Advanced Search
Article Contents

Determination of the Effect of Initial Inner-Core Structure on Tropical Cyclone Intensification and Track on a Beta Plane


doi: 10.1007/s00376-016-5241-9

  • The sensitivity of TC intensification and track to the initial inner-core structure on a β plane is investigated using a numerical model. The results show that the vortex with large inner-core winds (CVEX-EXP) experiences an earlier intensification than that with small inner-core winds (CCAVE-EXP), but they have nearly the same intensification rate after spin-up. In the early stage, the convective cells associated with surface heat flux are mainly confined within the inner-core region in CVEX-EXP, whereas the vortex in CCAVE-EXP exhibits a considerably asymmetric structure with most of the convective vortices being initiated to the northeast in the outer-core region due to the β effect. The large inner-core inertial stability in CVEX-EXP can prompt a high efficiency in the conversion from convective heating to kinetic energy. In addition, much stronger straining deformation and PBL imbalance in the inner-core region outside the primary eyewall ensue during the initial development stage in CVEX-EXP than in CCAVE-EXP, which is conducive to the rapid axisymmetrization and early intensification in CVEX-EXP. The TC track in CVEX-EXP sustains a northwestward displacement throughout the integration, whereas the TC in CCAVE-EXP undergoes a northeastward recurvature when the asymmetric structure is dominant. Due to the enhanced asymmetric convection to the northeast of the TC center in CCAVE-EXP, a pair of secondary gyres embedded within the large-scale primary β gyres forms, which modulates the ventilation flow and thus steers the TC to move northeastward.
  • 加载中
  • Bui H. H., R. K. Smith, M. T. Montgomery, and J. Y. Peng, 2009: Balanced and unbalanced aspects of tropical cyclone intensification. Quart. J. Roy. Meteor. Soc., 135, 1715- 1731.10.1002/qj.5024b945f82eeff3ee9a50772da5e1b7dc5http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.502%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1002/qj.502/abstractNot Available
    Chan J. C. L., R. T. Williams, 1987: Analytical and numerical studies of the beta-effect in tropical cyclone motion. Part I: Zero mean flow. J. Atmos. Sci., 44, 1257- 1265.10.1175/1520-0469(1987)044<1257:AANSOT>2.0.CO;292a0d6677751d182bf80629bf2db502ahttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1987JAtS...44.1257Chttp://adsabs.harvard.edu/abs/1987JAtS...44.1257CNot Available
    DeMaria M., 1985: Tropical cyclone motion in a nondivergent barotropic model. Mon. Wea. Rev., 113, 1199- 1210.10.1175/1520-0493(1985)1132.0.CO;25ad3186ab10b98f6ce240527abd6408chttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1985MWRv..113.1199Dhttp://adsabs.harvard.edu/abs/1985MWRv..113.1199DNot Available
    DeMaria M., 1996: The effect of vertical shear on tropical cyclone intensity change. J. Atmos. Sci., 53, 2076- 2088.10.1175/1520-0469(1996)053<2076:TEOVSO>2.0.CO;2d9c321b4b09313000cf962d4ba0509afhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1996JAtS...53.2076Dhttp://adsabs.harvard.edu/abs/1996JAtS...53.2076DCiteSeerX - Scientific documents that cite the following paper: The effect of vertical shear on tropical cyclone intensity change
    DeMaria M., W. H. Schubert, 1984: Experiments with a spectral tropical cyclone model. J. Atmos. Sci., 41, 901- 924.10.1175/1520-0469(1984)041<0901:EWASTC>2.0.CO;27beb86b5-e3f7-4f22-b930-f57686e6dfac847ea51179edcd6e79f6ae3c050a9095http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1984JAtS...41..901Drefpaperuri:(75944f094136f84866466af4b3010adf)http://adsabs.harvard.edu/abs/1984JAtS...41..901DThe three-layer balanced axisymmetric tropical cyclone model presented by Ooyama (1969a) is generalized to three dimensions and the resultant primitive equations are solved using the Galerkin method with Fourier basis functions on a doubly-periodic mid-latitude (beta)-plane. The nonlinear terms are evaluated using the transform method where the necessary transforms are performed using FFT algorithms. The spectral equations are transformed so that the dependent variables represent the normal modes of the linearized equations. In this form, the application of nonlinear normal mode initialization is straightforward. The model is run with an axisymmetric initial condition of an f-plane and it is shown that many of the results presented by Ooyama (1969a) can be reproduced. The energy of the gravity and rotational modes are calculated and it is shown that the gravity mode energy is more than an order of magnitude smaller than the rotational mode energy. The model is then run on the (beta)-plane and it is shown that the tropical cyclone moves towards the northwest at about 2 ms('-1) and elongates towards the west and develops sharper geopotential gradients towards the east. The model is also run with a basic state wind profile and it is shown that large asymmetries develop. It is shown that the basic state wind in the upper layer can interact with the storm outflow to either increase or decrease the storm intensification rate, while the basic state in the lower layer can affect the intensification rate and the size of the model tropical cyclone. The effect of initialization procedures on a tropical cyclone forecast is also studied. The results from linear and nonlinear normal mode initialization and results from applying the nonlinear balance equation are compared. It is shown that the nonlinear normal mode initialization procedure results in much smaller track and intensity forecast errors, and prevents the excitation of spurious gravity waves. Several examples of tropical cyclone motion in the nondivergent barotropic model are presented. It is shown that the spectral truncation, horizontal diffusion coefficient and the tangential wind profile outside of the radius of maximum wind can each affect the track of a tropical cyclone in the barotropic model.
    Dudhia J., 1989: Numerical study of convection observed during the winter monsoon experiment using a mesoscale two-dimensional model. J. Atmos. Sci., 46, 3077- 3107.10.1175/1520-0469(1989)0462.0.CO;234a0f338a8622d0aee3c3811d44d3450http%3A%2F%2Fci.nii.ac.jp%2Fnaid%2F10013124897http://ci.nii.ac.jp/naid/10013124897Not Available
    Ebita A., Coauthors, 2011: The Japanese 55-year reanalysis "JRA-55": An interim report. SOLA, 7, 149- 152.10.2151/sola.2011-0384738102278a5a53c5f03b7e5a7def19chttp%3A%2F%2Fci.nii.ac.jp%2Fnaid%2F130004940944%2Fhttp://ci.nii.ac.jp/naid/130004940944/The carbon monoxide (CO) dehydrogenase of Oligotropha carboxidovorans is composed of an S-selanylcysteine-containing 88. 7-kDa molybdoprotein (L), a 17.8-kDa iron-sulfur protein (S), and a 30.2-kDa flavoprotein (M) in a (LMS)(2) subunit structure. The flavoprotein could be removed from CO dehydrogenase by dissociation with sodium dodecylsulfate. The resulting M(LS)(2)- or (LS)(2)-structured CO dehydrogenase species could be reconstituted with the recombinant apoflavoprotein produced in Escherichia coli. The formation of the heterotrimeric complex composed of the apoflavoprotein, the molybdoprotein, and the iron-sulfur protein involves structural changes that translate into the conversion of the apoflavoprotein from non-FAD binding to FAD binding. Binding of FAD to the reconstituted deflavo (LMS)(2) species occurred with second-order kinetics (k(+1) = 1350 M(-1) s(-1)) and high affinity (K(d) = 1.0 x 10(-9) M). The structure of the resulting flavo (LMS)(2) species at a 2.8-A resolution established the same fold and binding of the flavoprotein as in wild-type CO dehydrogenase, whereas the S-selanylcysteine 388 in the active-site loop on the molybdoprotein was disordered. In addition, the structural changes related to heterotrimeric complex formation or FAD binding were transmitted to the iron-sulfur protein and could be monitored by EPR. The type II 2Fe:2S center was identified in the N-terminal domain and the type I center in the C-terminal domain of the iron-sulfur protein.
    Emanuel K., C. DesAutels, C. Holloway, and R. Korty, 2004: Environmental control of tropical cyclone intensity. J. Atmos. Sci., 61, 843- 858.10.1175/1520-0469(2004)061<0843:ECOTCI>2.0.CO;2e8cd2a868331337e39c36a6b2f8f5fc9http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2004JAtS...61..843Ehttp://adsabs.harvard.edu/abs/2004JAtS...61..843EThe influence of various environmental factors on tropical cyclone intensity is explored using a simple coupled ocean atmosphere model. It is first demonstrated that this model is capable of accurately replicating the intensity evolution of storms that move over oceans whose upper thermal structure is not far from monthly mean climatology and that are relatively unaffected by environmental wind shear. A parameterization of the effects of environmental wind shear is then developed and shown to work reasonably well in several cases for which the magnitude of the shear is relatively well known. When used for real-time forecasting guidance, the model is shown to perform better than other existing numerical models while being competitive with statistical methods. In the context of a limited number of case studies, the model is used to explore the sensitivity of storm intensity to its initialization and to a number of environmental factors, including potential intensity, storm track, wind shear, upper-ocean thermal structure, bathymetry, and land surface characteristics. All of these factors are shown to influence storm intensity, with their relative contributions varying greatly in space and time. It is argued that, in most cases, the greatest source of uncertainty in forecasts of storm intensity is uncertainty in forecast values of the environmental wind shear, the presence of which also reduces the inherent predictability of storm intensity.
    Fang J., F. Q. Zhang, 2012: Effect of beta shear on simulated tropical cyclones. Mon. Wea. Rev., 140, 3327- 3346.10.1175/MWR-D-10-05021.11bdb43d2197ff9d6e5e82a440301fb1fhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2012MWRv..140.3327Fhttp://adsabs.harvard.edu/abs/2012MWRv..140.3327FThrough cloud-resolving simulations, this study examines the effect of 0205 on the evolution of tropical cyclones (TCs). It is found that the TC simulated on a 0205 plane with variable Coriolis parameter 04’ is weaker in intensity but larger in size and strength than the TC simulated on an 04’ plane with constant 04’. Such differences result mainly from the effect of the 0205 shear rather than from the variation of 04’ due to the latitudinal change of the TC position, as illustrated in a three-stage conceptual model developed herein. The first stage begins with the establishment of the 0205 shear and the emergence of asymmetries as the TC intensifies. The 0205 shear peaks in value during the second stage that subsequently leads to the formation of an extensive stratiform region outside of the primary eyewall. The evaporative cooling associated with the stratiform precipitation acts to sharpen the low-level equivalent potential temperature gradient into a frontlike zone outside of the eyewall region, which leads to the burst of convection outside of the primary eyewall. The third stage is characterized by a weakening 0205 shear and the corresponding TC vortex axisymmetrization and expansion. The convection on the inner edge of the stratiform region becomes more organized in the azimuthal direction and eventually causes the TC structure to evolve in a manner similar to the secondary eyewall formation and eyewall replacement usually observed in TCs. It is the active convect0102on outs0102de of the pr0102mary eyewall that contributes to a relatively weaker but larger TC on the 0205 plane than that on the 04’ plane.
    Fiorino M., R. L. Elsberry, 1989: Some aspects of vortex structure related to tropical cyclone motion.J. Atmos. Sci., 46, 975- 990.fd8d4f6dda4fbc011d56ad81d1ffeb52http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1989jats...46..975f/s?wd=paperuri%3A%282c203ac0ce16b83963880b69ce00676a%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1989jats...46..975f&ie=utf-8&sc_us=8782977554611181556
    Ge X. Y., W. Xu, and S. W. Zhou, 2015: Sensitivity of tropical cyclone intensification to inner-core structure. Adv. Atmos. Sci.,32(10), 1407-1418, doi: 10.1007/s00376-015-4286-5.10.1007/s00376-015-4286-50a3574fcfa37f8d32db70298466d5a96http%3A%2F%2Fwww.cnki.com.cn%2FArticle%2FCJFDTotal-DQJZ201510008.htmhttp://d.wanfangdata.com.cn/Periodical/dqkxjz-e201510008In this study, the dependence of tropical cyclone (TC) development on the inner-core structure of the parent vortex is examined using a pair of idealized numerical simulations. It is found that the radial profile of inner-core relative vorticity may have a great impact on its subsequent development. For a system with a larger inner-core relative vorticity/inertial stability, the conversion ratio of the diabatic heating to kinetic energy is greater. Furthermore, the behavior of the convective vorticity eddies is likely modulated by the system-scale circulation. For a parent vortex with a relatively higher inner-core vorticity and larger negative radial vorticity gradient, convective eddy formation and radially inward propagation is promoted through vorticity segregation. This provides a greater potential for these small-scale convective cells to self-organize into a mesoscale inner-core structure in the TC. In turn, convectively induced diabatic heating that is close to the center, along with higher inertial stability, efficiently enhances system-scale secondary circulation. This study provides a solid basis for further research into how the initial structure of a TC influences storm dynamics and thermodynamics.
    Hack J. J., W. H. Schubert, 1986: Nonlinear response of atmospheric vortices to heating by organized cumulus convection. J. Atmos. Sci., 43, 1559- 1573.89a46b81ed18ddc23d3c49d27845f920http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1986jats...43.1559h/s?wd=paperuri%3A%285587dbe5d07de08acb8388e94e595d5a%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1986jats...43.1559h&ie=utf-8&sc_us=10752594496012888995
    Holland, G. J., 1983: Tropical cyclone motion: Environmental interaction plus a beta effect. J. Atmos. Sci., 40, 328- 342.10.1175/1520-0469(1983)040<0328:TCMEIP>2.0.CO;2aaa135dbb3ac71426e4d7dbc54651782http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1983JAtS...40..328Hhttp://adsabs.harvard.edu/abs/1983JAtS...40..328HThe dynamics of tropical cyclone motion are investigated by solving the vergent barotropic vorticity equation on a beta plane. Two methods of solution are presented: a direct analytic solution for a constant basic current, and a simple numerical solution for a more general condition. These solutions indicate that cyclone motion can be accurately prescribed by a non linear combination of two processes: an interaction between the cyclone and its basic current (the well known steering concept), and an interaction with the earth's vorticity field which causes a westward deviation from the pure steering flow. The nonlinear manner in which these two processes combine together with the effect of asymmetries in the steering current raise some interesting questions on the way in which cyclones of different characteristics interact with their environment, and has implications for tropical cyclone forecasting and the manner in which forecasting techniques are derived. (Author)*TROPICAL CYCLONES
    Hong S.-Y., J. Dudhia, and S.-H. Chen, 2004: A revised approach to ice microphysical processes for the bulk parameterization of clouds and precipitation. Mon. Wea. Rev., 132, 103- 120.10.1175/1520-0493(2004)1322.0.CO;27913d9ed85b1a9bbcdcb88db96a17cbbhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2004MWRv..132..103Hhttp://adsabs.harvard.edu/abs/2004MWRv..132..103HNot Available
    Huang Y.-H., M. T. Montgomery, and C.-C. Wu, 2012: Concentric eyewall formation in Typhoon Sinlaku (2008). Part II: Axisymmetric dynamical processes. J. Atmos. Sci., 69, 662- 674.9f9fbbd87ad9075a02f2f76fdd45ed64http%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FXREF%3Fid%3D10.1175%2FJAS-D-11-0114.1/s?wd=paperuri%3A%28d95333a528b369c1d55b844138546f7e%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FXREF%3Fid%3D10.1175%2FJAS-D-11-0114.1&ie=utf-8&sc_us=15849624915402601324
    Li T., X. Y. Ge, M. Peng, and W. Wang, 2012: Dependence of tropical cyclone intensification on the Coriolis parameter. Tropical Cyclone Research and Review, 1, 242- 253.857b513b44111f4633eb39b479bb6439http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F271217113_Dependence_of_tropical_cyclone_intensification_on_the_Coriolis_parameterhttp://www.researchgate.net/publication/271217113_Dependence_of_tropical_cyclone_intensification_on_the_Coriolis_parameter
    Madala R. V., S. A. Piacsek, 1975: Numerical simulation of asymmetric hurricanes on a 尾-plane with vertical shear. Tellus, 27, 453- 468.10.1111/j.2153-3490.1975.tb01699.x2a2cc7e9-e538-4b2e-8022-d8fc0590cdee2b4144ad26e593a9ff9694bbaebfef93http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1111%2Fj.2153-3490.1975.tb01699.x%2Fpdfrefpaperuri:(da9f8458afbe4d845c693d784a852783)http://onlinelibrary.wiley.com/doi/10.1111/j.2153-3490.1975.tb01699.x/pdfAbstract A three-layer, semi-implicit model was developed to simulate moving and asymmetric hurricanes on a β-plane, using Kuo's method of cumulus parametrization. Sensible and latent heat transfer from ocean to atmosphere was included implicitly in the model. In order to predict hurricane movement over a large area and yet resolve finer details near the eye, a multi-grid network was used with a movable finest grid of 20 km mesh size, surrounded by two coarse grid nets with mesh spacings of 60 km and 180 km, respectively. A comparison of the results with f -plane calculations shows that the vortex on the β-plane intensified at a slower rate before the storm stage, but at the same rate thereafter. The β-plane hurricane was asymmetric throughout its life cycle, and these asymmetrics looked similar to those observed in real hurricanes. The vortex moved on the β-plane with a phase velocity of 4.3 km hour 611 for the westerly and 3.3 km hour 611 for the northerly components. The model was also integrated for two cases where the initial vortex was superimposed on a vertically varying basic current. Results showed that the strength of the simulated hurricane depends very much upon the magnitude of the vertical shear of the basic current; for a large shear (≥ 15 m sec 611 /12 km) the vortex failed to intensify into a hurricane. Computations showed that the pressure weighted mean of the basic current between the surface and 12 km level agreed very well with the magnitude of the steering current. As a result of the interaction between the hurricane circulation and the basic current, the hurricane moved in an oscillatory path with an amplitude of 30 km and a period of about 20 hours.
    Mlawer E. J., S. J. Taubman, P. D. Brown, M. J. Iacono, and S. A. Clough, 1997: Radiative transfer for inhomogeneous atmospheres: RRTM, a validated correlated-k model for the longwave. J. Geophys. Res., 102, 16 663- 16 682.10.1029/97JD00237b0b4ef8d-8ae2-48ea-bd48-f3cc1fc4cb61bf5f762e845a497b1ec8058223fb6df8http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F97JD00237%2Fpdfrefpaperuri:(98daaed043b544401196cd274fa354f5)http://onlinelibrary.wiley.com/doi/10.1029/97JD00237/pdfA rapid and accurate radiative transfer model (RRTM) for climate applications has been developed and the results extensively evaluated. The current version of RRTM calculates fluxes and cooling rates for the longwave spectral region (10-3000 cm) for an arbitrary clear atmosphere. The molecular species treated in the model are water vapor, carbon dioxide, ozone, methane, nitrous oxide, and the common halocarbons. The radiative transfer in RRTM is performed using the correlated-k method: the k distributions are attained directly from the LBLRTM line-by-line model, which connects the absorption coefficients used by RRTM to high-resolution radiance validations done with observations. Refined methods have been developed for treating bands containing gases with overlapping absorption, for the determination of values of the Planck function appropriate for use in the correlated-k approach, and for the inclusion of minor absorbing species in a band. The flux and cooling rate results of RRTM are linked to measurement through the use of LBLRTM, which has been substantially validated with observations. Validations of RRTM using LBLRTM have been performed for the midlatitude summer, tropical, midlatitude winter, subarctic winter, and four atmospheres from the Spectral Radiance Experiment campaign. On the basis of these validations the longwave accuracy of RRTM for any atmosphere is as follows: 0.6 W m(relative to LBLRTM) for net flux in each band at all altitudes, with a total (10-3000 cm) error of less than 1.0 W mat any altitude; 0.07 K dfor total cooling rate error in the troposphere and lower stratosphere, and 0.75 K din the upper stratosphere and above. Other comparisons have been performed on RRTM using LBLRTM to gauge its sensitivity to changes in the abundance of specific species, including the halocarbons and carbon dioxide. The radiative forcing due to doubling the concentration of carbon dioxide is attained with an accuracy of 0.24 W m, an error of less than 5%. The speed of execution of RRTM compares favorably with that of other rapid radiation models, indicating that the model is suitable for use in general circulation models.
    Noh Y., W. G. Cheon, S. Y. Hong, and S. Raasch, 2003: Improvement of the K-profile model for the planetary boundary layer based on large eddy simulation data. Bound.-Layer Meteor., 107, 401- 427.10.1023/A:1022146015946ebfc6da0-44fb-434f-9c8b-7b341d91258f2b395fb205157effe9f061a3ffca68b8http%3A%2F%2Fwww.springerlink.com%2Fcontent%2Fm7847n667m47v41w%2Frefpaperuri:(18193937665365080c4f347114f8bfc9)http://www.springerlink.com/content/m7847n667m47v41w/Modifications of the widely used K-profile model of the planetary boundary layer (PBL), reported by Troen and Mahrt (TM) in 1986, are proposed and their effects examined by comparison with large eddy simulation (LES) data. The modifications involve three parts. First, the heat flux from the entrainment at the inversion layer is incorporated into the heat and momentum profiles, and it is used to predict the growth of the PBL directly. Second, profiles of the velocity scale and the Prandtl number in the PBL are proposed, in contrast to the constant values used in the TM model. Finally, non-local mixing of momentum was included. The results from the new PBL model and the original TM model are compared with LES data. The TM model was found to give too high PBL heights in the PBL with strong shear, and too low heights for the convection-dominated PBL, which causes unrealistic heat flux profiles. The new PBL model improves the predictability of the PBL height and produces profiles that are more realistic. Moreover, the new PBL model produces more realistic profiles of potential temperature and velocity. We also investigated how each of these three modifications affects the results, and found that explicit representation of the entrainment rate is the most critical.
    Peng M. S., B.-F. Jeng, and R. T. Williams, 1999: A numerical study on tropical cyclone intensification. Part I: Beta effect and mean flow effect. J. Atmos. Sci., 56, 1404- 1423.10.1175/1520-0469(1999)056<1404:ANSOTC>2.0.CO;22cc173dc96db276af8552dd2243ec987http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1999JAtS...56.1404Phttp://adsabs.harvard.edu/abs/1999JAtS...56.1404PThe effect of planetary vorticity gradient (beta) and the presence of a uniform mean flow on the intensification of tropical cyclones are studied using a limited-area primitive equation model. The most intense storm evolves on a constant-f plane with zero-mean flow and its structure is symmetric with respect to the vortex center. The presence of an environmental flow induces an asymmetry in a vortex due to surface friction. When f varies the vortex is distorted by the beta gyres. Fourier analysis of the wind field shows that a deepening cyclone is associated with a small asymmetry in the low-level wavenumber-one wind field. A small degree of asymmetry in the wind field allows a more symmetric distribution of the surface fluxes and low-level moisture convergence. On the other hand, a weakening or nonintensifying cyclone is associated with a larger asymmetry in its wavenumber-one wind field. This flow pattern generates asymmetric moisture convergence and surface fluxes and a phase shift may exist between their maxima. The separation of the surface flux maximum and the lateral moisture convergence reduces precipitation and inhibits the development of the tropical cyclone. Since the orientation of the asymmetric circulation induced by beta is in the southeast to northwest direction, the asymmetry induced by a westerly flow partially cancels the beta effect asymmetry while that of an easterly flow enhances it. Therefore, in a variable-f environment, westerly flows are more favorable for tropical cyclone intensification than easterly flows of the same speed.
    Powell M. D., 1990: Boundary layer structure and dynamics in outer hurricane rainbands. Part II: Downdraft modification and mixed layer recovery. Mon. Wea. Rev., 118, 918- 938.10.1175/1520-0493(1990)118<0918:BLSADI>2.0.CO;2893738d4-4334-409a-9d20-a49769c8af1ca6fb3b61dee1110cfbed91eed1089fedhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1990MWRv..118..918Prefpaperuri:(e04632faf93de8b371790b637bd58fdc)http://adsabs.harvard.edu/abs/1990MWRv..118..918PNot Available
    Rappaport, E. N., Coauthors, 2009: Advances and challenges at the National Hurricane Center. Wea.Forecasting, 24, 395- 419.10.1175/2008WAF2222128.14d87fc2a18c68ff8a54160435ca8b554http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2009WtFor..24..395Rhttp://adsabs.harvard.edu/abs/2009WtFor..24..395RNot Available
    Schecter D. A., D. H. Dubin, 1999: Vortex motion driven by a background vorticity gradient. Phys. Rev. Lett., 83, 2191.10.1103/PhysRevLett.83.21917593a288bb667c502ec2dd22cc2c129ahttp%3A%2F%2Fjournals.aps.org%2Fprl%2Fabstract%2F10.1103%2FPhysRevLett.83.2191http://journals.aps.org/prl/abstract/10.1103/PhysRevLett.83.2191The motion of self-trapped vortices on a background vorticity gradient is examined numerically and analytically. The vortices act to level the local background vorticity gradient. Conservation of momentum dictates that positive vortices (“clumps”) and negative vortices (“holes”) react oppositely: clumps move up the gradient, whereas holes move down the gradient. A linear analysis gives the trajectory of small clumps and holes that rotate against the local shear. Prograde clumps and holes are always nonlinear, and move along the gradient at a slower rate. This rate vanishes when the background shear is sufficiently large.
    Schubert W. H., J. J. Hack, 1982: Inertial stability and tropical cyclone development. J. Atmos. Sci., 39, 1687- 1697.366ee78577768dcd8a2284c049291b21http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1982JAtS...39.1687S/s?wd=paperuri%3A%28c535c9fc85eeee321243d5e8fcaf2f7d%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1982JAtS...39.1687S&ie=utf-8&sc_us=14768741958169196800
    Shapiro L. J., H. E. Willoughby, 1982: The response of balanced hurricanes to local sources of heat and momentum. J. Atmos. Sci., 39, 378- 394.d50ade92889a7e5b59e5c0d404874eb9http%3A%2F%2Fadsabs.harvard.edu%2Fcgi-bin%2Fnph-data_query%3Fbibcode%3D1982JAtS...39..378S%26db_key%3DPHY%26link_type%3DABSTRACT/s?wd=paperuri%3A%2872d15a9d4b30ef748d25a1f693c16640%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fadsabs.harvard.edu%2Fcgi-bin%2Fnph-data_query%3Fbibcode%3D1982JAtS...39..378S%26db_key%3DPHY%26link_type%3DABSTRACT&ie=utf-8&sc_us=1931974973090579848
    Terwey W. D., M. T. Montgomery, 2008: Secondary eyewall formation in two idealized, full-physics modeled hurricanes. J. Geophys. Res., 113,D12112, doi: 10.1029/2007JD008897.10.1029/2007JD00889714acb099f6136edef9d1594483717c56http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2007JD008897%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1029/2007JD008897/abstractPrevailing hypotheses for secondary eyewall formation are examined using data sets from two high-resolution mesoscale numerical model simulations of the long-time evolution of an idealized hurricane vortex in a quiescent tropical environment with constant background rotation. The modeled hurricanes each undergo a secondary eyewall cycle, casting doubt on a number of other authors' hypotheses for secondary eyewall formation due to idealizations present in the simulation formulations. A new hypothesis for secondary eyewall formation is proposed here and is shown to be supported by these high-resolution numerical simulations. The hypothesis requires the existence of a region with moderate horizontal strain deformation and a sufficient low-level radial potential vorticity gradient associated with the primary swirling flow, moist convective potential, and a wind-moisture feedback process at the air-sea interface to form the secondary eyewall. The crux of the formation process is the generation of a finite-amplitude lower-tropospheric cyclonic jet outside the primary eyewall with a jet width on the order of a local effective beta scale determined by the mean low-level radial potential vorticity gradient and the root-mean square eddy velocity. This jet is hypothesized to be generated by the anisotropic upscale cascade and axisymmetrization of convectively generated vorticity anomalies through horizontal shear turbulence and sheared vortex Rossby waves as well as by the convergence of system-scale cyclonic vorticity by the low-level radial inflow associated with the increased convection. Possible application to the problem of forecasting secondary eyewall events is briefly considered.
    Wang Y. Q., 1995: An inverse balance equation in sigma coordinates for model initialization. Mon. Wea. Rev., 123, 482- 488.10.1175/1520-0493(1995)1232.0.CO;2882ed6c5e98e4cfab4222d5cb02a5db8http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1995MWRv..123..482Whttp://adsabs.harvard.edu/abs/1995MWRv..123..482WNot Available
    Wang Y. Q., 2009: How do outer spiral rainbands affect tropical cyclone structure and intensity? J. Atmos. Sci., 66, 1250- 1273.10.1175/2008JAS2737.1cd29fe53-1c32-46e2-9bec-70af57f4ffcfcd8d5604045d66cacadd3ba570056c28http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2009JAtS...66.1250Wrefpaperuri:(af255fedfc79687b9a5f772c703f2709)http://adsabs.harvard.edu/abs/2009JAtS...66.1250WFrey MH, Payne DA.
    Wang Y. Q., G. J. Holland, 1996a: The beta drift of baroclinic vortices. Part I: Adiabatic vortices. J. Atmos. Sci., 53, 411- 427.10.1175/1520-0469(1996)053<0411:TBDOBV>2.0.CO;2c22daabe-0f19-4848-b2f0-3e375b922abc1f64b9abd46dad136ed6d328d0f1eba2http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1996JAtS...53..411Wrefpaperuri:(2d61c8a3fd71c1625b4820cc0f2d98ea)http://adsabs.harvard.edu/abs/1996JAtS...53..411WThe dynamics of the movement of an initially axisymmetric baroclinic vortex embedded in an environment at rest on a beta plane is investigated with a three-dimensional primitive equation model. The study focuses on the motion and evolution of an adiabatic vortex and especially the manner in which vertical coupling of a tilted vortex influences its motion. The authors find that the vortex movement is determined by both the asymmetric flow over the vortex core associated with beta gyres and the flow associated with vertical projection of the tilted potential vorticity anomaly. The effects of vortex tilt can be large and complex. The secondary divergent circulation is found to be associated with the development of potential temperature anomalies required to maintain a balanced state. The processes involved strongly depend on the vertical structure, size, and intensity of the vortex together with external parameters such as the earth rotation and static stability of the environment. As a result, simple relationships between vortex motion and the vertical mean relative angular momentum are not always applicable.
    Wang Y. Q., G. J. Holland, 1996b: The beta drift of baroclinic vortices. Part II: Diabatic vortices. J. Atmos. Sci., 53, 3737- 3756.10.1175/1520-0469(1996)053<3737:TBDOBV>2.0.CO;2233417c2-699d-4fa3-8d4c-f4245a70787aae03128f53a36df09469a94ce91b379fhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1996JAtS...53.3737Wrefpaperuri:(f0aab038279efab2d5b651eff076eb20)http://adsabs.harvard.edu/abs/1996JAtS...53.3737WPart II. Focuses on the beta drift of baroclinic vortices. Emphasis on diabatic vortices; Use of a three-dimensional primitive equation model; Difference in the motion and evolution of diabatic and adiabatic votices.
    Wu C.-C., K. A. Emanuel, 1993: Interaction of a baroclinic vortex with background shear: Application to hurricane movement. J. Atmos. Sci., 50, 62- 76.10.1175/1520-0469(1993)050<0062:IOABVW>2.0.CO;2deae02c7807095faed4f50a778912bebhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1993JAtS...50...62Whttp://adsabs.harvard.edu/abs/1993JAtS...50...62WABSTRACT Most extant studies of tropical cyclone movement consider a barotropic vortex on a plane. However, observations have shown that real tropical cyclones are strongly baroclinic, with broad anticyclones aloft. Also, the distribution of the large-scale potential vorticity gradient in the tropical atmosphere is very nonuniform. These properties may substantially influence the movement of such storms.Note that the anticyclone above a hurricane will interact with the lower hurricane vortex and induce storm motion. Such interaction can be caused by both the direct effect of ambient vertical shear and the effect of vertical variation of the background potential vorticity gradient. In this paper, an attempt to isolate the effect of background vertical shear is made. The hurricane is represented in a two-layer quasigeostrophic model as a point source of mass and zero potential vorticity air in the upper layer, collocated with a point cyclone in the lower layer. The model is integrated by the method of contour dynamics and contour surgery.The results show that Northern Hemisphere tropical cyclones should have a component of drift relative to the mean flow in a direction to the left of the background vertical shear. The effect of weak shear is also found to be at least as strong as the effect, and the effect is maximized by a certain optimal ambient shear. The behavior of the model is sensitive to the thickness ratio of the two layers and is less sensitive to the ratio of the vortices' horizontal scale to the radius of deformation. Storms with stronger negative potential vorticity anomalies tend to exhibit more vortex drift.
    Wu L. G., B. Wang, 2000: A potential vorticity tendency diagnostic approach for tropical cyclone motion. Mon. Wea. Rev., 128, 1899- 1911.10.1175/1520-0493(2000)128<1899:APVTDA>2.0.CO;29d964d6e-dc01-4967-8cd5-10c3296c7fff8174e6234ee758523a9fc39bdd707cc2http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2000MWRv..128.1899Wrefpaperuri:(4cfecb8bb923d194011f21f1937a6095)http://adsabs.harvard.edu/abs/2000MWRv..128.1899WIn order to understand the roles of various physical processes in baroclinic tropical cyclone (TC) motion and the vertical coupling between the upper- and lower-level circulations, a new dynamical framework is advanced. A TC is treated as a positive potential vorticity (PV) anomaly from environmental flows, and its motion is linked to the positive PV tendency. It is shown that a baroclinic TC moves to the region where the azimuthal wavenumber one component of the PV tendency reaches a maximum, but does not necessarily follow the ventilation flow (the asymmetric flow over the TC center). The contributions of individual physical processes to TC motion are equivalent to their contributions to the wavenumber one PV component of the PV tendency. A PV tendency diagnostic approach is described based on this framework. This approach is evaluated with idealized numerical experiments using a realistic hurricane model. The approach is capable of estimating TC propagation with a suitable accuracy and determining fractional contributions of individual physical processes (horizontal and vertical advection, diabatic heating, and friction) to motion. Since the impact of the ventilation flow is also included as a part of the influence of horizontal PV advection, this dynamical framework is more general and particularly useful in understanding the physical mechanisms of baroclinic and diabatic TC motion.
    Wu L. G., S. A. Braun, 2004: Effects of environmentally induced asymmetries on hurricane intensity: A numerical study. J. Atmos. Sci., 61, 3065- 3081.10.1175/JAS-3343.14ab519fb4a7f2ae4d674475b56d6e196http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2004JAtS...61.3065Whttp://adsabs.harvard.edu/abs/2004JAtS...61.3065WThe influence of uniform large-scale flow, the beta effect, and vertical shear of the environmental flow on hurricane intensity is investigated in the context of the induced convective or potential vorticity asymmetries in the core region with a hydrostatic primitive equation hurricane model. In agreement with previous studies, imposition of one of these environmental effects weakens the simulated tropical cyclones. In response to the environmental influence, significant wavenumber-1 asymmetries develop. Asymmetric and symmetric tendencies of the mean radial and azimuthal winds and temperature associated with the environment-induced convective asymmetries are evaluated. The inhibiting effects of environmental influences are closely associated with the resulting eddy momentum fluxes, which tend to decelerate tangential and radial winds in the inflow and outflow layers. The corresponding changes in the symmetric circulation tend to counteract the deceleration effect. The net effect is a moderate weakening of the mean tangential and radial winds. The reduced radial wind can be viewed as an anomalous secondary radial circulation with inflow in the upper troposphere and outflow in the lower troposphere, weakening the mean secondary radial circulation.
    Zhu T., D.-L. Zhang, 2006: The impact of the storm-induced SST cooling on hurricane intensity. Adv. Atmos. Sci. ,23, 14-22, doi:10.1007/s00376-006-0002-9.10.1007/s00376-006-0002-91267dc5780c303278f5f8c9e3bbac10ehttp%3A%2F%2Fd.wanfangdata.com.cn%2FPeriodical_dqkxjz-e200601002.aspxhttp://d.wanfangdata.com.cn/Periodical_dqkxjz-e200601002.aspxThe effects of storm-induced sea surface temperature (SST) cooling on hurricane intensity are investigated using a 5-day cloud-resolving simulation of Hurricane Bonnie (1998). Two sensitivity simulations are performed in which the storm-induced cooling is either ignored or shifted close to the modeled storm track. Results show marked sensitivity of the model-simulated storm intensity to the magnitude and relative position with respect to the hurricane track. It is shown that incorporation of the storm-induced cooling, with an average value of 1.3℃, causes a 25-hPa weakening of the hurricane, which is about 20hPa per 1℃ change in SST. Shifting the SST cooling close to the storm track generates the weakest storm,accounting for about 47% reduction in the storm intensity. It is found that the storm intensity changes are well correlated with the air-sea temperature difference. The results have important implications for the use of coupled hurricane-ocean models for numerical prediction of tropical cyclones.
  • [1] GE Xuyang, XU Wei, ZHOU Shunwu, 2015: Sensitivity of Tropical Cyclone Intensification to Inner-Core Structure, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 1407-1418.  doi: 10.1007/s00376-015-4286-5
    [2] HUANG Wei, LIANG Xudong, 2010: Convective Asymmetries Associated with Tropical Cyclone Landfall: beta-Plane Simulations, ADVANCES IN ATMOSPHERIC SCIENCES, 27, 795-806.  doi: 10.1007/s00376-009-9086-3
    [3] K. C. SZETO, Johnny C. L. CHAN, 2010: Structural Changes of a Tropical Cyclone during Landfall: β-plane Simulations, ADVANCES IN ATMOSPHERIC SCIENCES, 27, 1143-1150.  doi: 10.1007/s00376-009-9136-x
    [4] MA Shuqing, CHEN Hongbin, WANG Gai, PAN Yi, LI Qiang, 2004: A Miniature Robotic Plane Meteorological Sounding System, ADVANCES IN ATMOSPHERIC SCIENCES, 21, 890-896.  doi: 10.1007/BF02915591
    [5] 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
    [6] 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
    [7] Zhang Daizhou, Tanaka Hiroshi, Qin Yu, 1996: Internal Gravity Waves Generated by a Local Thermal Source in an Irrotational Zonal-Vertical Plane: Numerical Analysis, ADVANCES IN ATMOSPHERIC SCIENCES, 13, 124-132.  doi: 10.1007/BF02657033
    [8] Ren Shuzhan, 1991: Nonlinear Stability of Plane Rotating Shear Flow under Three-Dimensional Nondivergence Disturbances, ADVANCES IN ATMOSPHERIC SCIENCES, 8, 129-136.  doi: 10.1007/BF02658089
    [9] Luo Dehai, 1998: Topographically Forced Three-Wave Quasi-Resonant and Non-Resonant Interactions among Barotropic Rossby Waves on an Infinite Beta-Plane, ADVANCES IN ATMOSPHERIC SCIENCES, 15, 83-98.  doi: 10.1007/s00376-998-0020-x
    [10] MA Leiming, DUAN Yihong, ZHU Yongti, 2004: The Structure and Rainfall Features of Tropical Cyclone Rammasun (2002), ADVANCES IN ATMOSPHERIC SCIENCES, 21, 951-963.  doi: 10.1007/BF02915597
    [11] 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
    [12] Yaping WANG, Yongjie HUANG, Xiaopeng CUI, 2018: Impact of Mid- and Upper-Level Dry Air on Tropical Cyclone Genesis and Intensification: A Modeling Study of Durian (2001), ADVANCES IN ATMOSPHERIC SCIENCES, 35, 1505-1521.  doi: 10.1007/s00376-018-8039-0
    [13] Hepeng ZHENG, Yun ZHANG, Lifeng ZHANG, Hengchi LEI, Zuhang WU, 2021: Precipitation Microphysical Processes in the Inner Rainband of Tropical Cyclone Kajiki (2019) over the South China Sea Revealed by Polarimetric Radar, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 65-80.  doi: 10.1007/s00376-020-0179-3
    [14] 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
    [15] CHENG Xiaoping, FEI Jianfang, HUANG Xiaogang, ZHENG Jing, 2012: Effects of Sea Spray Evaporation and Dissipative Heating on Intensity and Structure of Tropical Cyclone, ADVANCES IN ATMOSPHERIC SCIENCES, 29, 810-822.  doi: 10.1007/s00376-012-1082-3
    [16] 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
    [17] TANG Xiaodong, TAN Zhemin, 2006: Boundary-Layer Wind Structure in a Landfalling Tropical Cyclone, ADVANCES IN ATMOSPHERIC SCIENCES, 23, 737-749.  doi: 10.1007/s00376-006-0737-3
    [18] Tian Yongxiang, Luo Zhexian, 1994: Vertical Structure of Beta Gyres and Its Effect on Tropical Cyclone Motion, ADVANCES IN ATMOSPHERIC SCIENCES, 11, 43-50.  doi: 10.1007/BF02656992
    [19] 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
    [20] LI Wei-Wei, WANG Chunzai, WANG Dongxiao, YANG Lei, DENG Yi, 2012: Modulation of Low-Latitude West Wind on Abnormal Track and Intensity of Tropical Cyclone Nargis (2008) in the Bay of Bengal, ADVANCES IN ATMOSPHERIC SCIENCES, 29, 407-421.  doi: 10.1007/s00376-011-0229-y

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 14 November 2015
Manuscript revised: 26 March 2016
Manuscript accepted: 14 April 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

Determination of the Effect of Initial Inner-Core Structure on Tropical Cyclone Intensification and Track on a Beta Plane

  • 1. Center for Monsoon System Research, Institute of Atmospheric Physics, Chinese Academy of Sciences, Beijing 100190

Abstract: The sensitivity of TC intensification and track to the initial inner-core structure on a β plane is investigated using a numerical model. The results show that the vortex with large inner-core winds (CVEX-EXP) experiences an earlier intensification than that with small inner-core winds (CCAVE-EXP), but they have nearly the same intensification rate after spin-up. In the early stage, the convective cells associated with surface heat flux are mainly confined within the inner-core region in CVEX-EXP, whereas the vortex in CCAVE-EXP exhibits a considerably asymmetric structure with most of the convective vortices being initiated to the northeast in the outer-core region due to the β effect. The large inner-core inertial stability in CVEX-EXP can prompt a high efficiency in the conversion from convective heating to kinetic energy. In addition, much stronger straining deformation and PBL imbalance in the inner-core region outside the primary eyewall ensue during the initial development stage in CVEX-EXP than in CCAVE-EXP, which is conducive to the rapid axisymmetrization and early intensification in CVEX-EXP. The TC track in CVEX-EXP sustains a northwestward displacement throughout the integration, whereas the TC in CCAVE-EXP undergoes a northeastward recurvature when the asymmetric structure is dominant. Due to the enhanced asymmetric convection to the northeast of the TC center in CCAVE-EXP, a pair of secondary gyres embedded within the large-scale primary β gyres forms, which modulates the ventilation flow and thus steers the TC to move northeastward.

1. Introduction
  • The prediction of tropical cyclone (TC) intensity change is a very challenging task, because it involves complex physics during TC evolution. A considerable body of research has been conducted to reveal the impacts on TC intensity change from multiscale interactions of different phenomena, such as sea surface temperature, vertical wind shear, environmental moisture, and cloud microphysics (e.g., DeMaria, 1996; Emanuel et al., 2004; Zhu and Zhang, 2006). Although there have been some improvements in TC intensity forecasting in recent years (Rappaport et al., 2009), our understanding and prediction of TC intensification remains limited.

    Previous studies have revealed TC intensity variation on different planes in the absence of environmental flows. For instance, (Li et al., 2012) elucidated the dependence of TC intensification on the Coriolis parameter. They reported that faster TC intensification in a lower planetary vorticity environment than a higher planetary vorticity environment can be attributed to the boundary layer imbalance in relation with the earth planetary vorticity f. (Madala and Piacsek, 1975) documented that a vortex on a β plane intensified at a slower rate than the one on an f plane before the storm stage, but at the same rate thereafter. This discrepancy in TC development arises from a larger inertial stability in the upper layer and the vertical shearing of centers in a β-plane simulation (DeMaria and Schubert, 1984). In addition, the β-induced flow pattern generates asymmetric moisture convergence and surface fluxes, and a phase shift between their maxima, which can inhibit TC development (Peng et al., 1999; Wu and Braun, 2004). Recently, (Ge et al., 2015) examined the influence of different inner-core structures on TC intensification on an f plane. They found that a vortex with a larger inner-core relative vorticity and inertial stability has a greater potential for conversion from diabatic heating to kinetic energy and small-scale convective vorticity aggregation, in contrast to that with a small inner-core vorticity and inertial stability. However, the role of the β effect in modulating TC structural evolution and intensification is not well understood, given distinct initial inner-core structures.

    On the other hand, a substantial amount of TC research has been devoted to improving our understanding of TC movement. To a large degree, TC movement results mainly from the advection of TC relative vorticity by the environmental steering flow. However, when the steering flow is weak, the β effect due to the latitudinal variation of the Coriolis force causes the TC track to deviate from that specified by steering flow. An initially symmetric vortex will develop a pair of counter-rotating β gyres due to Rossby wave dispersion. In a barotropic framework, (Chan and Williams, 1987) demonstrated that if only the β effect is retained, without any feedback to vortex flow, the vortex would only be stretched to the west, without any significant displacement. Only through the inclusion of the nonlinear term can the vortex move towards the northwest, and this northwestward movement increases with both the maximum wind speed and the radius of maximum wind (RMW). (Fiorino and Elsberry, 1989) decomposed the flow into an axisymmetric and an asymmetric component, and showed that the β effect first causes the development of asymmetric vorticity, which is then rotated by the symmetric vortex flow, forming a southwest-northeast-oriented asymmetric streamfunction after an initial adjustment. As a result, the asymmetric flow associated with the gradient of streamfunction can advect the vortex vorticity towards the northwest. For a baroclinic vortex, because the tangential wind decreases with height, the intensity of β gyres and the associated ventilation flow will decrease with height. Therefore, the so-called differential β effect in the vertical direction plays an important role in vortex movement (Wang and Holland, 1996a). Besides, the existence of vertical shear and diabetic heating in the TC core can produce large-scale anticyclonic circulation and an associated PV anomaly in the upper-tropospheric outflow layer, which can then affect the lower level vortex motion through the vertical penetration flow (Wu and Emanuel, 1993; Wang and Holland, 1996b). Although some studies have stated that TC motion is insensitive to the wind structure in the inner region in a barotropic framework (e.g. Holland, 1983; DeMaria, 1985), the question regarding how different inner-core structures of a baroclinic vortex in a full-physics model modulate the β effect and cause the discrepancy in TC displacement needs to be further addressed.

    The objective of this study is to examine the role of initial inner-core structure in TC intensification and track on a β plane. Section 2 presents the model configuration and experimental design. The effect of the inner-core structure on TC intensification is presented in section 3. Section 4 introduces the contribution of different inner-core structures to TC track. A summary is given in section 5.

2. Model configuration and experimental design
  • The WRF model, version 3.6, is adopted to investigate the sensitivity of TC intensification and track to the initial inner-core structure on a β plane. The environmental thermodynamic profiles are specified as the mean climatology in the TC peak region (5°-20°N, 130°-170°E) during July-September from 1979 to 2012, based on the Japanese 55-year Reanalysis (Ebita et al., 2011). The simulations are initialized with an axisymmetric vortex on a β plane centered at 18°N, with a variable Coriolis parameter, f=f0+β y in which f0 denotes the planetary vorticity at 18°N. The vortex is embedded in a quiescent environment over the ocean with a constant sea surface temperature of 29°C. The model contains three domains all with 271× 271 grid points and with horizontal grid spacing of 27, 9 and 3 km, respectively. The innermost domain is designed to move with the model TC center.

    The\,physics\;parameterizations\;include\,the\,WSM6\,scheme (Hong et al., 2004), the YSU (Yonsei University) PBL scheme (Noh et al., 2003), the Dudhia shortwave parameterization (Dudhia, 1989), and the RRTM longwave parameterization (Mlawer et al., 1997). Considering that the domains are large enough to cover the convection in the inner-core area and outer spiral rainbands, cumulus parameterization is not adopted in the two innermost meshes in this study. The TC center is defined as the center of axisymmetric circulation along which the azimuthal mean wind at the lowest model level reaches the maximum at a given time.

    Axisymmetric baroclinic vortices, with the same tangential wind profile outside the RMW but different ones inside the RMW, are superimposed on the horizontally uniform environment. Two idealized experiments are carried out with the model lowest-level tangential wind profile V(r) beyond the RMW, defined by \begin{equation} \label{eq1} V(r)\!=\!V_{m}\left(\dfrac{r}{r_{m}}\right)\left[\exp\left(1\!-\!\dfrac{r}{r_{m}}\right)\!-\!\dfrac{|r-r_{m}|}{R_0-r_{m}} \exp\left(1\!-\!\dfrac{R_0}{r_{m}}\right)\right],\quad r\!\ge\! r_{m} , (1)\end{equation} where V m=12 m s-1 is the maximum tangential wind at the RMW, r m=108 km, r is the radius, and R0=900 km is the radius at which the vortex wind vanishes. To discriminate the inner-core wind fields, the tangential wind profiles inside the RMW are specified as follows: \begin{eqnarray} \label{eq2} &&V_{m}\left(\dfrac{r}{r_{m}}\right)\exp\left(1-\dfrac{r}{r_{m}}\right) ,\quad r<r_{m} ,(2)\\ \label{eq3} &&V_{m}\left(\dfrac{r}{r_{m}}\right)^3\exp\left(1-\dfrac{r}{r_{m}}\right) ,\quad r<r_{m} , (3)\end{eqnarray} which denote the convex- and concave-shaped wind profiles within the RWM (Fig. 1a), corresponding to two experiments referred to as CVEX-EXP and CCAVE-EXP, respectively. In the vertical direction, the initial tangential wind profiles decrease sinusoidally with pressure to vanish at 100 hPa. The mass and thermodynamic fields associated with the vortex are obtained by solving the nonlinear balance equation (Wang, 1995).

3. Effect of the inner-core structure on TC intensification
  • Figure 2 exhibits the temporal evolution of TC intensity in terms of the minimum surface level pressure and maximum azimuthal mean wind speed at the lowest model level. The striking feature is that the TCs in the two experiments display distinct rapid intensification timings. Specifically, the vortex in CVEX-EXP commences intensification at 30 h of integration, while the rapid intensification of the storm in CCAVE-EXP starts at approximately 50 h. In the mature stage, the TCs in CVEX-EXP and CCAVE-EXP reach 928 hPa and 939 hPa at 84 h and 102 h, respectively. After that time, both of the TCs basically sustain their quasi-steady intensity, although the subsequent intensity seems to be oscillatory (Fig. 2a). The sensitivity of TC intensification to the initial inner-core structure is in agreement with the finding on an f plane in (Ge et al., 2015), who showed the delayed intensification of a vortex with a Rankine-type structure, as compared to one with a convex-shaped structure, in the inner-core region.

    Figure 1.  Radial profiles of (a) tangential wind (units: m s$^-1$) and (b) relative vorticity (units: $\times 10^-4$ s$^-1$) at the lowest model level of the initial vortex in CVEX-EXP (solid line) and CCAVE-EXP (dashed line).

    Figure 2.  Temporal evolution of the (a) minimum sea level pressure (units: hPa; thick line) and maximum azimuthal mean wind speed (units: m s$^-1$; thin line) at the lowest model level in CVEX-EXP (solid line) and CCAVE-EXP (dashed line), and (b) the TC track, with dots indicating the 12-h positions in CVEX-EXP (solid line) and CCAVE-EXP (dashed line).

    Figure 3.  Plan view of the 12-hourly simulated radar reflectivity (units: dB$Z$) at the height of 1 km, from 24 h to 84 h, in (a-f) CVEX-EXP and (g-l) CCAVE-EXP. The domain shown in each panel is 360 km $\times$ 360 km. The $x$-axis and $y$-axis denote the east-west and south-north distance, respectively.

    Figure 4.  The distribution of (a, b) surface heat flux (units: W m$^-2$) and (c, d) 10-m wind vector and speed (units: m s$^-1$), averaged between 36 and 48 h, in (a, c) CVEX-EXP and (b, d) CCAVE-EXP. The domain shown in each panel is 360 km $\times$ 360 km. The $x$-axis and $y$-axis denote the east-west and south-north distance, respectively.

    Figure 5.  Radius-height cross section of the azimuthal mean effective $\beta$ (units: $\times 10^-8$ m$^-1$ s$^-1$), averaged from 36 to 48 h, in (a) CVEX-EXP and (b) CCAVE-EXP. The regions with negative values are shaded.

    The sharp contrast in TC intensification can be easily inferred from the radar reflectivity at the lowest model level. In the early stage of 24 h, sporadic small-scale convective vortices emerge within a radius of 60 km in CVEX-EXP (Fig. 3a), while sparse convective cells are scattered outside 60 km in CCAVE-EXP (Fig. 3g). As time proceeds, the convective systems in CVEX-EXP begin to grow through the merger of small-scale vortices and are organized around the TC center in the inner-core region (Fig. 3b). In contrast, although the vortex in CCAVE-EXP lacks convective cells inside 60 km, the loosely organized convective vortices mainly spread from eastern to northern sections of the outer-core region (Fig. 3h). Until 48 h of simulation, the convective systems in CVEX-EXP are well-organized, such that the convective ring shrinks inward forming a clear eye region (Fig. 3c). Whereas, the vortex in CCAVE-EXP undergoes a slower development with substantial convective activity in the northeastern semicircle in the absence of an eye (Fig. 3i). During the subsequent simulation, the vortex in CVEX-EXP contracts tightly, developing a nearly axisymmetric convective eyewall with reduced outer spiral rainbands (Figs. 3d-f). By comparison, until the vortex in CCAVE-EXP develops a well-defined eye after 50 h (Fig. 3j), the storm begins to intensify rapidly. Different from the suppressed outer spiral rainbands in CVEX-EXP, the convective bands are filamented in the outer-core region, giving rise to active outer spiral rainbands and a relatively large inner-core size (Figs. 3j-l). The active spiral rainbands can block the boundary layer inflow, which converges the mass into the eyewall, thus dynamically decreasing the primary eyewall convection and weakening the TC. On the other hand, the convective overturning induced by spiral rainbands produces compensating subsidence that introduces low equivalent potential temperature air from the middle troposphere down to the inflow boundary layer, which can thermodynamically limit TC intensity. The hydrostatic adjustment associated with the net diabatic heating in the outer spiral rainband also causes a reduction in the pressure gradient across the eyewall, which can weaken TC intensity but increase the size of the TC inner core (e.g. Shapiro and Willoughby, 1982; Powell, 1990; Wang, 2009). Therefore, the TC in CCAVE-EXP, with active outer spiral rainbands, evolves into a relatively less intense vortex in the mature stage compared to that in CVEX-EXP, as shown in Fig. 2a.

    As depicted above, two striking differences during the initial spin-up stage between the two experiments can be observed: one is the lack of convective vortices in the inner-core region in CCAVE-EXP, compared to in CVEX-EXP; and the other is a more asymmetric structure in CCAVE-EXP, with vigorous outer convective rainbands in the northeastern quadrant. The discrepancies can be partly accounted for by the surface heat flux and wind field associated with the initial inner-core structure and β-induced asymmetry. Figure 4 depicts the distribution of surface heat flux and the wind field, averaged between 36 and 48 h of simulation. Consistent with the distribution of radar reflectivity in Fig. 3, the TC center is tightly surrounded by large surface heat flux within a radius of 60 km, due to the strong inner-core surface wind in CVEX-EXP (Figs. 4a and c). In sharp contrast, the surface heat flux positively proportional to the magnitude of inner-core wind speed is considerably suppressed in the CCAVE-EXP inner-core region, resulting in a reduction of inner-core convection (Figs. 4b and d). On the other hand, the surface heat flux and wind speed in CCAVE-EXP are highly asymmetric, with a maximum region located to the northeast outside the inner core. This is in good agreement with the findings of (Chan and Williams, 1987), who identified that the β-effect and nonlinear effect combined can create a wind speed maximum to the northeast of the vortex and, moreover, the asymmetry in the outer-core region is more evident than in the inner-core region. Comparatively, although the wind speed in CVEX-EXP appears somewhat asymmetrically, the well-organized eyewall and maximum wind enclose the TC center almost axisymmetrically (Fig. 4c).

    In addition to the wind-induced surface heat exchange (WISHE) emphasized above, two extra experiments are conducted, in which the latent heat flux and sensible heat flux at the surface are fixed at 100 W m-2 and 20 W m-2, respectively, throughout the integration, in such a way that the difference in vortex development can be further tested under the theory of conditional instability of the second kind (CISK). The results show that, given the same surface heat flux, the well-organized convective vortices in CVEX-EXP emerge to the northeast of the TC center, which is earlier than that in CCAVE-EXP. The vigorous convective vortices concentrate within a radius of 60 km in CVEX-EXP, while they are outside 60 km in CCAVE-EXP. In the absence of the WISHE mechanism, the extra experiments exhibit qualitatively similar results to those with WISHE involved, albeit with a relatively slow intensification rate due to the small magnitude of surface heat flux (not shown). It suggests that the CISK-type mechanism can also lead to the distinct vortex development given the different inner-core structure profiles. Besides, the convectively generated mesoscale positive vorticity anomalies tend to move up the ambient vorticity gradient (e.g., Schecter and Dubin, 1999; Ge et al., 2015). Figure 1b exhibits a salient difference in the initial profile of relative vorticity, in that the large vorticity magnitude and its negative radial gradient are within a radius of 70 km in CVEX-EXP (in contrast to those in CCAVE-EXP), which is favorable for vorticity aggregation and convergence, and subsequent formation of a self-amplified mesoscale core vortex. The stronger the inner-core background low-level vorticity, the greater the Ekman pumping effect and diabatic heating, and thus the earlier the TC intensification.

    Another important dynamic parameter——the so-called effective β——can be used to measure the extent to which the spiral convective band is facilitated in the inner-core region. The effective β is defined as \((-\partial\bar{q}/\partial r)(\bar\xi/\bar{q})\), with \(\bar{q}\) and \(\bar\xi=f+2\bar{V}_t/r\) denoting the azimuthal mean potential vorticity and modified Coriolis parameter, respectively, in which \(\bar{V}_t\) denotes the azimuthally-mean tangential wind. Previous studies indicate that a well-defined low-level β skirt favor the formation of a well-organized spiral rainband outside the eyewall, via the maintenance of long-lasting deep convection and the upscale transfer of energy from convective-scale motion (e.g., Terwey and Montgomery, 2008; Fang and Zhang, 2012). In Fig. 5 we can see that the low-level effective β averaged from 36 to 48 h in CVEX-EXP is almost positive radially outward from 15 km, with a maximum at a radius of 27 km. Therefore, the β skirt outside of the eyewall is sufficient to constrain the asymmetric flow and prompt the transfer of perturbation vorticity and kinetic energy from sporadic deep convection to azimuthal mean flow, which in turn promotes the axisymmetrization of asymmetric eddies and rapid intensification in CVEX-EXP (Fig. 5a). In contrast, the positive effective β only occurs outside 60 km in CCAVE-EXP and, moreover, with much smaller magnitude, suggesting that CCAVE-EXP has an unfavorable inner-core dynamic condition that is hostile to the robust activity of an inner spiral rainband, and thus delays intensification (Fig. 5b).

    The response of the secondary circulation to the heat source is sensitive to inertial stability, defined as $I=\sqrt{(f+2V/r)[f+\partial(rV)/r\partial r]}$. The vortex intensification becomes more efficient for a given heat source with increasing inertial stability (e.g. Schubert and Hack, 1982; Shapiro and Willoughby, 1982; Hack and Schubert, 1986). When the inertial stability is high, internal atmospheric heating can effectively warm the atmospheric column and lower the surface pressure. Physically, a larger inertial stability corresponds to a smaller Rossby deformation radius, causing the energy produced by the diabatic heating to be confined within a smaller radius, rather than dispersed by gravity waves, as the inertial stability is low. Figure 6 clearly shows that the inertial stability inside 60 km is much larger in CVEX-EXP than in CCAVE-EXP, which implies an increase in efficiency with which the latent heating can locally warm the troposphere and increase the tangential winds. Comparatively, the inertial stability outside 90 km in CCAVE-EXP is slightly larger, which may be attributable to the relatively large winds in the outer-core region to the northeast, related to β-induced asymmetry, as shown in Fig. 4d.

    Recent work has identified the role of the unbalanced PBL process in TC intensification (e.g., Bui et al., 2009; Huang et al., 2012). (Bui et al., 2009) found that PBL balanced theory underestimates the low-level radial inflow, and therefore the maximum azimuthal-mean tangential wind tendency. The accelerated radial inflow associated with the PBL imbalanced process can enhance the inward advection of high absolute angular momentum, and thus increase vortex intensity. A useful way to quantify the unbalanced component of boundary layer dynamics is to compute the agradient force (AF), defined as the sum of the azimuthally averaged radial pressure gradient force, the Coriolis force, and centrifugal forces:

    $$AF=-\dfrac{1}{\rho} \dfrac{\partial\bar{p}}{\partial r}+f\bar{v}+\dfrac{\bar{v}^2}{r},(4)$$

    where p is pressure, ρ is the air density, and v is the tangential wind. The overbar represents the azimuthal-mean component. AF<0 ( AF>0) corresponds to the subgradient (supergradient) radial inflow, indicating a tendency to accelerate (decelerate) the inflows toward the vortex center. Figure 7 shows the agradient force averaged below 2 km from 24 to 48 h. It displays clearly that, in CVEX-EXP, the strong outward-directed (positive) agradient force resides within 36 km, with the maximum at 20 km, while the inward-directed (negative) agradient force exists outside the radius of 36 km. This radial profile of agradient force implies the convergence of radial flow appears in the zone with the decelerated flow on its radially inward side, and the accelerated flow on its radially outward side. The enhanced accelerated radial inflows on the radially outward side also bring the high absolute angular momentum inward to contribute effectively to TC intensity. By comparison, in CCAVE-EXP, the positive and negative agradient forces are detached at a relatively larger radius. Moreover, the magnitude of agradient force is much smaller than that in CVEX-EXP. As a result, the attenuated convergence is situated at a larger radius, which intensifies the vortex less effectively, and thus delays the vortex intensification.

    Figure 6.  Radial profile of the inertial stability parameter normalized by the local Coriolis parameter, averaged between 950 and 700 hPa during 24-48 h, in CVEX-EXP (sold line) and CCAVE-EXP (dashed line).

    Figure 7.  Agradient force (units: m s$^-1$ h$^-1$) averaged below 2 km, from 24 to 48 h, in CVEX-EXP (solid line) and CCAVE-EXP (dashed line).

4. Effect of the inner-core structure on TC track
  • In addition to the sensitivity of TC intensification to the different inner-core structures, the resulting structure also modulates the TC track. As demonstrated in Fig. 2b, during the first 36 h, both of the TCs have similar northwestward displacement, with a tiny difference in that the TC track in CVEX-EXP is located slightly to the east of that in CCAVE-EXP. Interestingly, after 36 h of integration, the TC movement in CCAVE-EXP displays a more northward component than that in CVEX-EXP, and even makes a northeast turn during the period of 48-60 h. After 60 h, the TC translation in CCAVE-EXP shifts to the northwest, and then maintains an almost identical translation speed and direction to that in CVEX-EXP. Comparatively, the TC in CVEX-EXP sustains a basically uniform northwestward displacement. As a consequence, the TC track in CCAVE-EXP is displaced to the east of that in CVEX-EXP after 48 h. Therefore, the question naturally arises: what is responsible for the TC northeastward shift during the period from 36 to 60 h in CCAVE-EXP?

    A TC on a β plane is primarily steered by the ventilation flow between β gyres in the absence of environmental flow. Therefore, the sudden change of TC track in CCAVE-EXP can be accounted for by the steering flow related to the ventilation flow. In this study, the steering flow is calculated as the mean vector averaged within a radius of 300 km from the TC center between 850 and 300 hPa. Figure 8 displays the wavenumber-1 wind and geopotential height averaged between 850 and 300 hPa from 36 to 60 h. In CVEX-EXP (Fig. 8a), the wavenumber-1 structure features a dipole-like pattern with a positive (negative) anomaly of geopotential height and an anticyclonic (cyclonic) gyre to the east-northeast (west-southwest) of the TC center, which is in good agreement with the results of previous studies (e.g., Fiorino and Elsberry, 1989; Wang and Holland, 1996a). As a result, the wavenumber-1 asymmetry exhibits northwest-oriented ventilation flow with steering flow of approximately 1.4 m s-1 that coincides with the TC movement, as shown in Fig. 2b.

    In sharp contrast, the wavenumber-1 geopotential height in CCAVE-EXP clearly takes on the secondary gyres within a radius of 300 km, with a positive (negative) anomaly along with an anomalously anticyclonic (cyclonic) circulation to the southwest (northeast) of the TC center, which is embedded within the primary β gyres. It is this pair of secondary gyres that modifies the β effect and rotates the mean ventilation flow clockwise from northwest to northeast, steering the TC to move northeastward (Fig. 8b). After 60 h, accompanied by the TC intensification and axisymmetrization as described in section 3, the primary β gyres are dominant and then shift the TC translation from northeast to northwest. The initiation of the secondary gyres in CCAVE-EXP can be attributed to the asymmetric convection. As depicted in Figs. 3i and 4b and d, the strong convective vortices in CCAVE-EXP are mostly distributed between the radii of 60 and 120 km in the northeastern section of the TC, which can warm the local atmospheric column and lower the geopotential height by hydrostatic adjustment, such that the gyre structure is modified with an anomalously negative (positive) geopotential height to the northeast (southwest), centered at the radius of 90 km. After 60 h, the vortex in CCAVE-EXP intensifies and develops a more axisymmetric and vertically aligned structure. The decreased structural asymmetry, to a lesser extent, modifies the β effect such that the TC track is re-oriented northwestward, similar to that in CVEX-EXP afterward, as displayed in Fig. 2b.

    Figure 8.  The wavenumber-1 wind (vectors; units: m s$^-1$) and geopotential height (contours; units: m$^2$ s$^-2$) averaged between 850 and 300 hPa, from 36 to 60 h, in (a) CVEX-EXP and (b) CCAVE-EXP. The thick arrows indicate the direction and magnitude of steering flow, with a maximum value of 1.4 m s$^-1$. The domain shown in each panel is 1800 km $\times$ 1800 km. The $x$-axis and $y$-axis denote the east-west and south-north distance, respectively.

    Previous studies have pointed out that a baroclinic TC tends to move to the region where the azimuthal wavenumber-1 component of (potential) vorticity tendency reaches a maximum (e.g., Holland, 1983; Wu and Wang, 2000). To further understand the role of the asymmetric structure in the TC track, the simulated wavenumber-1 component of the vorticity tendency averaged between 850 and 300 hPa from 36 to 60 h is displayed in Fig. 9. As seen in Fig. 9a, the maximum vorticity tendency in CVEX-EXP is directed to the northwest, in alignment with the direction of TC movement and steering flow. The structure of the wavenumber-1 vorticity tendency is also consistent with the quasi-steady solution in streamfunction tendency analysis in (Fiorino and Elsberry, 1989). Similar to the structure in CCAVE-EXP as shown in Fig. 8b, the distribution of the wavenumber-1 vorticity tendency features a pair of northeast-southwest-oriented gyres within a radius of 300 km, with a positive (negative) anomaly to the northeast (northwest) of the TC center. As discussed above, the anomalously enhanced convective heating to the northeast acts as a major contributor to the local increasing of vorticity with time, favorable for the northeastward displacement of the TC during the spin-up period.

    Figure 9.  The wavenumber-1 wind (vectors; units: m s$^-1$) and vorticity tendency (shaded; units: $\times 10^-9$ s$^-2$) averaged between 850 and 300 hPa, from 36 to 60 h, in (a) CVEX-EXP and (b) CCAVE-EXP. The thick arrows indicate the direction of maximum vorticity tendency. The domain shown in each panel is 360 km $\times$ 360 km. The $x$-axis and $y$-axis denote the east-west and south-north distance, respectively.

5. Conclusion
  • The sensitivity of TC intensification and track to the initial inner-core structure is investigated using the WRF model. The initial symmetric vortices with the same outer-core but different inner-core structures (convex- versus concave-shaped inner-core wind profiles) are specified on a β plane. It is found that the vortex with a convex-shaped inner-core wind profile in CVEX-EXP intensifies earlier than that in CCAVE-EXP, but has a similar intensification rate after the intensification is initiated. In the final steady state, the TC in CCAVE-EXP is observed to be somewhat weaker than its counterpart in CVEX-EXP, due to the more active outer rainband in CCAVE-EXP that can act as an inhibitor for TC intensification from the dynamic and thermodynamic perspectives.

    In the early stage, the strong wind in the inner-core region in CVEX-EXP can trigger more surface heat flux, and then convective heating, within a radius of 60 km. Convectively generated small-scale vortices tend to aggregate rapidly in the region with strong ambient vorticity, and form a self-amplified vortex through a bottom-up development process. In contrast, the inner-core convection is less active in CCAVE-EXP, whereas most of the convective cells spread outside the inner-core region to the northeast of the TC center. The outer-core structural asymmetry in CCAVE-EXP can be ascribed to the β effect that generates a wind speed maximum to the northeast of the vortex. The well-defined low-level β skirt within a relatively small radius in CVEX-EXP can favor the formation of a well-organized convective band outside the eyewall, via the maintenance of long-lasting deep convection and the upscale transfer of energy from convective-scale motion, which accelerates TC intensification. Owing to the high inertial stability in the inner-core region in CVEX-EXP, the inner-core convective heating can effectively warm the atmospheric column and convert potential energy to kinetic energy. Besides, the striking PBL imbalance in CVEX-EXP in the early stage drives the significant convergence in conjunction with subgradient and supergradient flows, along with the radially inward transport of high angular momentum, assisting in the advancement of TC intensification. Although the TC in CCAVE-EXP has a delayed intensification, it also experiences a rapid intensification once the convective ring and the radius of maximum wind contract significantly in such a way that the efficiency with which the convective heating forces the secondary circulation becomes increased.

    In terms of TC track, the vortex in CVEX-EXP sustains a consistent northwestward displacement during the simulation. In contrast, the TC in CCAVE-EXP experiences a northeastward recurvature during the period from 36 to 60 h, and afterward restores to the northwest direction as TC intensification and axisymmetrization proceed. Different from the larger-scale primary gyres due to the meridional gradient of planetary vorticity in CVEX-EXP, a pair of smaller-scale secondary gyres in CCAVE-EXP is embedded within the primary β gyres, consisting of a cyclonic one to the northeast and an anticyclonic one to the southwest. The combination of primary and second gyres in CCAVE-EXP modifies the direction of ventilation flow, steering the TC to move northeastward when the asymmetric structure is significant. The northeast-southwest-oriented secondary gyres in CCAVE-EXP are closely related to the enhanced convective heating to the northeast of the TC center outside 60 km. As a consequence, the wavenumber-1 minimum geopotential height and maximum vorticity tendency correspond well with the TC translation direction.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return