Advanced Search
Article Contents

Asymmetric Distribution of Convection in Tropical Cyclones over the Western North Pacific Ocean


doi: 10.1007/s00376-016-5277-x

  • Forecasts of the intensity and quantitative precipitation of tropical cyclones (TCs) are generally inaccurate, because the strength and structure of a TC show a complicated spatiotemporal pattern and are affected by various factors. Among these, asymmetric convection plays an important role. This study investigates the asymmetric distribution of convection in TCs over the western North Pacific during the period 2005-2012, based on data obtained from the Feng Yun 2 (FY2) geostationary satellite. The asymmetric distributions of the incidence, intensity and morphology of convections are analyzed. Results show that the PDFs of the convection occurrence curve to the azimuth are sinusoidal. The rear-left quadrant relative to TC motion shows the highest occurrence rate of convection, while the front-right quadrant has the lowest. In terms of intensity, weak convections are favored in the front-left of a TC at large distances, whereas strong convections are more likely to appear to the rear-right of a TC within a 300 km range. More than 70% of all MCSs examined here are elongated systems, and meso-β enlongated convective systems (MβECSs) are the most dominant type observed in the outer region of a TC. Smaller MCSs tend to be more concentrated near the center of a TC. While semi-circular MCSs [MβCCSs, MCCs (mesoscale convective complexes)] show a high incidence rate to the rear of a TC, elongated MCSs [MβECSs, PECSs (persistent elongated convective systems)] are more likely to appear in the rear-right quadrant of a TC within a range of 400 km.
  • 加载中
  • Anderson C. J., R. W. Arritt, 1998: Mesoscale convective complexes and persistent elongated convective systems over the United States during 1992 and 1993. Mon. Wea. Rev., 126, 578- 599.fbcd771cfc330e92c6966096e7cff180http%3A%2F%2Fadsabs.harvard.edu%2Fcgi-bin%2Fnph-data_query%3Fbibcode%3D1998MWRv..126..578A%26db_key%3DPHY%26link_type%3DABSTRACThttp://xueshu.baidu.com/s?wd=paperuri%3A%28986d8c733543b686bd2bd01262e73930%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fadsabs.harvard.edu%2Fcgi-bin%2Fnph-data_query%3Fbibcode%3D1998MWRv..126..578A%26db_key%3DPHY%26link_type%3DABSTRACT&ie=utf-8&sc_us=3250868996915071335
    Augustine J. A., K. W. Howard, 1988: Mesoscale convective complexes over the United States during 1985. Mon. Wea. Rev., 116, 685- 701.10.1175/1520-0493(1988)116<0685:MCCOTU>2.0.CO;2453ffcf055e2f965a6529ff73f4178cbhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1988MWRv..116..685Ahttp://adsabs.harvard.edu/abs/1988MWRv..116..685AAbstract Infrared imagery from GOES was used to document mesoscale convective complexes (MCCs) over the United States during 1986 and 1987. A near-record 58 MCCs occurred in 1986, and 44 occurred in 1987. Although these totals were above average relative to MCC numbers of the 7 years prior to 1985, seasonal distributions for both years were atypical. Particularly, each had an extended period (3 weeks) when no MCCs occurred in late spring and early summer, a time when the mean MCC seasonal distribution peaks. This peculiarity was investigated by comparing mean large-scale surface and upper-air environments of null- and active-MCC periods of both years. Results confirmed the primary importance to MCC development of strong low-level thermal forcing, as well as proper vertical phasing of favorable lower- and midtropospheric environments. A cursory survey of MCCs documented outside of the United States reveals that MCCs, or MCC-type storms, are a warm-season phenomenon of midlatitude, subtropical, and low-lat...
    Barnes G. M., E. J. Zipser, D. Jorgensen, and F. Marks, 1983: Mesoscale and convective structure of a hurricane rainband. J. Atmos. Sci., 40, 2125- 2137.10.1175/1520-0469(1983)0402.0.CO;21d86c9125e56ca4250b292af6fd9ff7ehttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1983JAtS...40.2125Bhttp://adsabs.harvard.edu/abs/1983JAtS...40.2125BThe mesoscale thermodynamic, kinematic, and radar structure of a Hurricane Floyd rainband observed on 7 September 1981 is presented. Data are from 26 aircraft passes through the rainband from 150 to 6400 m. A composite technique which presents rainband structure as a function of distance from the storm circulation center reveals inflow from the outer edge of the band and a partial barrier to this flow below 3 km. In the direction parallel to rainband orientation, radar reveals cellular reflectivity structure on the upwind and central portions of the rainband; the frequency of cellular precipitation decreases in favor of stratiform precipitation further downwind as the band spirals gradually towards the eyewall. In the radial direction, a decrease of 12 K in -, is observed across the rainband in the subcloud layer. Convective scale up- and downdrafts that are associated with cellular reflectivity structure are hypothesized to be responsible for the thermodynamic modification of the cloud and subcloud layers.
    Barnes G. M., J. F. Gamache, M. A. LeMone, and G. J. Stossmeister, 1991: A convective cell in a hurricane rainband. Mon. Wea. Rev., 119, 776- 794.10.1175/1520-0493(1991)1192.0.CO;29c120f6f37053a4bec722e978fe88507http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1991MWRv..119..776Bhttp://adsabs.harvard.edu/abs/1991MWRv..119..776BOn 10 October 1983 the two NOAA WP-3D aircraft completed a mission designed to provide airborne Doppler radar data for a convective cell embedded in a weak rainband on the trailing side of Hurricane Raymond. Comparisons of the wind field produced from the pseudo-dual-Doppler radar technique with in situ wind measurements suggest that the larger convective-scale feature may be resolved if the sampling time is kept to a minimum. The convective cell was found to move downband faster than any environmental winds but slightly slower than the winds found in the reflectivity core that delineates the cell. In the core of the cell the tangential wind is increased and the radial inflow turns to outflow with respect to the circulation center. The flow field demonstrates that the downband stratiform portion of a rainband is not from cells currently active since the updraft detrains upwind relative to the cell but rather it is due to the fallout from ice particles placed into the upper troposphere by clouds that have since dissipated. The mass flux of this cell is estimated to be 5%-10% of the mass flux accomplished by an eyewall of a moderate tropical cyclone. This finding supports the concept that large, convectively active rainbands have a major effect on the subcloud layer air flowing toward the eyewall.
    Blank M. L., J. F. Gamache, F. D. Marks, C. E. Samsury, and H. E. Willoughby, 2002: Eastern Pacific hurricanes Jimena of 1991 and Olivia of 1994: The effect of vertical shear on structure and intensity. Mon. Wea. Rev., 130, 2291- 2312.10.1175/1520-0493(2002)1302.0.CO;281f65226a8dd11da7a97d0e8513ea865http%3A%2F%2Fci.nii.ac.jp%2Fnaid%2F80015453904http://ci.nii.ac.jp/naid/80015453904Abstract Shear is a key inhibitor of tropical cyclone intensification. Although its signature is readily recognized in satellite imagery and theoretical or modeling studies provide some insight, detailed observations have been limited. Airborne radar and in situ observations in Hurricanes Jimena of 1991 and Olivia of 1994 are a step toward better understanding. Each storm was observed on two consecutive days. Initially, both had small eyes, 16-18-km radius, and maximum winds of 6557 m s611 over sea surface temperatures (SST) >28C in easterly environmental shear. Jimena maintained constant intensity or weakened gradually for 2 days in 13-20 m s611 easterly shear. Olivia intensified in 8 m s611 shear on the first day. Overnight, the shear diminished to reverse and became westerly. On the second day, Olivia weakened as the shear increased to >15 m s611 from the west, the storm moved over cooler SST, and became surrounded by dryer air. As convection weakened and the outer rainbands ceased to be effective barriers...
    Boyd J. P., 2001: Chebyshev and Fourier Spectral Methods. 2nd ed., Dover Publications, Inc, 44 pp.10.1007/BF010248322222163197a49c1789db2d2ddba43390http%3A%2F%2Fwww.ams.org%2Fmathscinet-getitem%3Fmr%3D1874071http://www.ams.org/mathscinet-getitem?mr=1874071Completely revised text focuses on use of spectral methods to solve boundary value, eigenvalue, and time-dependent problems, but also covers Hermite, Laguerre, rational Chebyshev, sinc, and spherical harmonic functions, as well as cardinal functions, linear eigenvalue problems, matrix-solving methods, coordinate transformations, methods for unbounded intervals, spherical and cylindrical geometry, and much more. 7 Appendices. Glossary. Bibliography. Index. Over 160 text figures.Boyd, John P
    Burpee R. W., M. L. Black, 1989: Temporal and spatial variations of rainfall near the centers of two tropical cyclones. Mon. Wea. Rev., 117, 2204- 2218.10.1175/1520-0493(1989)1172.0.CO;27923545177889c538278c2f2ab40d027http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1989MWRv..117.2204Bhttp://adsabs.harvard.edu/abs/1989MWRv..117.2204BAbstract The Hurricane Research Division collected radar reflectivity data with a portable recorder attached to National Weather Service (NWS) WSR-57 radar as Hurricanes Alicia of 1983 and Elena of 1985 approached the coastline of the United States. The reflectivity data were used to estimate rain rates for the eyewall region, including the rain-free eye, and the rainbands in the annular area outside the eyewall, but within 75 km of the center of the eye. The rain rates include reflectivity corrections that were based upon the variation of average returned power with range in four hurricanes This study examines the temporal and spatial variations of rain rates in the cores of Hurricanes Alicia and Elena. In Alicia, variations of area-averaged rain rate (R) in the eyewall region were caused by the growth and decay of mesoscale convective areas. In Elena, the life cycles of individual convective cells also accounted for large changes in the eyewall R. In both hurricanes, the time series of R in the rainband...
    Carbone R. E., J. D. Tuttle, D. A. Ahijevych, and S. B. Trier, 2002: Inferences of predictability associated with warm season precipitation episodes. J. Atmos. Sci., 59, 2033- 2056.10.1175/1520-0469(2002)0592.0.CO;28856fd28ec3701ad597b9156f3635154http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2002JAtS...59.2033Chttp://adsabs.harvard.edu/abs/2002JAtS...59.2033CCiteSeerX - Scientific documents that cite the following paper: 2002: Inferences of Predictability Associated with Warm Season Precipitation Episodes
    Carvalho L. M. V., C. Jones, 2001: A satellite method to identify structural properties of mesoscale convective systems based on the maximum spatial correlation tracking technique (MASCOTTE). J. Appl. Meteor., 40, 1683- 1701.10.1175/1520-0450(2001)0402.0.CO;29b782e7bcc02b699e5e675d6436fb46bhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2001JApMe..40.1683Chttp://adsabs.harvard.edu/abs/2001JApMe..40.1683CA simple, fully automated, and efficient method to determine the structural properties and evolution (tracking) of cloud shields of convective systems (CS) is described. The method, which is based on the maximum spatial correlation tracking technique (MASCOTTE), is a new alternative to the existent techniques available for studies that monitor the evolution of CS using satellite images. MASCOTTE provides as CS structural properties the following parameters: mean and variance of brightness temperature, horizontal area, perimeter, minimum brightness temperature, fractional convective area, center of gravity, and fragmentation. The fragmentation parameter has the potential to monitor the evolution of the CS. A new way of estimating the orientation and eccentricity of CS is proposed and is based on the empirical orthogonal function analysis of CS pixel coordinates. The method includes an accurate detection of splitting and merging of convective systems, which is a critical step in the automated satellite CS life cycle determination. Based on the magnitudes of the spatial correlation between consecutive satellite images and the changes in horizontal areas of CS, MASCOTTE provides a simple and skillful technique to track the evolution of CS life cycles. The MASCOTTE methodology is applied to infrared satellite images during seven consecutive days of the Wet-Season Atmospheric Mesoscale Campaign of the Large-Scale Biosphere-Atmosphere Experiment and ground validation experiment of the Tropical Rainfall Measuring Mission in the Brazilian state of Rondia in the Amazon basin. The results indicate that MASCOTTE is a valuable approach to understanding the variability of CS.
    Chan J. C. L., K. S. Liu, S. E. Ching, and E. S. T. Lai, 2004: Asymmetric distribution of convection associated with tropical cyclones making landfall along the south China coast. Mon. Wea. Rev., 132, 2410- 2420.10.1007/s00703-006-0244-17dbeb86fd7740de77cb17bbf3a17fcaahttp%3A%2F%2Flink.springer.com%2Farticle%2F10.1007%2Fs00703-006-0244-1http://link.springer.com/article/10.1007/s00703-006-0244-1This study examines the convection distribution associated with 18 TCs that made landfall along the South China coast during 1995 and 2005. Cloud-top temperatures from high-resolution satellite imageries of the Geosynchronous Meteorological Satellite 5 are used as proxy of strong convection. It is found that convection tends to be enhanced on the western side of the TC as it makes landfall in 10 of the cases, in agreement with the conclusion of some previous studies. Four cases have stronger convection on the eastern side. This eviation- from the general rule appears to be related to the TCs being more slow-moving or their interaction of the TC with another land surface prior to its making landfall along the South China coast. For the remaining cases in which no significant asymmetries in convection can be identified, the vertical wind shear appears to enhance convection on the east side.
    Chen L. S., Z. Y. Meng, 2001: An overview on tropical cyclone research progress in China during the past ten years. Chinese J. Atmos. Sci., 25( 4), 420- 432. (in Chinese)9062e850033bbe76b9e3238f4065b78ehttp%3A%2F%2Fen.cnki.com.cn%2FArticle_en%2FCJFDTOTAL-DQXK200103012.htmhttp://en.cnki.com.cn/Article_en/CJFDTOTAL-DQXK200103012.htmAt the beginning of the 1990s, several large-scape tropical cyclone (TC) field experiments were launched internationally and domestically. Much valuable intensive observational data have been obtained. With these data, a series of studies on tropical cyclone especially on abnormal typhoons were carried out in China. Under the national key project related to tropical cyclone such as," Typhoon scientific operational experiment and studies on its synoptic and dynamic theories" and international cooperative project on typhoon named " SPECTRUM", TC sudden change phenomena and forecast techniques were investigated with emphasis in China. During the past 10 years, much improvements have been achieved in sudden changes of tropical cyclone motion, its structure and intensity, sudden intensification of typhoon caused heavy rain, TC track forecast methods etc. This paper will summarize briefly these developments.
    Chen S. S., J. A. Knaff, and F. D. Marks, 2006: Effects of vertical wind shear and storm motion on tropical cyclone rainfall asymmetries deduced from TRMM. Mon. Wea. Rev., 134( 11), 3190- 3208.f1a8d844a5fadaaaf81ef87ea78f8401http%3A%2F%2Fwww.bioone.org%2Fservlet%2Flinkout%3Fsuffix%3Di1551-5036-67-sp1-109-Chen1%26dbid%3D16%26doi%3D10.2112%252FSI_67_8%26key%3D10.1175%252FMWR3245.1http://xueshu.baidu.com/s?wd=paperuri%3A%28dd55ae827766f893be008cf310a409ba%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fwww.bioone.org%2Fservlet%2Flinkout%3Fsuffix%3Di1551-5036-67-sp1-109-Chen1%26dbid%3D16%26doi%3D10.2112%252FSI_67_8%26key%3D10.1175%252FMWR3245.1&ie=utf-8&sc_us=6627478276274112225
    Corbosiero K. L., J. Molinari, 2002: The effects of vertical wind shear on the distribution of convection in tropical cyclones. Mon. Wea. Rev., 130( 8), 2110- 2123.e22cc79d0c229d4224dff62a608075e5http%3A%2F%2Fadsabs.harvard.edu%2Fcgi-bin%2Fnph-data_query%3Fbibcode%3D2002MWRv..130.2110C%26db_key%3DPHY%26link_type%3DEJOURNALhttp://xueshu.baidu.com/s?wd=paperuri%3A%286cb3f7e9e9217596e45ee93403a962db%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fadsabs.harvard.edu%2Fcgi-bin%2Fnph-data_query%3Fbibcode%3D2002MWRv..130.2110C%26db_key%3DPHY%26link_type%3DEJOURNAL&ie=utf-8&sc_us=8130994636086979390
    Corbosiero K. L., J. Molinari, 2003: The relationship between storm motion, vertical wind shear, and convective asymmetries in tropical cyclones. J. Atmos. Sci., 60( 3), 366- 376.10.1175/1520-0469(2003)0602.0.CO;20a6ba1070a9ecb832ce5f4b6e6de8cf1http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2003JAtS...60..366Chttp://adsabs.harvard.edu/abs/2003JAtS...60..366CAbstract The influence of the direction of storm motion on the azimuthal distribution of electrified convection in 35 Atlantic basin tropical cyclones from 1985 to 1999 was examined using data from the National Lightning Detection Network. In the inner 100 km, flashes most often occurred in the front half of storms, with a preference for the right-front quadrant. In the outer rainbands (r = 100-300 km), flashes occurred predominantly to the right of motion, although the maximum remained in the right-front quadrant. The results are shown to be consistent with previous studies of asymmetries in rainfall, radar reflectivity, and vertical motion with respect to tropical cyclone motion. The motion effect has been attributed to the influence of asymmetric friction in the tropical cyclone boundary layer. The authors previously found a strong signature in the azimuthal distribution of lightning with respect to vertical wind shear. Because both effects show clearly, vertical wind shear and storm motion must themse...
    Demaria M., J. A. Knaff, and C. Sampson, 2007: Evaluation of long-term trends in tropical cyclone intensity forecasts. Meteor. Atmos. Phys., 97, 19- 28.10.1007/s00703-006-0241-4ffae03528c8bd1c2a7ad5eed49371785http%3A%2F%2Flink.springer.com%2Farticle%2F10.1007%2Fs00703-006-0241-4http://link.springer.com/article/10.1007/s00703-006-0241-4The National Hurricane Center and Joint Typhoon Warning Center operational tropical cyclone intensity forecasts for the three major northern hemisphere tropical cyclone basins (Atlantic, eastern North Pacific, and western North Pacific) for the past two decades are examined for long-term trends. Results show that there has been some marginal improvement in the mean absolute error at 24 and 4865h for the Atlantic and at 7265h for the east and west Pacific. A new metric that measures the percent variance of the observed intensity changes that is reduced by the forecast (variance reduction, VR) is defined to help account for inter-annual variability in forecast difficulty. Results show that there have been significant improvements in the VR of the official forecasts in the Atlantic, and some marginal improvement in the other two basins. The VR of the intensity guidance models was also examined. The improvement in the VR is due to the implementation of advanced statistical intensity prediction models and the operational version of the GFDL hurricane model in the mid-1990s. The skill of the operational intensity forecasts for the 5-year period ending in 2005 was determined by comparing the errors to those from simple statistical models with input from climatology and persistence. The intensity forecasts had significant skill out to 9665h in the Atlantic and out to 7265h in the east and west Pacific. The intensity forecasts are also compared to the operational track forecasts. The skill was comparable at 1265h, but the track forecasts were 2 to 5 times more skillful by 7265h. The track and intensity forecast error trends for the two-decade period were also compared. Results showed that the percentage track forecast improvement was almost an order of magnitude larger than that for intensity, indicating that intensity forecasting still has much room for improvement.
    Frank W. M., E. A. Ritchie, 1999: Effects of environmental flow upon tropical cyclone structure.Mon. Wea. Rev., 127( 9), 2044- 2061.10.1175/1520-0493(1999)1272.0.CO;2dca0dabfd3ad2709c4b756436bc71a0fhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1999MWRv..127.2044Fhttp://adsabs.harvard.edu/abs/1999MWRv..127.2044FAbstract Numerical simulations of tropical-cyclone-like vortices are performed to analyze the effects of unidirectional vertical wind shear and translational flow upon the organization of convection within a hurricanecore region and upon the intensity of the storm. A series of dry and moist simulations is performed using the Pennsylvania State University ational Center for Atmospheric Research Mesoscale Model version 5 (MM5) with idealized initial conditions. The dry simulations are designed to determine the patterns of forced ascent that occur as the vortex responds to imposed vertical wind shear and translational flow, and the mechanisms that modulate the vertical velocity field are explored. The moist simulations are initialized with the same initial conditions as the dry runs but with a cumulus parameterization and explicit moisture scheme activated. The moist simulations are compared to the dry runs in order to test the hypothesis that the forced vertical circulation modes modulate the convection...
    Frank W. M., E. A. Ritchie, 2001: Effects of vertical wind shear on the intensity and structure of numerically simulated hurricanes. Mon. Wea. Rev., 129( 9), 2249- 2269.5f1c156264b7fcd36d588789473feb48http%3A%2F%2Fadsabs.harvard.edu%2Fcgi-bin%2Fnph-data_query%3Fbibcode%3D2001MWRv..129.2249F%26db_key%3DPHY%26link_type%3DABSTRACThttp://xueshu.baidu.com/s?wd=paperuri%3A%28a4082a554b992bcf9b4aef8e3ef598a9%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fadsabs.harvard.edu%2Fcgi-bin%2Fnph-data_query%3Fbibcode%3D2001MWRv..129.2249F%26db_key%3DPHY%26link_type%3DABSTRACT&ie=utf-8&sc_us=16582585971990525617
    Houze, R. A. Jr., D. C. Wilton, B. F. Smull, 2007: Monsoon convection in the Himalayan region as seen by the TRMM Precipitation Radar. Quart. J. Roy. Meteor. Soc., 133, 1389- 1411.10.1002/qj.106028aa5645765e6ed5d8ba5e4acee6346http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.106%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1002/qj.106/fullThree-dimensional structure of summer monsoon convection in the Himalayan region and its overall variability are examined by analyzing data trom the Iropical Raintall Measuring Mission (TRMM) Precipitation Radar over the June-September seasons of 2002 and 2003. Statistics are compiled for both convective and stratiform components of the observed radar echoes. Deep intense convective echoes (40 dBZ echo reaching heights >10 km) occur primarily just upstream (south) of and over the lower elevations of the Himalayan barrier, especially in the northwestern concave indentation of the barrier. The deep intense convective echoes are vertically erect, consistent with the relatively weak environmental shear. They sometimes extend above 17 km, indicating that exceptionally strong updraughts loft graupel to high altitudes. Occasionally, scattered isolated deep intense convective echoes occur over the Tibetan Plateau. Wide intense convective echoes (40 dBZ echo >1000 km
    Jirak I. L., W. R. Cotton, and R. L. McAnelly, 2003: Satellite and radar survey of mesoscale convective system development. Mon. Wea. Rev., 131( 10), 2428- 2449.10.1175/1520-0493(2003)131<2428:SARSOM>2.0.CO;2767268647886f416135924604bd27a32http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2003MWRv..131.2428Jhttp://adsabs.harvard.edu/abs/2003MWRv..131.2428JAn investigation of several hundred mesoscale convective systems (MCSs) during the warm seasons (April-August) of 1996-98 is presented. Circular and elongated MCSs on both the large and small scales were classified and analyzed in this study using satellite and radar data. The satellite classification scheme used for this study includes two previously defined categories and two new categories: mesoscale convective complexes (MCCs), persistent elongated convective systems (PECSs), meso-circular convective systems (MCSs), and meso- elongated convective systems (MCSs). Around two-thirds of the MCSs in the study fell into the larger satellite-defined categories (MCCs and PECSs). These larger systems produced more severe weather, generated much more precipitation, and reached a peak frequency earlier in the convective season than the smaller, meso-systems. Overall, PECSs were found to be the dominant satellite-defined MCS, as they were the largest, most common, most severe, and most prolific precipitation-producing systems. In addition, 2-km national composite radar reflectivity data were used to analyze the development of each of the systems. A three-level radar classification scheme describing MCS development is introduced. The classification scheme is based on the following elements: presence of stratiform precipitation, arrangement of convective cells, and interaction of convective clusters. Considerable differences were found among the systems when categorized by these features. Grouping systems by the interaction of their convective clusters revealed that more than 70% of the MCSs evolved from the merger of multiple convective clusters, which resulted in larger systems than those that developed from a single cluster. The most significant difference occurred when classifying systems by their arrangement of convective cells. In particular, if the initial convection were linearly arranged, the mature MCSs were larger, longer-lived, more severe, and more effective at producing precipitation than MCSs that developed from areally arranged convection.
    Johnson R. H., S. L. Aves, P. E. Ciesielski, and T. D. Keenan, 2005: Organization of oceanic convection during the onset of the 1998 East Asian summer monsoon. Mon. Wea. Rev., 133( 2), 131- 148.10.1175/MWR-2843.1a26cfcb2a19fb97f9a4dbaceb3cf5ea1http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2005MWRv..133..131Jhttp://adsabs.harvard.edu/abs/2005MWRv..133..131JAbstract The organizational modes of convection over the northern South China Sea (SCS) during the onset of the summer monsoon are documented using radar and sounding data from the Mayune 1998 South China Sea Monsoon Experiment (SCSMEX). The onset occurred in mid-May with a rapid increase in deep convection over a 10-day period, accompanied by a major shift in the circulation over the east Asian region. Analysis of Bureau of Meteorology Research Centre (BMRC) radar data from Dongsha Island reveals a wide range of organizational modes of convection over the northern SCS. Proximity sounding data indicate that lower- and middle-level vertical wind shears exerted a dominant control over the orientation of convective lines within mesoscale convective systems in this region, as has been found in the Australian monsoon region and the equatorial western Pacific. The results are consistent with the conceptual model of LeMone et al. based on the Tropical Ocean Global Atmosphere Coupled Oceantmosphere Response E...
    Kane, R. J. Jr., C. R. Chelius, J. M. Fritsch, 1987: Precipitation characteristics of mesoscale convective weather systems. J. Appl. Meteor., 26, 1345- 1357.10.1175/1520-0450(1987)0262.0.CO;26f24fb94025ffb71b6e3c3c2b4697bc7http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1987JApMe..26.1345Khttp://adsabs.harvard.edu/abs/1987JApMe..26.1345KPrecipitation from 74 mesoscale convective complexes is examined to determine the total precipitation, areal extent, and characteristic precipitation pattern of an average convective complex. The relationship between the average precipitation pattern and the track of the centroid of the satellite-observed, cold-cloud shield is determined as an aid to forecasting. The amount and spatial distribution of precipitation during each stage (i.e., initiation, maturation and dissipation) of the average convective system's life cycle are presented, as well as the precipitation patterns for systems that form in particular synoptic environments. The precipitation characteristics of MCCs are compared to those from 32 other convective weather systems that are similar to MCCs but do not meet all the MCC-definition criteria.
    Lee C. S., B. F. Chen, and R. L. Elsberry, 2012: Long-lasting convective systems in the outer region of tropical cyclones in the western North Pacific. J. Geophys. Res., 39, L21812.d952ebc7628e8d097d56ad7965b692dfhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2012GL053685%2Ffullhttp://xueshu.baidu.com/s?wd=paperuri%3A%28a1161a80875972d6b98103f113d555b0%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2012GL053685%2Ffull&ie=utf-8&sc_us=11216735214274718622
    Lee C. S., K. K. W. Cheung, J. S. N Hui, and R. L. Elsberry, 2008: Mesoscale features associated with tropical cyclone formations in the Western North Pacific. Mon. Wea. Rev., 136, 2006- 2022.10.1175/2007MWR2267.1f291a8796294ddddb1d68dee81d3444bhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2008MWRv..136.2006Lhttp://adsabs.harvard.edu/abs/2008MWRv..136.2006LThe mesoscale features of 124 tropical cyclone formations in the western North Pacific Ocean during 1999-2004 are investigated through large-scale analyses, satellite infrared brightness temperature (TB), and Quick Scatterometer (QuikSCAT) oceanic wind data. Based on low-level wind flow and surge direction, the formation cases are classified into six synoptic patterns: easterly wave (EW), northeasterly flow (NE), coexistence of northeasterly and southwesterly flow (NEW), southwesterly flow (SW), monsoon confluence (MC), and monsoon shear (MS). Then the general convection characteristics and mesoscale convective system (MCS) activities associated with these formation cases are studied under this classification scheme. Convection processes in the EW cases are distinguished from the monsoon-related formations in that the convection is less deep and closer to the formation center. Five characteristic temporal evolutions of the deep convection are identified: (i) single convection event, (ii) two convection events, (iii) three convection events, (iv) gradual decrease in TB, and (v) fluctuating TB, or a slight increase in TB before formation. Although no dominant temporal evolution differentiates cases in the six synoptic patterns, evolutions ii and iii seem to be the common routes taken by the monsoon-related formations. The overall percentage of cases with MCS activity at multiple times is 63%, and in 35% of cases more than one MCS coexisted. Most of the MC and MS cases develop multiple MCSs that lead to several episodes of deep convection. These two patterns have the highest percentage of coexisting MCSs such that potential interaction between these systems may play a role in the formation process. The MCSs in the monsoon-related formations are distributed around the center, except in the NEW cases in which clustering of MCSs is found about 100-200 km east of the center during the 12 h before formation. On average only one MCS occurs during an EW formation, whereas the mean value is around two for the other monsoon-related patterns. Both the mean lifetime and time of first appearance of MCS in EW are much shorter than those developed in other synoptic patterns, which indicates that the overall formation evolution in the EW case is faster. Moreover, this MCS is most likely to be found within 100 km east of the center 12 h before formation. The implications of these results to internal mechanisms of tropical cyclone formation are discussed in light of other recent mesoscale studies.
    Li Q. Q., Y. H. Duan, 2013: Sensitivity of quasi-periodic outer rainband activity of tropical cyclones to the surface entropy flux. Acta Meteorologica Sinica, 27( 5), 636- 657.10.1007/s13351-013-0502-3bd4ab277fbf21c53a5124d2b3997b225http%3A%2F%2Fwww.cqvip.com%2FQK%2F88418X%2F201305%2F47640916.htmlhttp://d.wanfangdata.com.cn/Periodical/qxxb-e201305003The influence of outer-core surface entropy fluxes(SEFs)on tropical cyclone(TC)outer rainband activity is investigated in this study with a fully compressible,nonhydrostatic model.A control simulation and two sensitivity experiments with the outer-core SEF artificially increased and decreased by 20%respectively were conducted to examine the quasi-periodic outer rainband behavior.Larger negative horizontal advection due to the greater radial wind and the positive contribution by asymmetric eddies leads to a longer period of outerrainband activity in the SEF-enhanced experiment.The well-developed outer rainbands in the control and SEF-reduced simulations significantly limit the TC intensity,whereas such an intensity suppression influence is not pronounced in the SEF-enhanced experiment.As diabatic heating in outer rainbands strengthens the outer-core tangential wind,the quasi-periodic activity of outer rainbands contributes to the quasi-periodic variations of the inner-core size of the TCs.
    Lonfat, M, F. D. Marks, S. S. Chen, 2004: Precipitation distribution in tropical cyclones using the tropical rainfall measuring mission (TRMM) Microwave Imager: A global perspective. Mon. Wea. Rev., 132( 7), 1645- 1660.347eb0738a0a959a254e5c3dd8c618f0http%3A%2F%2Fwww.bioone.org%2Fservlet%2Flinkout%3Fsuffix%3Di1551-5036-67-sp1-109-Lonfat1%26dbid%3D16%26doi%3D10.2112%252FSI_67_8%26key%3D10.1175%252F1520-0493%282004%291322.0.CO%253B2http://xueshu.baidu.com/s?wd=paperuri%3A%28351df68365eab9fac2da1189798738b4%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fwww.bioone.org%2Fservlet%2Flinkout%3Fsuffix%3Di1551-5036-67-sp1-109-Lonfat1%26dbid%3D16%26doi%3D10.2112%252FSI_67_8%26key%3D10.1175%252F1520-0493%282004%291322.0.CO%253B2&ie=utf-8&sc_us=13096078972896950794
    Luo Y. L., R. H. Zhang, W. M. Qian, Z. Z. Luo, and X. Hu, 2011: Intercomparison of deep convection over the Tibetan Plateau-Asian Monsoon Region and subtropical North America in boreal summer using CloudSat/CALIPSO Data. J.Climate, 24, 2164- 2177.10.1175/2010JCLI4032.1bc23e0a1ffaa510210f301aa40ca3192http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2010AGUFM.A11D0074Lhttp://adsabs.harvard.edu/abs/2010AGUFM.A11D0074LDeep convection in the Tibetan Plateau-southern Asian monsoon region (TP-SAMR) is analyzed using CloudSat and Cloud-Aerosol Lidar and Infrared Pathfinder Satellite Observations (CALIPSO) data for the boreal summer season (June-August) from 2006 to 2009. Three subregions are definedhe TP, the southern slope of the plateau (PSS), and the SAMRnd deep convection properties (such as occurrence frequency, internal vertical structure, system size, and local environment) are compared among these subregions. To cast them in a broader context, four additional regions that bear some similarity to the TP-SAMR are also discussed: East Asia (EA), tropical northwestern Pacific (NWP), and western and eastern North America (WNA and ENA, respectively). The principal findings are as follows: 1) Compared to the other two subregions of the TP-SAMR, deep convection over the TP is shallower, less frequent, and embedded in smaller-size convection systems, but the cloud tops are more densely packed. These characteristics of deep convection over the TP are closely related to the unique local environment, namely, a significantly lower level of neutral buoyancy (LNB) and much drier atmosphere. 2) In a broader context in which all seven regions are brought together, deep convection in the two tropical regions (NWP and SAMR; mostly over ocean) is similar in many regards. A similar conclusion can be drawn among the four subtropical continental regions (TP, EA, WNA, and ENA). However, tropical oceanic and subtropical land regions present some significant contrasts: deep convection in the latter region occurs less frequently, has lower cloud tops but comparable or slightly higher tops of large radar echo (e.g., 0 and 10 dBZ), and is embedded in smaller systems. The cloud tops of the subtropical land regions are generally more densely packed. Hence, the difference between the TP and SAMR is more of a general contrast between subtropical land regions and tropical oceanic regions during the boreal summer. 3) Deep convection over the PSS possesses some uniqueness of its own because of the distinctive terrain (slopes) and moist low-level monsoon flow. 4) Results from a comparison between the daytime (1:30 p.m.) and nighttime (1:30 a.m.) overpasses are largely consistent with researchers' general understanding of the diurnal variation of tropical and subtropical deep convection.
    Maddox R. A., 1980: Mesoscale convective complexes. Bull. Amer. Meteor. Soc., 61, 1374- 1387.2e172054-7cd8-460f-a4b0-4c6b265664fd36bcf1b059922cd6608b9b43eef0ce5fhttp%3A%2F%2Fwww.mendeley.com%2Fresearch%2Fmesoscale-convective-complexes-6%2Frefpaperuri:(69df345c180d3808b3acab0596548a6d)http://www.mendeley.com/research/mesoscale-convective-complexes-6/
    May P. T., 1996: The organization of convection in the rainbands of Tropical Cyclone Laurence. Mon. Wea. Rev., 124( 5), 807- 815.10.1175/1520-0493(1996)1242.0.CO;20497a2404af49de21e5776f435e9428dhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1996MWRv..124..807Mhttp://adsabs.harvard.edu/abs/1996MWRv..124..807MA slow-moving weak tropical cyclone passed near Darwin, Australia, in December 1990. Rainbands were observed by a Doppler weather radar and a 50-MHz wind profiler for over 24 h. The principal bands were seen to be organized on two distinct scales. Bands of stratiform precipitation formed at a radius of about 100 km from the center of the storm and moved outward at about 6 m s. These decayed after they moved past Darwin over land. A distinct midlevel jet extended along the bands. Within the bands, convective lines formed at regular intervals, propagated against and outward with respect to the mean flow, and acted as a partial barrier to the radial inflow. Deep, active convection was confined to these lines. The vertical motion in the convection showed a distinct acceleration above the freezing level with measured updrafts of up to 10 m s. The convection elevated the tropopause height over the rainband. It is hypothesized that an inertia-gravity wave propagating from near the storm eye was responsible for triggering the convection within the lines. This hypothesis, although difficult to test, accounts for the propagation characteristics of the convective lines and offers an explanation of why similar features have not been seen in more intense storms.
    Merrill, R. T, 1988: Environmental influences on hurricane intensification. J. Atmos. Sci., 45( 11), 1678- 1687.10.1175/1520-0469(1988)0452.0.CO;29cd46d8e708cabfcad12497a7c948a95http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1985PhDT.......145Mhttp://adsabs.harvard.edu/abs/1985PhDT.......145MAbstract Although driven by internal processes, hurricanes are also regulated by conditions in their oceanic and atmospheric surroundings. Sea surface temperature determines an upper bound on the intensity of hurricanes, but most never reach this potential, apparently because of adverse atmospheric conditions. Winds measured by satellite cloud tracking, commercial aircraft, and rawinsondes are composited using a rotated coordinate system designed to preserve the asymmetries in the upper-tropospheric environment. Composites of upper-tropospheric environmental flows for intensifying and nonintensifying hurricanes for a five-year period are compared. Nonintensifying composites indicate stronger mean environmental flow relative to the hurricane motion, unidirectional flow over and near the hurricane center, and slightly weaker radial outflow and/or more pronounced anticyclonic flow surrounding the center in the upper troposphere.
    Miller B. I., 1958: Rainfall rates in Florida hurricanes. Mon. Wea. Rev., 86, 258- 264.10.1175/1520-0493(1958)0862.0.CO;22fffe263aa798ff64c46891b00a4ea21http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1958MWRv...86..258Mhttp://adsabs.harvard.edu/abs/1958MWRv...86..258MNot Available
    Rodgers E. B., S. W. Chung, and H. F. Pierce, 1994: A satellite observational and numerical study of precipitation characteristics in western North Atlantic tropical cyclones. J. Appl. Meteor., 33( 3), 129- 139.10.3724/SP.J.1077.2011.002717ca0b27f1e0bd4af74eb7e89045ef9d5http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1994JApMe..33..129Rhttp://adsabs.harvard.edu/abs/1994JApMe..33..129RAbstract Special Sensor Microwave/Imager (SSM/I) observations were used to examine the spatial and temporal changes of the precipitation characteristics of tropical cyclones. SSM/I observations were also combined with the results of a tropical cyclone numerical model to examine the role of inner-core diabatic heating in subsequent intensity changes of tropical cyclones. Included in the SSM/I observations were rainfall characteristics of 18 named western North Atlantic tropical cyclones between 1987 and 1989. The SSM/I rain-rate algorithm that employed the 85-GHz channel provided an analysis of the rain-rate distribution in greater detail. However, the SSM/I algorithm underestimated the rain rates when compared to in situ techniques but appeared to be comparable to the rain rates obtained from other satellite-borne passive microwave radiometers. The analysis of SSM/I observations found that more intense systems had higher rain rates, more latent heat release, and a greater contribution from heavier rain to...
    Shapiro L. J., 1983: The asymmetric boundary layer flow under a translating hurricane. J. Atmos. Sci., 40( 8), 1984- 1998.10.1175/1520-0469(1983)0402.0.CO;268420d0bf743638b9a63153fbc383a2ahttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1983JAtS...40.1984Shttp://adsabs.harvard.edu/abs/1983JAtS...40.1984SAn investigation is made of the role of the translation of a hurricane in determining the distribution of boundary layer winds and in the organization of convection. A slab boundary layer model of constant depth is used to analyze the steady flow under a specified translating symmetric vortex in gradient balance. A truncated spectral formulation is used, including asymmetries through wavenumber 2. The role of linear and nonlinear asymmetric effects in the determination of the boundary layer response is diagnosed. These effects am relevant to relatively slowly and rapidly translating hurricanes, respectively.The analysis is compared to observations of Hurricanes Frederic of 1979 and Allen of 1980, as well as to other observational and theoretical cures. Allen's translation speed was approximately twice that of Frederic. It is found that the simple boundary layer formulation simulates the qualitative features of the wind field observed in Frederic. The distribution of convection in Frederic and Allen compares favorably with boundary layer convergence diagnosed from the model.
    Wang Y. Q., 2009: How do outer spiral rainbands affect tropical cyclone structure and intensity? J. Atmos. Sci., 66, 1250- 1273.10.1175/2008JAS2737.1cd8d5604045d66cacadd3ba570056c28http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2009JAtS...66.1250Whttp://adsabs.harvard.edu/abs/2009JAtS...66.1250WA long-standing issue on how outer spiral rainbands affect the structure and intensity of tropical cyclones is studied through a series of numerical experiments using the cloud-resolving tropical cyclone model TCM4. Because diabatic heating due to phase changes is the main driving force of outer spiral rainbands, their effect on the tropical cyclone structure and intensity is evaluated by artificially modifying the heating and cooling rate due to cloud microphysical processes in the model. The view proposed here is that the effect of diabatic heating in outer spiral rainbands on the storm structure and intensity results mainly from hydrostatic adjustment; that is, heating (cooling) of an atmospheric column decreases (increases) the surface pressure underneath the column. The change in surface pressure due to heating in the outer spiral rainbands is significant on the inward side of the rainbands where the inertial stability is generally high. Outside the rainbands in the far field, where the inertial stability is low and internal atmospheric heating is mostly lost to gravity wave radiation and little is left to warm the atmospheric column and lower the local surface pressure, the change in surface pressure is relatively small. This strong radially dependent response reduces the horizontal pressure gradient across the radius of maximum wind and thus the storm intensity in terms of the maximum low-level tangential wind while increasing the inner-core size of the storm. The numerical results show that cooling in the outer spiral rainbands maintains both the intensity of a tropical cyclone and the compactness of its inner core, whereas heating in the outer spiral rainbands decreases the intensity but increases the size of a tropical cyclone. Overall, the presence of strong outer spiral rainbands limits the intensity of a tropical cyclone. Because heating or cooling in the outer spiral rainbands depends strongly on the relative humidity in the near-core environment, the results have implications for the formation of the annular hurricane structure, the development of concentric eyewalls, and the size change in tropical cyclones.
    Wingo M. T., D. J. Cecil, 2010: Effects of vertical wind shear on tropical cyclone precipitation.Mon. Wea. Rev., 138( 4), 645- 662.10.1175/2009MWR2921.144a8b93c9bc7a15621ac92034692b7b4http%3A%2F%2Fwww.cabdirect.org%2Fabstracts%2F20103129515.htmlhttp://www.cabdirect.org/abstracts/20103129515.htmlAbstract The response of the precipitation field for tropical cyclones in relation to the surrounding environmental vertical wind shear has been investigated using 6520 000 snapshots of passive-microwave satellite rain rates. Composites of mean rain rates, 95th percentile rain rates, and rain coverage were constructed to compare how the spatial distribution of the precipitation was organized under varying environmental shear. Results indicated that precipitation is displaced downshear and to the left (right for Southern Hemisphere) of the shear vector. The amplitude of this displacement increases with stronger shear. The majority of the asymmetry found in the mean rain rates is accounted for by the asymmetry in the occurrence of heavy rain. Although rain is common in all quadrants of the sheared tropical cyclones, heavy rain (8 mm h611 at the 6525-km scale) is comparatively rare in the upshear-right quadrant. It is shown that the effect that shear has on the rain field is nearly instantaneous. Strong wester...
    Wu C. C., T. H. Yen, Y. H. Kuo, and W. Wang, 2002: Rainfall simulation associated with Typhoon Herb (1996) near Taiwan. Part I: The topographic effect. Wea.Forecasting, 17( 5), 1001- 1015.10.1175/1520-0434(2003)0172.0.CO;21f69329d23d292946f3d9e71038cfbdehttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2002WtFor..17.1001Whttp://adsabs.harvard.edu/abs/2002WtFor..17.1001WIn this study, a series of numerical experiments are performed to examine the ability of a high-resolution mesoscale model to predict the track, intensity change, and detailed mesoscale precipitation distributions associated with Typhoon Herb (1996), which made landfall over Taiwan. The fifth-generation Pennsylvania State University-National Center for Atmospheric Research Mesoscale Model (MM5), with a 2.2-km horizontal grid spacing, successfully simulates the mesoscale rainfall distribution associated with Herb, and the predicted maximum 24-h rainfall of 1199 mm accounts for about 70% of the observed amount of 1736 mm at Mount A-Li. It is shown that, with an accurate track simulation, the ability of the model to simulate successfully the observed rainfall is dependent on two factors: the model's horizontal grid spacing and its ability to describe the Taiwan terrain. The existence of the Central Mountain Range has only a minor impact on the storm track, but it plays a key role in substantially increasing the total rainfall amounts over Taiwan. The analysis presented here shows that the model and terrain resolutions play a nearly equivalent role in the heavy precipitation over Mount A-Li. The presence of maximum vertical motion and heating rate in the lower troposphere, above the upslope mountainous region, is a significant feature of forced lifting associated with the interaction of the typhoon's circulation and Taiwan's mountainous terrain. Overall, Typhoon Herb is a case in point to indicate the intimate relation between Taiwan's topography and the rainfall distribution associated with a typhoon at landfall.
    Yang X. R., J. F. Fei, X. G. Huang, X. P. Cheng, L. W. V. Carvalho, and H. R. He, 2015: Characteristics of mesoscale convective systems over China and its vicinity using geostationary satellite FY2. J.Climate, 28, 4890- 4907.10.1175/JCLI-D-14-00491.1011de2daaade7c90c4a1651ea7790620http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2015JCli...28.4890Yhttp://adsabs.harvard.edu/abs/2015JCli...28.4890YNot Available
    Zheng Y. G., J. Chen, and P. J. Zhu, 2008: Climatological distribution and diurnal variation of mesoscale convective systems over China and its vicinity during summer. Chin. Sci. Bull., 53( 10), 1574- 1586.10.1007/s11434-008-0116-9acd920c76e21caadef6faf173899cc13http%3A%2F%2Flink.springer.com%2F10.1007%2Fs11434-008-0116-9http://www.cnki.com.cn/Article/CJFDTotal-JXTW200810018.htmThe climatological distribution of mesoscale convective systems (MCSs) over China and its vicinity during summer is statistically analyzed, based on the 10-year (1996-2006, 2004 excluded) June-August infrared TBB (Temperature of black body) dataset. Comparing the results obtained in this paper with the distribution of thunderstorms from surface meteorological stations over China and the distribution of lightning from low-orbit satellites over China and its vicinity in the previous studies, we find that the statistic characteristics of TBB less than -52 can better represent the spatiotemporal distribution of MCSs over China and its vicinity during summer.The spreading pattern of the MCSs over this region shows three transmeridional bands of active MCSs, with obvious fluctuation of active MCSs in the band near 30N. It can be explained by the atmospheric circulation that the three bands of active MCSs are associated with each other by the summer monsoon over East Asia. We focus on the diurnal variations of MCSs over different underlying surfaces, and the result shows that there are two types of MCSs over China and its vicinity during summer. One type of MCSs has only one active period all day long (single-peak MCSs), and the other has multiple active periods (multi-peak MCSs). Single-peak MCSs occur more often over plateaus or mountains, and multi-peak MCSs are more common over plains or basins. Depending on lifetimes and active periods, single-peak MCSs can be classified as Tibetan Plateau MCSs, general mountain MCSs, Ryukyu MCSs, and so on. The diurnal variation of multi-peak MCSs is very similar to that of MCCs (mesoscale convective complexes), and it reveals that multi-peak MCSs has longer life cycle and larger horizontal scale, becomes weaker after sunset, and develops again after midnight. Tibetan Plateau MCSs and general mountain MCSs both usually develop in the afternoon, but Tibetan Plateau MCSs have longer life cycle and more active MCSs. Ryukyu MCSs generally develop after midnight, last longer time, and also have more active MCS. The abundant moisture and favorable large-scale environment over Indian monsoon surge areas lead to active MCSs and MCSs almost at any hour all day during summer. Due to local mountain-valley breeze circulation over the Sichuan Basin, MCSs are developed remarkably more often during the nighttime, and again there are also more active MCSs. Because of local prominent sea-land breeze circulation over Guangxi and Guangdong, the MCSs over this region propagate from sea to land in the afternoon and from land to sea after midnight. The statistic characteristics of TBB less than -52 clearly display the different climatological characteristics of MCSs owing to the thermal difference among water, land and rough terrain. Not only the large-scale atmospheric circulation but also the local atmospheric circulation caused by the thermal difference among water, land and rough terrain, to a great extent, determines the climatological distribution of MCSs over China and its vicinity during summer.
  • [1] FU Danhong, GUO Xueliang, 2006: A Cloud-resolving Study on the Role of Cumulus Merger in MCS with Heavy Precipitation, ADVANCES IN ATMOSPHERIC SCIENCES, 23, 857-868.  doi: 10.1007/s00376-006-0857-9
    [2] QIN Xiaohao, MU Mu, 2014: Can Adaptive Observations Improve Tropical Cyclone Intensity Forecasts?, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 252-262.  doi: 10.1007/s00376-013-3008-0
    [3] HUANG Hong, JIANG Yongqiang, CHEN Zhongyi, LUO Jian, WANG Xuezhong, 2014: Effect of Tropical Cyclone Intensity and Instability on the Evolution of Spiral Bands, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 1090-1100.  doi: 10.1007/s00376-014-3108-5
    [4] Chang-Hoi HO, Joo-Hong KIM, Hyeong-Seog KIM, Woosuk CHOI, Min-Hee LEE, Hee-Dong YOO, Tae-Ryong KIM, Sangwook PARK, 2013: Technical Note on a Track-pattern-based Model for Predicting Seasonal Tropical Cyclone Activity over the Western North Pacific, ADVANCES IN ATMOSPHERIC SCIENCES, 30, 1260-1274.  doi: 10.1007/s00376-013-2237-6
    [5] MA Zhanhong, FEI Jianfang, HUANG Xiaogang, CHENG Xiaoping, 2014: Impacts of the Lowest Model Level Height on Tropical Cyclone Intensity and Structure, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 421-434.  doi: 10.1007/s00376-013-3044-9
    [6] GAO Feng*, Peter P. CHILDS, Xiang-Yu HUANG, Neil A. JACOBS, and Jinzhong MIN, 2014: A Relocation-based Initialization Scheme to Improve Track-forecasting of Tropical Cyclones, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 27-36.  doi: 10.1007/s00376-013-2254-5
    [7] Bo SUN, 2018: Asymmetric Variations in the Tropical Ascending Branches of Hadley Circulations and the Associated Mechanisms and Effects, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 317-333.  doi: 10.1007/s00376-017-7089-z
    [8] JIANG Zhina, WANG Xin, WANG Donghai, 2015: Exploring the Phase-Strength Asymmetry of the North Atlantic Oscillation Using Conditional Nonlinear Optimal Perturbation, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 671-679.  doi: 10.1007/s00376-014-4094-3
    [9] CHEN Shumin, Youyu LU, LI Weibiao, WEN Zhiping, 2015: Identification and Analysis of High-Frequency Oscillations in the Eyewalls of Tropical Cyclones, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 624-634.  doi: 10.1007/s00376-014-4063-x
    [10] Yanying CHEN, Ning JIANG, Yang AI, Kang XU, Longjiang MAO, 2023: Influences of MJO-induced Tropical Cyclones on the Circulation-Convection Inconsistency for the 2021 South China Sea Summer Monsoon Onset, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 262-272.  doi: 10.1007/s00376-022-2103-5
    [11] Yan ZHENG, Liguang WU, Haikun ZHAO, Xingyang ZHOU, Qingyuan LIU, 2020: Simulation of Extreme Updrafts in the Tropical Cyclone Eyewall, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 781-792.  doi: 10.1007/s00376-020-9197-4
    [12] Meng Zhiyong, Chen Lianshou, Xu Xiangde, 2002: Recent Progress on Tropical Cyclone Research in China, ADVANCES IN ATMOSPHERIC SCIENCES, 19, 103-110.  doi: 10.1007/s00376-002-0037-5
    [13] P. VINAY KUMAR, Gopa DUTTA, M.V. RATNAM, E. KRISHNA, B. BAPIRAJU, B. Venkateswara RAO, Salauddin MOHAMMAD, 2016: Impact of Cyclone Nilam on Tropical Lower Atmospheric Dynamics, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 955-968.  doi: 10.1007/s00376-016-5285-x
    [14] Kelvin T. F. CHAN, Johnny C. L. CHAN, 2016: Sensitivity of the Simulation of Tropical Cyclone Size to Microphysics Schemes, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 1024-1035.  doi: 10.1007/s00376-016-5183-2
    [15] ZENG Zhihua, Yuqing WANG, DUAN Yihong, CHEN Lianshou, GAO Zhiqiu, 2010: On Sea Surface Roughness Parameterization and Its Effect on Tropical Cyclone Structure and Intensity, ADVANCES IN ATMOSPHERIC SCIENCES, 27, 337-355.  doi: 10.1007/s00376-009-8209-1
    [16] Zhehan CHEN, Qingqing LI, 2021: Re-examining Tropical Cyclone Fullness Using Aircraft Reconnaissance Data, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 1596-1607.  doi: 10.1007/s00376-021-0282-0
    [17] GE Xuyang, MA Yue, ZHOU Shunwu, Tim LI, 2014: Impacts of the Diurnal Cycle of Radiation on Tropical Cyclone Intensification and Structure, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 1377-1385.  doi: 10.1007/s00376-014-4060-0
    [18] Chenxi WANG, Zhihua ZENG, Ming YING, 2020: Uncertainty in Tropical Cyclone Intensity Predictions due to Uncertainty in Initial Conditions, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 278-290.  doi: 10.1007/s00376-019-9126-6
    [19] Zhenhua HUO, Wansuo DUAN, Feifan ZHOU, 2019: Ensemble Forecasts of Tropical Cyclone Track with Orthogonal Conditional Nonlinear Optimal Perturbations, ADVANCES IN ATMOSPHERIC SCIENCES, 36, 231-247.  doi: 10.1007/s00376-018-8001-1
    [20] Tong XIE, Liguang WU, Yecheng FENG, Jinghua YU, 2024: Alignment of Track Oscillations during Tropical Cyclone Rapid Intensification, ADVANCES IN ATMOSPHERIC SCIENCES, 41, 655-670.  doi: 10.1007/s00376-023-3073-y

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 30 December 2015
Manuscript revised: 18 July 2016
Manuscript accepted: 25 July 2016
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

Asymmetric Distribution of Convection in Tropical Cyclones over the Western North Pacific Ocean

  • 1. Institute of Meteorology and Oceanography, PLA University of Science and Technology, Nanjing, Jiangsu 211101, China

Abstract: Forecasts of the intensity and quantitative precipitation of tropical cyclones (TCs) are generally inaccurate, because the strength and structure of a TC show a complicated spatiotemporal pattern and are affected by various factors. Among these, asymmetric convection plays an important role. This study investigates the asymmetric distribution of convection in TCs over the western North Pacific during the period 2005-2012, based on data obtained from the Feng Yun 2 (FY2) geostationary satellite. The asymmetric distributions of the incidence, intensity and morphology of convections are analyzed. Results show that the PDFs of the convection occurrence curve to the azimuth are sinusoidal. The rear-left quadrant relative to TC motion shows the highest occurrence rate of convection, while the front-right quadrant has the lowest. In terms of intensity, weak convections are favored in the front-left of a TC at large distances, whereas strong convections are more likely to appear to the rear-right of a TC within a 300 km range. More than 70% of all MCSs examined here are elongated systems, and meso-β enlongated convective systems (MβECSs) are the most dominant type observed in the outer region of a TC. Smaller MCSs tend to be more concentrated near the center of a TC. While semi-circular MCSs [MβCCSs, MCCs (mesoscale convective complexes)] show a high incidence rate to the rear of a TC, elongated MCSs [MβECSs, PECSs (persistent elongated convective systems)] are more likely to appear in the rear-right quadrant of a TC within a range of 400 km.

1. Introduction
  • In recent years, while forecasts of tropical cyclone (TC) tracks have improved significantly, there are still many uncertainties in forecasts of TC intensity and quantitative precipitation (Demaria et al., 2007). This is mainly due to the fact that the strength and structure of TCs show a complicated spatiotemporal pattern and are affected by various factors. Among these, asymmetric convection plays an important role. Mass convergence, downdrafts, strong rainfall, and latent heat from energy release, which are caused by the convective activities of TCs, can significantly alter the strength and structure of TCs to a large degree (Merrill, 1988; Chen and Meng, 2001).

    In a mature TC, convective activities are mostly concentrated in the eyewall, inner spiral rainbands, and outer spiral rainbands. The eyewall is the area of strongest convection, comprising cumulonimbus clouds embedded with hot vortex towers. Inner spiral rainbands are the MCSs located within the inner core of a TC, which is the zone defined by the triple radius of maximum wind (Wang, 2009). In most cases, inner spiral rainbands cannot be observed from satellite cloud images, mainly because they are hidden by overlying cirrus clouds. In contrast to inner spiral rainbands, outer spiral rainbands usually possess broad stratiform precipitation, and occur alongside various organized convection cells (Barnes et al., 1983, 1991; May, 1996). All of these convection systems show distinct asymmetric structures.

    Observational and numerical studies have been conducted on the asymmetric structures in TC convections. (Miller, 1958) observed a maximum of rainfall in the front-right quadrant (viewing in the direction of motion) from a 16-storm composite around Florida, USA. (Rodgers et al., 1994) obtained a similar conclusion by observing fast-moving TCs through satellites over the North Atlantic Ocean. However, other studies (Burpee and Black, 1989; Corbosiero and Molinari, 2003) have shown that convective activities are scattered. (Burpee and Black, 1989) constructed radial profiles of rainfall and observed that the precipitation maxima of Hurricane Allen (1980) appeared in the front-right quadrant in the core (30 km <r<100 km, r means the radius from a TC center), whereas the precipitation maxima of Hurricane Frederick (1979) and Hurricane Alicia (1983) were more prominent to the left of the TC motion. This discrepancy suggests that there is a significant difference regarding the asymmetric distribution of different storms. Owing to a shortage of data, most studies of TCs over the western North Pacific Ocean have investigated the convection distribution near the TC's landfall. All of these studies agreed with the results obtained for TCs over the North Atlantic, with maximal precipitation tending to appear in the front-right quadrant of a TC. However, there are significant differences between the environmental conditions of the two oceans. So, are the previously determined results applicable to all TCs over the western North Pacific? And what is the typical convection distribution in the ocean away from the coasts? A general investigation of the convection distribution is required. Ground-based radar data have a limited spatial survey range; data obtained from polar orbiting satellites have a poor temporal resolution; and numerical modeling results show little universality. Therefore, in order to meet the statistical requirements for such an investigation, data obtained from geostationary satellites, with high temporal and high spatial resolution, are used in the present study.

    Previous studies have linked the asymmetric structures in TC convections to many factors, including environmental vertical wind shear (Wingo and Cecil, 2010), TC motion (Shapiro, 1983), and the underlying surface (Wu et al., 2002). In a moving storm, friction-induced asymmetric boundary layer convergence results in stronger winds in the lower troposphere to the right of the storm path (Shapiro, 1983; Rodgers et al., 1994; Corbosiero and Molinari, 2003; Lonfat et al., 2004). Accordingly, in the inner core, the front-right quadrant is more favorable for the formation of convection. Environmental vertical wind shear is also an important factor affecting the asymmetric structure of convection in the inner core. Several studies have indicated that a strong convective system is predominantly located in the downshear-left quadrant (Frank and Ritchie, 1999, 2001; Blank et al., 2002; Corbosiero and Molinari, 2002), consistent with the largest vertical motion. However, only a small number of studies have been conducted on convection in the outer region of a TC (outside the inner core), and the two regions have notably different asymmetric characteristics. (Corbosiero and Molinari, 2002) observed a maximum occurrence of lightning in the downshear-right quadrant of the outer region, on the basis of data from the NLDN for Atlantic hurricanes and considered environmental shear as a dominant factor controlling convection activities. The same conclusion was also found in the research of (Chen et al., 2006).

    The majority of previous observational and modeling studies have focused on convection in the core of TCs near the coast (e.g., Chan et al., 2004). In this study, the generality and complexity of convection asymmetry in the outer region of TCs over both the open sea and the land in the western North Pacific region is investigated, for the period 2005-2012. Black body temperature (TBB) data from satellites are used to determine the location, intensity and geometry morphology of convections. Convection distributions from the entire database, based on radial and azimuthal variations for various TC intensities, are discussed in sections 3 and 4 from the perspective of convection occurrence frequency and intensity, respectively. In section 5, quasi-circular and elongated MCSs are classified, and the spatial distribution of different categories of convective cells is analyzed. A summary and conclusions are provided in section 6.

2. Data and methodology
  • 2.1.1. Satellite data

    TBB data were obtained from infrared images taken by the FY 2series Geostationary Meteorological Satellites C and E (http://satellite.nsmc.org.cn/portalsite/default.aspx), with a temporal interval of 1 h and a spatial resolution of 0.1°× 0.1°, and used as a measure of convection. In general, a lower TBB value implies a higher cloud-top height, and as cumulonimbus clouds rise high in strong convections, in a sense the TBB data can directly reflect the strength of convection (Zheng et al., 2008). When comparing satellite images with TMI rainfall data, results showed that convective precipitation (>10 mm h-1) matched well with low TBB values (<-52°C). TRMM data with a high spatial resolution can acquire a TC's 3D structure, but TRMM data have poor spatiotemporal continuity. However, in one orbit, TRMM cannot always catch a whole TC system. Therefore, TBB data have been widely applied to statistical studies on convection.

    A "threshold method", in which a suitable maximum TBB threshold is empirically selected to identify convective clouds, is typically adopted. In their study into the formation of TCs, (Lee et al., 2008) defined areas where the TBB was lower than the thresholds of -32°C, -60°C and -75°C as moderate cumulus convection (MC), deep convection (DC), and extreme convection (EC), respectively, in vertical hot towers. In this paper, we use Lee's definition in order to classify convections. Given that high cirrus clouds mainly occur in the inner core (Li and Duan, 2013), we focus on convections in the outer region, defined by a circle with radius ranging between 100 and 1000 km, and whose center lies at the center of the TC.

    2.1.2. Best-track data

    The best-track dataset for TCs over the western North Pacific used in this study was compiled by the Shanghai Typhoon Institute (www.typhoon.gov.cn). The six-hourly track and intensity analyses were recorded. In order to match the temporal resolution of the TBB data, linear interpolations of two closest best-track records between standard hours are calculated. In this paper, TCs are classified into six levels follows the national standard of China ICS 07. 060 (Appendix A) , in which levels 1 to 6 respectively represent: tropical depression (TD), tropical storm (TS), severe tropical storm (STS), typhoon (TY), severe typhoon (STY), and super typhoon (Super TY). Excluding the cases that are outside the range of the basin, as well as twin typhoons, a total of 136 TCs were recorded during 2005-2012.

    2.1.3. Reanalysis data

    NCEP-NCAR (2005-2012) 1°× 1° global reanalysis data with six-hourly resolution are used in this study. Vertical wind shear is defined as the average difference in horizontal wind vectors between 200 hPa and 850 hPa within a 1000-km radius from the center of a TC.

  • In this section, we discuss the spatial distribution of convection from the perspectives of occurrence frequency, intensity, and the morphology of convection systems.

    The characteristics of convection are herein described in storm-relative coordinates. Azimuthal variation is depicted in the clockwise direction and divided into 36 equal parts from 0° to 360°, in which 0° represents the direction of storm motion. Radial variation is divided into 10 equal parts outward from the storm center, up to a 1000-km radius. Thus, the TC is segmented into 36× 10 angular rings. As each angular ring differs in size for different radius values, the conditional probability flux density functions (PFDFs) of convection occurrence are calculated to describe the occurrence of MC, DC and EC. The number of grids whose values are below the TBB threshold divided by the total number of grids denotes the PFDF within each angular ring, averaged over time. Next, in order to be more intuitive, the PDFs of convection occurrence are determined by considering only azimuthal variation, and the asymmetric distributions of varying grades of TCs for different radius values are compared.

    The distribution of convections based on their intensity, given by TBB values, is discussed. Azimuthal asymmetry is discussed in section 3; thus, we calculate the PDFs of TBB values, representing the intensity of convection, considering only radial distance. In order to determine the asymmetry of TBB values, first-order Fourier coefficients are computed using the data from satellite grids (Boyd, 2001; Lonfat et al., 2004): \begin{eqnarray} \label{eq1} a_1&=&\sum_i{T_i\cos\theta_i} ,(1)\\ \label{eq2} b_1&=&\sum_i{T_i\sin\theta_i} , (2)\end{eqnarray} where Ti is the TBB value of each individual and θi is the phase angle of the value relative to storm motion. The spatial structure of the first-order asymmetry (M1) can be represented by \begin{equation} \label{eq3} M_1=(a_1\cos\theta+b_1\sin\theta)/T , (3)\end{equation} where T is the mean TBB calculated over the entire annulus.

    The methods for identifying MCSs from infrared satellite data are the same as those employed in (Yang et al., 2015). First, a predetermined TBB value and a cloud area threshold need to be provided (Lee et al., 2012), for which a detailed description is given in section 5.1. Secondly, all convection clouds are identified in a single image. Then, the "maximum spatial correlation tracking technique" developed by (Carvalho and Jones, 2001) is used to track target systems between consecutive images. Once an MCS is identified, several features of the MCS can be computed, including mean and minimum cloud shield temperatures, the area of the MCS, the number of cold clusters embedded in the cloud systems, and its eccentricity, displacement, and duration. Finally, after automated identification, a visual check is performed to ensure the accuracy of the results.

3. Spatial distribution of convection occurrence
  • Figure 1 illustrates the spatial distribution of convection occurrence frequency (PFDF) over a span of eight years. The left column represents MC (TBB <-32°C), the middle column represents DC (TBB <-60°C), and the right column represents EC (TBB <-75°C). In general, DCs are distributed within a 500-km radius around the center of a TC, while ECs are distributed within a 300 km-radius from the center. It is apparent that MC occurs more frequently than DC. Irrespective of the intensity of convection, when a TC is weak, the front-left quadrant of the TC is more favorable for the formation of convections. With the development of the TC, the frequency of convection occurrence increases; the high-value zone rotates anticlockwise from the front-left quadrant to the rear-left quadrant and the distribution becomes more asymmetrical. When TC transforms into Super TY, convection is concentrated towards the rear. In contrast, the front-right quadrant of the TC shows an inhibition of convection. It should be noted that when a TC matures into an STY, a closed low-value zone appears near the core of the TC in the front-right quadrant, which is equivalent to the average location of the typhoon eye.

    Figure .  Spatial distribution of the integral ratio of convection occurrence. Left column represents moderate convection, middle column represents deep convection, and right column represents extreme convection. In essence, this ratio is the conditional probability flux density function. The colored shading indicates the frequency of convection occurring at each grid. The ordinate unit is km.

    To be more precise, the distribution of convection is discussed by considering only the azimuthal variation. Figure 2a shows the PDF curves of convection occurrence for all TCs. Irrespective of the intensity of convection, the PDFs are approximately sine curves and display a wavenumber-1 pattern. Fitting the experimental data using the least-squares method reveals that all of the results pass the significance tests. According to Fig. 2a, stronger convection accompanies larger amplitude, which implies a higher asymmetry. Convection dominates in the rear-left quadrant (180°-300°) of a TC, compared to the front-right quadrant (40°-90°). Each curve has two maxima, at 200° and at 280°, but only one minimum.

    The PDFs of different levels of TC are shown in Figs. 2b-g. In the TD period, the distribution of convection varies over a smaller range. Two peaks are observed at 180° and 330°, and two troughs exist in the rear-left (230°) and front-right quadrants (70°) of the TC. At the beginning of TC formation, multiple convections merge from all directions, and the curves vary within a small range. With the development of a TC, the convection asymmetry changes are significantly more dramatic. Two peaks merge into one; meanwhile, the location of the peak moves from the left-rear to the rear. By comparing the trends of the four functions, it can be determined that the more severe the convective activities are, the more sensitive they are to larger amplitudes of curves and stronger asymmetry. Thus, more severe convections are more sensitive to the azimuth around the core of a TC. In Fig. 2, the yellow, green and blue curves show similar tendencies, in contrast to that for MC. Before a TC matures into an STY, MC is more likely to form in the front-left quadrant of a TC, while DC and EC are more likely to form in the left quadrant. When TCs mature into STY or Super TY, all forms of convective activity converge toward the direction opposite to TC movement.

    Figure 2.  PDF curves of convection occurrence based on azimuthal variations, in which (a) shows the convection occurrence PDF curve with azimuth for all TCs. PDFs of different grades of TCs are shown in (b)-(g).

    Figure 3.  (a) Shear direction distribution relative to TC motion over the western North Pacific. (b) Sketched relationship between shear and convection-prone areas.

    Figure 4.  PDF curves of convection occurrence based on azimuthal variations and PDFs of convection occurrence by azimuth for varying radial distances from the center of a TC. Left column represents moderate convection, and right column represents deep convection. The blue, green, red, purple, yellow and black curves represent PDFs within 100 km, 100-200 km, 200-300 km, 300-400 km, 400-500 km, 500-600 km and 600-700 km from the center, respectively.

    As mentioned in the introduction, the effects of asymmetric friction suggest that the front-right quadrant is more favorable for the formation of convection in the inner core. However, this is not supported by our results, indicating that the movement of TCs is not a crucial factor influencing the distribution of convection.

    One possible explanation for the observed distribution is vertical wind shear. We present in Fig. 3a the average shear direction relative to TC motion in the western North Pacific, and comparison with Fig. 2a reveals that the convection-prone zone is located to the average downshear left. In particular, a stronger convection-prone zone is located closer to the direction of shear. When TCs became stronger with counterclockwise circulation enhancing, convections are more likely to be located in the average downshear left. Therefore, the maximum of convection incidence shifts from the left-front to the rear of TCs. The specific influence of shear on convection will be discussed in a future study.

    Figure 4 illustrates the PDF curves of convection occurrence based on azimuthal variations for different radius values. The left column represents MC, and the right column represents DC. As only a few convections lie in the periphery, for this analysis only convections within a 700-km radius of the center of a TC are considered for simplicity. As can be seen from the images in Fig. 4, the larger the radius, the greater the change in amplitude, which means that the distribution of convections is more asymmetrical. As the radius increases, the peak of the convection occurrence rate first rotates clockwise at a radius of 100 km, before rotating anticlockwise between radii of 200 km and 700 km from the center. However, after a TC matures into an STY, the peak rotates in an opposite direction to the previous developmental stages. The transformation of a cyclonic outflow in the core of the TC into an anticyclone outflow is a particular phenomenon of a mature TC. This change is consistent with the movement of the convection peak when a TC matures. In summary, the environmental flow significantly influences the asymmetric structure of convection.

4. Spatial distribution of convection intensity
  • The distributions of different convection intensities, given by TBB values, corresponding to different radial variations are discussed in this section. The radial variation is divided into 20 km-wide annuli outward from the center of a TC, up to a 1000-km radius. This method is identical to that employed by (Lonfat et al., 2004). Figure 5 illustrates the unconditional azimuthal mean TBB values as a function of the radial distance from the center of a TC. The averaged TBB values reach a minimum at a radial distance of 100 km, which is the average position of the eyewall.

    The change in the PDFs of convection intensity as a function of radial distance for different grades of TC is shown in Fig. 6. It is worth noting that a bright band with a TBB value of -95°C is observed near the core of the TC in every image. This bright band may represent high-altitude cirrus clouds, and a weak TC is likely to be influenced by cirrus clouds. (Li and Duan, 2013) pointed out that a TC's inner core is more easily affected by cirrus, but the outer region is basic insusceptible, relatively speaking. Moreover, cirrus's lifespan is short (about one hour), and on a small scale. After time-averaging, the influence of cirrus can be neglected, as in our longstanding and large-scale survey samples.

    In Fig. 6, it can be seen that the area within a radius of 300 km is essentially a deep convective zone, whereas outside of this zone is dominated by MC. When TCs develop into STY (Fig. 6e), low TBB values appear at a radius of 100 km and, in particular, a closed low-value center is observed in Super TYs (Fig. 6f). (Lonfat et al., 2004) suggested that the width of the PDF is a good indication of the degree of convection symmetry. Thus, from the images it appears that, although convections are weak, their distribution broadens with elevated asymmetry at large distances from the center (r>300 km).

    A first-order Fourier transformation was calculated according to Eqs. (2)-(4), in order to analyze the asymmetric spatial distribution of convection intensity. As shown in Fig. 7, positive values indicate a highly asymmetric distribution while negative values indicate a relatively low asymmetric distribution. In Fig. 7, it is evident that convection intensity is more uniform near the center of the TC, as the convection encircles the entire storm eye. Maximum asymmetry is observed in the rear-right quadrant, while minimum asymmetry is observed in the front-left, within a radius of 300-500 km. As evident from Fig. 3a, a stronger convection is more likely to deflect away from the direction of downshear. These results are in accordance with the distribution convection occurrence described in the preceding paragraphs. Generally, with increasing radius, TCs have the greatest asymmetry in the front-left quadrant, as far as 500 km from the center.

    Figure 5.  Radial profile of azimuthally averaged TBB values.

    Figure 6.  PDFs of TBB values with varying radial distance. The PDFs of different grades of TCs are shown in (a)-(f).

    Figure 7.  Normalized phase maximum of first-order convection asymmetry.

5. Spatiotemporal distribution of outer MCSs
  • Under different circumstances, MCSs take on a variety of morphologies, such as a quasi-circular shape, a persistent elongated shape, or a convective line (e.g., Carbone et al., 2002; Johnson et al., 2005). Several previous studies have used multiple satellite observations to identify the various MCS morphologies over land, and their results indicate that the spatiotemporal properties and characteristics of individual MCSs differ significantly (Houze et al., 2007; Luo et al., 2011; Yang et al., 2015). Determining the variations in form and associated distribution regularities in outer-MCSs of TCs is applicable to the aims of this study.

  • The criteria for identifying MCSs using geostationary satellite images have long been contentious. A diversity of research objectives has resulted in non-uniform classification criteria for MCSs. However, all criteria developed to date have been based on the mesoscale convective complex (MCC) survey compiled by (Maddox, 1980) (Yang et al., 2015). In his research, MCCs are defined with specific shapes, durations, area thresholds, and the lowest limit of TBB. In subsequent studies, scholars have found that Maddox's definitions were too rigorous, and could not be meaningfully applied to other cases. Therefore, they assumed -52°C as the TBB threshold, and discovered that the centroid of the -52°C cold-cloud-shield is closely related to precipitation (Kane et al., 1987; Augustine and Howard, 1988). Other studies found that MCCs represent only a part of the MCSs; (Anderson and Arritt, 1998) classified another type of MCS, termed persistent elongated convective systems (PECSs), as the major convective system occurring over the North American continent. Further to this, considering the scale of convections, (Jirak et al., 2003) classified MCSs into four types: MCCs, PECSs, MβCCSs (meso-β circular convective systems), and MβECSs (meso-β elongated convective systems). They noted significant differences between various morphologies of MCSs in terms of both spatiotemporal distribution and environmental conditions. This classification has been widely used in census studies of MCSs over land; however, such studies into TCs remain relatively limited.

    Figure 8.  Identified MCSs in satellite images of typhoon Matsa from 1200 to 1400 UTC and 1900 to 2100 UTC 12 August 2005. The shaded region is the area with an IR brightness temperature less than $-32^\circ$C.

    Figure 9.  (a) Distribution of MCS occurrence. (b) Distribution of MCS areal extent (units: 100 km$^2$). (c) Averaged TBB of MCSs (units: $^\circ$C). (d) Lifespan of MCSs.

    Figure 10.  Scatterplot showing the locations of MCSs. Each black point represents a single MCS. The ordinate units are 1 km.

    A TC is an MCS in itself, and the rainbands and vortices inside the TC cannot be identified if the -52°C threshold is adopted. Therefore, the identification threshold must be revised in order to be successfully applied in our study. Through comparative analysis and multi-group tests, the thresholds of TBB and minimum area were changed to -62°C and 12 000 km2, respectively, which proved to be the optimal results from the tests in order to identify MCSs from a TC. Furthermore, we defined the range of outer-MCSs within a radius of 1000 km from the center. A part of the results of these analyses is presented in Fig. 8, which demonstrates that the algorithm has a preferable automatic identification result, in that it can isolate the rainbands and vortices from a TC.

    As the thresholds have been modified from previously published classification criteria, it is also imperative that the various MCSs are redefined. In view of this, the sizes of the MCSs must be scaled down, as shown in Table 1.

  • In order to focus only on the outer-MCSs, the MCSs found embedded within the core of a TC are first eliminated. Figures 9a and b separately show the distribution of outer-MCSs' occurrence and average area. Owing to the larger radius, both the numbers and sizes of MCS are greater in the outer region than at the center of the TC. Large MCSs lie in the rear-left quadrant of the TC according to these images. Figure 9c shows the averaged TBB values of MCSs. In general, the rear-left region is favorable for lower TBB, as well as strong MCSs. However, both the maximum and minimum average TBB appear near the center of the TC, within 300 km of the center, implying that the intensity of MCSs is more uniform in the outer region. Additionally, it can be noted from Figs. 9a-c that the maximum value alternates with the minimum value in each image, which may be due to the fact that, in the convection of a TC, the ascending and descending motions change discontinuously. Figure 9d shows the duration of the MCSs. From this, it can be seen that most of the identified MCSs survive for only one hour, and that only a few MCSs meet the previously published time criteria.

    The proportions of different categories of MCSs listed in Table 1 are computed. The results reveal that the outer-MCSs are dominated by MβECSs (46%), followed by PECSs (24%), while MCCs make up the smallest number (11%). Nearly 70% of the identified MCSs fall into the elongated categories, and 65% of the MCSs are small-scale. This general pattern is consistent with the results observed from MCSs over land (Yang et al., 2015).

    The scatterplot in Fig. 10 presents the spatial distribution of the various MCSs, with each point marked in black indicating a single MCS. In order to quantitatively describe the asymmetrical distribution of MCSs, a TC is segmented into 8× 5 angular rings. With an increase in the radius of the TC, the area of angular rings becomes larger and the number of MCSs in the outer annulus increases. In view of this, we compute the density of MCSs in each sector. Taking the innermost zone as a benchmark, the magnitude of each kind of MCS in subsequent angular rings from the center is divided by 3, 5, 7 and 9, respectively (Fig. 11). As shown in Fig. 10, four types of MCSs have a homogeneous distribution. In general, meso-β MCSs (MβCCs, MβECSs) tend to cluster near the core of a TC, while α-mesoscale MCSs are scattered in the outer region. For semi-circular MCSs (MβCCSs, MCCs) with large eccentricity, the rear of the TC is a notably high incidence area. In contrast, elongated MCSs (MβECSs, PECSs) are more likely to be observed in the rear-right quadrant of a TC, within 400 km from the center. With increasing radius, the high incidence area for such convections rotates clockwise from the rear to the rear-left, which is consistent with the clockwise outward propagation of rainbands. The causes of these different distributions may be linked to environmental conditions, and the specific controlling mechanisms therefore require further research.

    Figure 11.  Spatial distribution of the density of each type of MCS. Warmer colors indicate larger densities, and colder colors indicate smaller densities.

6. Summary and conclusions
  • The aim of this study was to investigate the asymmetric distribution of convections in TC formations over the western North Pacific between 2005 and 2012. Infrared imagery from FY2 and the Shanghai Typhoon Institute best-track dataset were employed to determine the incidence, intensity and morphology of the convective activities in TCs.

    The results of these investigations demonstrate that the convections in TCs have a clear asymmetric distribution. In general, the PDF curves of convection occurrence based on azimuthal variation for all TCs are approximately sinusoidal. The rear-left area relative to TC motion has the highest rate of convection occurrence, while the front-right area has the lowest. With the development of a TC, the asymmetry of convection increases slightly in the outer region, and the maximum incidence shifts anticlockwise from the front-left to the rear of the TC. In general, the convection asymmetry of a TC has larger amplitudes with respect to vertical wind shear than to TC motion. We found that the convection-prone zone is located in the downshear left region and, in particular, a stronger convection-prone zone is found closer to the direction of shear. When comparing PDFs at different TC radii, the results show that the distribution of asymmetric convections is proportional to the radius. With an increasing radius, the peak of the convection PDF first rotates in an anticlockwise direction in the inner core, and then in a clockwise direction outwards from the center when the TC is mature. This shift is consistent with the upper flow field.

    The spatial distribution of convection intensity can be divided into an azimuthal average and a wavenumber-1 asymmetry. Weak convections are more widely observed in the front-left of a TC at large distances. In contrast, strong convections are more likely to appear to the rear-left of a TC, within 300 km of the center. The wavenumber-1 asymmetry amplitudes are much smaller near the center, and the maximum intensity of convection is observed in the rear-right quadrant, while the minimum intensity is observed to the front-left, within a radius of 300-500 km.

    Using infrared satellite imagery, outer-MCSs were classified into four types on the basis of MCS activity and morphology: MCCs, PECSs, MβCCSs, and MβECSs. More than 70% of all MCSs examined here are elongated systems, and MβECSs are the most dominant type in the outer region of a TC. Smaller sized MCSs are more concentrated near the center of a TC, and semi-circular MCSs (MβCCSs, MCCs) with large eccentricity are concentrated to the rear of a TC. In contrast, elongated MCSs (MβECSs, PECSs) are more likely to appear to the rear-right of a TC, within a 400 km radius, and subsequently propagate outwards in the clockwise direction.

    The present study provides new insights into the convection asymmetry in the outer region of TCs of the western North Pacific. However, this research does not attempt to investigate the factors influencing such a distribution and the dynamical processes involved. Further studies should therefore focus on the effects of environmental vertical shear, TC motion, and the underlying surface, on the characteristics of the convective activity. TCs should be classified according to their maturation time and landing time, and the environmental factors influencing the various types of outer-MCSs should be further investigated. In general, a better understanding of the distribution of convective activities within TCs will result in more accurate forecasts of heavy rainfall events.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return