Modeling the Effects of Land Sea Roughness Contrast on Tropical Cyclone Winds

Similar documents
Super-parameterization of boundary layer roll vortices in tropical cyclone models

The dynamics of heat lows over flat terrain

10.6 The Dynamics of Drainage Flows Developed on a Low Angle Slope in a Large Valley Sharon Zhong 1 and C. David Whiteman 2

DUE TO EXTERNAL FORCES

SENSITIVITY OF DEVELOPING TROPICAL CYCLONES TO INITIAL VORTEX DEPTH AND THE HEIGHT OF ENVIRONMENTAL DRY AIR

ATMS 310 Tropical Dynamics

2. External nfluences

Sea and Land Breezes METR 4433, Mesoscale Meteorology Spring 2006 (some of the material in this section came from ZMAG)

Winds and Ocean Circulations

The Relationship between Storm Motion, Vertical Wind Shear, and Convective Asymmetries in Tropical Cyclones

Conditions for Offshore Wind Energy Use

Lecture 22: Ageostrophic motion and Ekman layers

+ R. gr T. This equation is solved by the quadratic formula, the solution, as shown in the Holton text notes given as part of the class lecture notes:

THE EFFECT OF THE OCEAN EDDY ON TROPICAL CYCLONE INTENSITY

Gravity waves in stable atmospheric boundary layers

August 1990 H. Kondo 435. A Numerical Experiment on the Interaction between Sea Breeze and

A R e R v e iew e w on o n th t e h e Us U e s s e s of o Clou o d u - (S ( y S s y t s e t m e )-Re R sol o ving n Mod o e d ls Jeff Duda

Chapter 2. Turbulence and the Planetary Boundary Layer

Atmospheric Waves James Cayer, Wesley Rondinelli, Kayla Schuster. Abstract

Ermenek Dam and HEPP: Spillway Test & 3D Numeric-Hydraulic Analysis of Jet Collision

Effect of Orography on Land and Ocean Surface Temperature

2.4. Applications of Boundary Layer Meteorology

Lecture 7. More on BL wind profiles and turbulent eddy structures. In this lecture

The comparative effects of frictional convergence and vertical wind shear on the interior asymmetries of a hurricane

EXTREME WIND GUSTS IN LARGE-EDDY SIMULATIONS OF TROPICAL CYCLONES

INTRODUCTION * Corresponding author address: Michael Tjernström, Stockholm University, Department of Meteorology, SE-

Idealized Numerical Modeling of a Land/Sea Breeze

The impacts of explicitly simulated gravity waves on large-scale circulation in the

An Analysis of the South Florida Sea Breeze Circulation: An Idealized Study

Boundary layer meso-scale flows. Hannu Savijärvi University of Helsinki

YAMEI XU YUQING WANG. (Manuscript received 8 December 2012, in final form 14 June 2013) ABSTRACT

ATS150: Global Climate Change. Oceans and Climate. Icebergs. Scott Denning CSU 1

Small- and large-scale circulation

A Theory for Strong Long-Lived Squall Lines Revisited

8.4 COASTAL WIND ANOMALIES AND THEIR IMPACT ON SURFACE FLUXES AND PROCESSES OVER THE EASTERN PACIFIC DURING SUMMER

Thorsten Mauritsen *, Gunilla Svensson Stockholm University, Stockholm, Sweden

It is advisable to refer to the publisher s version if you intend to cite from the work.

Air Pressure and Wind

Darwin s mid-evening surge

PHSC 3033: Meteorology Air Forces

Local Winds. Please read Ahrens Chapter 10

Tropical-cyclone flow asymmetries induced by a uniform flow revisited

The dryline is a mesoscale phenomena whose development and evaluation is strongly linked to the PBL.

Undertow - Zonation of Flow in Broken Wave Bores

CHAPTER 8 WIND AND WEATHER MULTIPLE CHOICE QUESTIONS

VORTEX EVOLUTION BY STRAINING: A MECHANISM FOR DOMINANCE OF STRONG INTERIOR ANTICYCLONES

PGF. Pressure Gradient. Wind is horizontal movement of the air or other word air in motion. Forces affecting winds 2/14/2017

THEORETICAL EVALUATION OF FLOW THROUGH CENTRIFUGAL COMPRESSOR STAGE

Chapter. Air Pressure and Wind

Undertow - Zonation of Flow in Broken Wave Bores

McKnight's Physical Geography 11e

Forest Winds in Complex Terrain

NUMERICAL INVESTIGATION OF THE FLOW BEHAVIOUR IN A MODERN TRAFFIC TUNNEL IN CASE OF FIRE INCIDENT

PUBLICATIONS. Journal of Advances in Modeling Earth Systems

Analysis of NAM Forecast Wind Shear for Dissipation of Mesoscale Convective Systems

National Weather Service

Mesoscale Meteorology

Review of Equivalent Neutral Winds and Stress

Standard atmosphere Typical height (m) Pressure (mb)

Atomspheric Waves at the 500hPa Level

The evolution of vortices in vertical shear. 11: Large-scale asymmetries

SUPPLEMENTARY INFORMATION

6.28 PREDICTION OF FOG EPISODES AT THE AIRPORT OF MADRID- BARAJAS USING DIFFERENT MODELING APPROACHES

Ocean Circulation. Si Hui Lee and Frances Wen. You can access ME at

Synoptic features of orographically enhanced heavy rainfall on the east coast of Korea associated with Typhoon Rusa (2002)

ABSTRACT INTRODUCTION

The Air-Sea Interaction. Masanori Konda Kyoto University

Observations and Modeling of Coupled Ocean-Atmosphere Interaction over the California Current System

Atmospheric & Ocean Circulation-

Wednesday, September 27, 2017 Test Monday, about half-way through grading. No D2L Assessment this week, watch for one next week

Mountain Forced Flows

AIRFLOW GENERATION IN A TUNNEL USING A SACCARDO VENTILATION SYSTEM AGAINST THE BUOYANCY EFFECT PRODUCED BY A FIRE

Flow and Mixing in the Liquid between Bubbles

Wind: Small Scale and Local Systems Chapter 9 Part 1

18.1 Understanding Air Pressure 18.1 Understanding Air Pressure Air Pressure Defined Measuring Air Pressure Air pressure barometer

2. THE NEW ENGLAND AIR QUALITY STUDY

Typhoon Vamei: An Equatorial Tropical Cyclone Formation

Meteorology. Circle the letter that corresponds to the correct answer

THE RESPONSE OF THE GULF OF MAINE COASTAL CURRENT SYSTEM TO LATE-SPRING

ESCI 107/109 The Atmosphere Lesson 9 Wind

Mesoscale air-sea interaction and feedback in the western Arabian Sea

AERODYNAMICS I LECTURE 7 SELECTED TOPICS IN THE LOW-SPEED AERODYNAMICS

Alongshore wind stress (out of the page) Kawase/Ocean 420/Winter 2006 Upwelling 1. Coastal upwelling circulation

Scales of Atmospheric Motion. The atmosphere features a wide range of circulation types, with a wide variety of different behaviors

Idealized WRF model sensitivity simulations of sea breeze types and their effects on offshore windfields: Supplementary material

Dynamics and variability of surface wind speed and divergence over mid-latitude ocean fronts

Influence of enhanced convection over Southeast Asia on blocking ridge and associated surface high over Siberia in winter

Numerical Simulations of the Effects of Coastlines on the Evolution of Strong, Long-Lived Squall Lines

Gravity Waves in Shear and

Summary of Lecture 10, 04 March 2008 Introduce the Hadley circulation and examine global weather patterns. Discuss jet stream dynamics jet streams

Why must hurricanes have eyes?

Wind Flow Validation Summary

The total precipitation (P) is determined by the average rainfall rate (R) and the duration (D),

The Coriolis force, geostrophy, Rossby waves and the westward intensification

Seasonal and interannual variation of currents in the western Japan Sea: numerical simulation in comparison with infrared satellite imagery

Organized Deep Cumulus Convection Over the South China Sea and its Interaction with Cold Surges

3. Climatic Variability. El Niño and the Southern Oscillation Madden-Julian Oscillation Equatorial waves

Local vs. Remote SST Forcing in Shaping the Asian-Australian Monsoon Variability

A Numerical Simulation of Convectively Induced Turbulence (CIT) above Deep Convection

Xiaoli Guo Larsén,* Søren Larsen and Andrea N. Hahmann Risø National Laboratory for Sustainable Energy, Roskilde, Denmark

Transcription:

SEPTEMBER 2007 W O N G A N D C H A N 3249 Modeling the Effects of Land Sea Roughness Contrast on Tropical Cyclone Winds MARTIN L. M. WONG AND JOHNNY C. L. CHAN Laboratory for Atmospheric Research, Department of Physics and Materials Science, City University of Hong Kong, Hong Kong, China (Manuscript received 30 May 2006, in final form 5 January 2007) ABSTRACT The fifth-generation Pennsylvania State University National Center for Atmospheric Research (PSU NCAR) Mesoscale Model (MM5) is used to simulate tropical cyclone (TC) wind distribution near landfall. On an f plane at 15 N, the effects of the different surface roughness between the land and sea on the wind asymmetry is examined under a strong constraint of a dry atmosphere and time-invariant axisymmetric mass fields. The winds are found to adjust toward a steady state for prelandfall (50, 100, and 150 km offshore), landfall, and postlandfall (50, 100, and 150 km inland) TC positions. The TC core is asymmetric even when it lies completely offshore or inland. The surface (10 m) wind asymmetry at the core for pre- (post) landfall position is apparently related to the acceleration (deceleration) of the flow that has just moved over the sea (land) as a response to the sudden change of surface friction. For prelandfall TC positions, the resulted strong surface inflow to the left and front left (relative to the direction pointing from sea to land) also induces a tangential (or total) wind maxima at a smaller radius, about 90 downstream of the maximum inflow, consistent with the absolute angular momentum advection (or work done by pressure). The surface maximum wind is of similar magnitude as the gradient wind. There is also a small region of weak outflow just inside the wind maxima. For postlandfall TC positions, inflow is weakened to the right and rear right associated with the onshore flow. Both onshore and offshore flows affect the surface wind asymmetry of the core in the landfall case. Above the surface and near the top of the planetary boundary layer (PBL), the wind is also asymmetric and a strongly supergradient tangential wind is primarily maintained by vertical advection of the radial wind. Much of the steady-state vertical structure of the asymmetric wind is similar to that forced by the motion-induced frictional asymmetry, as found in previous studies. The associated asymmetry of surface and PBL convergences has radial dependence. For example, the landfall case has stronger PBL convergence to the left for the 0 50-km core region, due to the radial inflow, but to the right for the 100 500-km outer region, due to the tangential wind convergence along the coastline. The strong constraint is then removed by considering an experiment that includes moisture, cumulus heating, and the free adjustments of mass fields. The TC is weakening and the sea level pressure has a slightly wavenumber-1 feature with larger gradient wind to the right than to the left, consistent with the drift toward the land. The asymmetric features of the wind are found to be very similar to those in the conceptual experiments. 1. Introduction Corresponding author address: Johnny Chan, Dept. of Physics and Materials Science, City University of Hong Kong, Tat Chee Ave., Kowloon, Hong Kong, China. E-mail: johnny.chan@cityu.edu.hk Tropical cyclone (TC) landfall processes have been an important research problem because of the potentially huge destruction of lives and properties when the intense rainfall, severe winds, and storm surge batter coastal regions. Strongly asymmetric structures of wind and rainfall usually accompany landfall. Previous studies have documented the asymmetric convective activities during TC landfall. Generally speaking, rainfall patterns of landfalling TCs vary widely from case to case because various factors (e.g., land surface properties, topography, vertical wind shear) could drive the asymmetries. Some idealized modeling studies (e.g., Chan and Liang 2003) have been carried out to highlight the nonuniform surface fluxes as a factor to generate the asymmetries. A more detailed summary of those results can be found in Chan et al. (2004). Some observational studies have documented the wind distribution of landfalling TCs (e.g., Powell 1982, 1987; Powell et al. 1991; Powell and Houston 1996, DOI: 10.1175/JAS4027.1 2007 American Meteorological Society

3250 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S VOLUME 64 1998). It was noted that maximum wind and gusts tend to occur near the most convective regions (e.g., Parrish et al. 1982). Possible mechanisms include the downward transport of high-velocity air by precipitation-induced downdrafts, the spreading of downbursts along the ground, and the formation of mesovortices (Willoughby and Black 1996). Powell (1987) argued that the wind asymmetry depends on the combined effects of land sea roughness differences, background environmental flow, and storm translation. The wind asymmetry has also been studied in a few idealized numerical simulations. Tuleya and Kurihara (1978) found that the roughness contrast between land and sea creates quasi-steady convergence (divergence) and negative (positive) relative vorticity zones along the coastline where the flow is onshore (offshore). Tuleya et al. (1984) noted the possibility of a temporary displacement of the surface circulation center from the pressure center at landfall. Jones (1987) found that the increase of rainfall to the left of the landfalling TC relative to a nonlandfalling one is related to the enhanced relative inflow in the left-front quadrant near landfall. Kepert (2002) noted that the asymmetric structure of Hurricane Danny at landfall is similar to that induced by motion, where, in each case, it is forced by frictional asymmetry. Recently, Kepert (2006a,b) has studied Hurricanes Georges and Mitch and also found close analogy between the motion-induced and landfallinduced asymmetry. There would also be differences between the structures forced by motion and land friction (Kepert 2002) because the asymmetric friction would be discontinuous rather than smooth, as in the motion-induced case, and the asymmetry of friction is much stronger for landfall. Furthermore, the asymmetry in friction would not cover the whole storm if the TC center is some distance from the landfall position. In an attempt to study the vortex drift associated with the asymmetric wind and convection near landfall, Wong and Chan (2006, hereafter WC06) found that the planetary boundary layer (PBL) convergence of the TC core could become strongly asymmetric even when the core TC circulation is completely away from the coast. They proposed the surface-induced asymmetric convergence as an important factor that affects the asymmetric convective activities of the TC core. Previous studies have not focused on such features, and a more thorough investigation has yet to be done. This paper therefore focuses on understanding the asymmetric wind distribution of landfalling TCs. The hypothesis that the adjustment of air moving away from a roughness discontinuity could produce an asymmetry in the TC core is also addressed. Idealized numerical TABLE 1. List of numerical experiments. The labels S and L stand for sea and land, respectively. Expt S150 S100 S50 LS L50 L100 L150 Position of TC center 150 km offshore 100 km offshore 50 km offshore At the coast 50 km inland 100 km inland 150 km inland experiments are performed with the fifth-generation Pennsylvania State University (PSU) National Center for Atmospheric Research (NCAR) Mesoscale Model (MM5: Dudhia 1993). The numerical experiments and simplifications of the model physics are described in section 2. Results on the wind asymmetries and the associated convergences are shown in section 3. Section 4 presents comparisons with a full-physics experiment, followed by the discussion and conclusions in section 5. 2. Model and design of experiments a. MM5 The experimental design is the same as in WC06. The triply nested square domains have horizontal resolutions of 45, 15, and 5 km and span 6750, 2250, and 1200 km, respectively. There are 26 vertical layers. An f plane at 15 N is used. The surface fluxes and vertical diffusion are determined from a modified Mellor Yamada level-2.5 PBL scheme (Janjić 1994). For the full-physics experiment, an explicit moisture prediction scheme (Dudhia 1989) and the Betts and Miller (1986) cumulus parameterization are also used. The specified vortex is also the same, which includes the horizontal variation of the tangential wind profile first used by Wang and Li (1992) and the vertical variation of the wind used by Wong and Chan (2004). The vortex is then spun up for a 24-h period after which it attains an intensity of 976 hpa. The axisymmetric part of the mass and wind fields then form the initial conditions of the conceptual experiments (see below for details) with the TC center placed at the center of the model domain. In all the experiments, a portion of the surface is specified to be land of roughness length 0.5 m (sea roughness depends on wind strength) so that the TC is located at 50, 100, 150 km onshore and inland, and right on the straight coastline, giving a total of seven experiments (Table 1). The mass fields are held fixed during the 48-h simulation without feed-

SEPTEMBER 2007 W O N G A N D C H A N 3251 back from moisture or heat sources. Two more experiments with sea-only and land-only surfaces are also performed as control. As it turns out, a quasi-steady state is reached for all the experiments. The hourly results of the second day (t 24 to t 48 h) are averaged and analyzed. The use of the time-invariant axisymmetric mass fields during the numerical integration implies that the intensity of the TC is unchanged, the mass adjustment is negligible, and there is no TC movement. The results of the full-physics experiment RD of WC06 1 will illustrate that mass adjustment is not severe, and the results of the conceptual experiments would still be applicable to a nonsteady condition of a weakening TC (see section 4). Although TC motion is not considered, we expect that the roughness contrast would be the most dominant reason for PBL asymmetry at landfall except for unusually fast movement. There are other reasons why we do not simply analyze the results of the RD experiment. The asymmetry in vertical motion in the RD experiment could result from vortex-scale vertical wind shear (see WC06 and Kepert 2006b) and not just the adjustment of air moving from one surface to another. By fixing the mass fields, we expect to remove the shear effect and determine more clearly the effect of roughness contrast on the asymmetries. Moreover, with the same mass fields for all of the conceptual experiments, comparisons of the magnitudes of winds among the experiments could be made. If the mass fields are allowed to change as in the RD experiment, the TC eventually evolves into different intensities at different times, making comparisons difficult. Analytical solutions of the PBL flow over a homogeneous surface have been successfully obtained by Kepert (2001) for stationary and moving TCs. It might be possible that a similar analytical solution can be obtained for a stationary TC on a land sea interface. However, the derivation would be much more difficult, so a numerical solution is opted instead. b. Physical processes examined in the idealized experiments 1 In the RD experiment of WC06, the mass fields are free to change. The TC center, which is initially 150 km offshore, drifts forward left (relative to the onshore direction) toward the rough and dry land surface. For the experiments with the axisymmetric mass fields invariant in time and no moisture variables, only the momentum variables are allowed to change. After initial adjustments, a steady state is reached. The equations for the diagnoses of the radial wind u and tangential wind are u u r u u du w r z dt f r 1 p r D u, u r d w r z dt f r u D, where D u and D are the diffusion terms that strongly depend on vertical diffusion (which includes the surface fluxes of momentum), and all the other symbols carry their usual meaning. The vertical motion w is related to the horizontal wind components through the continuity equation. In the absence of diffusion (i.e., D u D 0), the gradient wind solution (u w 0, g ) satisfies (2.1) and (2.2). As a consequence, gradient balance is generally observed above the PBL where the vertical diffusion is small (Willoughby 1990). When the vertical diffusion is not negligible, as for the case of the PBL, u and w become nonzero and would deviate from the gradient wind. Willoughby (1990) argued that, theoretically, axisymmetric supergradient flow could be realized by the inflow in the PBL, but observations do not support the systematic departure. On the other hand, the derivation and numerical simulation of Kepert (2001) and Kepert and Wang (2001) showed that a supergradient wind, primarily maintained by vertical advection of the radial wind, exists near the top of the PBL where the vertical gradient of the radial wind is large. In the following section we will also discuss the wind distribution in terms of its deviation from the gradient wind. 3. Results for the conceptual experiments a. Surface wind asymmetry The surface wind in this study is referred to the 10-m level wind, which is a model output variable based on the surface layer similarity theory. The maximum surface gradient wind speed is about 31 m s 1 at a radius of 50 km. As expected, the surface wind field for the sea-only experiment (Fig. 1a) is axisymmetric and characterized by the presence of inflow and a reduction of the tangential wind speed to a value less than that of the gradient wind. The maximum surface tangential and total winds are 24 and 26 m s 1 respectively, both found at a radius of 50 km. The wind reduction factors, computed as the ratio between the wind speeds and the gradient wind speed, are 0.77 and 0.84, respectively. For the land-only experiment the inflow angle and the speed reduction are larger (Fig. 1b), and the radius of maximum wind (RMW) is smaller, at 40 km. The surface wind reduction factors for the tangential and total

3252 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S VOLUME 64 FIG. 1. Tangential (contour) and radial (shaded) surface winds for the (a) sea-only and (b) land-only experiments. The TC center is located at the origin. (Unit: m s 1.) winds are, respectively, 0.43 and 0.57. These factors then decrease to nearly constant values at large radii. In addition, the maximum surface inflow is slightly larger in the land-only experiment (10.8 versus 9.6 m s 1 ) but is weaker at larger radii. For the experiments with a land sea interface, the most salient characteristic of the surface wind field is, as expected, the smaller surface wind speed (both tangential and radial components) over the land than the sea surface (Figs. 2a g). Some other important features also require explanations. First, the winds near the TC core are asymmetric. When the TC is 150 km from the coast before landfall (Fig. 2a), the eye/eyewall region is located completely over the sea. However, the radial inflow is clearly stronger to the left/front left (relative to the direction facing land) of the TC and the tangential wind is also asymmetric and strongest to the rear left/rear. This implies that the local wind is not determined solely by the local surface roughness. On the contrary, the radial inflow for the postlandfall cases (L50, L100, L150) is weaker on the right/rear-right side where the flow is onshore and toward the TC (Figs. 2e g, respectively), although the asymmetry of the core tangential wind is less clear. At the time of landfall, radial inflow is strong (weak) to the rear left (front right) of the TC (Fig. 2d). In this case, the offshore (onshore) flow is toward the rear-left (front right) quadrant. The maximum tangential wind occurs over the sea as the front half of the TC is over the land surface. Second, for prelandfall (S150, S100, S50) and landfalling (LS) cases (Figs. 2a d, respectively), the strongest inflow near the TC and the maximum tangential wind that occurs approximately 90 downstream at a smaller radius are associated with the offshore flow. This may seem counterintuitive as the offshore flow originates from the land surface at smaller wind speeds. The dynamics of this result is discussed further below. Third, a comparison of the magnitude of the flows with the control experiments could reveal the importance of the transitional effect near the coast. In regions where the flow has moved over the same surface, the wind carries the characteristics of that surface. For example, in the LS case, the tangential and radial flows more than 100 km offshore are close to those in the sea-only experiment (cf. Figs. 1a and 2d), while the flows more than 100 km inland are close to those in the land-only experiment (cf. Figs. 1b and 2d). In the transition region, the flow never quite attains either the sea-only or land-only states because the boundary layer adjustment time 1/I (where I is the inertial stability; see Eliassen and Lystad 1977) is of similar order as the rotational period r/ near the storm core. The maximum inflow for the LS case (Fig. 2d) has a magnitude 15 m s 1 and is actually larger than those in both the sea-only and land-only experiments. Moreover, the maximum tangential wind speed reaches over 25 m s 1 and is also larger than both control experiments. For the sea-only experiment, the surface tangential wind reduction factor (Fig. 3a) reaches 0.8 in only a small region, but that for the LS experiment has a region of near 1 (Fig. 3b), significantly larger than the control experiments. These asymmetric features can be explained by (2.1) and (2.2). For the surface wind, which is often of primary interest, the sudden additional force (and acceleration) due to a change of the roughness would be in the direction of the wind for the offshore flow and in the opposite direction of the wind for the onshore flow.

SEPTEMBER 2007 W O N G A N D C H A N 3253 FIG. 2. As in Fig. 1 except for the experiments with a land sea interface (see Table 1). The vertical dashed line indicates the coastline with land (sea) to its left (right).

3254 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S VOLUME 64 FIG. 3. Surface tangential wind reduction factor for the (a) sea-only and the (b) LS experiments. The TC center is at the origin. Values larger than 1 are shaded and the vertical dashed line in (b) indicates the coastline with land (sea) to its left (right). Therefore, if the land surface is rough, the inflow angle is large and the offshore flow would undergo strong inflow acceleration. This would explain why the maximum surface radial inflow associated with the offshore flow for prelandfall and landfalling cases (Figs. 2a d) could be stronger than both control experiments (Figs. 1a,b). Furthermore, the maximum total wind could be associated with the offshore flow if the air could move very close to the TC center and attain more kinetic energy. In fact, a contraction of the left (the direction relative to that pointing from sea to land) eyewall of Hurricane Danny in 1997 with associated maximum wind and rainfall was observed (Blackwell 2000). This inflow acceleration could also explain why the surface tangential wind is comparable to the gradient wind in the conceptual landfall experiments (Fig. 3). For small friction over sea (assumed negligible to facilitate the following discussion), the air that has just moved offshore is subjected to inward acceleration until the Coriolis and centrifugal forces balance the pressure gradient force in (2.1). However, the radial flow is still negative, which means that the tangential wind and total wind continue to increase according to (2.2). At this point, the total wind is already stronger than the (local) gradient wind owing to the presence of the inflow. Gradually the radial acceleration becomes positive (or the tangential wind has become supergradient) and the radial inflow ceases at some small radius. Note that there is also a region of outflow near the maximum tangential wind at a slightly smaller radius (e.g., Fig. 2b), consistent with the negative sign of d /dt downstream of the area of maximum surface tangential wind [see (2.2)]. To elaborate further, we consider the trajectory of the flow in the lowest model level ( 0.995; 43 m) in the LS case. Four streamlines are considered (Fig. 4). Each of the streamlines represents air movement for a 2-h period with the air located at 150 km either rear, right, front, or left of the TC center at the middle of the 2-h period. The results show that the onshore (offshore) flow associated with the right (left) path is accompanied by stronger tangential deceleration (inflow acceleration) than inflow deceleration (tangential ac- FIG. 4. Four streamlines in the LS case, at the lowest model level ( 0.995; 43 m). The streamlines represent airflow for a period of 2 h and the air is located at 150 km to the rear, right, front, and left of the TC center (TC symbol) at the middle of the period (marked by ), respectively. Dots indicate air position at half-hourly intervals. The land surface is shaded.

SEPTEMBER 2007 W O N G A N D C H A N 3255 FIG. 5. (a) Tangential and (b) inflow accelerations (i.e., d /dt and du/dt) along the four streamlines in Fig. 4; (c) and (d) the respective accelerations due to vertical diffusion, which includes surface friction (unit is 10 3 ms 2 ). The value of 0 on the time axis corresponds to the middle of the 2-h period. celeration) (Figs. 5a and 5b, respectively). The acceleration is primarily related to the change of surface roughness. Decomposition of the vertical diffusion of momentum into tangential and inflow components show that the evolution is consistent with the accelerations (Figs. 5c and 5d, respectively). For the left path, the tangential component of vertical diffusion is actually positive when the air is near the coast, which means that the tangential momentum flux from the upper layer is stronger than the tangential momentum flux at the surface (i.e., surface stress) and results in an acceleration. As noted by Kepert (2002), an analogy between the motion-induced asymmetry could be made in order to explain the asymmetric features. For example, in the LS case the maximum inflow occurred at the rear left, and could be realized by motion from right to left. The asymmetry has, however, higher wavenumber ( 1) components. This is partly due to the configuration of the land sea interface and, partly, the large frictional asymmetry that causes significant deviation from the linear regime in deriving the solution in Kepert (2001). Moreover, for prelandfall and postlandfall TC positions the analogy is not obvious. For example, for prelandfall the inflow is strongest at the left or front left. Note also that the maximum tangential wind that occurred at inner radii has been explained as enhanced angular momentum advection by Kepert and is similar to the explanation of the maximum total wind in terms of the work done by pressure in the above.

3256 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S VOLUME 64 b. Wind asymmetry above the surface 2 The difference between the maximum total wind and maximum tangential wind is found to be small. The study of the wind above the surface is also of great importance because the higher winds could be brought down to the surface as gusts, and the relationship between upper-level and surface winds has always been used for the estimation of surface winds based on upper-level measurement (e.g., radar, reconnaissance flight). The steady-state structure of the asymmetric flow is required to bring inertial waves to a halt (Kepert 2001). To begin, we first investigate the maximum tangential wind above the surface for the sea-only and land-only experiments. The supergradient nature of the tangential flow near the top of the PBL of a TC is simulated (Kepert 2001; Kepert and Wang 2001). The maximum tangential 2 wind of 39 (42) m s 1 for the sea- (land-) only experiment is located at 0.94 ( 0.885), or 522 (1024) m MSL, about 50 km from the TC center (Figs. 6a and 6c). It is about 24% (37%) supergradient, higher than the 10% 25% range found in Kepert and Wang (2001). The vortex is inertially stable in the current study, and the higher winds would be related to the different PBL parameterization used here because additional sensitivity experiments show that the strength of the supergradient wind is sensitive to the choice of PBL parameterization (not shown). The figures should then be interpreted with caution. Vertical advection of the radial wind in (2.1) is found to be the dominant contributor to the supergradient flow (not shown). For the LS case, the height, depth, and strength of the azimuthal mean supergradient jet is in between the sea-only and land-only experiments (Fig. 6e). At 100 km from the TC center, the supergradient wind has become less significant for all cases (Figs. 6b,d,f). The jet is clearly lower for the sea-only than the landonly experiment (cf. Figs. 6a and 6c). Moreover, the PBL is shallower at 50-km radius than at 100-km radius if we consider the depth of the radial inflow, and it is most evident in the sea-only case (cf. Figs. 6a and 6b). These results are consistent with Kepert (2001), who found that the jet and the PBL depth would be scaled as (2K/I) 1/2, K being the vertical diffusivity and I the inertial stability. Note that one could also define the PBL top to be the height at which the turbulent kinetic energy has dropped below a threshold value, similar to Gayno et al. (1994). Contrary to the above depth scale, this definition gives a higher PBL height at 50 than at 100 km from the TC center due to stronger shear production of turbulent kinetic energy. This seemingly contrasting result is merely a matter of definition. The winds above the surface layer for the LS case are also asymmetric. The asymmetries of tangential and radial winds rotate anticyclonically with height (Fig. 7). At 0.915, radial inflow is strongest to the front left and has become weaker (Fig. 7c). As at the surface, the region of maximum tangential wind is about 90 downstream of the region of maximum inflow so that the region of maximum wind is approximately in between. Such a rotation is similar to that found in Kepert (2001) for a moving cyclone. The flow has become symmetric at 0.825 (Fig. 7f). Kepert (2001) and Kepert and Wang (2001) studied the dynamics of supergradient winds. The linear effect is similar to the classical Ekman spiral, where supergeostrophic wind occurs near the top of the PBL. Vertical advection is found to increase the level of supergradient winds. As in Kepert (2001) and Kepert and Wang (2001), we would like to see how the supergradient flow is maintained. At 0.915 for the LS case, maximum supergradienticity (the excess of the Coriolis and centrifugal forces over the pressure gradient force) occurs to the left (Fig. 8a). The effects of radial advection and vertical diffusion are found to be small (not shown), so the tangential and vertical advections explain most of the asymmetries (Fig. 8b). Moreover, vertical advection is more dominant over the tangential advection (Figs. 8d and 8c, respectively). c. Vorticity and divergence An accompanying consequence of the asymmetric wind is the asymmetric vorticity and convergence. Except for the strongly positive relative vorticity in the TC core, the surface relative vorticity is characterized by a band of negative (positive) relative vorticity associated with the onshore (offshore) wind (not shown), in agreement with Tuleya and Kurihara (1978). The vorticity change is related to the friction term in the vorticity equation. Since the surface convergence (divergence) along the coast coincides with the negative (positive) vorticity, the magnitude of the vorticity band decreases downstream of the coast. WC06 hypothesized that the enhanced inflow could drive the asymmetric convection. Moreover, the degree of asymmetry is stronger for prelandfall than for postlandfall TCs. As can be seen from the previous subsections, the inflow asymmetry would lead to asymmetric divergence at the TC core. However, the convergence is also contributed by the speed divergence of the tangential wind. Therefore, each of the effects has to be taken into account.

SEPTEMBER 2007 W O N G A N D C H A N 3257 FIG. 6. Azimuthal-mean tangential, radial, and gradient winds for the (a) sea-only, (c) land-only, and (e) LS experiments at a radius of 50 km from the respective TC center. The lowest level shown is 0.995 (43 m). (b), (d), (f) The results at a radius of 100 km from the respective TC center. We investigate, for each of the seven cases in Table 1, the phase of the wavenumber-1 (WN1) convergence (both surface and PBL) of the horizontal wind ( V h ) in three radial bands: 1) 0 50, 2) 50 100, and 3) 100 500 km from the TC center. For simplicity the PBL flow is defined to be the averaged flow in the layer 1.0 0.9. For the surface (Fig. 9, large symbols), stronger convergence occurs to the left in the

3258 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S VOLUME 64 FIG. 7. Tangential (contour) and radial (shaded) winds at various levels above the surface for the LS case: (a) 0.975, (b) 0.94, (c) 0.915, (d) 0.885, (e) 0.855, and (f) 0.825. The vertical dashed line indicates the coastline, with land (sea) to its left (right). 0 50-km band except for the LS case, the only case in which the coastline is within the radial band. In the 50 100-km band surface convergence is larger to the right for cases where the coastline is within the radial band (L50, LS, S50) and is shifted slightly to the front right for the other cases. Surface convergence is consistently stronger to the right in the 100 500-km radial band as the convergence (divergence) effect at the

SEPTEMBER 2007 W O N G A N D C H A N 3259 FIG. 8. Combination of various terms in (2.1) at 0.915 for the LS case: (a) Coriolis centrifugal pressure gradient, (b) tangential advection vertical advection, (c) tangential advection, and (d) vertical advection. Unit is 10 3 ms 2. Positive values are shaded. The TC center is at the origin and the vertical dashed line represents the coastline with land (sea) to its left (right). coastline is included in all experiments. Results using a stronger vortex are similar (not shown). The phase of PBL convergence (Fig. 9, small symbols) is similar to that of the surface except for the LS case, where the PBL convergence is also stronger to the left as in the other experiments. In other words, the stronger surface convergence associated with the onshore wind does not result in stronger PBL convergence to the right. This could be explained by referring back to Figs. 7a c, where the discontinuity in tangential wind along the coastline has become much weaker above the surface while the stronger inflow to the left (or front left) persists. The asymmetric convergence is similar to that forced by motion (Kepert 2001), which for the boundary layer convergence (and hence the vertical velocity forced by the boundary layer) in the 0 50-km band is to the left. However, there are also distinct differences. For example, the surface convergence is largest to the right and the convergence for the 100 500-km band has a reverse relationship. 4. Comparison with the full-physics experiment The results presented in section 3 assumed that the feedback from heat and moisture sources to the mass fields is small. This assumption may not be correct because most, if not all, TCs weaken upon landfall. Moreover, the asymmetric convergences and convective heating may distort the mass fields, and there may also be significant mass adjustments for the strong imbalance of the flow that has just moved over the surface of distinct roughness. However, Kepert (2002) has noted that the simulated wind field matches quite well with observations, even when Hurricane Danny is not in a steady state. In this section the similarities and differences between the results presented in the previous section and those of

3260 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S VOLUME 64 FIG. 9. Azimuthal location (relative to the direction pointing from sea to land) of the maximum wavenumber-1 component of the surface (large symbols) and PBL (small symbols) convergence in the 0 50-km (squares), 50 100-km (triangles), and 100 500-km (circles) radial bands. The results of a particular experiment is shown at the corresponding x position (e.g., L150 results along x 150 km). the RD experiment of WC06 are discussed. The RD experiment is initiated with an intense TC 150 km offshore. Because of a large-scale asymmetry generated, the vortex is found to drift toward the coast. The focus in this section is on the effects of feedback on the mass fields and the wind in nonsteady intensity situations. a. Sea level pressure and wind distributions Past modeling and observational studies have shown that the surface pressure distribution remains fairly symmetric despite significant asymmetry of the wind (e.g., Tuleya and Kurihara 1978). The results of the RD experiment at t 116 h and t 122 h indicate that the sea level pressure had not undergone large asymmetric transition (Figs. 10a and 10b), and there is also no noticeable discontinuity of the pressure gradient across the coast. As revealed by the gradient winds (Figs. 10c and 10d), the pressure gradient is slightly higher to the right than to the left, probably related to the tendency of increasing surface pressure due to the convergence/ divergence along the coastline at the storm scale, and is consistent with the landward drift of the vortex found in WC06. The sea level pressure field of the RD experiment would therefore favor stronger winds to the right near landfall. However, the asymmetric surface inflow and tangential wind are found to be similar to those in the conceptual experiments (Figs. 10e and 10f; cf. Fig. 2). This shows that the effect of the asymmetric mass fields on the surface wind asymmetry is small compared to that forced by the land sea interface. Note also that the TC is weakening, with central pressure rising from 910 to 923 hpa during this period, and the TC is much more intense than that in the conceptual experiments. This also supports the qualitative results obtained in the conceptual experiments, which are applicable to more general conditions. To see more clearly the surface wind asymmetry and the similarities to the results in the conceptual experiments, the hourly results of the 144-h simulation are further examined. Just prior to landfall, the surface tangential wind is a maximum in the rear left (Fig. 11a), while the maximum surface inflow occurs to the left/ front left (Fig. 11c). In other words, the maximum surface tangential wind occurs about 90 downstream of the maximum surface radial inflow, and both are associated with the offshore flow as in the conceptual experiments. Moreover, the radii of maximum tangential and maximum radial winds are smaller ( 20 km) just prior to landfall (Figs. 11b and 11d, respectively), which is also consistent with the arguments in section 3a concerning (2.3) and (2.4). The maximum surface gradient wind occurs to the right (Fig. 11e). After landfall, the vortex core breaks down with maximum surface gradient wind located in outer bands (Fig. 11f). Furthermore, the surface wind reduction factor of the tangential wind increases when the TC is close to the coast (Fig. 12). It increases from about 0.65 when the TC is about 150 km from the coast to over 0.8 just prior to landfall. The results are also similar to those in the conceptual experiments (cf. Fig. 3). b. Asymmetric convection WC06 showed that the asymmetries of rainfall at inner and outer radii differ. Rainfall is larger over the left or front left within 100 km from the TC surface center over the 6-day simulation but, when the averaging distance is 300 km, rainfall is actually smallest to the left and strongest to the front, then gradually to the front right as the TC approaches the coast and finally moved over land. These results agree fairly well with the PBL convergence shown in Fig. 9 in the current study in two ways. First, the inner (0 50 km) and outer (100 500 km) PBL convergences are out of phase. Second, the phase of the PBL convergence in the middle (50 100 km) radial band rotates from front right to right as the vortex approaches the coast from the sea.

SEPTEMBER 2007 W O N G A N D C H A N 3261 FIG. 10. (a) Sea level pressure (hpa), (c) surface gradient wind speed (m s 1 ), and (e) surface radial (shaded) and tangential (contour) winds (m s 1 ) at t 116 h in the RD experiment of WC06; (b), (d), (f) respective results at t 122 h. The TC center is at the origin and the vertical dashed line represents the coastline with land (sea) to its left (right). 5. Summary and discussion We have studied the asymmetric wind distribution of TCs forced by the different surface roughness between land and sea using MM5. Under a strong constraint that the axisymmetric mass fields are time invariant during the model integration and a straight coast separating land and sea, the center of the TCs are placed directly at the coast and at positions corresponding to prelandfall and postlandfall positions. The cur-

3262 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S VOLUME 64 FIG. 11. (a) Azimuthal location (relative to the direction pointing from sea to land) and (b) radial distance of the maximum surface tangential wind in the RD experiment of WC06. Each symbol represents the result at a particular hour of the 6-day (144 h) simulation. (c), (d) Respective results for the surface inflow; (e), (f) the surface gradient wind.

SEPTEMBER 2007 W O N G A N D C H A N 3263 FIG. 12. Ratio of the maximum surface tangential wind to the gradient wind in the RD experiment of WC06. The vertical dashed line marks the time of landfall at t 122 h. rent study could be regarded as an extension of the study of Kepert (2001) and Kepert and Wang (2001), but using a surface of nonuniform roughness. The results for the surface wind indicate that, apart from the onshore (offshore) convergence (divergence), significant asymmetry could also be triggered in the TC core. For prelandfall TCs, the acceleration of the offshore flow in the left-front quadrant of the TC, which is shown to be related to vertical diffusion, has a radial component that enhances the inflow. Work is done by the pressure gradient force and results in strong tangential wind and maximum wind associated with the offshore flow, which is located approximately in the left-rear and left quadrants. The maximum tangential wind is also shown to be stronger than the gradient wind. However, it should be noted that in reality the maximum wind for a landfalling TC does not necessarily have to be stronger than that without landfall since some TCs also become less intense as they approach the coast. For postlandfall TCs, radial acceleration of the onshore flow causes a reduced inflow in the right rear quadrant. The winds above the surface layer are also asymmetric and the tangential component could also be supergradient. We have shown that vertical advection of the radial wind is the most important contribution to the supergradient tangential wind, in agreement with Kepert and Wang (2001). Although many of the results are analogous to the motion-induced asymmetry, there are also differences. First, the motion-induced asymmetry is dominated by the wavenumber-1 component for slow translation speed or the linear approximation (e.g., Shapiro 1983; Kepert 2001), but the land sea-induced asymmetry has forcing of higher ( 1) wavenumbers and strong nonlinear interactions. Second, it is shown that the maximum surface and PBL convergences occur in distinct quadrants between inner and outer regions. While radial inflow gives stronger convergence to the left of the TC in the inner region (0 50 km), tangential acceleration (deceleration) gives stronger convergence to the right in the outer region (100 500 km). Such a radial dependence would not be true for motion-induced asymmetry. Two processes could make a steady-state impossible. One is due to the mass-adjustment and intensity changes of the vortex, which involves a change of the mass fields and would occur even without land. We have shown in the RD experiment, even when it is not in steady state and we allow the free adjustment of mass, that the basic features in the conceptual experiments remain. These results are found to be similar to those from the full-physics version of the MM5, which suggests that in landfall situations the different roughness between land and sea leads to strong wind asymmetries that can be largely explained by the net acceleration associated with a deviation from gradient balance. Another is due to the motion of the TC, which is not considered in this study. For a slow-moving vortex as in the RD experiment, the motion has negligible effect and the asymmetry is virtually related to the land sea roughness contrasts alone. Kepert (2006b) has also shown with observations that the results remain true in the case of the slow-moving Hurricane Mitch. For larger translational speed, the modification to the land sea-induced asymmetries would be related to the direction and speed of movement. For example, if the TC is moving from right to left (relative to the direction facing land in the Northern Hemisphere), the core asymmetries would be enhanced because the effect of motion could promote asymmetries of a similar azimuthal phase. On the other hand, the core asymmetries would be reduced (or even reversed in azimuthal phase for fast movement) if the TC is moving from left to right. In reality a landfalling TC would also interact with topographic features of various complexity. An obvious follow-up work of the current study would be to consider topographic effects, perhaps in idealized simulations as in the current study or using real terrain data (e.g., South China).

3264 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S VOLUME 64 Acknowledgments. This research is sponsored by the Research Grant Council of the Hong Kong Special Administrative Region, China Grant CityU 100203. Thanks are due to Jeff Kepert and the anonymous reviewers for their helpful reviews. REFERENCES Betts, A. K., and M. Miller, 1986: A new convective adjustment scheme. Part II: Single column tests using GATE wave, BOMEX, ATEX and arctic air-mass data sets. Quart. J. Roy. Meteor. Soc., 112, 693 709. Blackwell, K. G., 2000: The evolution of Hurricane Danny (1997) at landfall: Doppler-observed eyewall replacement, vortex contraction/intensification, and low-level wind maxima. Mon. Wea. Rev., 128, 4002 4016. Chan, J. C. L., and X. Liang, 2003: Convective asymmetries associated with tropical cyclone landfall. Part I: f plane simulations. J. Atmos. Sci., 60, 1560 1576., K. S. Liu, S. E. Ching, and E. S. T. Lai, 2004: Asymmetric distribution of convection associated with tropical cyclones making landfall along the South China coast. Mon. Wea. Rev., 132, 2410 2420. Dudhia, J., 1989: Numerical study of convection observed during the Winter Monsoon Experiment using a mesoscale twodimensional model. J. Atmos. Sci., 46, 3077 3107., 1993: A nonhydrostatic version of the Penn State NCAR Mesoscale Model: Validation tests and simulation of an Atlantic cyclone and cold front. Mon. Wea. Rev., 121, 1493 1513. Eliassen, A., and M. Lystad, 1977: The Ekman layer of a circular vortex: A numerical and theoretical study. Geophys. Norv., 31, 1 16. Gayno, G. A., N. L. Seaman, A. M. Lario, and D. R. Stauffer, 1994: Forecasting visibility using a 1.5-order closure boundary layer scheme in a 12-km non-hydrostatic model. Preprints, 10th Conf. on Numerical Weather Prediction, Portland, OR, Amer. Meteor. Soc., 18 20. Janjić, Z. I., 1994: The step-mountain eta coordinate model: Further developments of the convection, viscous sublayer, and turbulence closure schemes. Mon. Wea. Rev., 122, 927 945. Jones, R. W., 1987: A simulation of hurricane landfall with a numerical model featuring latent heating by the resolvable scales. Mon. Wea. Rev., 115, 2279 2297. Kepert, J. D., 2001: The dynamics of boundary layer jets within the tropical cyclone core. Part I: Linear theory. J. Atmos. Sci., 58, 2469 2484., 2002: The impact of landfall on tropical cyclone boundary layer winds. Extended Abstracts, 25th Conf. on Hurricanes and Tropical Meteorology, San Diego, CA, Amer. Meteor. Soc., 335 336., 2006a: Observed boundary layer wind structure and balance in the hurricane core. Part I: Hurricane Georges. J. Atmos. Sci., 63, 2169 2193., 2006b: Observed boundary layer wind structure and balance in the hurricane core. Part II: Hurricane Mitch. J. Atmos. Sci., 63, 2194 2211., and Y. Wang, 2001: The dynamics of boundary layer jets within the tropical cyclone core. Part II: Nonlinear enhancement. J. Atmos. Sci., 58, 2485 2501. Parrish, J. R., R. W. Burpee, F. D. Marks Jr., and R. Grebe, 1982: Rainfall patterns observed by digitized radar during the landfall of Hurricane Frederic (1979). Mon. Wea. Rev., 110, 1933 1944. Powell, M. D., 1982: The transition of the Hurricane Frederic boundary-layer wind field from the open Gulf of Mexico to landfall. Mon. Wea. Rev., 110, 1912 1932., 1987: Changes in the low-level kinematic and thermodynamic structure of Hurricane Alicia (1983) at landfall. Mon. Wea. Rev., 115, 75 99., and S. H. Houston, 1996: Hurricane Andrew s landfall in South Florida. Part II: Surface wind fields and potential realtime applications. Wea. Forecasting, 11, 329 349., and, 1998: Surface wind fields of 1995 Hurricanes Erin, Opal, Luis, Marilyn, and Roxanne at landfall. Mon. Wea. Rev., 126, 1259 1273., P. P. Dodge, and M. L. Black, 1991: The landfall of Hurricane Hugo in the Carolinas: Surface wind distribution. Wea. Forecasting, 6, 379 399. Shapiro, L. J., 1983: The asymmetric boundary layer flow under a translating hurricane. J. Atmos. Sci., 40, 1984 1998. Tuleya, R. E., and Y. Kurihara, 1978: A numerical simulation of the landfall of tropical cyclones. J. Atmos. Sci., 35, 242 257., M. A. Bender, and Y. Kurihara, 1984: A simulation study of the landfall of tropical cyclones. Mon. Wea. Rev., 112, 124 136. Wang, B., and X. Li, 1992: The beta drift of three-dimensional vortices: A numerical study. Mon. Wea. Rev., 120, 579 593. Willoughby, H. E., 1990: Gradient balance in tropical cyclones. J. Atmos. Sci., 47, 265 274., and P. G. Black, 1996: Hurricane Andrew in Florida: Dynamics of a disaster. Bull. Amer. Meteor. Soc., 77, 543 549. Wong, M. L. M., and J. C. L. Chan, 2004: Tropical cyclone intensity in vertical wind shear. J. Atmos. Sci., 61, 1859 1876., and, 2006: Tropical cyclone motion in response to land surface friction. J. Atmos. Sci., 63, 1324 1337.