On the Use of Two-Dimensional Incompressible Flow to Study Secondary Eyewall Formation in Tropical Cyclones

Similar documents
Using NOGAPS Singular Vectors to Diagnose Large-scale Influences on Tropical Cyclogenesis

Secondary eyewall formation in tropical cyclones

15B.7 RESPONSE OF CONVECTION TO HURRICANE-LIKE HORIZONTAL AND VERTICAL SHEARS

Spiral Rainbands in a Numerical Simulation of Hurricane Bill (2009). Part II: Propagation of Inner Rainbands

Precipitation Structure and Processes of Typhoon Nari (2001): A Modeling Propsective

VORTEX INTERACTIONS AND THE BAROTROPIC ASPECTS OF CONCENTRIC EYEWALL FORMATION

Cloud-Resolving Simulations of West Pacific Tropical Cyclones

Convection and Shear Flow in TC Development and Intensification

Simulated Interactions Between Symmetric Tropical Cyclone-like Vortices

How Do Outer Spiral Rainbands Affect Tropical Cyclone Structure and Intensity?*

A New Ocean Mixed-Layer Model Coupled into WRF

8B.2 MULTISCALE OBSERVATIONS OF TROPICAL CYCLONE STRUCTURE USING AIRBORNE DOPPLER COMPOSITES. Miami, FL. Miami, FL

Structure and Formation of an Annular Hurricane Simulated in a Fully Compressible, Nonhydrostatic Model TCM4*

8C.6 Idealized simulations of the impact of dry Saharan air on Atlantic hurricanes

Impacts of environmental humidity on concentric eyewall structure

Advanced Research WRF High Resolution Simulations of Hurricanes Katrina, Rita and Wilma (2005)

Western North Pacific Typhoons with Concentric Eyewalls

AMERICAN METEOROLOGICAL SOCIETY

Lectures on Tropical Cyclones

Inner core dynamics: Eyewall Replacement and hot towers

1. Introduction. In following sections, a more detailed description of the methodology is provided, along with an overview of initial results.

Effect of the Initial Vortex Structure on Intensification of a Numerically Simulated Tropical Cyclone

1. INTRODUCTION. investigating the differences in actual cloud microphysics.

Vortex Rossby Waves in a Numerically Simulated Tropical Cyclone. Part II: The Role in Tropical Cyclone Structure and Intensity Changes*

15.6 A NUMERICAL MODELING STUDY OF THE MICROPHYSICAL PROCESSES LEADING TO TROPICAL CYCLOGENESIS UNDER DIFFERENT ENVIRONMENTAL CONDITIONS

Convection and Shear Flow in TC Development and Intensification

Hurricanes are intense vortical (rotational) storms that develop over the tropical oceans in regions of very warm surface water.

Mélicie Desflots* RSMAS, University of Miami, Miami, Florida

TROPICAL CYCLONE GENESIS IN TC-LAPS: THE IMPORTANCE OF SUFFICIENT NET DEEP CONVECTION AND SYSTEM SCALE CYCLONIC ABSOLUTE VORTICITY

Isentropic Analysis applied to Convection, Hurricanes and Walker Circulation

Typhoon Relocation in CWB WRF

Tropical Cyclone Genesis and Sudden Changes of Track and Intensity in the Western Pacific

Effects of Convective Heating on Movement and Vertical Coupling of Tropical Cyclones: A Numerical Study*

Vortex Interactions and Barotropic Aspects of Concentric Eyewall Formation

A Numerical Study of the Impact of Vertical Shear on the Distribution of Rainfall in Hurricane Bonnie (1998)

Tropical Cyclone Intensification from Asymmetric Convection: Energetics and Efficiency

WMO Training Course on Tropical Cyclones La Réunion (September 2015)

Vortex Rossby Waves in a Numerically Simulated Tropical Cyclone. Part I: Overall Structure, Potential Vorticity, and Kinetic Energy Budgets*

Incorporation of 3D Shortwave Radiative Effects within the Weather Research and Forecasting Model

sheared tropical cyclones

Tropical cyclone energy dispersion under vertical shears

The role of mid-level vortex in the intensification and weakening of tropical cyclones

The Structure and Evolution of Hurricane Elena (1985). Part II: Convective Asymmetries and Evidence for Vortex Rossby Waves

P1.1 BAROCLINICITY INFLUENCES ON STORM DIVERGENCE IN THE SUBTROPICS

The variability and predictability of axisymmetric hurricanes. in statistical equilibrium

PUBLICATIONS. Journal of Advances in Modeling Earth Systems

DISTRIBUTION STATEMENT A: Distribution approved for public release; distribution is unlimited.

Advisor: Prof. Hung Chi Kuo Speaker: Yu Ming Tsai Date: 2008 May 6. Outline

Influences of Asymmetric Flow Features on Hurricane Evolution

Interactions between Simulated Tropical Cyclones and an Environment with a Variable Coriolis Parameter

Barotropic instability in the tropical cyclone outer region

The Formation of Tropical Cyclones 1

Secondary eyewall formation in two idealized, full-physics modeled hurricanes

Hurricane Rainband and Intensity Experiment 2005: RAINEX. Principal Investigators:

Impact of cloud-radiative processes on hurricane track

10D.2 Methods for Introducing Vortical Hot Tower Heating in Numerical Models: Retrieving Latent Heat

Numerical Simulation of Torrential Rainfall and Vortical Hot Towers in a Midlatitude Mesoscale Convective System

LARGE-SCALE WRF-SIMULATED PROXY ATMOSPHERIC PROFILE DATASETS USED TO SUPPORT GOES-R RESEARCH ACTIVITIES

Track sensitivity to microphysics and radiation

Dependence of tropical cyclone intensification on the Coriolis parameter

地球系统科学前沿讲座 台风研究现状和问题 林岩銮

Department of Meteorology, University of Utah, Salt Lake City, UT. Second revision. Submitted to Monthly Weather Review.

Post Processing of Hurricane Model Forecasts

A Tropical Cyclone with a Very Large Eye

Impact of subgrid-scale vertical turbulent mixing on eyewall asymmetric structures and mesovortices of hurricanes

Four- dimensional climate data sets of the AMMA Special Observing Period #3

Numerical Simulation of Hurricane Bonnie (1998). Part II: Sensitivity to Varying Cloud Microphysical Processes

Convective-Scale Variations in the Inner-Core Rainbands of a Tropical Cyclone

Improved Tropical Cyclone Boundary Layer Wind Retrievals. From Airborne Doppler Radar

EVOLUTION OF FINITE VORTICITY STRIPS USING A NON-DIVERGENT, BAROTROPIC MODEL

Simulating the formation of Hurricane Katrina (2005)

Tropical Cyclone Intensification

Sensitivity of intensifying Atlantic hurricanes to vortex structure

Isentropic Analysis of a Simulated Hurricane

Xuanli LI and Zhaoxia PU. Department of Atmospheric Sciences, University of Utah, Salt Lake City, Utah, USA

A New Method for Representing Mixed-phase Particle Fall Speeds in Bulk Microphysics Parameterizations

Tropical cyclone genesis efficiency: mid-level versus bottom vortex

Sensitivity of intensifying Atlantic hurricanes to vortex structure

Corresponding author address: Yu Wang, 3141 Turlington Hall, University of Florida, Gainesville, FL

Tropical cyclones in ver/cal shear: dynamic, kinema/c, and thermodynamic aspects of intensity modification

8.2 Numerical Study of Relationships between Convective Vertical Velocity, Radar Reflectivity Profiles, and Passive Microwave Brightness Temperatures

Vortex Rossby Waves and Hurricane Evolution in the Presence of Convection and Potential Vorticity and Hurricane Motion

Tropical Cyclone Formation

Initialization of Tropical Cyclone Structure for Operational Application

Numerical Study of Precipitation Intensification and Ice Phase Microphysical Processes in Typhoon Spiral Band

High-Resolution Simulation of Hurricane Bonnie (1998). Part I: The Organization of Eyewall Vertical Motion

Impacts of Turbulence on Hurricane Intensity

WRF Model Simulated Proxy Datasets Used for GOES-R Research Activities

INTERNAL CONTROL OF HURRICANE INTENSITY VARIABILITY: THE DUAL NATURE OF POTENTIAL VORTICITY MIXING

Topic 1: Tropical Cyclone Structure and Intensity Change

A modeling study of Typhoon Nari (2001) at landfall: 2. Structural changes and terrain induced asymmetries

NOTES AND CORRESPONDENCE. The Effects of Dissipative Heating on Hurricane Intensity

Evolution of Multiscale Vortices in the Development of Hurricane Dolly (2008)

3A.6 HURRICANES IVAN, JEANNE, KARL (2004) AND MID-LATITUDE TROUGH INTERACTIONS

Weather Research and Forecasting Model. Melissa Goering Glen Sampson ATMO 595E November 18, 2004

The Properties of Convective Clouds Over the Western Pacific and Their Relationship to the Environment of Tropical Cyclones

Abstract. 1. Introduction

SIMULATION OF ATMOSPHERIC STATES FOR THE CASE OF YEONG-GWANG STORM SURGE ON 31 MARCH 2007 : MODEL COMPARISON BETWEEN MM5, WRF AND COAMPS

Impact of different cumulus parameterizations on the numerical simulation of rain over southern China

Improving Air-Sea Coupling Parameterizations in High-Wind Regimes

Transcription:

VOLUME 67 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 DECEMBER 2010 On the Use of Two-Dimensional Incompressible Flow to Study Secondary Eyewall Formation in Tropical Cyclones YUMIN MOON, DAVID S. NOLAN, AND MOHAMED ISKANDARANI Rosenstiel School of Marine and Atmospheric Science, University of Miami, Miami, Florida (Manuscript received 28 July 2010, in final form 7 September 2010) ABSTRACT Previous studies have offered hypotheses for the mechanisms that lead to secondary eyewall formation in tropical cyclones by using two-dimensional incompressible flow. Those studies represented the convectioninduced vorticity field as either large but weak vortices that are the same sign as the tropical cyclone core or as purely asymmetric vorticity perturbations that are an order of magnitude weaker than the core. However, both observations and full-physics simulations of tropical cyclones indicate that the convection-induced vorticity field should also include clusters of small vorticity dipoles whose magnitude is comparable to that of the high-vorticity core. Results of numerical simulations indicate that the interaction between the tropical cyclone core vortex and the convection-induced small vorticity dipoles of considerable strength in twodimensional flow does not lead to coherent concentric vorticity ring formation. The axisymmetrization process under the simplification of two-dimensional incompressible flow appears to be incomplete for describing secondary eyewall formation. 1. Introduction The exact mechanisms that lead to secondary eyewall formation in tropical cyclones remain unclear. The complete understanding of this process is critical in predicting the future evolution of tropical cyclones, as a period of significant intensity change typically occurs after secondary eyewall formation (e.g., Willoughby et al. 1982; Kossin and Sitkowski 2009; Kuo et al. 2009). Several hypotheses have been proposed (e.g., Willoughby et al. 1984; Nong and Emanuel 2003; Terwey and Montgomery 2008; Judt and Chen 2010), and a number of them use two-dimensional incompressible flow (Kuo et al. 2004, hereafter K04; Kuo et al. 2008, hereafter K08; Martinez et al. 2010, hereafter MBY10). K04 proposed that the interaction between a small strong vortex and a large weak vortex could result in a concentric vorticity ring structure. K04 was an extension of the work by Dritschel and Waugh (1992, hereafter DW92) who examined the interaction of two circular Corresponding author address: Yumin Moon, University of Miami, RSMAS/MPO, 4600 Rickenbacker Causeway, Miami, FL 33149. E-mail: ymoon@rsmas.miami.edu vortex patches having equal uniform vorticity but unequal size. The initial configuration of the two vortices (see Fig. 1 of DW92) is controlled by two parameters: the ratio of the radii of the smaller vortex to the larger vortex R 2 /R 1 and the dimensionless gap between the vortices D/R 1 where D5d 2 (R 1 1 R 2 )andd is the initial separation distance between the vortex centers. In the parameter space considered by DW92 (0, R 2 /R 1, 1 and 0,D/R 1, 1.6), the stronger vortex (hereafter R 1 ) is larger than the weaker vortex (hereafter R 2 ). DW92 found five different regimes of the interaction: no merger, partial/complete merger, and partial/complete strainingout. In the complete straining-out (CSO) regime (see Fig. 3b of DW92), the stronger vortex completely shears apart the smaller weaker vortex and turns it into thin filaments that wrap around the stronger vortex but do not get incorporated into the stronger vortex. Noting that the vorticity ring produced in the CSO regime by DW92 was much too thin to be identified with that observed in the outer eyewall of a tropical cyclone, K04 extended DW92 by examining the interaction with an additional parameter: the ratio of uniform vorticity in the weaker vortex to the stronger vortex z 2 /z 1 (see Fig. 2 of K04). Now the problem is controlled by three parameters. In the parameter space considered by K04 DOI: 10.1175/2010JAS3615.1 Ó 2010 American Meteorological Society 3765

3766 JOURNAL OF THE ATMOSPHERIC SCIENCES (1, R2/R1, 4, 0, D/R1, 4 and 0.1, z2/z1, 1), the stronger vortex is usually smaller and represents the rapidly rotating tropical cyclone core. The weaker vortex is larger and represents the vorticity field induced by moist convection outside the core. K04 found a set of parameters under which the core vortex completely shears apart the large weak vortex and turns it into thin filaments. These filaments wrap around the core vortex, become barotropically unstable, break down and mix to form a concentric ringlike structure (see section 3a for more details). K08 extended the results of K04 by representing the core vortex with a modified Rankine profile, which is more consistent with observations (e.g., Mallen et al. 2005), and reached similar conclusions. Both K04 and K08 represented the convection-induced vorticity field as a large vortex of weak vorticity that is the same sign as the core vortex. This characterization appears to be based on radar observations of tropical cyclones that had a large area of peripheral convection that eventually wrapped around the core to produce a secondary eyewall [e.g., Typhoon Lekima (2001) from Fig. 1 of K04]. This large area of convection appeared as a region of uniform reflectivity value, and K04 assumed that reflectivity could be used as a proxy for the convection-induced vorticity. However, although reflectivity offers useful information about the vigor of convection and the rainfall rate associated with it, it does not provide an accurate picture of the convection-induced vorticity field. In a recent study by MBY10, it was proposed that secondary eyewall formation in tropical cyclones could result from the axisymmetrization of purely asymmetric wavenumber-4 (n 5 4) disturbances by the high-vorticity core of tropical cyclones. Unlike K04 and K08, MBY10 represented the hurricane core as a region of low vorticity surrounded by a ring of elevated vorticity, and a skirt of vorticity was added outside the core, consistent with observations (Mallen et al. 2005). The n 5 4 vorticity asymmetries used to represent the convectioninduced vorticity field are broad (;40 km in width) and their peak values are approximately an order of magnitude smaller (;15%) than the vorticity in the hurricane core region (see Fig. 2 of MBY10). MBY10 chose n 5 4 asymmetries because they dominate the wave activity and represent the fastest growing modes for the particular axisymmetric vortex used in that study. However, in contrast to the choices by K04, K08, and MBY10, observations (e.g., Fig. 9 of Powell 1990; Figs. 6 and 7 of Hence and Houze 2008; Fig. 15 of Didlake and Houze 2009) of tropical cyclones indicate that a better representation of the convection-induced vorticity field should also include positive negative vorticity dipoles that have magnitudes comparable to that of the VOLUME 67 FIG. 1. (a) A horizontal cross section of vertical component of vorticity (s21) at z 5 4.40 km from a WRF simulation of Hurricane Bill (2009) and (b) a convective stratiform partitioning applied at the same time as in (a) by using the method described in Braun et al. (2010). high-vorticity core vortex (z2/z1 0.5) and are smaller in size than the core vortex (R2/R1, 1). Full-physics numerical simulations also support this notion. Shown in Fig. 1a is a horizontal cross section at z 5 4.40 km of the vertical component of vorticity (z 5 y/ x 2 u/ y) from a full-physics simulation of Hurricane Bill (2009) using the Weather Research Forecasting model version 3.1.1 (WRF; Skamarock et al. 2008). Initial and boundary conditions from the Geophysical Fluid Dynamics Laboratory (GFDL) hurricane forecasting model were used, along with WRF single-moment 6-class (WSM-6) microphysics (Hong and Lim 2006), the Yonsei University (YSU) planetary boundary layer (Noh et al. 2003; Hong et al. 2006), the Goddard shortwave (Chou et al. 1998),

DECEMBER 2010 M O O N E T A L. 3767 and Rapid Radiative Transfer Model (RRTM) longwave (Mlawer et al. 1997) parameterization schemes. Figure 1b shows a convective stratiform partitioning applied at the same time as in Fig. 1a by using the method described in Braun et al. (2010), which is similar to that by Tao et al. (1993). Convection outside the tropical cyclone core appears mostly convective in the upwind region of a spiral rainband (east and southeast of the core) while an extensive region of stratiform convection is present to the north of the core. Although different types of convection and anomalies (e.g., environment induced and stormmotion induced) coexist in tropical cyclones outside the core, vorticity dipoles of different size and strength appear to be the dominant features of the convection-induced vorticity field. These vorticity dipoles are generated through the tilting of horizontal vorticity into the vertical direction by updrafts embedded within convection (e.g., Fig. 19 of Powell 1990; Fig. 10 of Montgomery et al. 2006). 2. Numerical model and initial conditions In this study, we use a fully nonlinear model of twodimensional, incompressible fluid to examine how a stronger, larger tropical cyclone core vortex interacts with convection-induced small vorticity dipoles whose magnitude is comparable to that of the high-vorticity tropical cyclone core. The numerical model is a spectral element model presented by Iskandarani et al. (1995) and Iskandarani (2008). The domain is a 400 km 3 400 km box that has periodic boundary conditions in the zonal direction and is discretized into 80 3 80 spectral elements of 5-km size. An eighth degree polynomial is used to approximate the solution in each element, yielding an average effective grid spacing of 625 m. A 3-s time step is used to integrate for 12 h with a third-order Runge Kutta scheme. A low value of viscosity (5 m 2 s 21 )isused. The high-vorticity core vortex is represented by a smoothed modified Rankine profile: 8 z core, 0 # r # r core d z z(r) 5 core S r r core 1 d r (11a) 1 0.5z 2d core (1 a) S r core 1 d r ><, r r core 2d core d # r # r core 1 d r (11a) >: 0.5z core (1 a) r $ r 1 d, core r core (1) where r is the radial distance from the center of the core vortex, r core is the radial width of the uniform vorticity region in the center of the modified Rankine profile, z core is the value of the uniform vorticity, and a is the decay parameter. The cubic Hermite polynomial, S(x) 5 1 2 3x 2 1 2x 3, that satisfies S(0) 5 1, S(1) 5 0, and S9(0) 5 S9(1) 5 0, is used over the transition zone of 2d to smooth the sharp vorticity decrease outside r core. Also z core 5 0.008 s 21, r core 5 15 km, a 5 0.5, d 5 5km are used in (1), making the peak wind speed equal to 56.2 m s 21 at a radius of maximum wind (RMW) of 16.6 km (see Fig. 2a). This is in good agreement with observations that the average intensity around the time of secondary eyewall formation is ;60 m s 21 (Kossin and Sitkowski 2009), and that the typical RMW of a tropical cyclone with the maximum wind of 60 m s 21 is 15 km (Kimball and Mulekar 2004). Vortex patches of a convection-induced vorticity dipole are represented by a smoothed Rankine profile: z(r) 56z dipole exp r 6, (2) a where r is the radial distance from the center of a patch, z dipole is the magnitude of uniform vorticity in the patch, and a is a parameter that controls the radial width of the patch. Vorticity decreases to 0.1% of the peak magnitude at approximately r 5 1.38a. 3. Results a. A like-signed convection-induced vortex We first reproduce the formation of a concentric vorticity ring structure from the interaction between a tropical cyclone core vortex and a like-signed, convection-induced vortex. The core vortex is initially centered at x 5260 km and y 5 0 km. The convection-induced vortex is created by using (2) with z dipole 5 0.002 s 21 and a 5 45 km and is initially centered at x 5 60 km and y 5 0 km (see Fig. 2b). For this section only, the model is integrated for 24 h. As briefly described in section 1, the core vortex shears apart the convection-induced vortex and turns it into thin filaments that wrap around the core vortex (Fig. 2c). Vorticity filaments then become barotropically unstable (Fig. 2d) and break down and mix to form a concentric ringlike structure of considerable width (Fig. 2e). This vorticity ring is stabilized by the core vortex-induced differential rotation, which opposes the self-imposed differential rotation of the ring and prevents phase locking between the ring edges (Kossin et al. 2000). This concentric ring

3768 JOURNAL OF THE ATMOSPHERIC SCIENCES VOLUME 67 FIG. 2. (a) Radial profiles of tangential velocity (solid, m s21) and vorticity (dash dotted, 3104 s21) of the highvorticity core vortex at t 5 0 h and azimuthally averaged tangential velocity at t 5 24 h (dashed). Vorticity field (31023 s21) from the interaction between the core vortex and the like-signed large but weak convection vortex at t 5 (b) 0, (c) 6, (d) 12, (e) 18, and (f) 24 h. See section 3a for details. structure is clearly separated from the high-vorticity core by a region of lower vorticity (Fig. 2f), and it is clearly associated with a secondary wind maximum (dashed line in Fig. 2a). b. A single vorticity dipole The interaction between the core vortex and the convection-induced vorticity field as represented by a

DECEMBER 2010 M O O N E T A L. 3769 FIG. 3. Vorticity field (310 23 s 21 ) from the interaction between the core vortex and a positive negative vorticity dipole at (from top to bottom) t 5 0, 2, 4, and 6 h. See section 3b for details.

FIG. 4. Vorticity field (31023 s21) from the interaction between the core vortex and multiple vorticity dipoles at (from left to right) t 5 0, 3, 6, and 12 h. See section 3c for details. 3770 JOURNAL OF THE ATMOSPHERIC SCIENCES VOLUME 67

DECEMBER 2010 M O O N E T A L. 3771 positive negative vorticity dipole is now examined. Hereafter, the core vortex is initially centered at x 5 0 km and y 5 0 km, and z dipole 5 0.004 s 21 and a 5 5 km are used in (2) to create a vorticity dipole. In the left column of Fig. 3, the negative and positive vorticity patches are initially centered at x 5 50 and 65 km and y 5 0 km, respectively, while the right column of Fig. 3 has them initially centered at x 5 50 km and y 567.5 km, respectively. The separation distance between the patches is about 3 km. The main difference between these two configurations is the orientation of the initial self-induced motion of the dipole with the circular advection by the core vortex. Because the separation distance between the patches of the dipole is small, both the positive and negative vorticity patches induce considerable cyclonic and anticyclonic flows on their respective counterparts. The resulting self-induced motion of the dipole is equal to the induced flow of each vorticity patch on the other (Batchelor 1967, chapter 7.3). The self-induced motion is initially directed parallel to the circular advection in the left column of Fig. 3 but perpendicular to the circular advection in the right column of Fig. 3. The core vortex again completely shears apart the positive vorticity patch and turns it into filaments. However, the negative vorticity patch continuously interacts with positive vorticity filaments to prevent them from forming a coherent ringlike structure. Vorticity filaments gradually decay because of diffusion. In both cases as well as their permutations that have the self-induced motion in the opposite direction, no concentric ringlike structure is formed. c. Multiple vorticity dipoles There is a large amount of convection around the core of tropical cyclones, and within convection exist numerous updrafts and downdrafts. Therefore, a more realistic representation of the convection-induced vorticity field is to use multiple vorticity dipoles. Three cases are shown. The top row of Fig. 4 is initialized with four groups of two of the same vorticity dipole used in Fig. 3. The middle row of Fig. 4 is initialized with a spiral band of vorticity dipoles, as observations (e.g., Senn and Hiser 1959; Barnes et al. 1983) show that convection outside the core of tropical cyclones is typically organized in spiral shapes. The radially inner vorticity patches spiral radially inward from r 5 65 to 50 km in the counterclockwise direction while the outer vorticity patches spiral radially inward from r 5 80 to 65 km. The separation distance among the vorticity patches is 3 km. The bottom row of Fig. 4 is initialized with the same spiral band of vorticity dipoles used in the middle row of Fig. 4 on top of a spiral band of weak positive vorticity. This spiral band of weak positive vorticity represents the enhanced background vorticity from stretching driven by net condensational heating within a spiral rainband. The spiral band of weak positive vorticity is constructed from 50 overlapping circular patches of positive vorticity, similar to the method in which spiral rainband diabatic heating was constructed in Moon and Nolan (2010). Each circular patch of positive vorticity is created by using (2) with z dipole 5 0.00015 s 21 and a 5 20 km, and its center location spirals radially inward from r 5 72.5 to 57.6 km in the counterclockwise direction. In all cases described above, the high-vorticity core vortex shears apart the like-signed positive vorticity patches of the dipoles and turns them into filaments that wrap around the core vortex. However, negative vorticity patches continuously interact with positive vorticity filaments to prevent them from forming a coherent concentric ringlike structure. It is possible to argue that a thin concentric vorticity ring structure is formed for all cases at t 5 6 h (third column of Fig. 4) and 12 h (fourth column of Fig. 4). However, Fig. 5 confirms that the azimuthally-averaged tangential velocity field around the center of the core vortex does not exhibit secondary wind maxima that could be associated with secondary eyewall formation in tropical cyclones. Just as K04 stated for the DW92 results, the concentric ring is too thin to be identified as representative of secondary eyewall formation. 4. Summary and discussion The formation of a concentric vorticity ring structure from the interaction between the high-vorticity core vortex and the positive negative vorticity dipoles has been examined in the framework of two-dimensional, incompressible flow. In contrast to previous studies, the vorticity field induced by peripheral moist convection outside the rapidly rotating tropical cyclone core was represented by vorticity dipoles that have half the magnitudeofthecorevortexbutaremuchsmaller.this representation is more consistent with observations and full-physics simulations of tropical cyclones. In addition to the cases shown in section 3, many other combinations of vorticity dipoles were examined. However, in all cases considered in this study, no coherent concentric ringlike structure of vorticity was formed. The favorable results from previous studies using twodimensional, incompressible flow to study secondary eyewall formation in tropical cyclones suggested that the essential mechanisms of the phenomenon could be fully captured by two-dimensional dynamics. The effects of convection were represented as two-dimensional vorticity perturbations in this approach. If the above mechanisms are valid, then the concentric vorticity ring structure should be produced as the end state of two-dimensional

3772 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 67 numerical simulations over a large parameter space, as a large fraction (;60% 70%) of tropical cyclones of sufficient strength (i.e., maximum wind speed of ;60 m s 21 or higher) are observed to form secondary eyewalls at some time during their lifetime (Kossin and Sitkowski 2009; Kuo et al. 2009). The lack of a coherent concentric vorticity ring in the final state when vorticity dipoles are used as initial conditions indicates that the essential mechanisms of secondary eyewall formation are not fully captured by the two-dimensional framework. The results of this study suggest that the axisymmetrization process under the simplification of two-dimensional incompressible flow is incomplete for describing secondary eyewall formation. Recent observational and numerical simulation studies (e.g., Houze et al. 2007; Terwey and Montgomery 2008; Houze 2010; Judt and Chen 2010) emphasize the importance of three-dimensional processes in secondary eyewall formation. Acknowledgments. Support from the National Science Foundation through Grants ATM-0756308 for D. S. Nolan and Y. Moon and OCE-0622662 for M. Iskandarani is gratefully acknowledged. Constructive reviews by Drs. Jim Kossin, Shigeo Yoden, and two anonymous reviewers and helpful comments from Dr. Brian E. Mapes are appreciated. We thank Mr. Munehiko Yamaguchi for help with initially setting up the numerical model. Calculations were performed on the High Performance Computing core at the Center for Computation Science of the University of Miami. REFERENCES FIG. 5. Radial profiles of azimuthally-averaged tangential velocity (m s 21 ) of (a) top, (b) middle, and (c) bottom rows of Fig. 4 at t 5 6 (solid) and 12 h (dashed). Barnes, G. M., E. J. Zipser, D. Jorgensen, and F. Marks Jr., 1983: Mesoscale and convective structure of a hurricane rainband. J. Atmos. Sci., 40, 2125 2137. Batchelor, G. K., 1967: An Introduction to Fluid Dynamics. Cambridge University Press, 615 pp. Braun, S. A., M. T. Montgomery, K. J. Mallen, and P. D. Reasor, 2010: Simulation and interpretation of the genesis of Tropical Storm Gert (2005) as part of the NASA Tropical Cloud Systems and Processes Experiment. J. Atmos. Sci., 67, 999 1025. Chou, M.-D., M. J. Suarez, C.-H. Ho, M. M.-H. Yan, and K.-T. Lee, 1998: Parameterizations for cloud overlapping and shortwave single-scattering properties for use in general circulation and cloud ensemble models. J. Climate, 11, 202 214. Didlake, A. C., Jr., and R. A. Houze Jr., 2009: Convective-scale downwdrafts in the principal rainband of Hurricane Katrina (2005). Mon. Wea. Rev., 137, 3269 3293. Dritschel, D. G., and D. W. Waugh, 1992: Quantification of the inelastic interaction of unequal vortices in two-dimensional vortex dynamics. Phys. Fluids, 4A, 1737 1744. Hence, D. A., and R. A. Houze Jr., 2008: Kinematic structure of convective-scale elements in the rainbands of Hurricanes Katrina and Rita (2005). J. Geophys. Res., 113, D15108, doi:10.1029/2007jd009429.

DECEMBER 2010 M O O N E T A L. 3773 Hong, S.-Y., and J.-O. J. Lim, 2006: The WRF single-moment 6-class microphysics scheme (WSM6). J. Korean Meteor. Soc., 42, 129 151., Y. Noh, and J. Dudhia, 2006: A new vertical diffusion package with an explicit treatment of entrainment processes. Mon. Wea. Rev., 134, 2318 2341. Houze, R. A., Jr., 2010: Clouds in tropical cyclones. Mon. Wea. Rev., 138, 293 334., S. S. Chen, B. F. Smull, W.-C. Lee, and M. M. Bell, 2007: Hurricane intensity and eyewall replacement. Science, 315, 1235 1239. Iskandarani, M., 2008: Simulating hydrostatic and non-hydrostatic oceanic flows. Int. J. Numer. Methods Fluids, 58, 1135 1146., D. B. Haidvogel, and J. P. Boyd, 1995: A staggered spectral element model with application to the oceanic shallow water equations. Int. J. Numer. Methods Fluids, 20, 393 414. Judt, F., and S. S. Chen, 2010: Convectively generated potential vorticity in rainbands and formation of secondary eyewall in Hurricane Rita of 2005. J. Atmos. Sci., 67, 3581 3599. Kimball, S. K., and M. S. Mulekar, 2004: A 15-year climatology of North Atlantic tropical cyclones. Part I: Size parameters. J. Climate, 17, 3555 3575. Kossin, J. P., and M. Sitkowski, 2009: An objective model for identifying secondary eyewall formation in hurricanes. Mon. Wea. Rev., 137, 876 892., W. H. Schubert, and M. T. Montgomery, 2000: Unstable interactions between a hurricane s primary eyewall and a secondary ring of enhanced vorticity. J. Atmos. Sci., 57, 3893 3917. Kuo, H.-C., L.-Y. Lin, C.-P. Chang, and R. T. Williams, 2004: The formation of concentric vorticity structures in typhoons. J. Atmos. Sci., 61, 2722 2734., W. H. Schubert, C.-L. Tsai, and Y.-F. Kuo, 2008: Vortex interactions and barotropic aspects of concentric eyewall formation. Mon. Wea. Rev., 136, 5183 5198., C.-P. Chang, Y.-T. Yang, and H.-J. Jiang, 2009: Western North Pacific typhoons with concentric eyewalls. Mon. Wea. Rev., 137, 3758 3770. Mallen, K. J., M. T. Montgomery, and B. Wang, 2005: Reexamining the near-core radial structure of the tropical cyclone primary circulation: Implications for vortex resiliency. J. Atmos. Sci., 62, 408 425. Martinez, Y., G. Brunet, and M. K. Yau, 2010: On the dynamics of two-dimensional hurricane-like concentric rings vortex formation. J. Atmos. Sci., 67, 3253 3268. Mlawer, E. J., S. J. Taubman, P. D. Brown, M. J. Iacono, and S. A. Clough, 1997: Radiative transfer for inhomogeneous atmospheres: RRTM, a validated correlated-k model for the longwave. J. Geophys. Res., 102 (D14), 16 663 16 682. Montgomery, M. T., M. E. Nicholls, T. A. Cram, and A. B. Saunders, 2006: A vortical hot tower route to tropical cyclogenesis. J. Atmos. Sci., 63, 355 386. Moon, Y., and D. S. Nolan, 2010: The dynamic response of the hurricane wind field to spiral rainband heating. J. Atmos. Sci., 67, 1779 1805. 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, 421 427. Nong, S., and K. A. Emanuel, 2003: A numerical study of the genesis of concentric eyewalls in hurricane. Quart. J. Roy. Meteor. Soc., 129, 3323 3338. Powell, M. D., 1990: Boundary layer structure and dynamics in outer hurricane rainbands. Part I: Mesoscale rainfall and kinematic structure. Mon. Wea. Rev., 118, 891 917. Senn, H. V., and H. W. Hiser, 1959: On the origin of hurricane spiral rain bands. J. Meteor., 16, 419 426. Skamarock, W. C., and Coauthors, 2008: A description of the advanced research WRF version 3. NCAR Tech. Note 4751STR, 113 pp. Tao, W.-K., J. Simpson, C.-H. Sui, B. Ferrier, S. Lang, J. Scala, M.-D. Chou, and K. Pickering, 1993: Heating, moisture, and water budgets of tropical and midlatitude squall lines: Comparisons and sensitivity to longwave radiation. J. Atmos. Sci., 50, 673 690. Terwey, W. D., and M. T. Montgomery, 2008: Secondary eyewall formation in two idealized, full-physics modeled hurricanes. J. Geophys. Res., 113, D12112, doi:10.1029/2007jd008897. Willoughby, H. E., J. A. Clos, and M. G. Shoreibah, 1982: Concentric eye walls, secondary wind maxima, and the evolution of the hurricane vortex. J. Atmos. Sci., 39, 395 411., H. L. Jin, S. J. Lord, and J. M. Piotrowicks, 1984: Hurricane structure and evolution as simulated by an axisymmetric, nonhydrostatic numerical model. J. Atmos. Sci., 41, 1169 1186.