demonstrated in the evolution of the maximum tangential velocity at 10 m for CNTRL and NoEvp (figure not shown). If the time when the 10-m maximum tangential velocity exceeds 17.5 m s−1 was taken as the TC genesis time, then genesis took place at about 0400 UTC 13 and 1600 UTC 11 October in CNTRL and NoEvp, respectively. This indicates that, although the thermodynamic structure associated with the MV destabilized the atmosphere and facilitated the development of the convection when the MV was well-defined, the processes leading to the formation of the MV acted to hinder the formation of Megi through the strong divergence taking place near the surface.

5.   Concluding remarks
  • With a successful high-resolution simulation, the formation of Megi was investigated in this study with a focus on the evolution of the MV. Associated with the passage of a Kelvin wave, there was an outbreak of deep convection in the pre-Megi disturbance. The deep convection and subsequent stratiform precipitation induced an increase of the mid-level vorticity and hence the formation of the MV, which was the result of the mid-level convergence associated with the stratiform precipitation together with the upward vorticity transport related to the convection. As the MV reached the maximum intensity, the near-surface circulation under it was anticyclonic and the lower troposphere was warm and dry, which were unfavorable for convective activity. However, the subsequent adjustment of the dynamic and thermodynamic fields in the pre-Megi disturbance promoted the occurrence of the shallow convection. The shallow convection humidified and cooled the lower troposphere and induced mid-level divergence leading to the weakening of the MV. As a result, an MV coupled with the low-level cold core developed. The cold core favored the convection by increasing the CAPE. As the second episode of deep convection occurred, the MV experienced fluctuating growth along with the diurnal cycle of deep convection. It usually strengthened in the deep convection and the following stratiform precipitation phase while weakening in the shallow convection phase.

    The MV played dual roles in the formation of Megi. On the one hand, the formation of MV corresponded to the remarkable decrease in low-level vorticity, which was unfavorable for the intensification of the near-surface cyclonic circulation. On the other hand, the cold core coupled with the MV benefited the accumulation of the CAPE, which was favorable for the convection. The sensitivity experiment with the evaporative cooling turned off indicated that the processes leading to the formation of MV retarded the genesis process of Megi. Evidence for this is that the disturbance in the sensitivity experiment intensified much quicker than that in the control experiment. It is worth noting that the result is not contradictory to the widely accepted view that the MV is a precursor for the formation of a TC. This work not only pays attention to the role of the MV in the period when it is well-defined, but also to the impact of the MV forming processes on the formation of Megi. From Figs. 4b and 5a, we can see that the formation of the MV (~0600−1800 UTC 9 October) was accompanied by low-level divergence and a decrease in low-level vorticity. From this perspective, the formation of the MV hindered the genesis. However, if we only focus on the period when the MV was well-defined, we can see that the MV was always situated in the center area of pre-Megi (Figs. 3d-l), and the coupled low-level cold core under it favored the accumulation of CAPE, which was important to the convection and then the spin-up of low-level cyclonic circulation. Since the MV was always situated in the pre-Megi’s center area after it formed, it could be regarded as the precursor disturbance of Megi, just as it is in the literature.

    In this work, we found that the fluctuating growth of the MV after it formed was primarily induced by the diurnal cycle of the deep convection. Although the diurnal cycle of the deep convection in tropical cyclogenesis has been mentioned in many previous works, the reason responsible for this phenomenon is still not clear. The preliminary results derived from the sensitivity experiment without the solar radiation indicate that, even though the solar radiation was totally turned off, the diurnal cycle of the deep convection could still happen in the pre-Megi disturbance. Therefore, this issue is an interesting topic that deserves further discussion in the future. In addition, the sensitivity experiment performed in this work was not really physically realistic because there is always some degree of evaporative cooling in the real atmosphere. It is necessary to design more reasonable sensitivity experiments to discuss the influence of the MV on tropical cyclogenesis in the future.

    Acknowledgements. The authors are grateful for the constructive comments provided by Dr. Fuqing ZHANG and the three anonymous reviewers. This work was supported in part by the National Key Research and Development Program of China (Grant No. 2017YFC150 1601) and the National Natural Science Foundation of China (Grant No. 41875067). Computations were performed at the Texas Advanced Computing Center and at the High Performance Computing Center of Nanjing University.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return