Advanced Search
Article Contents

Effects of Surface Flux Parameterization on the Numerically Simulated Intensity and Structure of Typhoon Morakot (2009)


doi: 10.1007/s00376-015-4202-z

  • The effects of surface flux parameterizations on tropical cyclone (TC) intensity and structure are investigated using the Advanced Research Weather Research and Forecasting (WRF-ARW) modeling system with high-resolution simulations of Typhoon Morakot (2009). Numerical experiments are designed to simulate Typhoon Morakot (2009) with different formulations of surface exchange coefficients for enthalpy (C K) and momentum (C D) transfers, including those from recent observational studies based on in situ aircraft data collected in Atlantic hurricanes. The results show that the simulated intensity and structure are sensitive to C K and C D, but the simulated track is not. Consistent with previous studies, the simulated storm intensity is found to be more sensitive to the ratio of C K/C D than to C K or C D alone. The pressure-wind relationship is also found to be influenced by the exchange coefficients, consistent with recent numerical studies. This paper emphasizes the importance of C D and C K on TC structure simulations. The results suggest that C D and C K have a large impact on surface wind and flux distributions, boundary layer heights, the warm core, and precipitation. Compared to available observations, the experiment with observed C D and C K generally simulated better intensity and structure than the other experiments, especially over the ocean. The reasons for the structural differences among the experiments with different C D and C K setups are discussed in the context of TC dynamics and thermodynamics.
  • 加载中
  • Bao J.-W., S. G. Gopalakrishnan, S. A. Michelson, F. D. Marks, and M. T. Montgomery, 2012: Impact of physics representations in the HWRFX on simulated hurricane structure and pressure-wind relationships. Mon. Wea. Rev., 140, 3278- 3299.
    Beljaars A. C. M., P. Viterbo, 1998: Role of the boundary layer in a numerical weather prediction model. Clear and Cloudy Boundary Layers, A. A. M. Holtslag, and P. G. Duynkerke, Eds., Royal Netherlands Academy of Arts and Sciences, Amsterdam, 287- 304.f2b22d51-a29b-489a-aecd-c6adcba2c0cf/s?wd=paperuri%3A%28996da912208f796063774e7bfcc2da79%29&filter=sc_long_sign&sc_ks_para=q%3DThe%20role%20of%20the%20boundary%20layer%20in%20a%20Numerical%20Weather%20Prediction%20model&tn=SE_baiduxueshu_c1gjeupa&ie=utf-8
    Bell M. M., M. T. Montgomery, and K. A. Emanuel, 2012: Air-sea enthalpy and momentum exchange at major hurricane wind speeds observed during CBLAST. J. Atmos. Sci., 69, 3197- 3222.10.1175/JAS-D-11-0276.10ddc6ad9-8481-4a41-8a05-e5d5ac81b73cb6ed799e78f30abc6518ab41df96947dhttp://www.researchgate.net/publication/258659563_Air-Sea_Enthalpy_and_Momentum_Exchange_at_Major_Hurricane_Wind_Speeds_Observed_during_CBLASThttp://www.researchgate.net/publication/258659563_Air-Sea_Enthalpy_and_Momentum_Exchange_at_Major_Hurricane_Wind_Speeds_Observed_during_CBLASTQuantifying air--sea exchanges of enthalpy and momentum is important for understanding and skillfully predicting tropical cyclone intensity, but the magnitude of the corresponding wind speed--dependent bulk exchange coefficients is largely unknown at major hurricane wind speeds greater than 50 m s-1. Since direct turbulent flux measurements in these conditions are extremely difficult, the momentum and enthalpy fluxes were deduced via absolute angular momentum and total energy budgets. An error analysis of the methodology was performed to quantify and mitigate potentially significant uncertainties resulting from unresolved budget terms and observational errors. An analysis of six missions from the 2003 Coupled Boundary Layers Air--Sea Transfer (CBLAST) field program in major hurricanes Fabian and Isabel was conducted using a new variational technique. The analysis indicates a near-surface mean drag coefficient CD of 2.4 X 10-3 with a 46% standard deviation and a mean enthalpy coefficient CK of 1.0 X 10-3 with a 40% standard deviation for wind speeds between 52 and 72 m s-1. These are the first known estimates of CK and the ratio of enthalpy to drag coefficient CK/CD in major hurricanes. The results suggest that there is no significant change in the magnitude of the bulk exchange coefficients estimated at minimal hurricane wind speeds, and that the ratio CK/CD does not significantly increase for wind speeds greater than 50 m s-1.
    Black, P. G., Coauthors, 2007: Air-sea exchange in hurricanes: Synthesis of observations from the coupled boundary layer air-sea transfer experiment. Bull. Amer. Meteor. Soc., 88, 357- 374.10.1175/BAMS-88-3-35762bda2e2-8c7e-41f4-b9b9-847f3f77fd710f346f2a4be77c59504c05d738c0c4d9http://www.researchgate.net/publication/215877065_Air--sea_exchange_in_hurricanes_Synthesis_of_observations_from_the_coupled_boundary_layer_air--sea_transfer_experiment?ev=auth_pubhttp://www.researchgate.net/publication/215877065_Air--sea_exchange_in_hurricanes_Synthesis_of_observations_from_the_coupled_boundary_layer_air--sea_transfer_experiment?ev=auth_pubAbstract The Coupled Boundary Layer Air–Sea Transfer (CBLAST) field program, conducted from 2002 to 2004, has provided a wealth of new air–sea interaction observations in hurricanes. The wind speed range for which turbulent momentum and moisture exchange coefficients have been derived based upon direct f lux measurements has been extended by 30% and 60%, respectively, from airborne observations in Hurricanes Fabian and Isabel in 2003. The drag coefficient ( C D ) values derived from CBLAST momentum flux measurements show C D becoming invariant with wind speed near a 23 m s 611 threshold rather than a hurricane-force threshold near 33 m s 611 . Values above 23 m s 611 are lower than previous open-ocean measurements. The Dalton number estimates ( C E ) derived from CBLAST moisture flux measurements are shown to be invariant with wind speeds up to 30 m s 611 , which is in approximate agreement with previous measurements at lower winds. These observations imply a C E / C D ratio of approximately 0.7, suggesting that additional energy sources are necessary for hurricanes to achieve their maximum potential intensity. One such additional mechanism for augmented moisture flux in the boundary layer might be “roll vortex” or linear coherent features, observed by CBLAST 2002 measurements to have wavelengths of 0.9–1.2 km. Linear features of the same wavelength range were observed in nearly concurrent RADARSAT Synthetic Aperture Radar (SAR) imagery. As a complement to the aircraft measurement program, arrays of drifting buoys and subsurface floats were successfully deployed ahead of Hurricanes Fabian (2003) and Frances (2004) [16 (6) and 38 (14) drifters (floats), respectively, in the two storms]. An unprecedented set of observations was obtained, providing a four-dimensional view of the ocean response to a hurricane for the first time ever. Two types of surface drifters and three types of floats provided observations of surface and sub-surface oceanic currents, temperature, salinity, gas exchange, bubble concentrations, and surface wave spectra to a depth of 200 m on a continuous basis before, during, and after storm passage, as well as surface atmospheric observations of wind speed (via acoustic hydrophone) and direction, rain rate, and pressure. Float observations in Frances (2004) indicated a deepening of the mixed layer from 40 to 120 m in approximately 8 h, with a corresponding decrease in SST in the right-rear quadrant of 3.2°C in 11 h, roughly one-third of an inertial period. Strong inertial currents with a peak amplitude of 1.5 m s 611 were observed. Vertical structure showed that the critical Richardson number was reached sporadically during the mixed-layer deepening event, suggesting shear-induced mixing as a prominent mechanism during storm passage. Peak significant waves of 11 m were observed from the floats to complement the aircraft-measured directional wave spectra.
    Braun S. A., W.-K. Tao, 2000: Sensitivity of high-resolution simulations of hurricane Bob (1991) to planetary boundary layer parameterizations. Mon. Wea. Rev., 128, 3941- 3961.10.1175/1520-0493(2000)1292.0.CO;235929886-7863-433a-943d-1cf7783536d8d8ff85e7457cdd89e9cfc139efaee6d4http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F242581818_Sensitivity_of_High-Resolution_Simulations_of_Hurricane_Bob_%281991%29_to_Planetary_Boundary_Layer_Parameterizationsrefpaperuri:(1fd0fedfedfbec6a59e20bb8309a370c)http://www.researchgate.net/publication/242581818_Sensitivity_of_High-Resolution_Simulations_of_Hurricane_Bob_(1991)_to_Planetary_Boundary_Layer_ParameterizationsPresents information on a study which described the sensitivity of high-resolution simulations of Hurricane Bob to planetary boundary layer parameterizations. Role of surface fluxes of sensible and latent heat in the development and maintenance of tropical cyclones; Physics options for the coarse-grid simulation; Coarse-grid results; Conclusions.
    Bryan G. H., 2012: Effects of surface exchange coefficients and turbulence length scales on the intensity and structure of numerically simulated Hurricanes. Mon. Wea. Rev., 140, 1125- 1143.10.1175/MWR-D-11-00231.1b516b4b6-ea9e-451d-bd57-f58391b0fbb3f55570939bdb806d1bad76b595fab795http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F258683536_Effects_of_Surface_Exchange_Coefficients_and_Turbulence_Length_Scales_on_the_Intensity_and_Structure_of_Numerically_Simulated_Hurricanesrefpaperuri:(ad83f7215cd6186c94cfe99bf579cc92)http://www.researchgate.net/publication/258683536_Effects_of_Surface_Exchange_Coefficients_and_Turbulence_Length_Scales_on_the_Intensity_and_Structure_of_Numerically_Simulated_HurricanesAbstract Using numerical simulations, this study examines the sensitivity of hurricane intensity and structure to changes in the surface exchange coefficients and to changes in the length scales of a turbulence parameterization. Compared to other recent articles on the topic, this study uses higher vertical resolution, more values for the turbulence length scales, a different initial environment (including higher sea surface temperature), a broader specification of surface exchange coefficients, a more realistic microphysics scheme, and a set of three-dimensional simulations. The primary conclusions from a recent study by Bryan and Rotunno are all upheld: maximum intensity is strongly affected by the horizontal turbulence length scale l h but not by the vertical turbulence length scale l υ , and the ratio of surface exchange coefficients for enthalpy and momentum, C k / C d , has less effect on maximum wind speed than suggested by an often-cited theoretical model. The model output is further evaluated against various metrics of hurricane intensity and structure from recent observational studies, including maximum wind speed, minimum pressure, surface wind–pressure relationships, height of maximum wind, and surface inflow angle. The model settings l h ≈ 1000 m, l υ ≈ 50 m, and C k / C d ≈ 0.5 produce the most reasonable match to the observational studies. This article also reconciles a recent controversy about the likely value of C k / C d in high wind speeds by noting that simulations in a study by Emanuel used relatively large horizontal diffusion and low sea surface temperature. The model in this study can produce category 5 hurricanes with C k / C d as low as 0.25.
    Carlson T. N., F. E. Boland, 1978: Analysis of urban-rural canopy using a surface heat flux/temperature model. J. Appl. Meteor., 17, 998- 1013.10.1175/1520-0450(1978)0172.0.CO;252ccaf6f-9cc6-43ce-9be9-dc98ab0ab46c314d35901941f55a502b7a7d401e79e7http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F234451527_Analysis_of_Urban-Rural_Canopy_Using_a_Surface_Heat_FluxTemperature_Modelrefpaperuri:(d91957d2135cb3202c8890b13904426c)http://www.researchgate.net/publication/234451527_Analysis_of_Urban-Rural_Canopy_Using_a_Surface_Heat_FluxTemperature_ModelAbstract A one-dimensional numerical model, capable of simulating surface temperature and heat flux, is described in terms of the effective atmospheric and terrain variables. The two model parameters which are most responsible for the formation of important temperature variations in the horizontal over the urban-rural complex are the thermal inertia (thermal property) and moisture availability, the former being most responsible for shaping the nighttime temperature pattern while the latter has a greater effect during the day. The controlling substrate variables are not easily determinable by direct measurement over a surface consisting of an inhomogeneous agglomerate of elements. We present one method whereby surface temperature, a more readily obtainable quantity, can be used in conjunction with the surface model to determine by numerical or graphical inversion of the latter the effective values of moisture availability and thermal inertia and thereby provide a quantitative framework for analysis of a rough surface and for an evaluation of the surface energy budget.
    Charnock H., 1955: Wind stress on a water surface. Quart. J. Roy. Meteor. Soc., 81, 639- 640.10.1002/qj.497080345140ee19794-b1cb-46e9-8397-ee819e20fed919fa6cb3ab6fb4e118315ba242a8c02bhttp://onlinelibrary.wiley.com/doi/10.1002/qj.49708135027/fullhttp://onlinelibrary.wiley.com/doi/10.1002/qj.49708135027/fullABSTRACT Values for the shear stress τ0 of the wind on a water surface, found in seven later experiments, are compared with the data of Francis (1951). Four sets of the new data show that the stress coefficient C = τ0/09u2 is not constant, but that it increases somewhat with windspeed u (09 = air density). Three sets merely show a large experimental scatter of C, and do not support any consistent law of drag. A theory is also discussed that τ0 is mainly caused by the drag of the small ripples, and not by the drag of the big waves.
    Chen S. H., W.-Y. Sun, 2002: A one-dimensional time dependent cloud model. J. Meteor. Soc.Japan, 80( 2), 99- 118.10.2151/jmsj.80.9988c85a7b-b4cd-476d-873b-5548843913682d93b119d0d43306b783d6f640cc5044http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F228377988_A_one-dimensional_time_dependent_cloud_modelrefpaperuri:(fb0bd712e71185e18f0ac5e1435a2166)http://www.researchgate.net/publication/228377988_A_one-dimensional_time_dependent_cloud_modelA one-dimensional prognostic cloud model has been developed for possible use in a Cumulus Parameterization Scheme (CPS). In this model, the nonhydrostatic pressure, entrainment, cloud microphysics, lateral eddy mixing and vertical eddy mixing are included, and their effects are discussed. The inclusion of the nonhydrostatic pressure can (1) weaken vertical velocities, (2) help the cloud develop sooner, (3) help maintain a longer mature stage, (4) produce a stronger overshooting cooling, and (5) approximately double the precipitation amount. The pressure perturbation consists of buoyancy pressure and dynamic pressure, and the simulation results show that both of them are important. We have compared our simulation results with those from Ogura and Takahashi's one-dimensional cloud model, and those from the three-dimensional Weather Research and Forecast (WRF) model. Our model, including detailed cloud microphysics, generates stronger maximum vertical velocity than Ogura and Takahashi's results. Furthermore, the results illustrate that this one-dimensional model is capable of reproducing the major features of a convective cloud that are produced by the three-dimensional model when there is no ambient wind shear.
    Chien F.-C., H.-C. Kuo, 2011: On the extreme rainfall of Typhoon Morakot (2009). J. Geophys. Res., 116,D05104, doi: 10.1029/2010JD015092.10.1029/2010JD015092bf33b49e-a2f1-44e4-b9ca-0cfa92063fea3fecc5631fd4ab59eeb29da237944830http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2010JD015092%2Fcitedbyrefpaperuri:(57e6d6c7d7577336221aa423a27b1c24)http://onlinelibrary.wiley.com/doi/10.1029/2010JD015092/citedby[1] Typhoon Morakot (2009), a devastating tropical cyclone (TC) that made landfall in Taiwan from 7 to 9 August 2009, produced the highest recorded rainfall in southern Taiwan in the past 50 years. This study examines the factors that contributed to the heavy rainfall. It is found that the amount of rainfall in Taiwan was nearly proportional to the reciprocal of TC translation speed rather than the TC intensity. Morakot's landfall on Taiwan occurred concurrently with the cyclonic phase of the intraseasonal oscillation, which enhanced the background southwesterly monsoonal flow. The extreme rainfall was caused by the very slow movement of Morakot both in the landfall and in the postlandfall periods and the continuous formation of mesoscale convection with the moisture supply from the southwesterly flow. A composite study of 19 TCs with similar track to Morakot shows that the uniquely slow translation speed of Morakot was closely related to the northwestward-extending Pacific subtropical high (PSH) and the broad low-pressure systems (associated with Typhoon Etau and Typhoon Goni) surrounding Morakot. Specifically, it was caused by the weakening steering flow at high levels that primarily resulted from the weakening PSH, an approaching short-wave trough, and the northwestward-tilting Etau. After TC landfall, the circulation of Goni merged with the southwesterly flow, resulting in a moisture conveyer belt that transported moisture-laden air toward the east-northeast. Significant mesoscale convection occurred on a long-lived east-west-oriented convergence line and on the mountain slope in southern Taiwan. This convective line was associated with large low-level moisture flux convergence caused by the northwesterly circulation of Morakot and the southwesterly flow. It is thus suggested that the long duration of Typhoon Morakot in the Taiwan area, the interaction of southwest monsoon and typhoon circulation, the mesoscale convection, and the presence of terrain are the key factors in generating the tremendous rainfall.
    Davis C., Coauthors, 2008: Prediction of landfalling hurricanes with the advanced hurricane WRF model. Mon. Wea. Rev., 136, 1990- 2005.10.1175/2007MWR2085.1a59b07e1-f0fc-4e47-afc5-1d5b1e2957f05184c98a0e3f868d1889c3c93f8080eahttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F253593423_Prediction_of_Landfalling_Hurricanes_with_the_Advanced_Hurricane_WRF_Modelrefpaperuri:(cba6c11de7a1102cd1a0ca0cf867e52a)http://www.researchgate.net/publication/253593423_Prediction_of_Landfalling_Hurricanes_with_the_Advanced_Hurricane_WRF_ModelABSTRACT Real-time forecasts of five landfalling Atlantic hurricanes during 2005 using the Advanced Research Weather Research and Forecasting (WRF) (ARW) Model at grid spacings of 12 and 4 km revealed performance generally competitive with, and occasionally superior to, other operational forecasts for storm position and intensity. Recurring errors include 1) excessive intensification prior to landfall, 2) insufficient momentum exchange with the surface, and 3) inability to capture rapid intensification when observed. To address these errors several augmentations of the basic community model have been designed and tested as part of what is termed the Advanced Hurricane WRF (AHW) model. Based on sensitivity simulations of Katrina, the inner-core structure, particularly the size of the eye, was found to be sensitive to model resolution and surface momentum exchange. The forecast of rapid intensification and the structure of convective bands in Katrina were not significantly improved until the grid spacing approached 1 km. Coupling the atmospheric model to a columnar, mixed layer ocean model eliminated much of the erroneous intensification of Katrina prior to landfall noted in the real-time forecast.
    DeCosmo J., K. B. Katsaros, S. D. Smith, R. J. Anderson, W. A. Oost, K. Bumke, and H. Chadwick, 1996: Air-sea exchange of water vapor and sensible heat: The Humidity Exchange over the Sea (HEXOS) results. J. Geophys. Res., 101, 12 001- 12 016.10.1029/95JC037963b9c2b5f-a18e-43b9-a486-ca952c86572d810aad6ff9d55b8e55f42c1d76e33e57http://onlinelibrary.wiley.com/doi/10.1029/95JC03796/abstracthttp://onlinelibrary.wiley.com/doi/10.1029/95JC03796/abstractSurface layer fluxes of sensible heat and water vapor were measured from a fixed platform in the North Sea during the Humidity Exchange over the Sea (HEXOS) Main Experiment (HEXMAX). Eddy wind stress and other relevant atmospheric and oceanic parameters were measured simultaneously and are used to interpret the heat and water vapor flux results. One of the main goals of the HEXOS program was to find accurate empirical heat and water vapor flux parameterization formulas for high wind conditions over the sea. It had been postulated that breaking waves and sea spray, which dominate the air-sea interface at high wind speeds, would significantly affect the air-sea heat and water vapor exchange for wind speeds above 15 m/s. Water vapor flux has been measured at wind speeds up to 18 m/s, sufficient to test these predictions, and sensible heat flux was measured at wind speeds up to 23 m/s. Within experimental error, the HEXMAX data do not show significant variation of the flux exchange coefficients with wind speed, indicating that modification of the models is needed. Roughness lengths for heat and water vapor derived from these direct flux measurements are slightly lower in value but closely parallel the decreasing trend with increasing wind speed predicted by the surface renewal model of Liu et al. [1979], created for lower wind speed regimes, which does not include effects of wave breaking. This suggests that either wave breaking does not significantly affect the surface layer fluxes for the wind speed range in the HEXMAX data, or that a compensating negative feedback process is at work in the lower atmosphere. The implication of the feedback hypothesis is that the moisture gained in the lower atmosphere from evaporation of sea spray over rough seas may be largely offset by decreased vapor flux from the air-sea interface.
    Donelan M. A., B. K. Haus, N. Reul, W. J. Plant, M. Stiassnie, H. C. Graber, O. B. Brown, and E. S. Saltzman, 2004: On the limiting aerodynamic roughness of the ocean in very strong wind. Geophys. Res. Lett., 31,L18306, doi: 10.1029/2004GL 019460.
    Drennan W. M., J. A. Zhang, J. R. French, C. McCormick, and P. G. Black, 2007: Turbulent fluxes in the hurricane boundary layer. Part II: Latent heat flux. J. Atmos. Sci., 64, 1103- 1115.10.1175/JAS3889.1a5fd1f18-2d79-44fd-8f6c-d864dffb4b7fe5765d5258018a67704e3e22110c00c1http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F228449591_Turbulent_fluxes_in_the_hurricane_boundary_layer._Part_II_Latent_heat_fluxrefpaperuri:(0c983f7b9d04ab2dde8398c347c02d8d)http://www.researchgate.net/publication/228449591_Turbulent_fluxes_in_the_hurricane_boundary_layer._Part_II_Latent_heat_fluxAbstract As part of the recent ONR-sponsored Coupled Boundary Layer Air–Sea Transfer (CBLAST) Departmental Research Initiative, an aircraft was instrumented to carry out direct turbulent flux measurements in the high wind boundary layer of a hurricane. During the 2003 field season flux measurements were made during Hurricanes Fabian and Isabel. Here the first direct measurements of latent heat fluxes measured in the hurricane boundary layer are reported. The previous wind speed range for humidity fluxes and Dalton numbers has been extended by over 50%. Up to 30 m s 611 , the highest 10-m winds measured, the Dalton number is not significantly different from the Humidity Exchange over the Sea (HEXOS) result, with no evidence of an increase with wind speed.
    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;2c241bbc0-6ba8-437a-b86c-3796ec18bb1b34a0f338a8622d0aee3c3811d44d3450http%3A%2F%2Fci.nii.ac.jp%2Fnaid%2F10013124897refpaperuri:(76536b43084d2ca4dd6ac23f1a23d059)http://ci.nii.ac.jp/naid/10013124897Abstract A two-dimensional version of the Pennsylvania State University mesoscale model has been applied to Winter Monsoon Experiment data in order to simulate the diurnally occurring convection observed over the South China Sea. The domain includes a representation of part of Borneo as well as the sea so that the model can simulate the initiation of convection. Also included in the model are parameterizations of mesoscale ice phase and moisture processes and longwave and shortwave radiation with a diurnal cycle. This allows use of the model to test the relative importance of various heating mechanisms to the stratiform cloud deck, which typically occupies several hundred kilometers of the domain. Frank and Cohen's cumulus parameterization scheme is employed to represent vital unresolved vertical transports in the convective area. The major conclusions are: Ice phase processes are important in determining the level of maximum large-scale heating and vertical motion because there is a strong anvil component. The heating is initiated by a thermodynamic adjustment that takes place after the air leaves the updrafts and is associated with the difference between water and ice saturation. Melting and evaporation contribute to a 1ocalized mesoscale subsidence in a 50 km region to the rear of the moving convective area. The cooling associated with this almost cancels the cumulus heating in the lower to midtroposphere. Radiative heating was found to be the main ascent-forcing influence at high levels occupied by the widespread cirrus outflow. Additionally, radiative clear-air cooling helped the convection by continuously destabilizing the troposphere and countering the warming effect of convective updrafts. The overall structure and development of the system were well simulated, particularly the growth near the coast, and the propagation and decay in the cooler boundary layer further off-shore, but the rainfall may have been underestimated because of the two-dimensional assumptions of the model.
    Dudhia, J., Coauthors, 2008: Prediction of Atlantic tropical cyclones with the Advanced Hurricane WRF (AHW) model. 28th Conf. on Hurricanes and Tropical Meteorology, Orlando, Florida, Preprint 18A. 2.cedcce8a-c9a9-4adb-b090-2b951b4a86d2/s?wd=paperuri%3A%28386dbb5c7ae7877cea0627e95a157023%29&filter=sc_long_sign&sc_ks_para=q%3DPREDICTION%20OF%20ATLANTIC%20TROPICAL%20CYCLONES%20WITH%20THE%20ADVANCED%20HURRICANE%20WRF%20%28AHW%29%20MODEL&tn=SE_baiduxueshu_c1gjeupa&ie=utf-8
    Emanuel K. A., 1986: An air-sea interaction theory for tropical cyclones. Part I: Steady-state maintenance. J. Atmos. Sci., 43, 585- 605.10.1175/1520-0469(1986)0432.0.CO;2af0a3265-005a-442d-9f3f-af90be5dccaf504439cc64b4832397ce434098772cbdhttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F215877085_An_air-sea_interaction_theory_for_tropical_cyclones._Part_I_Steady-state_maintenancerefpaperuri:(830a418b534890b226fee3a12303fa60)http://www.researchgate.net/publication/215877085_An_air-sea_interaction_theory_for_tropical_cyclones._Part_I_Steady-state_maintenanceAbstract Observations and numerical simulators of tropical cyclones show that evaporation from the sea surface is essential to the development of reasonably intense storms. On the other hand, the CISK hypothesis, in the form originally advanced by Charney and Eliassen, holds that initial development results from the organized release of preexisting conditional instability. In this series of papers, we explore the relative importance of ambient conditional instability and air-sea latent and sensible heat transfer in both the development and maintenance of tropical cyclones using highly idealized models. In particular, we advance the hypothesis that the intensification and maintenance of tropical cyclones depend exclusively on self-induced heat transfer from the ocean. In this sense, these storms may be regarded as resulting from a finite amplitude air-sea interaction instability rather than from a linear instability involving ambient potential buoyancy. In the present paper, we attempt to show that reasonably intense cyclones may be maintained in a steady state without conditional instability of ambient air. In Part II we will demonstrate that weak but finite-amplitude axisymmetric disturbances may intensify in a conditionally neutral environment.
    Emanuel K. A., 1995: Sensitivity of tropical cyclones to surface exchange coefficients and a revised steady-state model incorporating eye dynamics. J. Atmos. Sci., 52, 3969- 3976.10.1175/1520-0469(1995)0522.0.CO;226e5e6af-531d-4509-8a13-ed111ff2d09a9ce2c9bb8eda1ddf38eebf370b607942http://www.researchgate.net/publication/215877087_Sensitivity_of_tropical_cyclones_to_surface_exchange_coefficients_and_a_revised_steady-state_model_incorporating_eye_dynamicshttp://www.researchgate.net/publication/215877087_Sensitivity_of_tropical_cyclones_to_surface_exchange_coefficients_and_a_revised_steady-state_model_incorporating_eye_dynamicsAbstract Numerical and theoretical models of tropical cyclones indicate that the maximum wind speed in mature storms is sensitive to the ratio of the enthalpy and momentum surface exchange coefficients and that the spinup time of tropical cyclones varies inversely with the magnitude of these coefficients. At the same time, the Carnot cycle model developed by the author predicts that the central pressure of mature cyclones is independent of the magnitude of the exchange coefficients. The author presents numerical simulations that prove this last prediction false and suggest that the reason for this failure is the neglect of eye dynamics in the steady-state theory. On this basis, the existing theory is modified to account for eye dynamics, and the predictions of the revised theory are compared to the results of numerical simulations. Both the revised theory and the numerical modeling results, when compared to observations, suggest that the ratio of enthalpy to momentum exchange coefficients in real hurricanes lies in the range 0.75-1.5, contradicting published speculations about the behavior of this ratio at high surface wind speed.
    French J. R., W. M. Drennan, J. A. Zhang, and P. G. Black, 2007: Turbulent fluxes in the hurricane boundary layer. Part I: Momentum flux. J. Atmos. Sci., 64, 1089- 1102.10.1175/JAS3887.15d1f8c93-7f4c-4078-9b87-ff918a64a82ebaebbb88d10020b62bdc09f20b0d8d97http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F240686670_Turbulent_Fluxes_in_the_Hurricane_Boundary_Layer._Part_I_Momentum_Fluxrefpaperuri:(d9606d69e69040d8a8c6c3f787db99da)http://www.researchgate.net/publication/240686670_Turbulent_Fluxes_in_the_Hurricane_Boundary_Layer._Part_I_Momentum_FluxAbstract An important outcome from the ONR-sponsored Coupled Boundary Layer Air–Sea Transfer (CBLAST) Hurricane Program is the first-ever direct measurements of momentum flux from within hurricane boundary layers. In 2003, a specially instrumented NOAA P3 aircraft obtained measurements suitable for computing surface wind stress and ultimately estimating drag coefficients in regions with surface wind between 18 and 30 m s 611 . Analyses of data are presented from 48 flux legs flown within 400 m of the surface in two storms. Results suggest a roll-off in the drag coefficient at higher wind speeds, in qualitative agreement with laboratory and modeling studies and inferences of drag coefficients using a log-profile method. However, the amount of roll-off and the wind speed at which the roll-off occurs remains uncertain, underscoring the need for additional measurements.
    Garratt J. R., 1992: The Atmospheric Boundary Layer. Cambridge University Press,316 pp.
    Geernaert G. L., K. L. Davidson, S. E. Larsen, and T. Mikkelsen, 1988: Wind stress measurements during the Tower Ocean Wave and Radar Dependence Experiment. J. Geophys. Res., 93, 13 913- 13 923.10.1029/JC093iC11p139137b0619ae-7d1f-4e01-892b-04a2d32384a0744a576a46a66bd3a7ca8114fa00cbadhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2FJC093iC11p13913%2Ffullrefpaperuri:(704ec49bde76404ed958669227e0cdb0)http://onlinelibrary.wiley.com/doi/10.1029/JC093iC11p13913/fullDuring the Tower Ocean Wave and Radar Dependence Experiment (TOWARD), near-continuous measurements of the wind drag were conducted using the dissipation technique. An intercomparison between these measurements and direct stress magnitudes using a sonic anemometer was performed over 3 days of the experiment. The results indicated that the dissipation technique compared well to the directly determined values when conditions were steady and neutral; otherwise, the dissipation method performed poorly. When neutrally stratified data were used, the drag coefficient exhibited a systematic dependence on both surface tension and wave age.
    Green B. W., F. Q. Zhang, 2013: Impacts of air-sea flux parameterizations on the intensity and structure of tropical cyclones. Mon. Wea. Rev., 141, 2308- 2324.10.1175/MWR-D-12-00274.18d6d1c9a-5258-4848-86ab-9d98e25a589dc34dfa37050432560f73bda8770acc4ehttp://www.researchgate.net/publication/277449791_Impacts_of_AirSea_Flux_Parameterizations_on_the_Intensity_and_Structure_of_Tropical_Cyclones?ev=auth_pubhttp://www.researchgate.net/publication/277449791_Impacts_of_AirSea_Flux_Parameterizations_on_the_Intensity_and_Structure_of_Tropical_Cyclones?ev=auth_pubFluxes of momentum and moist enthalpy across the air-sea interface are believed to be one of the most important factors in determining tropical cyclone intensity. Because these surface fluxes cannot be directly resolved by numerical weather prediction models, their impacts on tropical cyclones must be accounted for through subgrid-scale parameterizations. There are several air-sea surface flux parameterization schemes available in the Weather Research and Forecasting (WRF) Model; these schemes differ from one another in their formulations of the wind speed-dependent exchange coefficients of momentum, sensible heat, and moisture (latent heat). The effects of surface fluxes on the intensity and structure of tropical cyclones are examined through convection-permitting WRF simulations of Hurricane Katrina (2005). It is found that the intensity (and, to a lesser extent, structure) of the simulated storms is sensitive to the choice of surface flux parameterization scheme. In agreement with recent studies, the drag coefficient CD is found to affect the pressure-wind relationship (between minimum sea level pressure and maximum 10-m wind speed) and to change the radius of maximum near-surface winds of the tropical cyclone. Fluxes of sensible and latent heat (i.e., moist enthalpy) affect intensity but do not significantly change the pressure-wind relationship. Additionally, when low-level winds are strong, the contribution of dissipative heating to calculations of sensible heat flux is not negligible. Expanding the sensitivity tests to several dozen cases from the 2008 to 2011 Atlantic hurricane seasons demonstrates the robustness of these findings.
    Hall J. D., M. Xue, L. K. Ran, L. M. Leslie, 2013: High-resolution modeling of Typhoon Morakot (2009): Vortex Rossby waves and their role in extreme precipitation over Taiwan. J. Atmos. Sci., 70, 163- 186.10.1175/JAS-D-11-0338.15e6de160-5ffe-4615-9b65-f7805f96276f1b14996c56101941ae7ccabfddad3be3http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F258794807_High-Resolution_Modeling_of_Typhoon_Morakot_%282009%29_Vortex_Rossby_Waves_and_Their_Role_in_Extreme_Precipitation_over_Taiwanrefpaperuri:(7b32a105c98d6f5107db1ffe6962cd3a)http://www.researchgate.net/publication/258794807_High-Resolution_Modeling_of_Typhoon_Morakot_(2009)_Vortex_Rossby_Waves_and_Their_Role_in_Extreme_Precipitation_over_TaiwanAbstract A high-resolution nonhydrostatic numerical model, the Advanced Regional Prediction System (ARPS), was used to simulate Typhoon Morakot (2009) as it made landfall over Taiwan, producing record rainfall totals. In particular, the mesoscale structure of the typhoon was investigated, emphasizing its associated deep convection, the development of inner rainbands near the center, and the resultant intense rainfall over western Taiwan. Simulations at 15- and 3-km grid spacing revealed that, following the decay of the initial inner eyewall, a new, much larger eyewall developed as the typhoon made landfall over Taiwan. Relatively large-amplitude wave structures developed in the outer eyewall and are identified as vortex Rossby waves (VRWs), based on the wave characteristics and their similarity to VRWs identified in previous studies. Moderate to strong vertical shear over the typhoon system produced a persistent wavenumber-1 (WN1) asymmetric structure during the landfall period, with upward motion and deep convection in the downshear and downshear-left sides, consistent with earlier studies. This strong asymmetry masks the effects of WN1 VRWs. WN2 and WN3 VRWs apparently are associated with the development of deep convective bands in Morakot southwestern quadrant. This occurs as the waves move cyclonically into the downshear side of the cyclone. Although the typhoon track and topographic enhancement contribute most to the record-breaking rainfall totals, the location of the convective bands, and their interaction with the mountainous terrain of Taiwan, also affect the rainfall distribution. Quantitatively, the 3-km ARPS rainfall forecasts are superior to those obtained from coarser-resolution models.
    Haus B. K., D. Jeong, M. A. Donelan, J. A. Zhang, and I. Savelyev, 2010: Relative rates of sea-air heat transfer and frictional drag in very high winds. Geophys. Res. Lett. 37,L07802, doi: 10.1029/2009GL042206.10.1029/2009GL0422064e86fe49-6f6a-445a-b1b6-24a7712489bb1cffdb84dc7a61a2acf3c9d93e3e3ef3http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2009GL042206%2Fpdfrefpaperuri:(87434d6e78a3294213503f22996ffc6b)http://onlinelibrary.wiley.com/doi/10.1029/2009GL042206/pdfABSTRACT Hurricanes are fueled by evaporation and convection from the ocean and they lose energy through the frictional drag of the atmosphere on the ocean surface. The relative rates of these processes have been thought to provide a limit on the maximum potential hurricane intensity. Here we report laboratory observations of these transfers for scaled winds equivalent to a strong Category 1 hurricane (38 ms-1). We show that the transfer coefficient ratio holds closely to a level of ˜0.5 even in the highest observed winds, where previous studies have suggested there is a distinct regime change at the air-sea interface. This value is well below the expected threshold value for intense hurricanes of 0.75. Recent three-dimensional model studies also find that the coefficient ratio can be much lower than 0.75, which suggests that other factors such as eyewall and/or vortex dynamics are responsible for the formation of very strong hurricanes.
    Hong S. Y., Y. Noh, and J. Dudhia, 2006: A new vertical diffusion package with an explicit treatment of entrainment processes. Mon. Wea. Rev., 134( 9), 2318- 2341.10.1175/MWR3199.179f98ee85a3853a6bfee0ec84e90c901http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F252170327_A_New_Vertical_Diffusion_Package_with_an_Explicit_Treatment_of_Entrainment_Processeshttp://www.researchgate.net/publication/252170327_A_New_Vertical_Diffusion_Package_with_an_Explicit_Treatment_of_Entrainment_ProcessesAbstract This paper proposes a revised vertical diffusion package with a nonlocal turbulent mixing coefficient in the planetary boundary layer (PBL). Based on the study of Noh et al. and accumulated results of the behavior of the Hong and Pan algorithm, a revised vertical diffusion algorithm that is suitable for weather forecasting and climate prediction models is developed. The major ingredient of the revision is the inclusion of an explicit treatment of entrainment processes at the top of the PBL. The new diffusion package is called the Yonsei University PBL (YSU PBL). In a one-dimensional offline test framework, the revised scheme is found to improve several features compared with the Hong and Pan implementation. The YSU PBL increases boundary layer mixing in the thermally induced free convection regime and decreases it in the mechanically induced forced convection regime, which alleviates the well-known problems in the Medium-Range Forecast (MRF) PBL. Excessive mixing in the mixed layer in the presence of strong winds is resolved. Overly rapid growth of the PBL in the case of the Hong and Pan is also rectified. The scheme has been successfully implemented in the Weather Research and Forecast model producing a more realistic structure of the PBL and its development. In a case study of a frontal tornado outbreak, it is found that some systematic biases of the large-scale features such as an afternoon cold bias at 850 hPa in the MRF PBL are resolved. Consequently, the new scheme does a better job in reproducing the convective inhibition. Because the convective inhibition is accurately predicted, widespread light precipitation ahead of a front, in the case of the MRF PBL, is reduced. In the frontal region, the YSU PBL scheme improves some characteristics, such as a double line of intense convection. This is because the boundary layer from the YSU PBL scheme remains less diluted by entrainment leaving more fuel for severe convection when the front triggers it.
    Hosomi T., 2005: Implementation of targeted moisture diffusion for the JMA Regional Spectral Model (RSM). CAS/JSC WGNE Research Activities in Atmospheric and Oceanic Modelling/WMO. , 35, 7- 8.be97e683-2551-40e0-ba83-715691c7ba226f9c3d485bc43ff41ee4a5a9bcd5297bhttp%3A%2F%2Fwww.wcrp-climate.org%2FWGNE%2FBlueBook%2F2005%2Findividual-articles%2F05_Hosomi_Takuya_jma_rsm_tmd.pdfrefpaperuri:(57fdf69fb5e1028815598792aae2e250)http://www.wcrp-climate.org/WGNE/BlueBook/2005/individual-articles/05_Hosomi_Takuya_jma_rsm_tmd.pdfJapan Meteorological Agency operates the Regional Spectral Model (RSM) twice a day. It covers East Asia and makes a forecast up to 51 hours. A forecast of the JMA Global Spectral Model (GSM) is used for a lateral boundary condition. An initial condition is analyzed by a
    Huang H.-L., M.-J. Yang, and C.-H. Sui, 2014: Water budget and precipitation efficiency of Typhoon Morakot (2009). J. Atmos. Sci., 71, 112- 129.10.1175/JAS-D-13-053.1430ced74-929d-4930-b696-ff6b8e1722aa0e50743106890727c9000b4de4c8bed4http%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FXREF%3Fid%3D10.1175%2FJAS-D-13-053.1refpaperuri:(b5e7f05aa58f3e42145f90e79a9ca277)http://onlinelibrary.wiley.com/resolve/reference/XREF?id=10.1175/JAS-D-13-053.1Abstract In this study, the Weather Research and Forecasting model, version 3.2, with the finest grid size of 1 km is used to explicitly simulate Typhoon Morakot (2009), which dumped rainfall of more than 2600 mm in 3 days on Taiwan. The model reasonably reproduced the track, the organization, the sizes of the eye and eyewall, and the characteristics of major convective cells in outer rainbands. The horizontal rainfall distribution and local rainfall maximum in the southwestern portion of the Central Mountain Range (CMR) are captured. The simulated rain rate and precipitation efficiency (PE) over the CMR are highly correlated. In the absence of terrain forcing, the simulated TC’s track is farther north and rainfall distribution is mainly determined by rainbands. The calculated rain rate and PE over the CMR during landfall are about 50% and 15%–20% less than those of the full-terrain control run, respectively. By following major convective cells that propagate eastward from the Taiwan Strait to the CMR, it is found that the PE and the processes of vapor condensation and raindrop evaporation are strongly influenced by orographic lifting; the PEs are 60%–75% over ocean and more than 95% over the CMR, respectively. The secondary increase of PE results from the increase of ice-phase deposition ratio when the liquid-phase condensation becomes small as the air on the lee side subsides and moves downstream. This nearly perfect PE over the CMR causes tremendous rainfall in southwestern Taiwan, triggering enormous landslides and severe flooding.
    JMA, 2007: Outline of the operational numerical weather prediction at the Japan Meteorological Agency. Appendix to WMO Numerical Weather Prediction Progress Report, Japan Meteorological Agency, Tokyo, 194 pp./s?wd=paperuri%3A%28788531031b89ca05cba38bad7aa316f1%29&filter=sc_long_sign&sc_ks_para=q%3DOutline%20of%20operational%20numerical%20weather%20prediction%20at%20Japan%20Meteorological%20Agency&tn=SE_baiduxueshu_c1gjeupa&ie=utf-8
    Large W. G., S. Pond, 1981: Open ocean momentum flux measurements in moderate to strong winds. J. Phys. Oceanogr., 11, 324- 336.49189dde-3478-4641-bcc6-5ced3167d6d385ffa5109ca53df3809ae633005221c3http%3A%2F%2Fwww.bioone.org%2Fservlet%2Flinkout%3Fsuffix%3Di1551-5036-24-sp3-159-Large1%26dbid%3D16%26doi%3D10.2112%252F06-0795.1%26key%3D10.1175%252F1520-0485%281981%29011%3C0324%253AOOMFMI%3E2.0.CO%253B2refpaperuri:(173ee46bbe9c0801121bf36602f46a6e)http://www.bioone.org/servlet/linkout?suffix=i1551-5036-24-sp3-159-Large1&dbid=16&doi=10.2112%2F06-0795.1&key=10.1175%2F1520-0485(1981)011%3C0324%3AOOMFMI%3E2.0.CO%3B2
    Li X., J. Ming, Y. Wang, K. Zhao, and M. Xue, 2013: Assimilation of T-TREC-retrieved wind data with WRF 3DVAR for the short-term forecasting of typhoon Meranti (2010) near landfall. J. Geophys. Res., 118, 10 361- 10 375.10.1002/jgrd.508150a097cb6-b94f-4b69-b61f-c9a3c8e18c5b90f01285b704be32a129061cd6171f05http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fjgrd.50815%2Fabstract%3Bjsessionid%3D1661830CC2477FB16C828F39AE713ECF.f04t03refpaperuri:(3539ed405e6e232a9a6bac2bf87961e1)http://d.wanfangdata.com.cn/Conference_WFHYXW561532.aspx?f=datatangABSTRACT extended Tracking Radar Echo by Correlation (TREC) technique, called T-TREC technique, has been developed recently to retrieve horizontal circulations within tropical cyclones (TCs) from single Doppler radar reflectivity (Z) and radial velocity (Vr, when available) data. This study explores, for the first time, the assimilation of T-TREC-retrieved winds for a landfalling typhoon, Meranti (2010), into a convection-resolving model, the WRF (Weather Research and Forecasting). The T-TREC winds or the original Vr data from a single coastal Doppler radar are assimilated at the single time using the WRF three-dimensional variational (3DVAR), at 8, 6, 4, and 2 h before the landfall of typhoon Meranti. In general, assimilating T-TREC winds results in better structure and intensity analysis of Meranti than directly assimilating Vr data. The subsequent forecasts for the track, intensity, structure and precipitation are also better, although the differences becomes smaller as the Vr data coverage improves when the typhoon gets closer to the radar. The ability of the T-TREC retrieval in capturing more accurate and complete vortex circulations in the inner-core region of TC is believed to be the primary reason for its superior performance over direct assimilation of Vr data; for the latter, the data coverage is much smaller when the TC is far away and the cross-beam wind component is difficult to analyze accurately with 3DVAR method.
    Lin Y.-L., R. D. Farley, and H. D. Orville, 1983: Bulk parameterization of the snow field in a cloud model. J. Climate Appl. Meteor., 22( 6), 1065- 1092.
    Liu Y. B., D.-L. Zhang, and M. K. Yau, 1999: A multiscale numerical study of Hurricane Andrew (1992). Part II: Kinematics and inner-core structures. Mon. Wea. Rev., 127, 2597- 2616.10.1175/1520-0493(1999)127<2597:AMNSOH>2.0.CO;2b0c4d84d-39fe-4658-8e91-1c9eaa6f653d92854b578b2d9d4412d2ef6605d09b62http%3A%2F%2Fci.nii.ac.jp%2Fnaid%2F80010744120refpaperuri:(e8dba250839fb18f58d52270a289bf35)http://ci.nii.ac.jp/naid/80010744120ABSTRACT Despite considerable research, understanding of the temporal evolution of the inner-core structures of hurricanes is very limited owing to the lack of continuous high-resolution observational data of a storm. In this study, the results of a 72-h explicit simulation of Hurricane Andrew (1992) with a grid size of 6 km are examined to explore the inner-core axisymmetric and asymmetric structures of the storm during its rapid deepening stage. Based on the simulation, a conceptual model of the axisymmetric structures of the storm is proposed. Most of the proposed structures confirm previous observations. The main ingredients include a main inflow (outflow) in the boundary layer (upper troposphere) with little radial flow in between, a divergent slantwise ascent in the eyewall, a penetrative dry downdraft at the inner edge of the eyewall, and a general weak subsiding motion in the eye with typical warming/drying above an inversion located near an altitude of about 2--3 km. The storm ...
    Malkus J. S., H. Riehl, 1960: On the dynamics and energy transformations in steady-state hurricanes. Tellus, 12, 1- 20.10.1111/j.2153-3490.1960.tb01279.xe59be9d0-ced8-4e17-afc1-d10e89ace9b8d47415c540430f79085fb251e533bc8bhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1111%2Fj.2153-3490.1960.tb01279.x%2Fpdfrefpaperuri:(1cc9aa1428cb77949faa3143473cc3ec)http://onlinelibrary.wiley.com/doi/10.1111/j.2153-3490.1960.tb01279.x/pdfFinally, rainfall, efficiency of work done by the storm, and kinetic energy budgets are examined in an attempt to understand the difference between the hurricane—a rare phenomenon—and the common sub-hurricane tropical storm.
    Ming J., Y. Q. Ni, and X. Y. Shen, 2009: The dynamical characteristics and wave structure of typhoon Rananim (2004). Adv. Atmos. Sci.,26, 523-542, doi: 10.1007/s00376-009-0523-0.
    Ming J., J. J. Song, B. J. Chen, and K. F. Wang, 2012: Boundary layer structure in typhoon Saomai (2006): Understanding the effects of exchange coefficient. J. Trop. Meteor., 18( 3), 195- 206.10.3969/j.issn.1006-8775.2012.02.00941ff03a3-df92-42f6-b45f-9aa75fb1b88980284d44fda421528b56279957e6498dhttp://www.cnki.com.cn/Article/CJFDTotal-RQXB201202011.htmhttp://www.cnki.com.cn/Article/CJFDTotal-RQXB201202011.htmRecent studies have shown that surface fluxes and exchange coefficients are particularly important to models attempting to simulate the evolution and maintenance of hurricanes or typhoons.By using an advanced research version of the Weather Research and Forecasting(ARW)modeling system,this work aims to study the impact of modified exchange coefficient on the intensity and structures of typhoon Saomai(2006)over the western North Pacific.Numerical experiments with the modified and unmodified exchange coefficients are used to investigate the intensity and structure of the storm,especially the structures of the boundary layer within the storm.Results show that,compared to the unmodified experiment,the simulated typhoon in the modified experiment has a bigger deepening rate after 30-h and is the same as the observation in the last 12-h.The roughness is leveled off when wind speed is greater than 30 m/s.The momentum exchange coefficient(CD)and enthalpy exchange coefficient(CK)are leveled off too,and CD is decreased more than CK when wind speed is greater than 30 m/s.More sensible heat flux and less latent heat flux are produced.In the lower level,the modified experiment has slightly stronger outflow,stronger vertical gradient of equivalent potential temperature and substantially higher maximum tangential winds than the unmodified experiment has.The modified experiment generates larger wind speed and water vapor tendencies and transports more air of high equivalent potential temperature to the eyewall in the boundary layer.It induces more and strong convection in the eyewall,thereby leading to a stronger storm.
    Mlawer E. J., S. J. Taubman, P. D. Brown, M. J. Iacono, and S. A. Clough, 1997: Radiative transfer for inhomogeneous atmosphere: RRTM, a validated correlated-k model for the longwave. J. Geophys. Res., 102, 16 663- 16 682.10.1029/97JD002372e06658e-dcf6-4eb8-a12a-08b5972b271fc69b740e6b8ec97bd8341a4efa8fd7a2http://onlinelibrary.wiley.com/doi/10.1029/97JD00237/pdfhttp://onlinelibrary.wiley.com/doi/10.1029/97JD00237/pdfABSTRACT A rapid and accurate radiative transfer model (RRTM) for climate applications 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-1) 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-2 (relative to LBLRTM) for net flux in each band at all altitudes, with a total (10-3000 cm-1) error of less than 1.0 W m-2 at any altitudes; 0.07 K d-1 for total cooling rate error in the troposphere and lower stratosphere, and 0.75 K d-1 in 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-2, 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.
    Montgomery M. T., R. K. Smith, and S. V. Nguyen, 2010: Sensitivity of tropical-cyclone models to the surface drag coefficient . Quart. J. Roy. Meteor. Soc.,136, 1945-1953, doi: 10.1002/qj.702.10.1002/qj.702659c8489-69a3-49ce-9e15-97f0d517158ac0aea88c37ee90d12995939355dcdcf8http://onlinelibrary.wiley.com/doi/10.1002/qj.702/pdfhttp://onlinelibrary.wiley.com/doi/10.1002/qj.702/pdfThe contribution of M. T. Montgomery to this article was prepared as part of his official duties as a United States Federal Government employee.
    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( 3), 401- 427.10.1023/A:102214601594650030b1a-fc69-48d5-921f-d9252d61dfa92b395fb205157effe9f061a3ffca68b8http%3A%2F%2Flink.springer.com%2Farticle%2F10.1023%2FA%3A1022146015946refpaperuri:(18193937665365080c4f347114f8bfc9)http://link.springer.com/article/10.1023/A:1022146015946Modifications 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.
    Nolan D. S., J. A. Zhang, and D. P. Stern, 2009: Evaluation of planetary boundary layer parameterizations in tropical cyclones by comparison of in situ data and high-resolution simulations of Hurricane Isabel (2003). Part I: Initialization, maximum winds, and the outer-core boundary layer structure. Mon. Wea. Rev., 137, 3651- 3674.
    Ooyama K., 1969: Numerical simulation of the life cycle of tropical cyclones. J. Atmos. Sci., 26, 3- 40.10.1175/1520-0469(1969)026<0003:NSOTLC>2.0.CO;2016d2522-6cd2-4d8b-b94e-1fdeb0332085895fda861cbdaeb2f7b42c91ed54e737http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F215877167_Numerical_simulation_of_the_life_cycle_of_tropical_cyclonesrefpaperuri:(59c5752acdb9443cdac53f609537932d)http://www.researchgate.net/publication/215877167_Numerical_simulation_of_the_life_cycle_of_tropical_cyclonesAbstract The tropical cyclone is a solitary creature of the tropical oceans accompanied by violent rotating winds and torrential rain. Observational studies and diagnostic analyses leave little doubt that the energy required for driving the vortex comes from the latent heat of condensation released by tall convective clouds around the center, and that the frictionally induced inflow in the vortex plays a major role in supporting the continued activity of convective clouds. This dual character with respect to important scales of motion poses a great difficulty in investigating the dynamics of tropical cyclones as time-dependent phenomena. However, in order to understand the large-scale aspects of tropical cyclones, one may formulate the role of convective clouds in terms of cyclone-scale variables with only implicit consideration of the dynamics of individual clouds. The present study is an attempt to understand the basic mechanism of tropical cyclones by constructing a numerical-dynamical model on such a basis. The model assumes that the large-scale hydrodynamical aspects of a tropical cyclone may be represented by an axisymmetric, quasi-balanced vortex in a stably stratified incompressible fluid, while the effects of moist convection may be formulated through the first law of thermodynamics applied to an implicit model of penetrative convective clouds. The airea exchange of angular momentum as well as latent and sensible heat is explicitly calculated in the model with the use of conventional approximations. Results of numerical integration show that the model is capable of simulating the typical life cycle of tropical cyclones, including the mature hurricane stage, with a remarkable degree of reality. The response of the model cyclone to changes in such parameters as the sea surface temperature, the coefficient of airea energy exchange, and the initial conditions is tested in a number of numerical experiments to show quite plausible results. A detailed diagnosis of the energy budget of the simulated tropical cyclone is also carried out. The rate of total rainfall, the production and dissipation of kinetic energy, and other energetic characteristics of the computed cyclone compare very well with available estimates for observed tropical cyclones. Because of the restrictive assumption of axisymmetry and other weak approximations, the model is not realistic enough to predict behavior of individual tropical cyclones in nature. The limitation of the present model in this regard is also discussed.
    Powell M. D., P. J. Vickery, and T. A. Reinhold, 2003: Reduced drag coefficient for high wind speeds in tropical cyclones. Nature, 422, 279- 283.10.1038/nature0148112646913014e5637-f005-4c58-9875-aa4510166466977c52112ed81307ce09c9ff8acab781http%3A%2F%2Fmed.wanfangdata.com.cn%2FPaper%2FDetail%2FPeriodicalPaper_PM12646913refpaperuri:(9f794ceb92e784a337ee22f4fed807a6)http://med.wanfangdata.com.cn/Paper/Detail/PeriodicalPaper_PM12646913The transfer of momentum between the atmosphere and the ocean is described in terms of the variation of wind speed with height and a drag coefficient that increases with sea surface roughness and wind speed. But direct measurements have only been available for weak winds; momentum transfer under extreme wind conditions has therefore been extrapolated from these field measurements. Global Positioning System sondes have been used since 1997 to measure the profiles of the strong winds in the marine boundary layer associated with tropical cyclones. Here we present an analysis of these data, which show a logarithmic increase in mean wind speed with height in the lowest 200 m, maximum wind speed at 500 m and a gradual weakening up to a height of 3 km. By determining surface stress, roughness length and neutral stability drag coefficient, we find that surface momentum flux levels off as the wind speeds increase above hurricane force. This behaviour is contrary to surface flux parameterizations that are currently used in a variety of modelling applications, including hurricane risk assessment and prediction of storm motion, intensity, waves and storm surges.
    Rosenthal S. L., 1971: The response of a tropical cyclone model to variations in boundary layer parameters, initial conditions, lateral boundary conditions and domain size. Mon. Wea. Rev., 99, 767- 777.10.1175/JCLI-3293.14184c72d-5ff7-4cde-8f33-1fa985e26c7cac5e8a56e6565a1c40fc3f604a2a8dbfhttp://www.researchgate.net/publication/249611701_Estimating_the_Uncertainty_in_a_Regional_Climate_Model_Related_to_Initial_and_Lateral_Boundary_Conditions?ev=auth_pubhttp://www.researchgate.net/publication/249611701_Estimating_the_Uncertainty_in_a_Regional_Climate_Model_Related_to_Initial_and_Lateral_Boundary_Conditions?ev=auth_pubThere is a growing demand for regional-scale climate predictions and assessments. Quantifying the impacts of uncertainty in initial conditions and lateral boundary forcing data on regional model simulations can potentially add value to the usefulness of regional climate modeling. Results from a regional model depend on the realism of the driving data from either global model outputs or global analyses; therefore, any biases in the driving data will be carried through to the regional model. This study used four popular global analyses and achieved 16 driving datasets by using different interpolation procedures. The spread of the 16 datasets represents a possible range of driving data based on analyses to the regional model. This spread is smaller than typically associated with global climate model realizations of the Arctic climate. Three groups of 16 realizations were conducted using the fifth-generation Pennsylvania State University ational Center for Atmospheric Research (PSU CAR) Mesoscale Model (MM5) in an Arctic domain, varying both initial and lateral boundary conditions, varying lateral boundary forcing only, and varying initial conditions only. The response of monthly mean atmospheric states to the variations in initial and lateral driving data was investigated. Uncertainty in the regional model is induced by the interaction between biases from different sources. Because of the nonlinearity of the problem, contributions from initial and lateral boundary conditions are not additive. For monthly mean atmospheric states, biases in lateral boundary conditions generally contribute more to the overall uncertainty than biases in the initial conditions. The impact of initial condition variations decreases with the simulation length while the impact of variations in lateral boundary forcing shows no clear trend. This suggests that the representativeness of the lateral boundary forcing plays a critical role in long-term regional climate modeling. The extent of impact of the driving data uncertainties on regional climate modeling is variable dependent. For some sensitive variables (e.g., precipitation, boundary layer height), even the interior of the model may be significantly affected.
    Rotunno R., K. A. Emanuel, 1987: An air-sea interaction theory for tropical cyclones. Part II: Evolutionary study using a nonhydrostatic axisymmetric numerical model. J. Atmos. Sci., 44, 542- 561.10.1175/1520-0469(1987)044<0542:AAITFT>2.0.CO;210f56979-7771-4e79-9655-57c861f7428fc30f99c2c09dabed521896d960950cbehttp://www.researchgate.net/publication/215877180_An_AirSea_interaction_theory_for_tropical_cyclones._Part_II_Evolutionary_study_using_a_nonhydrostatic_axisymmetric_numerical_modelhttp://www.researchgate.net/publication/215877180_An_AirSea_interaction_theory_for_tropical_cyclones._Part_II_Evolutionary_study_using_a_nonhydrostatic_axisymmetric_numerical_modelAbstract In Part I of this study an analytical model for a steady-state tropical cyclone is constructed on the assumption that boundary-layer air parcels are conditionally neutral to displacements along the angular momentum surfaces of the hurricane vortex. The reversible thermodynamics implied by this assumption allows the mature storm to be thought of as a simple Carnot engine, acquiring heat at the high-temperature ocean surface and losing heat near the low-temperature tropopause. Although the oceanic heat source is universally recognized as the sine qua non for the mature hurricane, there is also wide acceptance of conditional instability of the second kind (CISK) (which makes no specific reference to surface heat fluxes) as the formative mechanism. This ambivalence is seen in that all numerical-simulation studies find it essential to have transfer from the ocean surface yet all start from a conditionally unstable atmosphere. The hypothesis put forward in Part I, based on the steady-state theory, is that the truly important thermodynamic interaction, even in the developing star , is between vortex and ocean (as distinct from vortex and convection sustained by preexisting conditional instability as in the CISK theory) with cumulus convection rapidly redistributing heat acquired at the oceanic source upward and outward to the upper tropospheric sink. On this view, it is not the organization of convection that is needed per se, but the organization of surface heat flux. We have constructed a time-dependent nonhydrostatic axisymmetric numerical model with convection explicitly accounted for to examine this idea. The numerical experiments show that as a result of a finite-amplitude air-sea interaction instability a hurricane-like vortex may indeed amplify in an atmosphere which is neutral to cumulus convection and attain an intensity and structure which are in excellent agreement with the theoretical predictions of Part I. We examine in detail the model's heat budget which confirms the crucial importance of boundary-layer processes in controlling the structure and evolution of the vortex. We also confirm the conjecture made in Part I that, within a large-scale limit, the horizontal size of the mature tropical cyclone is determined by that of the initial disturbance.
    Schwartz C. S., Z. Q. Liu, Y. S. Chen, and X. Y. Huang, 2012: Impact of assimilating microwave radiances with a limited-area ensemble data assimilation system on forecasts of Typhoon Morakot. Wea.Forecasting, 27, 424- 437.10.1175/WAF-D-11-00033.1108c76c8-3f33-4014-a9d3-d73cb1792d58e1a18122a85e5ea7fdeb14b3226635b7http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F258724584_Impact_of_Assimilating_Microwave_Radiances_with_a_Limited-Area_Ensemble_Data_Assimilation_System_on_Forecasts_of_Typhoon_Morakotrefpaperuri:(dad552e05a0dd72a5b1b8e2983c56ded)http://www.researchgate.net/publication/258724584_Impact_of_Assimilating_Microwave_Radiances_with_a_Limited-Area_Ensemble_Data_Assimilation_System_on_Forecasts_of_Typhoon_MorakotTwo parallel experiments were designed to evaluate whether assimilating microwave radiances with a cyclic, limited-area ensemble adjustment Kalman filter (EAKF) could improve track, intensity, and precipitation forecasts of Typhoon Morakot (2009). The experiments were configured identically, except that one assimilated microwave radiances and the other did not. Both experiments produced EAKF analyses every 6 h between 1800 UTC 3 August and 1200 UTC 9 August 2009, and the mean analyses initialized 72-h Weather Research and Forecasting model forecasts. Examination of individual forecasts and average error statistics revealed that assimilating microwave radiances ultimately resulted in better intensity forecasts compared to when radiances were withheld. However, radiance assimilation did not substantially impact track forecasts, and the impact on precipitation forecasts was mixed. Overall, net positive results suggest that assimilating microwave radiances with a limited-area EAKF system is beneficial for tropical cyclone prediction, but additional studies are needed.
    Skamarock W.C., Coauthors, 2008: Description of the advanced research WRF version 3, Rep. NCAR/TN-475++STR, Natl. Cent. for Atmos. Res., Boulder, Colo., 113 pp.
    Smith R. K., M. T. Montgomery, and N. Van Sang, 2009: Tropical cyclone spin-up revisited. Quart. J. Roy. Meteor. Soc., 135, 1321- 1335.10.1002/qj.428d8c8c471-d68b-4814-a9f6-50629954f4604af3222696946267f233c55197ee2644http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.428%2Fpdfrefpaperuri:(c63b2f02173a285d0f4388d307750245)http://onlinelibrary.wiley.com/doi/10.1002/qj.428/pdfWe present numerical experiments to investigate axisymmetric interpretations of tropical cyclone spin-up in a three-dimensional model. Two mechanisms are identified for the spin-up of the mean tangential circulation. The first involves the convergence of absolute angular momentum above the boundary layer and is a mechanism to spin up the outer circulation, i.e. to increase the vortex size. The second involves the convergence of absolute angular momentum within the boundary layer and is a mechanism to spin up the inner core. It is associated with the development of supergradient wind speeds in the boundary layer. The existence of these two mechanisms provides a plausible physical explanation for certain long-standing observations of typhoons by Weatherford and Gray, which indicate that inner-core changes in the azimuthal-mean tangential wind speed often occur independently from those in the outer core. The unbalanced dynamics in the inner-core region are important in determining the maximum radial and tangential flow speeds that can be attained, and therefore important in determining the azimuthal-mean intensity of the vortex. We illustrate the importance of unbalanced flow in the boundary layer with a simple thought experiment. The analyses and interpretations presented are novel and support a recent hypothesis of the boundary layer in the inner-core region. Copyright 2009 Royal Meteorological Society
    Smith S. D., 1980: Wind stress and heat flux over the ocean in gale force winds. J. Phys. Oceanogr., 10, 709- 726.10.1175/1520-0485(1980)0102.0.CO;2efe69891-86b0-4f7c-b320-1687d39f7662bafbb26d39d4cec11b371c48279d2260http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F253496629_Wind_Stress_and_Heat_Flux_over_the_Ocean_in_Gale_Force_Windsrefpaperuri:(5b03b7ee69347df35f9970d87b7a3e40)http://www.researchgate.net/publication/253496629_Wind_Stress_and_Heat_Flux_over_the_Ocean_in_Gale_Force_WindsAbstract An offshore stable platform has been instrumented with wind turbulence, temperature and wave height sensors. Data from this platform have been analyzed by the eddy correlation method to obtain wind stress and heat flux at wind speeds from 6 to 22 m s 1 in a deep-water wave regime, significantly extending the range of available measurements. The sea surface drag coefficient increases gradually with increasing wind speed. Sensible heat fluxes have been observed over a much wider range than previously available. Heat flux coefficients are higher in unstable than stable conditions, but are not seen to increase with increasing wind speed.
    Stern D. P., D. S. Nolan, 2012: On the height of the warm core in tropical cyclones. J. Atmos. Sci., 69, 1657- 1680.10.1175/JAS-D-11-010.112848616-b05d-414d-9441-0cfefcec045cae7322f25d3fc14ff7f0e8af8db9fd35http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F258659553_On_the_Height_of_the_Warm_Core_in_Tropical_Cyclonesrefpaperuri:(2d62463265c6688e00f342ba0deb0e79)http://www.researchgate.net/publication/258659553_On_the_Height_of_the_Warm_Core_in_Tropical_CyclonesAbstract The warm-core structure of tropical cyclones is examined in idealized simulations using the Weather Research and Forecasting (WRF) Model. The maximum perturbation temperature in a control simulation occurs in the midtroposphere (5–6 km), in contrast to the upper-tropospheric (>10 km) warm core that is widely believed to be typical. This conventional view is reassessed and found to be largely based on three case studies, and it is argued that the “typical” warm-core structure is actually not well known. In the control simulation, the height of the warm core is nearly constant over a wide range of intensities. From additional simulations in which either the size of the initial vortex or the microphysics parameterization is varied, it is shown that the warm core is generally found at 4–8 km. A secondary maximum often develops near 13–14 km but is almost always weaker than the primary warm core. It is demonstrated that microwave remote sensing instruments are of insufficient resolution to detect this midlevel warm core, and the conclusions of some studies that have utilized these instruments may not be reliable. Using simple arguments based on thermal wind balance, it is shown that the height of the warm core is not necessarily related to either the height where the vertical shear of the tangential winds is maximized or the height where the radial temperature gradient is maximized. In particular, changes in the height of the warm core need not imply changes in either the intensity of the storm or in the manner in which the winds in the eyewall decay with height.
    Stull R. B., 1988: An Introduction to Boundary Layer Meteorology. Kluwer Academic,666 pp.10.1007/978-94-009-3027-8ac92f54cac6d8f5af55b7961e5a03734http%3A%2F%2Flink.springer.com%2F10.1007%2F978-94-009-3027-8http://link.springer.com/10.1007/978-94-009-3027-8Part of the excitement in boundary-layer meteorology is the challenge associated with turbulent flow - one of the unsolved problems in classical physics. An additional attraction of the filed is the rich diversity of topics and research methods that are collected under the umbrella-term of boundary-layer meteorology. The flavor of the challenges and the excitement associated with the study of the atmospheric boundary layer are captured in this textbook. Fundamental concepts and mathematics are presented prior to their use, physical interpretations of the terms in equations are given, sample data are shown, examples are solved, and exercises are included.The work should also be considered as a major reference and as a review of the literature, since it includes tables of parameterizatlons, procedures, filed experiments, useful constants, and graphs of various phenomena under a variety of conditions. It is assumed that the work will be used at the beginning graduate level for students with an undergraduate background in meteorology, but the author envisions, and has catered for, a heterogeneity in the background and experience of his readers.
    Wang C.-C., H.-C. Kuo, Y. H. Chen, H.-L. Huang, C.-H. Chung, and K. Tsuboki, 2012: Effects of asymmetric latent heating on typhoon movement crossing Taiwan: The case of Morakot (2009) with extreme rainfall. J. Atmos. Sci., 69, 3172- 3196.10.1175/JAS-D-11-0346.1e47d7693-e512-48b1-bba6-f530639b8cd1e5d2c37e7cf5bf1aa3df0ef2923b59dchttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F258659562_Effects_of_Asymmetric_Latent_Heating_on_Typhoon_Movement_Crossing_Taiwan_The_Case_of_Morakot_%282009%29_with_Extreme_Rainfallrefpaperuri:(712e465dd86acb386eaa1256c0abf7d8)http://www.researchgate.net/publication/258659562_Effects_of_Asymmetric_Latent_Heating_on_Typhoon_Movement_Crossing_Taiwan_The_Case_of_Morakot_(2009)_with_Extreme_RainfallAbstract Typhoon Morakot struck Taiwan during 6–9 August 2009, and it produced the highest rainfall (approaching 3000 mm) and caused the worst damage in the past 50 yr. Typhoon–monsoon flow interactions with mesoscale convection, the water vapor supply by the monsoon flow, and the slow moving speed of the storm are the main reasons for the record-breaking precipitation. Analysis of the typhoon track reveals that the steering flow, although indeed slow, still exceeded the typhoon moving speed by approximately 5 km h 611 (1 km h 611 = 0.28 m s 611 ) during the postlandfall period on 8 August, when the rainfall was the heaviest. The Cloud-Resolving Storm Simulator (CReSS) is used to study the dynamics of the slow storm motion toward the north-northwest upon leaving Taiwan. The control simulations with 3-km grid size compare favorably with the observations, including the track, slow speed, asymmetric precipitation pattern, mesoscale convection, and rainfall distribution over Taiwan. Sensitivity tests with reduced moisture content reveal that not only did the model rainfall decrease but also the typhoon translation speed increased. Specifically, the simulations consistently show a discernible impact on storm motion by as much as 50%, as the storms with full moisture move slower (~5 km h 611 ), while those with limited moisture (≤25%) move faster (~10 km h 611 ). Thus, in addition to a weak steering flow, the prolonged asymmetric precipitation in Typhoon Morakot also contributed to its very slow motion upon leaving Taiwan, and both lengthened the heavy-rainfall period and increased the total rainfall amount. The implications of a realistic representation of cloud microphysics from the standpoint of tropical cyclone track forecasts are also briefly discussed.
    Wang C.-C., H.-C. Kuo, T.-C. Yeh, C.-H. Chung, Y. H. Chen, S.-Y. Huang, Y. W. Wang, and C.-H. Liu, 2013: High-resolution quantitative precipitation forecasts and simulations by the Cloud-Resolving Storm Simulator (CReSS) for Typhoon Morakot (2009). J. Hydrol., 506, 26- 41.10.1016/j.jhydrol.2013.02.0188c8816ba-b3bf-4fda-a1fc-223ba182f0cd7d58d39655d9d86897d76fa8c355f5achttp%3A%2F%2Fwww.sciencedirect.com%2Fscience%2Farticle%2Fpii%2FS0022169413001303refpaperuri:(b6d4425fdbe9c3e1dbc453ffaf84e981)http://www.sciencedirect.com/science/article/pii/S0022169413001303Typhoon Morakot (2009) struck Taiwan during 7–9 August 2009, and brought extreme rainfall up to 285502mm and the worst damages in the past 5002years. The operational models showed deficiency and serious under-prediction in rainfall amount at real time. This study demonstrates that the Cloud-Resolving Storm Simulator (CReSS), a state-of-the-art, high-resolution model, at a grid size of 302km and starting as early as 0000 UTC 4 August, can successfully simulate and reproduce the event with high accuracy, including the distribution and timing of heavy rainfall in Taiwan. In the simulation starting at 0000 UTC 6 August, for example, the threat scores for 24-h rainfall for 8 August (with extreme amounts >145002mm) reach 0.8–0.4 even at thresholds of 100–50002mm. This result is only possible due to small track error and the phase-locking mechanism of the Taiwan topography to heavy rainfall.Furthermore, real-time forecast and hindcast integrations of the CReSS model show that high-quality quantitative precipitation forecasts (QPFs) with peak total amount 67–80% of the true value are also obtained from initial conditions at 0000 UTC 6 August, which is about 202days prior to the beginning of the heaviest rainfall in southern Taiwan. In these integrations, typhoon track errors in the global model forecasts used as boundary conditions are the major error source that prevent more ideal QPF results before and at 1200 UTC 5 August. When properly configured, it is believed that other similar cloud-resolving models can achieve comparable performance. Thus, the importance of and potential benefits from deterministic high-resolution forecasts is stressed, which may give an extended lead-time when the track error is small. With potentially longer time window for emergency action just prior to extreme rainfall events when it matters the most, such forecasts may ultimately lead to reduced losses in lives and properties.
    Xie B. G., F. Q. Zhang, 2012: Impacts of typhoon track and island topography on the heavy rainfalls in Taiwan associated with Morakot (2009). Mon. Wea. Rev., 140, 3379- 3394.10.1175/MWR-D-11-00240.10256cfb0-59d7-4004-86ff-19a7c66730d18731d68155128b7aaa5962d53591f146http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F258683550_Impacts_of_Typhoon_Track_and_Island_Topography_on_the_Heavy_Rainfalls_in_Taiwan_Associated_with_Morakot_%282009%29refpaperuri:(1fbff4dc801f3c9b6029654a7f753056)http://www.researchgate.net/publication/258683550_Impacts_of_Typhoon_Track_and_Island_Topography_on_the_Heavy_Rainfalls_in_Taiwan_Associated_with_Morakot_(2009)Cloud-resolving ensemble simulations and sensitivity experiments utilizing the Weather Research and Forecasting model (WRF) are performed to investigate the dynamics and predictability of the record-breaking rainfall and flooding event in Taiwan induced by Typhoon Morakot (2009). It is found that a good rainfall forecast foremost requires a good track forecast during Morakot's landfall. Given a good track forecast, interaction of the typhoon circulation with complex topography in southern Taiwan plays a dominant role in producing the observed heavy rainfalls. The terrain slope, strength of the horizontal winds, and mid-lower-tropospheric moisture content in the southwesterly upslope flow are the primary factors that determine the rainfall location and intensity, as elucidated by the idealized one-dimensional precipitation-rate forecast model. The typhoon circulation and the southwesterly monsoon flow transport abundant moisture into southern Taiwan, which produces the heavy rainfall through interactions with the complex high terrain in the area. In the meantime, as part of the south China monsoon, the southwesterly flow may be substantially enhanced by the typhoon circulation.
    Zhang D. L., Y. B. Liu, and M. K. Yau, 2001: A multiscale numerical study of Hurricane Andrew (1992). Part IV: Unbalanced flows. Mon. Wea. Rev., 129, 92- 107.10.1175/1520-0493(2001)1292.0.CO;22397b591-f615-4fe4-b1ea-8744b1c16f842afeaaf00825181f54f395b8d783044ehttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F260906221_A_Multiscale_Numerical_Study_of_Hurricane_Andrew_%281992%29._Part_IV_Unbalanced_Flowsrefpaperuri:(46700c8852a9433b026cea44b3c26785)http://www.researchgate.net/publication/260906221_A_Multiscale_Numerical_Study_of_Hurricane_Andrew_(1992)._Part_IV_Unbalanced_FlowsAbstract Despite considerable progress in understanding the hurricane vortex using balanced models, the validity of gradient wind balance in the eyewall remains controversial in observational studies. In this paper, the structure and development of unbalanced forces and flows in hurricanes are examined, through the analyses of the radial momentum and absolute angular momentum (AAM) budgets, using a high-resolution (i.e., x = 6 km), fully explicit simulation of Hurricane Andrew (1992). It is found from the radial momentum budgets that supergradient flows and accelerations, even after temporal and azimuthal averaging, are well organized from the bottom of the eye center to the upper outflow layer in the eyewall. The agradient accelerations are on average twice greater than the local Coriolis force, and caused mainly by the excess of the centrifugal force over the pressure gradient force. It is shown by the AAM budgets that supergradient flows could occur not only in the inflow region as a result of the inward AAM transport, but also in the outflow region through the upward transport of AAM. The eyewall is dominated by radial outflow in which the upward transport of AAM overcompensates the spindown effect of the outflow during the deepening stage. The intense upper outflow layer is generated as a consequence of the continuous outward acceleration of airflows in the eyewall updrafts. In spite of the pronounced agradient tendencies, results presented here suggest that the azimuthally averaged tangential winds above the boundary layer satisfy the gradient wind balance within an error of 10%. The analyses of instantaneous fields show pronounced asymmetries and well-organized wavenumber-2 structures of the agradient flows and forces in the form of azimuthally propagating vortex ossby waves in the eyewall. These waves propagate cyclonically downstream with a speed half the tangential winds near the top of the boundary layer and vertically upward. Agradient flows/forces and AAM transport in the eye are also discussed.
    Zhang D. L., H. Chen, 2012: Importance of the upper-level warm core in the rapid intensification of a tropical cyclone. Geophys. Res. Lett., 39,L02806, doi: 10.1029/2011GL 050578.10.1029/2011GL050578a88e92d6-924f-4194-99dc-640cfd3b692389354b7a46f7e2a84e1a79d871543b3ehttp://onlinelibrary.wiley.com/doi/10.1029/2011GL050578/fullhttp://onlinelibrary.wiley.com/doi/10.1029/2011GL050578/fullABSTRACT In this study, the rapid intensification (RI) of tropical cyclone is examined using a 72-h cloud-permitting prediction of Hurricane Wilma (2005) with a record-breaking intensity of 882 hPa. Results show the formation of an upper-level warm core from the descending air of stratospheric origin in the eye, which coincides with the onset of RI; it reaches the peak amplitude of more than 18 C from its initial conditions at the time of peak intensity. The descending air is associated with the detrainment of convective bursts in the eyewall, and it appears as (perturbation) cyclonic radial inflows above the upper outflow layer and causes the subsidence warming below. We hypothesize that the upper divergent outflow layer favors the generation of a warm core by protecting it from ventilation by environmental flows. Use of the hydrostatic equation shows that the warm core of stratospheric origin contributes more than twice as much as the lower-level warm column to the pressure change at the peak intensity of Wilma. Results suggest that more attention be paid to the magnitude of storm-relative flows and vertical wind shear in the upper troposphere, rather than just vertical shear in the typical 850-200 hPa layer, in order to reasonably predict the RI of tropical cyclones.
    Zhang F. Q., Y. H. Weng, Y.-H. Kuo, J. S. Whitaker, and B. G. Xie, 2010: Predicting Typhoon Morakot's catastrophic rainfall with a convection-permitting mesoscale ensemble system. Wea.Forecasting, 25, 1816- 1825.10.1175/2010WAF2222414.18a37c8fb-e0c6-4974-9e5b-a99bd1e8ab2f640e11a61dcb1f4573b809e5a92d0237http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F241301420_Predicting_Typhoon_Morakot%27s_Catastrophic_Rainfall_with_a_Convection-Permitting_Mesoscale_Ensemble_Systemrefpaperuri:(472ef020d7f056e61ee7c17b56252c1b)http://www.researchgate.net/publication/241301420_Predicting_Typhoon_Morakot's_Catastrophic_Rainfall_with_a_Convection-Permitting_Mesoscale_Ensemble_SystemThis study examines the prediction and predictability of the recent catastrophic rainfall and flooding event over Taiwan induced by Typhoon Morakot (2009) with a state-of-the-art numerical weather prediction model. A high-resolution convection-permitting mesoscale ensemble, initialized with analysis and flow-dependent perturbations obtained from a real-time global ensemble data assimilation system, is found to be able to predict this record-breaking rainfall event, producing probability forecasts potentially valuable to the emergency management decision makers and the general public. Since all the advanced modeling and data assimilation techniques used here are readily available for real-time operational implementation provided sufficient computing resources are made available, this study demonstrates the potential and need of using ensemble-based analysis and forecasting, along with enhanced computing, in predicting extreme weather events like Typhoon Morakot at operational centers.
    Zhang J. A., P. G. Black, J. R. French, and W. M. Drennan, 2008: First direct measurements of enthalpy flux in the hurricane boundary layer: the CBLAST results. Geophys. Res. Lett., 35(14), L14813, doi:10.1029/2008GL034374.10.1029/2008GL0343743b27a3ab-a15f-4b2e-9d6f-c75f443e575f20a15fc84983c853c0e2dfd3fc50cd56http://onlinelibrary.wiley.com/doi/10.1029/2008GL034374/citedbyhttp://onlinelibrary.wiley.com/doi/10.1029/2008GL034374/citedbyABSTRACT Hurricanes extract energy from the warm ocean through enthalpy fluxes. As part of the Coupled Boundary Layer Air-Sea Transfer (CBLAST) experiment, flights were conducted to measure turbulent fluxes in the high-wind boundary layer of hurricanes. Here we present the first field observations of sensible heat and enthalpy flux for 10m wind speeds to 30 ms-1. The analyses indicate no statistically significant dependence of these bulk exchange coefficients on wind speed. As a measure of hurricane development potential, we compute the mean ratio of the exchange coefficient for enthalpy to that for momentum and find it to be significantly below the lowest threshold estimated by previous investigators. This suggests that the enthalpy flux required for hurricane development may come from sources other than turbulent fluxes, such as lateral fluxes from the vortex warm core, or sea spray. Alternatively, it demands a re-evaluation of the theoretical models used to derive the threshold.
    Zhang J. A., R. F. Rogers, D. S. Nolan, and F. D. Marks, 2011: On the characteristic height scales of the hurricane boundary layer. Mon. Wea. Rev., 139, 2523- 2535.10.1175/MWR-D-10-05017.1b136b89b-307b-4d27-ab57-46ac71873cf4db0385e39e35a45d026e3b856b13cd60http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F259852999_On_the_Characteristic_Height_Scales_of_the_Hurricane_Boundary_Layerrefpaperuri:(5bd9ab6649365d4db9056cfe08b2de9e)http://www.researchgate.net/publication/259852999_On_the_Characteristic_Height_Scales_of_the_Hurricane_Boundary_LayerAbstract In this study, data from 794 GPS dropsondes deployed by research aircraft in 13 hurricanes are analyzed to study the characteristic height scales of the hurricane boundary layer. The height scales are defined in a variety of ways: the height of the maximum total wind speed, the inflow layer depth, and the mixed layer depth. The height of the maximum wind speed and the inflow layer depth are referred to as the dynamical boundary layer heights, while the mixed layer depth is referred to as the thermodynamical boundary layer height. The data analyses show that there is a clear separation of the thermodynamical and dynamical boundary layer heights. Consistent with previous studies on the boundary layer structure in individual storms, the dynamical boundary layer height is found to decrease with decreasing radius to the storm center. The thermodynamic boundary layer height, which is much shallower than the dynamical boundary layer height, is also found to decrease with decreasing radius to the storm center. The results also suggest that using the traditional critical Richardson number method to determine the boundary layer height may not accurately reproduce the height scale of the hurricane boundary layer. These different height scales reveal the complexity of the hurricane boundary layer structure that should be captured in hurricane model simulations.
    Zhang J. A., S. Gopalakrishnan, F. D. Marks, R. F. Rogers, and V. Tallapragada, 2012: A developmental framework for improving hurricane model physical parameterizations using aircraft observations. Trop. Cycl. Res. Rev.,1(5), 419-429, doi: 10.6057/2012TCRR04.01.10.6057/2012tcrr04.0111830ff2-6911-488c-bd6c-c874d8517369f4022e6f3b5ac08238abb9d4a9929a64http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F255729562_A_Developmental_Framework_for_Improving_Hurricane_Model_Physical_Parameterizations_Using_Aircraft_Observationsrefpaperuri:(2d2ffcb7fc3e32a9ee61e42e0fa96c42)http://www.researchgate.net/publication/255729562_A_Developmental_Framework_for_Improving_Hurricane_Model_Physical_Parameterizations_Using_Aircraft_ObservationsABSTRACT As part of NOAA650200s Hurricane Forecast Improvement Program (HFIP), this paper addresses the important role of aircraft observations in hurricane model physics validation and improvement. A model developmental framework for improving the physical parameterizations using quality-controlled and post- processed aircraft observations is presented, with steps that include model diagnostics, physics development, physics implementation and further evaluation. Model deficiencies are first identified through model diagnostics by comparing the simulated axisymmetric multi-scale structures to observational composites. New physical parameterizations are developed in parallel based on in-situ observational data from specially designed hurricane field programs. The new physics package is then implemented in the model, which is followed by further evaluation. The developmental framework presented here is found to be successful in improving the surface layer and boundary layer parameterization schemes in the operational Hurricane Weather Research and Forecast (HWRF) model. Observations for improving physics packages other than boundary layer scheme are also discussed.
  • [1] Hongxiong XU, Dajun ZHAO, 2021: Effect of the Vertical Diffusion of Moisture in the Planetary Boundary Layer on an Idealized Tropical Cyclone, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 1889-1904.  doi: 10.1007/s00376-021-1016-z
    [2] Hongxiong Xu, Dajun Zhao, 2023: Effect of the vertical diffusion of moisture in the planetary boundary layer on an idealized tropical cyclone, ADVANCES IN ATMOSPHERIC SCIENCES.  doi: 
    [3] Yali LUO, Jiahua ZHANG, Miao YU, Xudong LIANG, Rudi XIA, Yanyu GAO, Xiaoyu GAO, Jinfang YIN, 2023: On the Influences of Urbanization on the Extreme Rainfall over Zhengzhou on 20 July 2021: A Convection-Permitting Ensemble Modeling Study, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 393-409.  doi: 10.1007/s00376-022-2048-8
    [4] Meiying DONG, Chunxiao JI, Feng CHEN, Yuqing WANG, 2019: Numerical Study of Boundary Layer Structure and Rainfall after Landfall of Typhoon Fitow (2013): Sensitivity to Planetary Boundary Layer Parameterization, ADVANCES IN ATMOSPHERIC SCIENCES, 36, 431-450.  doi: 10.1007/s00376-018-7281-9
    [5] Lei LIU, Guihua WANG, Ze ZHANG, Huizan WANG, 2022: Effects of Drag Coefficients on Surface Heat Flux during Typhoon Kalmaegi (2014), ADVANCES IN ATMOSPHERIC SCIENCES, 39, 1501-1518.  doi: 10.1007/s00376-022-1285-1
    [6] Zhao Ming, 1987: ON THE PARAMETERIZATION OF THE VERTICAL VELOCITY AT THE TOP OF PLANETARY BOUNDARY LAYER, ADVANCES IN ATMOSPHERIC SCIENCES, 4, 233-239.  doi: 10.1007/BF02677070
    [7] MA Yaoming, Massimo MENENTI, Reinder FEDDES, 2010: Parameterization of Heat Fluxes at Heterogeneous Surfaces by Integrating Satellite Measurements with Surface Layer and Atmospheric Boundary Layer Observations, ADVANCES IN ATMOSPHERIC SCIENCES, 27, 328-336.  doi: 10.1007/s00376-009-9024-4
    [8] Xu Liren, Zhao Ming, 2000: The Influences of Boundary Layer Parameterization Schemes on Mesoscale Heavy Rain System, ADVANCES IN ATMOSPHERIC SCIENCES, 17, 458-472.  doi: 10.1007/s00376-000-0036-3
    [9] Fei Jianfang, Zhang Ming, Lu Hancheng, Yang Guoxiang, 1992: A Fine-Mesh Numerical Model with Detailed Boundary Layer Parameterization, ADVANCES IN ATMOSPHERIC SCIENCES, 9, 465-476.  doi: 10.1007/BF02677079
    [10] Jianjun LIU, Feimin ZHANG, Zhaoxia PU, 2017: Numerical Simulation of the Rapid Intensification of Hurricane Katrina (2005): Sensitivity to Boundary Layer Parameterization Schemes, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 482-496.  doi: 10.1007/s00376-016-6209-5
    [11] LIU Ximing, CHENG Xueling, WU Qiong, FU Minning, ZENG Qingcun, 2013: Some Characteristics of the Surface Boundary Layer of a Strong Cold Air Process over Southern China, ADVANCES IN ATMOSPHERIC SCIENCES, 30, 210-218.  doi: 10.1007/s00376-012-1223-8
    [12] Zhong Shiyuan, Zhou Mingyu, Li Xingsheng, 1987: A NUMERICAL STUDY ON THE MESO-SCALE POLLUTANT DISPERSION OVER A SLOPED SURFACE IN THE STABLE BOUNDARY LAYER, ADVANCES IN ATMOSPHERIC SCIENCES, 4, 300-312.  doi: 10.1007/BF02663600
    [13] Rui WANG, Yiting ZHU, Fengxue QIAO, Xin-Zhong LIANG, Han ZHANG, Yang DING, 2021: High-resolution Simulation of an Extreme Heavy Rainfall Event in Shanghai Using the Weather Research and Forecasting Model: Sensitivity to Planetary Boundary Layer Parameterization, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 98-115.  doi: 10.1007/s00376-020-9255-y
    [14] LIU Shuhua, YUE Xu, LIU Huizhi, HU Fei, 2004: Using a Modified Soil-Plant-Atmosphere Scheme (MSPAS) to Study the Sensitivity of Land Surface and Boundary Layer Processes to Soil and Vegetation Conditions, ADVANCES IN ATMOSPHERIC SCIENCES, 21, 717-729.  doi: 10.1007/BF02916369
    [15] Xuwei BAO, Leiming MA, Jianyong LIU, Jie TANG, Jing XU, 2018: Formation and Development of a Mountain-induced Secondary Center inside Typhoon Morakot (2009), ADVANCES IN ATMOSPHERIC SCIENCES, 35, 1160-1176.  doi: 10.1007/s00376-018-7199-2
    [16] Li Xingsheng, Yang Shuowen, 1986: A MODEL STUDY OF THE NOCTURNAL BOUNDARY LAYER, ADVANCES IN ATMOSPHERIC SCIENCES, 3, 59-71.  doi: 10.1007/BF02680045
    [17] Lin Naishi, Zhou Zugang, Zhou Liufei, 1998: An Analytical Study on the Urban Boundary Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 15, 258-266.  doi: 10.1007/s00376-998-0044-2
    [18] Zhao Ming, Xu Yinzi, Wu Rongsheng, 1989: The Wind Structure in Planetary Boundary Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 6, 365-376.  doi: 10.1007/BF02661542
    [19] Myoung-Hwan AHN, Eun-Ha SOHN, Byong-Jun HWANG, Chu-Yong CHUNG, Xiangqian WU, 2006: Derivation of Regression Coefficients for Sea Surface Temperature Retrieval over East Asia, ADVANCES IN ATMOSPHERIC SCIENCES, 23, 474-486.  doi: 10.1007/s00376-006-0474-7
    [20] Jian YUE, Zhiyong MENG, Cheng-Ku YU, Lin-Wen CHENG, 2017: Impact of Coastal Radar Observability on the Forecast of the Track and Rainfall of Typhoon Morakot (2009) Using WRF-based Ensemble Kalman Filter Data Assimilation, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 66-78.  doi: 10.1007/s00376-016-6028-8

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 02 April 2015
Manuscript revised: 02 June 2015
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

Effects of Surface Flux Parameterization on the Numerically Simulated Intensity and Structure of Typhoon Morakot (2009)

  • 1. Key Laboratory of Mesoscale Severe Weather/MOE and School of Atmospheric Sciences, Nanjing University, Nanjing 210023
  • 2. Hurricane Research Division, Atlantic Oceanographic and Meteorological Laboratory, National Oceanographic and Atmospheric Administration, Miami, FL, USA
  • 3. Rosenstiel School of Marine and Atmospheric Science, University of Miami, Miami, FL, USA

Abstract: The effects of surface flux parameterizations on tropical cyclone (TC) intensity and structure are investigated using the Advanced Research Weather Research and Forecasting (WRF-ARW) modeling system with high-resolution simulations of Typhoon Morakot (2009). Numerical experiments are designed to simulate Typhoon Morakot (2009) with different formulations of surface exchange coefficients for enthalpy (C K) and momentum (C D) transfers, including those from recent observational studies based on in situ aircraft data collected in Atlantic hurricanes. The results show that the simulated intensity and structure are sensitive to C K and C D, but the simulated track is not. Consistent with previous studies, the simulated storm intensity is found to be more sensitive to the ratio of C K/C D than to C K or C D alone. The pressure-wind relationship is also found to be influenced by the exchange coefficients, consistent with recent numerical studies. This paper emphasizes the importance of C D and C K on TC structure simulations. The results suggest that C D and C K have a large impact on surface wind and flux distributions, boundary layer heights, the warm core, and precipitation. Compared to available observations, the experiment with observed C D and C K generally simulated better intensity and structure than the other experiments, especially over the ocean. The reasons for the structural differences among the experiments with different C D and C K setups are discussed in the context of TC dynamics and thermodynamics.

1. Introduction
  • Tropical cyclones (TCs) obtain heat and moisture from the ocean and transfer momentum back to the ocean at the air-sea interface. Thus, surface fluxes of sensible and latent heat play a very important role in the development and maintenance of TCs. (Malkus and Riehl, 1960) found that strong surface fluxes can increase the equivalent potential temperature in the TC eyewall, which is linked to the decrease in minimum sea level pressure at the TC center (i.e., increase in TC intensity). Early theoretical studies found that the TC intensity is sensitive to the selection of exchange coefficients for air-sea momentum and heat transfers (Ooyama, 1969; Rosenthal, 1971; Emanuel, 1986). (Rotunno and Emanuel, 1987) showed that the increase in the exchange coefficient for enthalpy transfer (C K) and the decrease in the drag coefficient (C D) caused the increase in TC intensity in idealized numerical models. Based on comparisons of model predictions with observations for a number of hurricanes, (Emanuel, 1995) concluded that the TC intensity is most sensitive to the ratio of C K over C D, and this ratio must lie in the range 0.75-1.5 to achieve consistent model simulations with observations. Full-physical numerical model simulations of TCs also suggest that the TC intensity is sensitive to C D and C K (e.g., Braun and Tao, 2000; Nolan et al., 2009).

    Early observational studies (e.g., Smith, 1980; Large and Pond, 1981; Geernaert et al., 1988) investigated C D in weak to moderate wind conditions with 10 m wind speed (U10)<25 m s-1 in the open ocean. They found that C D increases almost linearly with U10. The relationship between C D and U10 has the form 1000C D=0.063U10+0.61, where both C D and U10 have been adjusted to neutral stability (Smith, 1980). However, results from the Coupled Boundary Layer and Air-Sea Transfer (CBLAST) experiment indicate that C D does not increase with U10 with no limitation (Black et al., 2007; French et al., 2007). (Donelan et al., 2004) estimated C D in a wave tank and found that C D reaches a maximum for U10 at 33 m s-1 and then levels off at higher wind speeds. They attributed this leveling-off of C D to the increasing sheltering of winds behind steeper waves at higher wind speeds. The above-mentioned studies support the result of (Powell et al., 2003), who estimated C D using the profile method by fitting hundreds of GPS dropsonde wind profiles collected in hurricanes to a logarithmic relationship with height. (Ming et al., 2012) showed that C D estimated based on observational data in typhoons also behaves similarly as in the (Powell et al., 2003) study.

    Due to instrumentation limitations associated with sea salt and/or sea spray, direct measurements of surface enthalpy fluxes have been very difficult over the ocean. The only available in situ observations over the past 30 years (DeCosmo et al., 1996; Drennan et al., 2007) show that the exchange coefficient for latent heat transfer (C E) does not depend on the wind speed. Similarly, the bulk sensible heat exchange coefficient (C H), or Stanton number, has also been found to be independent of wind speed. Note that C K is generally assumed to be equal to C E and C H, which is supported by the CBLAST result (Zhang et al., 2008). Wave tank observations also support the result from the CBLAST experiment, that C K is nearly constant with U10 up to 40 m s-1 (Haus et al., 2010). (Bell et al., 2012) also found that values of C K estimated based on budget analyses of dropsonde and radar data in the boundary layer of the eyewall region were close to those reported by (Zhang et al., 2008).

    Most previous numerical studies on the impact of surface exchange coefficients on TC intensity and structure focused on Atlantic hurricanes or used idealized simulations (e.g., Montgomery et al., 2010; Bryan, 2012; Bao et al., 2012; Green and Zhang, 2013). How sensitive the simulated intensity and structure in West Pacific typhoons are to C D and C K is not well documented. The objective of this study is to investigate the effects of C D and C K on the simulated intensity and structure of Typhoon Morakot (2009). Emphasis is placed on simulations of structure, such as boundary layer heights, the warm core, and precipitation. Typhoon Morakot (2009) possessed unusual structures, including a large eye and strong outer circulation; thus, this storm was atypical compared to most hurricanes in the Atlantic basin. This motivates us to evaluate the choice of C D and C K in the numerical simulations to reproduce the unusual structures of Typhoon Morakot (2009).

2. Typhoon Morakot (2009)
  • Typhoon Morakot (2009) generated from a tropical depression over the Philippine Sea on 2 August 2009. It gradually intensified to a tropical storm and was assigned the name Morakot on 3 August. After that, it moved westward towards Taiwan. Morakot (2009) intensified to typhoon intensity at 1800 UTC 5 August, with the location of its center at (23.1°N and 129.3°E) and a maximum wind speed of 33.4 m s-1. Morakot (2009) continued to move westward and made landfall in central Taiwan at 1200 UTC 7 August. One day later, Morakot (2009) turned to move northwestward and went back over water into the Taiwan Strait. Subsequently, Morakot (2009) weakened to a severe tropical storm and made its second landfall in Xiapu, Fujian Province, on the east coast of China at 0820 UTC 9 August (Fig. 1a). It gradually weakened as it moved northward. Roughly 24 hours later, it weakened to a depression during the landfall in Zhejiang Province, China. Typhoon Morakot (2009) brought record-breaking torrential rainfall over Taiwan (Chien and Kuo, 2011). This catastrophic storm sadly claimed more than 600 lives and caused more than 200 people to be classed as missing, creating a total cost of damage estimated at US $ 3.3 billion (Zhang et al., 2010). It also affected more than 11 million people throughout eastern China and damaged thousands of homes. As a result, Typhoon Morakot (2009) has received attention in several numerical studies.

    Figure 1.  The (a) tracks and time series of simulated (b) maximum surface wind speed and (c) minimum central pressure (hPa) of Morakot (2009) from 0000 UTC 5 August to 0000 UTC 10 August 2009 from the three experiments and the corresponding best-track analysis by JMA.

    (Schwartz et al., 2012) studied Typhoon Morakot (2009) using the Advanced Research Weather Research and Forecasting (WRF-ARW) modeling system with data assimilation of the microwave radiances using a cyclic, limited-area ensemble adjustment Kalman filter. They found that the track, intensity, and precipitation forecasts were improved after assimilating microwave radiances. (Xie and Zhang, 2012) performed ensemble simulations using WRF-ARW to investigate the dynamics and predictability of the record-breaking rainfall and flooding event in Taiwan induced by Typhoon Morakot (2009). They found that the large typhoon circulation and the southwesterly monsoon flow transported abundant moisture into southern Taiwan, which, along with the influence of the complex high terrain, produced the heavy rainfall.

    (Wang et al., 2012) used the Cloud-Resolving Storm Simulator (CReSS) to study the dynamics related to the motion of Typhoon Morakot (2009) over Taiwan. Their simulations showed that the reduced moisture content induced the decrease in the rainfall and the increase in the storm translation speed. They also pointed out that the asymmetric precipitation in Typhoon Morakot (2009) played an important role in its very slow motion upon leaving Taiwan, the lengthening of the heavy-rainfall period, and the increase of the total rainfall amount. Furthermore, their results emphasized the potential contribution of asymmetric heating to the slowdown of the typhoon motion in the presence of complex terrain or in a monsoon environment. (Wang et al., 2013) demonstrated that CReSS successfully simulated and reproduced both the distribution and timing of the heavy rainfall in Taiwan with high accuracy when Typhoon Morakot (2009) passed by. The real-time forecast integrations of the CReSS model also showed high-quality quantitative precipitation forecasts.

    (Hall et al., 2013) used the Advanced Regional Prediction System to simulate Typhoon Morakot (2009) when it made landfall over Taiwan. They investigated the mesoscale structure of Morakot (2009) and emphasized the role of deep convection on the rainfall simulation. They identified that relatively large-amplitude wave structures developed in the outer eyewall, known as vortex Rossby waves (VRWs). They found that the strong asymmetry of the convection was associated with wavenumber1 (WN1) VRWs, while the WN2 and WN3 VRWs were associated with the development of the deep convective band in Morakot's (2009) southwestern quadrant. (Huang et al., 2014) used the WRF model to explicitly simulate Typhoon Morakot (2009) and found that the simulated rain rate and precipitation efficiency (PE) over the Central Mountain Range (CMR) were highly correlated. They found also that the PE and the processes of vapor condensation and raindrop evaporation were strongly influenced by orographic lift. They also found that the increase in PE over the CMR compared with that over the ocean was due to an increase in the ice-phase deposition ratio when the liquid-phase condensation reduced as the air on the lee side subsided and moved downstream. They emphasized the effects of the terrain on the simulations of precipitation.

3. Experimental design
  • As mentioned above, the objective of this paper is to investigate the sensitivity of the simulated track, intensity and structure of Typhoon Morakot (2009) to the surface exchange coefficients. We used the ARW modeling system (version 3.2; Skamarock et al., 2008) to conduct numerical experiments. Three experiments were performed using the ARW model with two model domains. The model horizontal grid resolution was 4.5 km for the parent domain (D1) and 1.5 km for the nested domain (D2). The two domains covered 600× 400 and 421× 421 grid points, respectively (Fig. 1a). D2 was an automatic vortex-following moving nest and the center of the domain was always located at the center of the storm. In all the experiments, 35 vertical (σ) levels were used from the surface to the model top at 50 hPa.

    D1 was initialized at 0000 UTC 4 August 2009 and was run for 6 days. D2 was started at 0000 UTC 5 August after a one-day spin-up. The outer domain was run in a four-dimensional data assimilation (FDDA) mode to provide the best possible large-scale conditions for the inner domain. The inner nested domain was run without FDDA. The Japan Meteorological Agency (JMA) 6-hourly gridded regional analyses at 20× 20 km horizontal resolution with 20 pressure levels were used as the initial and boundary conditions for the ARW model. The JMA analyses were produced using a multivariate three-dimensional optimum interpolation method to combine the first-guess fields from JMA's regional spectral model (RSM) with observations from a variety of platforms (JMA, 2007; Hosomi, 2005). Note that the JMA analyses did not capture the real structure and intensity of Morakot (2009). Compared to the best track, the error of the minimum sea-level pressure was 13.7 hPa, and that of the maximum surface wind speed was 10 m s-1. The track of the storm in the JMA analyses also has a northwestward bias. Following (Ming et al., 2009), we appended a good vortex with the appropriate intensity (close to the best track) into the initial conditions. We ran D01 only from 0000 UTC to 1600 UTC 4 August, to make the storm intensity similar to the best track. The three-dimensional vortex was extracted and inserted back into the initial condition using the position of the best track, which corrected the bias of the storm center at the initial time. Under the new initial condition, the error of the minimum sea level pressure was 0.6 hPa and that of the maximum surface wind speed was 5.6 m s-1.

    The physics schemes used in the numerical experiments included the Purdue Lin microphysics scheme (Lin et al., 1983; Chen and Sun, 2002), the Yonsei University boundary layer scheme (Noh et al., 2003; Hong et al., 2006), the Rapid Radiative Transfer Model longwave radiation scheme (Mlawer et al., 1997), and the Dudhia shortwave radiation scheme (Dudhia, 1989). Cumulus parameterization schemes were not used for both domains.

    In this study, three experiments (see Table 1) were designed to examine the impact of C D and C K on the typhoon intensity and structure simulations. As mentioned earlier, surface fluxes were parameterized through C D and C K. In the surface-layer schemes of WRF-ARW, C D and C K were parameterized using surface momentum (z0) and thermal (z0 q) roughness lengths. The drag coefficient in the neutral condition was defined as:

    \begin{equation} \label{eq1} C_{ D}=\left(\dfrac{k}{\ln\frac{10}{z_0}}\right)^2 , (1)\end{equation}

    where k is the von Kármán constant. The enthalpy exchange coefficient was defined as:

    \begin{equation} \label{eq2} C_{ K}=\left(\dfrac{k}{\ln\frac{10}{z_0}}\right)\left(\dfrac{k}{\ln\frac{10}{z_{ 0q}}}\right) . (2)\end{equation}

    In the control experiment (referred to as CTL hereafter), z0 depended on the wind speed using the (Charnock, 1955) relationship, in the form of:

    \begin{equation} \label{eq3} z_0=0.0185(u_\ast^2/g)+1.59\times 10^{-5} ,(3) \end{equation}

    where u* is the friction velocity, g is the gravitational acceleration. This formula causes z0 to increase continuously with the wind speed, with no limit. For z0 q the Carlson-Boland scheme (Carlson and Boland, 1978) was used:

    \begin{equation} \label{eq4} \ln\dfrac{10}{z_{ 0q}}=\ln\left(\dfrac{10u_\ast k}{x_{ ka}}+\dfrac{10}{z_0}\right) , (4)\end{equation} where x ka is a constant that equals 2.4× 10-5. Since z0q varies more slowly with the friction velocity than z0, C K increases more slowly with the wind speed than C D (Dudhia et al., 2008). In the second experiment (referred to as TC1 hereafter), an alternate C D formulation based on the high-wind laboratory experiment of (Donelan et al., 2004) was adopted. The roughness length was defined as: \begin{equation} \label{eq5} z_0=10\exp\left(-\dfrac{10}{u_\ast^{-1/3}}\right) , (5)\end{equation}

    where 1.27× 10-7≤ z0 ≤ 2.85× 10-3 (units: m). Values of C D based on Eqs. (2) and (6) are smaller than those from the Charnok relationship. C D increases almost linearly with the 10 m wind speed, up to a maximum of 0.0024 at 35 m s-1 (Davis et al., 2008). In addition, the thermal roughness length was calculated using the ramped formula (Dudhia et al., 2008) as follows:

    \begin{equation} \ln\dfrac{10}{z_{0{ q}}}=\left\{ \begin{array}{l@{\quad}l} \ln(10^5) & u_{\ast}\le 1\\[2mm] \ln(10/[10^{-4}+10^{-3}(u_\ast-1)^2] & u_{\ast}\ge 1 \end{array} \right. . (6)\end{equation}

    This setup of z0 and z0 q causes C K to increase almost linearly with wind speed, and C K becomes larger than C D at wind speeds of >50 m s-1. Note that the ratio C K/C D exceeds 1 as the wind speed becomes larger than 50 m s-1, although the condition for the maximum surface wind speed of >50 m s-1 was never met in our simulations (Fig. 1b). In the third experiment (referred to as TC2 hereafter), the drag formulation was still based on the result of (Donelan et al., 2004), but the thermal roughness length was adopted based on the Garratt formula (Garratt, 1992, p. 102) as follows:

    \begin{eqnarray} \label{eq6} \ln\dfrac{10}{z_{0{ q}}}&=&\ln\dfrac{10}{z_0}+\ln\dfrac{z_0}{z_{0{ q}}} (7),\\ \label{eq7} \ln\dfrac{z_0}{z_{0{ q}}}&=&2.28R_{e^\ast}^{\frac{1}{4}}-2 ,(8)\\ \label{eq8} R_{e^\ast}&=&\dfrac{u_\ast z_0}{\gamma}(9) ,\\ \label{eq9} \gamma&=&[1.32+0.009(T-273.15)]10^{-5} , (10)\end{eqnarray}

    where T is air temperature at the surface, γ is the kinematic viscosity of air (units: m2 s-1). In this setup, C K is related to u* and T. C K decreases with increasing wind speed in smooth flow, while it is a constant in rough flow or high-wind conditions (Garratt, 1992, p. 102). Note that C K using the Garratt formula is closer to recent observations from CBLAST and the wave tank experiment mentioned in the introduction. All simulations in the three experiments were run from 0000 UTC 5 August to 0000 UTC 10 August 2009, which covered almost the entire lifecycle of Typhoon Morakot (2009), from the intensifying stage to the landfalling stage. All experiments were conducted with the same initial vortex and boundary conditions, so the differences in the experiments were solely related to the different formulas of the exchange coefficients used in the WRF-ARW model.

    The exchange coefficients and ratio of C K/C D as a function of the surface wind speed are shown in Fig. 2. In CTL, the Charnok relationship was used such that C D increased almost linearly with the 10 m wind speed, because of the increasing surface roughness. C K also increased slowly with the wind speed. In TC1 and TC2, when the Charnok relationship was changed to the new formulation of C D based on (Donelan et al., 2004), C D increased with the wind speed up to 33 m s-1, then levelled off. In CTL, C K was based on the Carlson-Boland formula and was larger than that based on the ramped formula used in TC1 for wind speeds of <30 m s-1. On the other hand, the C K quantities produced by these two formulas were close to one another for wind speeds of >30 m s-1. The Garratt formula for the thermal roughness length was used in TC2, where the C K was smaller than that in CTL and TC1 for wind speeds of >20 m s-1. In TC2, the ratio of C K/C D was larger than that in CTL for wind speeds of <20 m s-1, and marginally smaller for wind speeds of >20 m s-1. Note that the ratio of C K/C D in TC1 was larger than that in the other two experiments for all wind speeds.

    Figure 2.  The simulated (a) drag coefficient (C D), (b) enthalpy exchange coefficient (C K) and (c) exchange coefficient ratio (C K/C D) as a function of 10 m wind speed in the whole of domain 2 from the three experiments at 1800 UTC 6 August 2009.

4. Results
  • The simulated track of Typhoon Morakot (2009) from the three experiments from 0000 UTC 5 August to 0000 UTC 10 August are compared with the JMA best track in Fig. 1a. As mentioned earlier, Typhoon Morakot (2009) moved westward in the first two days, and turned northwestward after making landfall in Taiwan. It then continued to move northward, and made a second landfall in Fujian Province. The simulated storm tracks in the three experiments were quite similar, although they deviated from the best track. All three experiments reproduced the observed west-northwestward storm motion. Overall, the simulated tracks are not sensitive to the different formulas of exchange coefficients.

    The simulated maximum surface wind and minimum sea level pressure from the three experiments are compared with the best track from JMA in Figs. 1b and 1c. Although all three experiments were initialized with the same initial condition, the intensity forecasts were different, especially after the maximum wind speed reached 30 m s-1 at the forecast time of 0000 UTC 6 August. In the next two days, all three experiments simulated maximum surface wind speed that was larger than observed. The maximum surface wind speed simulated in CTL and TC2 were closer to the best track. The difference in the simulated minimum sea level pressure, on the other hand, was clearer than that in the maximum surface wind speed among the three experiments. The simulated minimum sea level pressure in TC2 was closest to the best track among the three experiments.

    Figure 3 plots the relationship between the maximum wind speed and minimum sea level pressure from the three experiments and the best track of JMA. It appears that the pressure-wind relationship is sensitive to C D and C K. The result is consistent with a recent numerical study carried out by (Green and Zhang, 2013), who found that the pressure-wind relationship was sensitive to the selection of different surface layer schemes in WRF simulations of Hurricane Katrina (2005). Our result is also consistent with (Bao et al., 2012), who ran different surface-layer physics in idealized simulations of the Hurricane Weather and Research Forecasting (HWRF) model and confirmed that the pressure-wind relationship was sensitive to C D and C K.

    Figure 3.  Time-radius Hovmöller plots of azimuthally averaged tangential wind (units: m s-1) at 2 km altitude from (a) CTL, (b) TC1 and (c) TC2, and radial wind (units: m s-1) at 250 m altitude from (d) CTL, (e) TC1 and (f) TC2. The thick lines depict the RMW at 2 km altitude and the maximum inflow at 0.25 km altitude.

    The simulated storm intensity and pressure-wind relationship in TC2 were closest to the best track among the three experiments, especially for wind speeds of >30 m s-1. This result is encouraging because the C D and C K used in TC2 were closer to recent observations than those in the other two experiments. Our result generally indicates that when C K is larger, the storm obtains more sensible and latent fluxes from the underlying ocean, resulting in larger surface wind speed and lower central pressure. On the other hand, when C D is larger, surface friction is larger, which tends to reduce the surface wind speed. It was pointed out by (Montgomery et al., 2010) that larger surface friction can also lead to a larger gradient wind imbalance in the boundary layer, which could lower the central pressure. This explains why the simulated minimum sea level pressure in CTL was lower than that in TC2 but the maximum wind speed in CTL was smaller than that in TC2.

    Figure 4.  Scatter plot of minimum sea level pressure vs maximum surface wind speed. The blue diamonds represent CTL, the red circles represent TC1, and the green crosses represent TC2. The open squares represent the pressure-wind relationship from JMA.

  • Time-radius Hovmöller diagrams of azimuthally averaged tangential wind velocity (V t) at the altitude of 2 km and radial wind velocity (V r) at the altitude of 250 m, from 0000 UTC 6 August to 1200 UTC 7 August, are displayed in Figs. 4a-c and 4d-f, respectively. As can be seen, the evolution of the axisymmetric V t and V r in the three experiments was similar. In the first 6 h of the simulations, Morakot (2009) was a weak storm that had a maximum V t of 25 m s-1 and minimum V r of -10 m s-1, located at the radius of 170 km. Later, V t increased gradually with time and the radius of maximum tangential wind speed (RMW) became smaller. The magnitude of V r doubled from -10 to -20 m s-1 in the next 24 h and the radius of the peak inflow also contracted (Figs. 4d-f) in response to the intensification of the storm. Inward from the RMW, the radial velocity decelerated rapidly at a larger rate than outside the RMW. After another 16 h, the storm center was close to Taiwan such that the magnitudes of both V t and V r decreased. Among the three experiments, the peak V t and V r in TC1 were the largest, mainly because the ratio of C K/C D used in TC1 was larger than that in CTL and TC2.

    The time-height evolution of the mean temperature anomalies (referred to as warm-core anomalies hereafter) from 0000 UTC 6 August to 1200 UTC 7 August is plotted in Fig. 5 for CTL, TC1 and TC2. Following (Liu et al., 1999) and (Li et al., 2013), we first calculated the mean temperature within the region of 630× 630 km from the typhoon's minimum surface pressure center at each vertical level. Then, the temperature anomaly was obtained by subtracting the mean temperature from the temperature at each grid point and each level. The warm-core anomaly was then defined as the average value of the temperature anomalies within the region of 300× 300 km from the storm center at each level. The height of the peak mean temperature anomaly represents the warm core height. It is evident from Fig. 5 that the peak warm core anomaly in CTL and TC2 was smaller than that in TC1, indicating that the warm-core anomaly was correlated with the storm intensity. According to the hydrostatic balance, the lower the minimum sea level pressure the larger the warm-core anomaly (Zhang and Chen, 2012). Thus, the warm-core anomaly in TC2 was the smallest among the three experiments as the storm intensity in TC2 was the lowest. On the other hand, the warm-core height was not correlated with the storm intensity, because the warm-core height was located at 8-12 km for the three experiments.

  • The horizontal distributions of 10 m wind speed valid at 1800 UTC 6 August 2009 are shown in Figs. 6a-c for CTL, TC1 and TC2, respectively. It is evident that the maximum surface wind speed in TC1 was larger than that in the other two experiments. The eyewall region, which covered the maximum wind speed, was broader in TC1 than in CTL and TC2. Comparing TC1 and TC2, the result suggests that increasing C K alone would increase the storm intensity in terms of the maximum wind speed in the eyewall region. The surface wind distribution in TC2 was similar to that in CTL, but the maximum wind speed in the right-rear quadrant was slightly larger in TC2 than in CTL. Although C K in TC1 was close to that in CTL, the maximum surface wind speed in TC1 was larger than that in CTL, especially for wind speeds of >25 m s-1. This difference was mainly due to the fact that the C D used in TC1 was smaller than that in CTL.

    Large values of latent fluxes were found in the eyewall and primary rain band regions where surface wind speeds were also large (Figs. 6d-f). The maximum latent heat flux in TC2 was much smaller than that in the other two experiments, mainly because the C K used in TC2 was smaller than that in CTL and TC1 (Fig. 2). The difference in the sensible heat fluxes among the three experiments (Figs. 6g-i) was much smaller than the difference in the latent heat fluxes, although the same formula was used for calculating both the latent and sensible heat fluxes. It appears that the simulated maximum sensible heat flux in TC1 was the largest among the three experiments, while that in TC2 was the smallest. The asymmetric distribution of the sensible heat flux in TC2 was closer to that in CTL than in TC1. Overall, the result (Fig. 6) implies that C K and C D alone have opposite effects on surface enthalpy flux, and C K influences the enthalpy flux more than C D. A larger C K induces more sensible and latent fluxes, which support more energy for a storm to intensify. On the other hand, a larger C D induces larger surface friction, which reduces the surface wind speed and in turn reduces the sensible and latent fluxes because these fluxes are also a function of the wind speed.

    Figure 5.  Time-height diagrams of temperature deviation (units: K) from the three experiments: (a) CTL; (b) TC1; (c) TC2. The average was computed within the area of 300× 300 km from the surface minimum pressure center for simulations. The anomalies were obtained by subtracting the averaged temperature within the region at every height level.

    Figure 6.  The model-simulated 10 m wind speed (color scale; units: m s-1) and wind vectors (arrows; units: m s-1) from (a) CTL, (b) TC1 and (c) TC2; latent heat flux (color scale; units: W m-2) and wind vectors (arrows; units: m s-1) from (d) CTL, (e) TC1 and (f) TC2; and sensible heat flux (color scale; units: W m-2) and wind vectors (arrows; units: m s-1) from (g) CTL, (h) TC1 and (i) TC2 at 1800 UTC 6 August 2009. The large vector indicates the motion of the storm, and the thin crosses divide the storm into four quadrants.

    Figure 7.  Azimuthally averaged radius-height cross sections of tangential wind (units: m s-1) from (a) CTL, (b) TC1 and (c) TC2; radial wind (units: m s-1) from (d) CTL, (e) TC1 and (f) TC2; and the gradient force imbalance (F PEC, units: m s-1 h-1) from (g) CTL, (h) TC1 and (i) TC2 at 1800 UTC 6 August 2009. The thick lines in (a-c) depict the height of the maximum wind speed varying with radius, and the thick dashed lines in (d-f) depict the inflow layer height, defined as the height where the radial wind speed is 10% of the peak inflow.

    Figure 7 shows the radius-height plots of the azimuthally averaged tangential and radial velocities for all three experiments at 1800 UTC 6 August 2009. The magnitudes of the tangential and radial velocities in TC1 were generally larger than those in the other two experiments, consistent with the simulated storm intensities. However, the peak tangential wind speed in TC2 was smaller than that in CTL, which was not consistent with the intensity difference between these two experiments in terms of the maximum surface wind, as TC2 had a larger maximum surface wind speed. This discrepancy can be explained by the role of C D in regulating the boundary layer dynamics. (Montgomery et al., 2010) pointed out that an increase in C D leads to an increase in storm intensity in terms of maximum tangential wind speed in the boundary layer, although it reduces the surface wind speed through surface friction. Our result is consistent with that of (Montgomery et al., 2010), indicating that surface flux parameterization affects the vertical structure of wind velocities above the surface layer.

    In many PBL schemes used in full-physics numerical models, one of the crucial elements is the boundary layer height, because it is coupled with the energy transport from the surface layer to the boundary layer and above (e.g., Beljaars and Viterbo, 1998; Noh et al., 2003). The boundary layer height is also a key variable that regulates the vertical distribution of turbulent fluxes and helps determine where turbulent fluxes tend to become negligible (Stull, 1988). Following (Zhang et al., 2011), the kinematic boundary layer height is defined by the height of maximum tangential wind speed (h vtm). Inflow layer depth (\(h_\inf\)), defined as the height where the inflow reduces to 10% of the peak value, also represents the kinematic boundary layer height. In all three experiments, h vtm decreased with decreasing radius toward the storm center (Figs. 7a-c). This behavior is consistent with the result given by (Zhang et al., 2011), who composited hundreds of dropsondes data collected from 13 hurricanes to study the characteristics of hurricane boundary layer heights. Within the radius of 125 km, h vtm in TC2 was smaller than that in the other two experiments, and was closer to observations (Zhang et al., 2011, Fig. 5a).

    Figure 8.  (a) U.S. Air Force Defense Meteorological Satellite Program (DMSP) polar-orbiting satellite F-17 microwave imagery of polarization corrected temperature with horizontal polarization with 91 GHz at 2112 UTC 6 August, and simulated composite radar reflectivity (units: dBZ) at 2100 UTC 6 August 2009 from (b) CTL, (c) TC1 and (d) TC2. (e) The observed composite reflectivity and simulated composite reflectivity from experiment (f) CTL, (g) TC1 and (h) TC2 at 1200 UTC 7 August 2009. (i) The observed composite reflectivity and simulated composite reflectivity from experiment (j) CTL, (k) TC1 and (l) TC2 at 0000 UTC 8 August 2009.

    The difference in the inflow layer depth (\(h_\inf\)) among the three experiments was much larger than that in h vtm (Figs. 7d-f). It appears that \(h_\inf\) in TC1 was the highest. All three experiments captured the decrease of \(h_\inf\) with deceasing radius, consistent with observations. Furthermore, h vtm was smaller than \(h_\inf\) in all three experiments, which was also consistent with observations. Overall, the magnitude of \(h_\inf\) in TC2 was closest to observations (Zhang et al., 2011, Fig. 5b). The above results indicate that boundary layer heights are tied to the surface flux parameterization.

    According to (Zhang et al., 2001), the momentum equation of the radial wind velocity in the cylindrical coordinates system can be written as

    \begin{equation} \label{eq10} \dfrac{dV_r}{dt}=-\dfrac{1}{\rho}\dfrac{\partial p}{\partial r}+\dfrac{V_t^2}{r}+fV_t+2\Omega\cos\phi w\cos\lambda+U_{ d} , (11)\end{equation}

    where w is vertical wind velocity; Ω is the angular velocity and φ is the latitude; r is the radius from the center; Λ is the azimuthal angle. Equation 11 states that the radial acceleration is determined by the radial pressure gradient force (F P; first term on the right-hand side of the equation), the centrifugal force (F E; second term), the Coriolis force (F C; third and fourth terms), and diffusion (U d; last term). The degree of gradient force imbalance or net agradient force (F PEC) is evaluated by adding F P, F E, and F C together (Figs. 7g-i). Firstly, F PEC was larger in TC1 than in the other two experiments, supporting the fact that the simulated storm in TC1 was stronger, because the storm tended to spin-up faster when F PEC was larger (Smith et al., 2009). Although the C K/C D was alike in CTL and TC2, F PEC was larger in CTL than that in TC2. This was mainly due to the different C D used in those two experiments. Following the dynamical explanation of (Montgomery et al., 2010), the agradient tendencies near the surface caused the inflowing rings of boundary-layer air to converge farther inwards in the storm center before rising out of the boundary layer and ascending into the eyewall updraught, resulting in enhanced maximum tangential wind speed. Our result is consistent with this argument (Fig. 7). It also suggests that intense positive supergradient acceleration occurs in the vicinity of the maximum tangential wind speed and is associated with the outflow jet above the boundary layer.

  • Next, we investigate the simulated radar reflectivity and precipitation in the three experiments. Figure 8a shows the observed U.S. Air Force Defense Meteorological Satellite Program polar-orbiting satellite F-17 microwave imagery of polarization corrected temperature with a horizontal polarization at 91 GHz valid at 2112 UTC 6 August. The simulated composite radar reflectivity valid at 2100 UTC 6 August from CTL, TC1 and TC2 are shown in Figs. 8b-d, respectively. Due to the interaction of the typhoon circulation with the monsoon flow and vertical wind shear (Wang et al., 2012), the storm became asymmetric. All three experiments captured the asymmetric distribution of radar reflectivity and reproduced the unclosed eyewall in the southern part of the storm. The simulated reflectivity in TC2 was only slightly closer to observations than in the other two experiments, as it captured broader high reflectivity area in the southern eyewall. Otherwise, the overall rainfall structure was similar in all experiments.

    The observed and simulated radar reflectivity composites at 1200 UTC 7 August and 0000 UTC 8 August are shown in Figs. 8e-h and Figs. 8i-l, respectively. At 1200 UTC 7 August, Typhoon Morakot (2009) was located on the east side of Taiwan before landfall. It is evident from the observation (Fig. 8e) that the strongest reflectivity was located on the southwest side of the storm. This asymmetric rainfall pattern was captured by all the experiments. The simulated reflectivity in the eyewall region was stronger in CTL and TC1 than in TC2 (Figs. 8f-h). The simulated reflectivity in TC2 was slightly closer to observations than in the other two experiments because the unclosed eyewall, with the principal rainband located in the southern part of the storm and the strong echo over Taiwan, were captured in TC2. At 0000 UTC 8 August, the observed precipitation pattern became more asymmetric than in earlier periods. The strong echoes were observed on the south side of Taiwan (Fig. 8i). In CTL and TC1, the simulated reflectivity was still stronger than that in TC2 (Figs. 8j-l). Again, TC2 performed slightly better than in the other two experiments.

    Overall, the total precipitation was strongest in TC1 and weakest in CTL, especially after 1200 UTC 6 August (Fig. 9a). Note that the accumulation period of precipitation is 1 h. As TC1 simulated the strongest storm while CTL simulated the weakest storm, this result suggests that the total precipitation is correlated with the storm intensity. The 10 m domain-averaged divergences of moisture flux (note that negative values represent convergence of the moisture flux) are shown in Fig. 9b. It is evident that the convergence of moisture flux was correlated with the total precipitation. This result is not surprising, as the low-level moisture was the main source of the rainfall. Nonetheless, the result suggests that surface flux parameterizations have a substantial impact on precipitation simulations.

    Although the simulated rainfall over the ocean was strongly tied to surface flux parameterization, interestingly, we found that the rainfall over land (i.e., Taiwan) was much less sensitive to the surface flux parameterization. Figure 10 compares the 12 h accumulated precipitation from CTL, TC1 and TC2 together with the objective analyses of rainfall measured by automatic weather stations over Taiwan, valid at 0600 UTC and 1800 UTC 7 August. Prior to the landfall of Morakot (2009) in Taiwan (1800 UTC 6 August to 0600 UTC 7 August), the observation (Fig. 10a) shows two regions of strong precipitation across the island: one on the north side of Taiwan and the other on the south side over high mountains. All three experiments simulated these two regions of strong rainfall, although the simulated rainfall was much stronger than the observed valued. In particular, all three simulations over-predicted the precipitation from central to southern Taiwan. However, the difference in the precipitation distribution in the three experiments was very small over the whole island.

    Figure 9.  Averaged (a) accumulated rainfall and (b) 10 m divergence of moisture flux from experiment CTL, TC1 and TC2. The average was computed within the area of 300× 300 km from the surface minimum pressure center.

    Figure 10.  12 h accumulated precipitation (units: mm) valid at 0600 UTC 7 August 2009 from (a) automatic weather station hourly observations, (b) CTL, (c) TC1 and (d) TC2 over Taiwan; and 12 h accumulated precipitation valid at 1800 UTC 7 August 2009 from (e) automatic weather station hourly observations, (f) CTL, (g) TC1 and (h) TC2 over Taiwan.

    The above result suggests that the precipitation was less sensitive to the exchange coefficients over land than over the ocean in Morakot (2009). Over the ocean, the role of the exchange coefficient in precipitation was an indirect one, through influencing the (horizontal) moisture flux convergence (and ultimately rainfall) and through affecting the intensity of the storm. For the precipitation over land, the moisture came from the ocean, even though the end precipitation fell over land. In Typhoon Morakot (2009), the effect of terrain (i.e., forced uplift) played a dominant role in the distribution of the precipitation over Taiwan (Hall et al., 2013; Wang et al., 2013), which is likely the main reason for the similar rainfall simulations among the three experiments.

5. Summary and discussion
  • In this study, three numerical experiments were performed with the WRF-ARW model to study the impact of surface flux parameterizations on the structure and intensity of Typhoon Morakot (2009). The initial conditions of the three experiments were all from the JMA RSM analysis field. The simulated track and intensity of Morakot (2009) were verified against the best track. Different formulas of momentum and heat roughness lengths were tested in sensitivity experiments that governed the behavior of the surface exchange coefficients for momentum and heat transfers. The results showed that the simulated track was not sensitive to the exchange coefficients, but the simulated intensity and structure were.

    Our results indicate that the surface exchange coefficients are key factors for the simulation of surface wind speed and fluxes. The effect of C K on the surface enthalpy flux is straightforward because of the linear relationship between these two parameters. On the other hand, the effect of C D on the enthalpy flux takes place via the surface wind speed. When the C D is small, the maximum surface wind speed tends to be larger due to reduced surface friction. In turn, the enthalpy flux becomes larger because of the larger wind speed. Overall, we found C K had a larger impact on the enthalpy flux simulation than C D.

    Consistent with previous studies (e.g., Emanuel, 1995), the simulated storm intensity was found to be more sensitive to the ratio of C K/C D than to C K or C D alone. According to the idealized numerical simulation given by (Montgomery et al., 2010), C K/C D should have a critical value for the intensification of storm. If C D is too large, the storm will not intensify. When C K/C D is larger, the simulated storm is stronger and vice versa. In the CTL experiment, the intensity simulation is comparable to that in TC2, because a similar C K/C D was used in these two experiments.

    The pressure-wind relationship was also found to be sensitive to C D and C K, consistent with recent numerical studies of Atlantic hurricanes (Bao et al., 2012; Green and Zhang, 2013). Overall, the simulated intensity and pressure-wind relationship in TC2 was closest to the best track than those in CTL and TC1. This result is encouraging because the C K and C D used in TC2 were close to recent field and wave tank observations. This result is also consistent with that of (Zhang et al., 2012), who showed that observation-based surface layer and boundary layer physics led to improvements in the operational HWRF model and better intensity forecasts.

    Our results also indicate that simulated structures, such as the surface wind distribution, boundary layer heights, warm-core anomaly and height, and precipitation are affected by C D and C K. Compared to the dropsonde observations from (Zhang et al., 2011), the simulated kinematic boundary layer heights in TC2 are closer to observations than the other two experiments. The warm-core anomaly is tied to the storm intensity but not the warm-core height, consistent with Stern and Nolan (2012). The difference in the rainfall over the ocean is consistent with the difference in storm intensity, which can be explained by the difference in the convergence of moisture flux in the boundary layer. Over land, the simulated rainfall is much less sensitive to C D and C K than over the ocean, which we attribute to the dominance of the terrain effect on the precipitation in Typhoon Morakot (2009), as pointed out by (Wang et al., 2013) and (Hall et al., 2013).

    We also conducted dynamical analyses to investigate why C D and C K affect the vertical structure of wind velocities in the boundary layer. Consistent with (Montgomery et al., 2010), we found that a larger drag coefficient can lead to a larger gradient wind imbalance in the boundary layer (Fig. 7). As a result of the larger agradient forcing, the boundary-layer air converged farther inward near the storm center before rising out of the boundary layer and ascending into the eyewall updraft. The end result was enhanced maximum tangential wind speed, despite the loss of absolute angular momentum en route.

    In this study, we focused on investigating the sensitivity of the simulated intensity and structure of Typhoon Morakot (2009) to the surface exchange coefficients only, while keeping the rest of the model physics the same. We note that other parts of the model physics (e.g., planetary boundary layer parameterization and radiation parameterization) may also be important for TC simulations. Future work will evaluate the impact of other aspects of model physics on numerical simulations of TC structure and intensity change.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return