dependence of tropical cyclone intensification on the...

12
LI ET AL. MAY 2012 1 Dependence of tropical cyclone intensification on the Coriolis parameter TIM LI AND XUYANG GE Department of Meteorology and IPRC, University of Hawaii, Honolulu, Hawaii MELINDA PENG Naval Research Laboratory, Monterey, California WEI WANG Shanghai Minhang Meteorology Bureau, Shanghai, China (Manuscript received , in final form ) ABSTRACT The dependence of tropical cyclone (TC) intensification on the Coriolis parameter was investigated in an idealized hurricane model. By specifying an initial balanced vortex on an f-plane, we observed faster TC de- velopment under lower planetary vorticity environment than under higher planetary vorticity environment. The diagnosis of the model outputs indicates that the distinctive evolution characteristics arise from the extent to which the boundary layer imbalance is formed and maintained in the presence of surface friction. Under lower planetary vorticity environment, stronger and deeper subgradient inflow develops due to Ekman pumping ef- fect, which leads to greater boundary layer moisture convergence and condensational heating. The strengthened heating further accelerates the inflow by lowing central pressure further. This positive feedback loop eventually leads to distinctive evolution characteristics. The outer size (represented by the radius of gale-force wind) and the eye of the final TC state also depend on the Coriolis parameter. The TC tends to have larger (smaller) outer size and eye under higher (lower) planetary vorticity environment. Whereas the radius of maximum wind or the eye size in the current setting is primarily determined by inertial stability, the TC outer size is mainly controlled by environmental absolute angular momentum. Key words: Corresponding author address: Tim Li, Department of Meteorology and IPRC, University of Hawaii, Honolulu, Hawaii, USA. Email: timli@ hawaii.edu DOI: 1. Introduction The fact that tropical cyclones (TCs) typically form a few degrees away from the equator (Gray 1968) implies that a non-zero Coriolis parameter is necessary to spin up a vor- tex. Beyond this low limit of planetary vorticity, however, it is not clear how dynamically TC development depends on the Coriolis parameter. From a pure dynamics point of view, the planetary vorticity may modulate the radius of Rossby deformation, the latter of which may determine the geostrophic adjustment and thus affect the final state of bal- anced flows in response to thermal forcing (Schubert and Hack 1982; Nolan 2007). Using a simplified model, De- Maria and Pickle (1988) found that the development into a hurricane is slower under a higher planetary vorticity en- vironment. They attributed the slower intensification to the greater radius of inflow away from the vortex center. Using an axisymmetric model on an f-plane, Bister (2001) found that an initial weak vortex takes twice the time to develop into hurricane intensity at 30°N than at 10°N. The study indicated that the inner core convection remains weaker for a long time in a higher Coriolis parameter environment and thus slows down the moistening of the inner region. Wang and Zhou (2008) found that, climatologically, TCs formed in November in western North Pacific have a great- er chance of rapid intensification (RI) than those formed in August, even though the TC frequency in August is greater than that in November. Given that the mean genesis loca- tion in August is nearly 8° latitude north of that in Novem- ber, it is likely that the planetary vorticity, in addition to other environmental factors, may play a role in determining

Upload: others

Post on 20-Jul-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Dependence of tropical cyclone intensification on the ...iprc.soest.hawaii.edu/users/li/www/Lietal2012.pdfRossby deformation, the latter of which may determine the geostrophic adjustment

Li et aL.May 2012 1

Dependence of tropical cyclone intensification on the Coriolis parameter

TiM Li and Xuyang ge

Department of Meteorology and IPRC, University of Hawaii, Honolulu, Hawaii

MeLinda Peng

Naval Research Laboratory, Monterey, California

Wei Wang

Shanghai Minhang Meteorology Bureau, Shanghai, China

(Manuscript received , in final form )

aBStRaCt

The dependence of tropical cyclone (TC) intensification on the Coriolis parameter was investigated in an idealized hurricane model. By specifying an initial balanced vortex on an f-plane, we observed faster TC de-velopment under lower planetary vorticity environment than under higher planetary vorticity environment. The diagnosis of the model outputs indicates that the distinctive evolution characteristics arise from the extent to which the boundary layer imbalance is formed and maintained in the presence of surface friction. Under lower planetary vorticity environment, stronger and deeper subgradient inflow develops due to Ekman pumping ef-fect, which leads to greater boundary layer moisture convergence and condensational heating. The strengthened heating further accelerates the inflow by lowing central pressure further. This positive feedback loop eventually leads to distinctive evolution characteristics.

The outer size (represented by the radius of gale-force wind) and the eye of the final TC state also depend on the Coriolis parameter. The TC tends to have larger (smaller) outer size and eye under higher (lower) planetary vorticity environment. Whereas the radius of maximum wind or the eye size in the current setting is primarily determined by inertial stability, the TC outer size is mainly controlled by environmental absolute angular momentum.

Key words:

Corresponding author address: Tim Li, Department of Meteorology and IPRC, University of Hawaii, Honolulu, Hawaii, USA. Email: [email protected]

DOI:

1. IntroductionThe fact that tropical cyclones (TCs) typically form a few

degrees away from the equator (Gray 1968) implies that a non-zero Coriolis parameter is necessary to spin up a vor-tex. Beyond this low limit of planetary vorticity, however, it is not clear how dynamically TC development depends on the Coriolis parameter. From a pure dynamics point of view, the planetary vorticity may modulate the radius of Rossby deformation, the latter of which may determine the geostrophic adjustment and thus affect the final state of bal-anced flows in response to thermal forcing (Schubert and Hack 1982; Nolan 2007). Using a simplified model, De-Maria and Pickle (1988) found that the development into

a hurricane is slower under a higher planetary vorticity en-vironment. They attributed the slower intensification to the greater radius of inflow away from the vortex center. Using an axisymmetric model on an f-plane, Bister (2001) found that an initial weak vortex takes twice the time to develop into hurricane intensity at 30°N than at 10°N. The study indicated that the inner core convection remains weaker for a long time in a higher Coriolis parameter environment and thus slows down the moistening of the inner region.

Wang and Zhou (2008) found that, climatologically, TCs formed in November in western North Pacific have a great-er chance of rapid intensification (RI) than those formed in August, even though the TC frequency in August is greater than that in November. Given that the mean genesis loca-tion in August is nearly 8° latitude north of that in Novem-ber, it is likely that the planetary vorticity, in addition to other environmental factors, may play a role in determining

Page 2: Dependence of tropical cyclone intensification on the ...iprc.soest.hawaii.edu/users/li/www/Lietal2012.pdfRossby deformation, the latter of which may determine the geostrophic adjustment

VoLuMe 22 TROPICAL CyCLONE RESEARCH AND REvIEW

the RI preference.Motivated by the aforementioned observational and

modeling studies, we intend to investigate specific dynamic processes through which TC development depends on the Coriolis parameter. The specific science questions we at-tempt to address in this study are: How does the environ-mental planetary vorticity affect the TC intensification? What are structural differences in inner core, boundary layer thickness and outer size for TCs simulated with dif-ferent planetary vorticity parameters?

The paper is organized as the follows: In section 2, the model and experiments are described. In section 3 we analyze the cause of different TC development behaviors from the boundary layer gradient wind imbalance perspec-tive. A positive feedback among the radial inflow, moisture convergence and surface evaporation, diabatic heating and surface pressure drop is illustrated. Finally, the conclusion and discussion are given in the last section.

2. Model experiments The numerical model used in this study is the triply nest-

ed, movable mesh primitive equation model-TCM3 (Wang 2001). The model consists of 25 layers in the vertical from σ=0 to σ=1 (here σ is normalized pressure relevant to surface pressure) with higher resolutions in the planetary boundary layer (PBL). The model physics include a modi-fied Monin-Obukhov scheme for the surface flux calcula-tions (Fairall et al. 1996), an E-ε turbulence closure scheme for sub-grid scale vertical mixing above the surface layer (Langland and Liou 1996), an explicit treatment of mixed phase cloud microphysics (Lin et al. 1983), and a fourth-order horizontal diffusion with a deformation-dependent diffusion coefficient. The model prognostic variables con-sist of zonal and meridional winds, surface pressure, tem-perature, turbulence kinetic energy and its dissipation rate, and mixing ratios of water vapor, cloud water, rain water, cloud ice, snow and graupel.

The model has been used for many studies on TC (e.g., Wang 2001, 2002; Li et al. 2003, 2006; Ge et al. 2007, 2008; Li 2012). The model is configured with three nested domains: a coarse mesh of 22.5 km resolution and two 2-way nested domains of 7.5 and 2.5 km resolution, re-spectively. The domains have a square shape, with a side of 5400, 1800, and 450 km respectively. A strong damping is specified within a sponge layer near the model lateral boundaries to minimize wave reflections. No cumulus pa-rameterization was used in the inner two domains. The sea surface temperature is set to be a constant of 29°C. The en-vironmental temperature and relative humidity profiles are the same as that in Holland (1997), representing the mean summertime conditions over the western Pacific.

The initial vortex is an axisymmetric vortex with a maxi-mum tangential wind speed of 15 ms-1 at a radius of 100 km. The strength of the tangential wind decreases with the height and vanishes at 100 hPa. Given the specified tan-

gential wind field, the mass and thermodynamic fields are then obtained based on a nonlinear balance equation so that the initial vortex satisfies both the hydrostatic and gradient wind balance.

To investigate the sensitivity of TC intensification to the Coriolis parameter, a series of experiments were designed by running the model on an f-plane centered at different lat-itudes of 5, 7.5, 10, 12.5, 15, 20, 25 and 30°N, respectively. The experiments are identified as F05, F075, and F30, etc. For each experiment, the model was run for 5 days in a resting environment.

3. ResultsThe time evolutions of the center minimum sea level

pressure (CMSLP) and maximum tangential winds (Vmax) in the eight experiments are shown Fig. 1. In all the experi-ments, the initial weak vortex develops into the typhoon strength within 5 days. Note that, however, the timing of rapid intensification depends on the Coriolis parameter. The rapid intensification is referred to as the rapid drop of CMSLP or rapid increase of maximum tangential wind. A TC develops much earlier under lower planetary vorticity

Fig. 1. Time evolution of the center minimum sea level pres-sure (unit: hPa, top panel) and Vmax (unit: ms-1; bottom panel) in F05 (dashed), F075 (red), F10 (green), F125 (solid curve with plus mark), F15 (open circle), F20 (filled circle), F25 (open square) and F30 (filled square).

Page 3: Dependence of tropical cyclone intensification on the ...iprc.soest.hawaii.edu/users/li/www/Lietal2012.pdfRossby deformation, the latter of which may determine the geostrophic adjustment

Li et aL.May 2012 3

environment than under planetary vorticity environment. For instance, the CMSLP starts to drop rapidly shortly af-ter hour 24 in the F05 experiment, and it reaches a quasi-steady state around hour 60. As the Coriolis parameter value increases, the timing of RI is delayed. For instance, the RI of TC in F30 starts after hour 48 and reaches a ma-ture state after hour 84.

a. Structure differences between high and low Coriolis parameter cases

What causes the distinctive difference in the timing of TC intensification among different Coriolis parameter values? To address this question we primarily examine the different development characteristics in two extreme cases (F05 and F30 cases). One of most striking differences is the distinc-tive evolutions of the boundary layer radial inflow. Figure 2 shows the time-radius cross section of the azimuthal mean radial flow at 500 m. Note that the radial inflow sets up much earlier in F05 than in F30. For example, the inflow with a speed of 2 m/s has already set up at r=25 km prior to hour 24, whereas in F30 it does not establish until hour 48.

The establishment of the boundary layer radial flow is crucial for subsequent TC development. The increase of boundary layer inflow leads to the strengthening of moisture convergence into the TC core region (figure not shown), which can further impact the condensational heating in the core region. The strengthening of the radial

inflow accelerates the development of local vorticity (and thus tangential wind) through a vorticity stretching effect. This effect can be clearly seen from maximum wind speed difference between F05 and F30 at hour 12-24 (Fig. 1). For example, at hour 24 the wind speed in F05 is approximate-ly twice as large as that in F30. The wind speed difference may lead to a marked difference in the surface latent heat flux (or surface evaporation), the latter of which further impacts the moisture and condensational heating in the TC core region.

To verify this wind speed effect, we examine the time-radius cross section of the azimuthal mean surface heat flux difference between F05 and F30 (Fig. 3). It is clear that due to the greater surface wind speed, the surface latent heat flux in the TC core region is much greater in F05 than in F30. As discussed by Emanuel (1988), the wind induced surface heat exchange is crucial for rapid TC intensifica-tion.

In addition to affecting the boundary layer moisture con-vergence and surface evaporation, the strengthened radial inflow also enhance the secondary circulation and directly affect vertical motion. The enhanced upward motion may further strengthen upward transport of moisture, leading to enhanced condensational latent heat release.

The increase of heating in the TC core region can further enhance the radial inflow through the decrease of the cen-tral minimum pressure at the surface. Such a positive feed-

Fig. 2. Time-radius cross section of the radial flow (unit: ms-1) at z=500m in F05 (left) and F30 (right).

Page 4: Dependence of tropical cyclone intensification on the ...iprc.soest.hawaii.edu/users/li/www/Lietal2012.pdfRossby deformation, the latter of which may determine the geostrophic adjustment

VoLuMe 24 TROPICAL CyCLONE RESEARCH AND REvIEW

momentum equation, i.e.

where p is the pressure, ρ is the air density and other quantities are as defined above. If F = 0, the tangential flow is in exact gradient wind balance; if F <0, this flow is sub-gradient, which means that there is a tendency to enhance the inflow toward the vortex center; and if F >0, it is super-gradient, which means that there is a tendency to enhance the outflow away from the vortex center.

The radius–height cross-sections of the azimuthal mean net radial force field (F) in the F05 and F30 experiments at hours 45, 60 and 120 are shown in Fig. 5. At hours 45 and 60, the most prominent regions of inflow acceleration indeed occur in the atmospheric boundary layer. The su-pergradient flow emanates inside of and slightly above the subgradient flow (Fig. 5c). During the TC mature phase (at hour 120), the supergradient flow is co-located with strong ascending motion at low level (Fig. 5e-f).

To illustrate distinctive evolution features between F05 and F30, we plotted the time-radius cross section of the azimuthal mean net radial force field at z=500 m (Fig. 6). Note that in F05, the negative F (i.e., subgradient flow) develops much earlier. It becomes evident around hour 24, and continues to strengthen during the intensification period. In contrast, the negative inflow tendency shows a significant delay (by about 24 hours) in F30.

Given the similar evolution characteristics between the radial flow and the net radial force (Figs. 2 and 6), it is like-ly that difference in the radial wind between F05 and F30 is primarily attributed to the extent by which the boundary layer imbalance is triggered. What causes the difference in the boundary layer imbalance between F05 and F30? We argue that the following factors may contribute. Firstly, a lower Coriolis parameter by definition corresponding to a smaller inertial stability may allow low-level inflow pen-etrating closer to the vortex center. Secondly, the strength and depth of the boundary layer inflow associated with the breakdown of the gradient wind balance depends on the Coriolis parameter. It has been shown that the Ekman pumping velocity (or boundary layer convergent flow) is reversely proportional to square root of f (Holton 1992). This means that for a specified cyclonic vorticity at top of the boundary layer, friction induced inflow is stronger when f is smaller. In addition, the Coriolis parameter also controls the depth of the boundary layer. Figure 7 shows the structure difference of the azimuthal mean boundary layer inflow between F05 and F30 during TC intensification and mature phases. Note that the depth of the inflow layer, as inferred from the Vr =-3 ms-1 contour, are much greater in F05 than in F30. For instance, at hour 60, the height of Vr =-3 ms-1 contour in the inner-core region near the RMW reaches 2.5 km in F05, but it is only about 1 km in F30.

Fig. 3. Time-radius cross section of the surface heat flux difference (unit: wm-2) between F05 and F30 (F05 minus F30).

back loop may be inferred from the distinctive structure evolutions of tangential wind and diabatic heating fields between the F05 and F30 cases. Figure 4 illustrates the radial-height cross section of symmetric component of the tangential wind and condensational heating rate during the intensification period. As the vortex intensifies, inner core diabatic heating tends to concentrate towards the vortex center. It is noted that during the intensification period the heating source is located near the radius of maximum wind (RMW) in both the cases, which may warm core develop-ment (vigh and Schubert 2009). In F05, the diabatic heat-ing near the RMW is stronger and develops earlier, com-pared to that in F30. Thus, the evolutions of both the radial and tangential wind components and the heating patterns under different planetary vorticity conditions are consistent with the distinctive CMSLP evolutions shown in Fig. 1.

b. imbalances in the boundary layer flowIn the analysis above, we attribute the distinctive devel-

opment characteristics with different Coriolis parameter to the difference in the boundary layer inflow. A natural ques-tion is what causes the radial wind difference? As we know, initially the radial flow is set to be zero for all the experi-ments. The triggering of the radial flow lies in the break-ing of a gradient wind balance due to surface friction. To measure how differently the boundary layer gradient wind imbalance is excited under different planetary vorticity conditions, we examine a net radial force field, following Smith and vogl (2008). The net radial force field is defined as the residual term of the gradient wind balance, reflecting a difference between the local radial pressure gradient and the sum of the centrifugal and Coriolis forces in the radial

Page 5: Dependence of tropical cyclone intensification on the ...iprc.soest.hawaii.edu/users/li/www/Lietal2012.pdfRossby deformation, the latter of which may determine the geostrophic adjustment

Li et aL.May 2012 5

Fig. 4. Radial-height cross section of symmetric component of tangential wind (contour) and condensational heating rate (shaded) in F05 (left) and F30 (right).

Even in the mature phase at hour 120, the difference in the inflow depth is still clearly seen.

Does the inflow depth difference appear during the initial development stage? Figure 8 shows the time-height cross section of the radial wind at r = 20 km. It shows clearly that the difference of the inflow layer appears even in the initial development period, say, prior to hour 24. The dependence of boundary layer depth on f can be readily interpreted in the context of the Ekman pumping theory (Holton 1992).

The Ekman layer depth may be define as ,

where Az is an eddy viscosity or diffusivity, and f is the

Coriolis parameter. The formula above indicates that a lower Coriolis parameter leads to a deeper Ekman layer.

A stronger, deeper inflow under lower planetary vorticity environment, in turn, brings about larger moisture conver-gence, leading to stronger condensational heating in the TC core region. The stronger heating under lower planetary vorticity environment is inferred from the time-radius cross section of rainfall field (Fig. 9). Note that the rainfall rate in the inner-core region is much greater in F05 than that in F30, particularly during initial development stage, whereas the rainfall rates in F10 and F15 are somehow in between. It is also noted that the radius of the maximum rainfall in F05 is about 12.5 km, which is smaller than that in F30

Page 6: Dependence of tropical cyclone intensification on the ...iprc.soest.hawaii.edu/users/li/www/Lietal2012.pdfRossby deformation, the latter of which may determine the geostrophic adjustment

VoLuMe 26 TROPICAL CyCLONE RESEARCH AND REvIEW

Fig. 5. Radius-height cross section of the net radial force field (ms-2) at hour 45 (top), 60 (middle) and 120 (bottom) from the F05 (left) and F30 (right) experiments. In (e) and (f), the red lines denote the vertical velocity (ms-1) at the same time.

(about 20 km). This confirms the inertial stability effect hypothesis that a TC under lower planetary vorticity en-vironment may have a relatively small eye, due to weaker inertial stability.

In summary, the dependence of timing of TC intensifica-tion on f is primarily attributed to the extent to which the boundary layer imbalance is triggered and maintained. The imbalance directly affects the strength of radial inflow.

Under lower planetary vorticity environment, surface fric-tion induced Ekman flow tends to be stronger and deeper, which results in greater moisture convergence and diabatic heating. The strengthened heating may further enhance the boundary layer inflow. Through this positive feedback loop, a vortex develops faster. Besides, a lower Coriolis param-eter allows greater contraction of the RMW due to weaker inertial stability.

Page 7: Dependence of tropical cyclone intensification on the ...iprc.soest.hawaii.edu/users/li/www/Lietal2012.pdfRossby deformation, the latter of which may determine the geostrophic adjustment

Li et aL.May 2012 7

Fig. 6. Time-radius cross section of the net radial force field (ms-2) in F05 (left) and F30 (right).

Fig. 7. Radial-height cross section of the radial flow (ms-1) field in the lower 3 km in F05 (left) and F30 (right).

Page 8: Dependence of tropical cyclone intensification on the ...iprc.soest.hawaii.edu/users/li/www/Lietal2012.pdfRossby deformation, the latter of which may determine the geostrophic adjustment

VoLuMe 28 TROPICAL CyCLONE RESEARCH AND REvIEW

c. Difference in TC sizeIs the final TC size dependent on the Coriolis parameter?

To address this question, we examine the simulated TC structures at the mature stage. Figure 10 shows the radius-height cross section of the azimuthal mean tangential wind speed at hour 120. As expected, in all the sensitivity experi-ments the maximum tangential wind speed appears at low levels. The TC size as defined by the radius of gale-force wind (approximately 20 ms−1 contour line at the surface) displays a greater dependence on the Coriolis parameter, that is, a TC formed under a higher Coriolis parameter has a larger size than that under a lower Coriolis parameter. For example, the radius of the 20 ms−1 contour at the mature stage is about 100km in F05, whereas it increases to 200km in F30.

To give a physical interpretation on the dependence, we apply the absolute angular momentum conservation concept. The absolute angular momentum is defined as

, where v is the symmetric component of

tangential wind, r is the radius and f is the Coriolis parameter. Based on the definition, for a higher value of the Coriolis parameter, environmental absolute angular momentum should be greater, and vice versa.

Figure 11 shows the radius-height cross section of the symmetric component of the absolute angular momentum

at different development stages. Generally, during the TC intensification period, the absolute angular momentum moves inward (outward) in the lower (upper) troposphere. The contours have a small angle to the horizontal at low-levels where the radial inflow is strong. The contours slope upwards and outwards in the region above the inflow layer where the TC flow is upward and outward, consistent with Smith and Montgomery (2008). In the absence of friction, it is materially conserved above the boundary layer. There-fore, the tangential wind speed will increase in the region where the mean absolute angular momentum isopleths move inwards. As a consequence, larger environmental mean absolute angular momentum results in a larger TC outer size. Compared to F05 (left panels), it is evident that the environmental absolute angular momentum is much greater in F30.

To examine closely the inner core structure, we plotted the azimuthally mean tangential wind speed and vertical velocity at hour 120 within 50 km radius in Fig. 12. The eyewall tilts radially outward with height, especially above z=5 km. Meanwhile, the strong upward motion coincides with the radius of maximum tangential wind. It is noted that the radius of the eyewall is about 12.5 km in F05, whereas it is about 22.5 km in F30. In general, with other environmental conditions fixed, the inner eyewall size de-pends on the Coriolis parameter, with a smaller f leading to a smaller eye. The cause of the dependence is speculated to be related to inertial stability. A smaller f leads to a weaker inertial stability. As a result, the low-level inflow can pen-etrate further into the vortex center region. In addition, TC inner-core size may be sensitive to the radial distribution of the surface entropy flux (Xu and Wang 2010). The mecha-nism responsible for the expansion in vortex outer size was discussed in previous studies (e.g., Ooyama 1969; Wil-loughby 1988, 1995; Smith et al. 2010). However, the inner core process may be independent of the spin-up of the outer circulation, because the boundary-layer inflow is primarily affected by the gradient wind imbalance brought about by the surface friction (Smith and Montgomery 2008). This implies that different dynamic processes may determine the inner eyewall size and the outer size. The TC outer size is largely determined by the outer environmental absolute angular momentum above the boundary layer. The spin-up of the inner core, on the other hand, is associated with the convergence of absolute angular momentum in the bound-ary layer. The convergence may be produced by increasing system-scale radial buoyancy gradients associated with deep, inner-core convection in the presence of enhanced surface moisture fluxes (Bui et al. 2009).

4. Conclusion and discussionIn this study, the dependence of tropical cyclone inten-

sification on latitude at a constant f-plane is examined. By specifying different Coriolis parameters in a resting envi-ronment, we found that vortex intensification under lower

Fig. 8. Time-height cross section of the radial flow (unit: ms-1) field at r=20 km.

Page 9: Dependence of tropical cyclone intensification on the ...iprc.soest.hawaii.edu/users/li/www/Lietal2012.pdfRossby deformation, the latter of which may determine the geostrophic adjustment

Li et aL.May 2012 9

Fig. 9. Time-radius cross section of 3-hourly accumulated rainfall (unit: mm) in F05, F10, F15 and F30 experiments.

Fig. 10. Radial-height cross section of symmetric component of the tangential wind field (unit: ms-1) in F05, F10, F15 and F30 experiments.

Page 10: Dependence of tropical cyclone intensification on the ...iprc.soest.hawaii.edu/users/li/www/Lietal2012.pdfRossby deformation, the latter of which may determine the geostrophic adjustment

VoLuMe 210 TROPICAL CyCLONE RESEARCH AND REvIEW

Fig. 11. Radial-height cross section of symmetric component of absolute angular momentum (unit: 1×105 m2s-1) in F05 (left) and F30 (right) at hour 24, 48, 72 and 120.

planetary vorticity environment is faster than under higher planetary vorticity environment. The distinctive develop-ment behaviors are attributed to the extent to which the boundary layer gradient wind balance is broken. Given an initial balanced vortex, surface friction destroys the gra-dient wind balance, leading to subgradient inflow in the boundary layer. The friction induced inflow is stronger and deeper under lower planetary vorticity environment, which brings about greater moisture convergence and leads to greater condensational heating in the TC core region. The strengthened heating lowers the central surface pressure, which further enhances the radial inflow. Through this posi-tive feedback loop, the vortex spins up at a faster rate under lower planetary vorticity environment.

Besides the distinctive development behavior, the Corio-lis parameter also controls TC size. It is found that the

TC has a larger (smaller) outer size under higher (lower) planetary vorticity environment, which is consistent with previous model results (DeMaria and Pickle 1988) and ob-servations (Merrill 1984). The dependence of the TC outer size on the Coriolis parameters is possibly attributed to the effect of environmental absolute angular momentum. The eye size or the RMW is also found to be f-dependent. it is smaller (larger) under lower (higher) planetary vorticity en-vironment. This dependence is possibly related the effect of inertial stability, which regulates the extent to which low-level inflow can penetrate into the core region.

In this study, we mainly focus on the timing of TC inten-sification and associated structural change. The dependence of maximum intensity on the environmental planetary vorticity, on the other hand, is another important issue. Our sensitivity numerical experiments show that at the final

Page 11: Dependence of tropical cyclone intensification on the ...iprc.soest.hawaii.edu/users/li/www/Lietal2012.pdfRossby deformation, the latter of which may determine the geostrophic adjustment

Li et aL.May 2012 11

equilibrium state, maximum TC intensity appears in the F075 and F10 experiments (i.e., red and green curves in Fig. 1). This indicates that the maximum intensity appears at intermediate background rotation environment. This suggests an optimal background rotation rate for the oc-currence of maximum TC intensity. The result is consistent with Smith et al. (2010) from an axisymmetric version of the minimal three-level hurricane model. However, in that study the maximum intensity is not a function of f beyond 15oN. The dynamical cause of this optimal latitude selec-tion for TC maximum intensity is unclear at the moment. This topic requires further investigation.

AcknowledgementThis work was supported by ONR Grants N000140810256

and N000141010774, and by International Pacific Research Center that is partially sponsored by the Japan Agency for Marine-Earth Science and Technology (JAMSTEC), NASA (NNX07AG53G) and NOAA (NA17RJ1230). This is SOEST contribution number 8687 and IPRC contribution number 893.

ReferenceBister, M., 2001: Effect of Peripheral Convection on Tropical Cy-

clone Formation. J. Atmos. Sci, 58, 3463-3476. Bui, H. B., R. K. Smith, M. T. Montgomery, and J. Peng, 2009:

Balanced and unbalanced aspects of tropical-cyclone intensifi-

cation. Q. J. R. Meteorol. Soc., 135, 1715-1731.DeMaria, M., and J. D. Pickle, 1988: A simplified system of equa-

tions for simulation of tropical cyclones. J. Atmos. Sci, 45, 1542–1554.

Emanuel, K. A., 2005: Genesis and maintenance of “Mediterranean hurricanes.” Advances in Geosciences, 2, 217-220.

Fairall, C. W., E. F. Bradley, D. P. Rogers, J. B. Edson, and G. S. young, 1996: Bulk parameterisation of air–sea fluxes for Tropical Ocean–Global Atmosphere Coupled Ocean Atmo-sphere Response Experiment. J. Geophys. Res., 101 (C), 3747–3764.

Ge, X. y, T. Li, and S. Peng, 2012: Tropical cyclone genesis effi-ciency: Mid-level versus bottom vortex. J. Tropical Meteorol-ogy, in press.

Ge, X., T. Li, and X. Zhou, 2007: Tropical cyclone energy disper-sion under vertical shears. Geophys. Res. Lett., 34, L23807, doi:10.1029/2007GL031867.

Ge, X., T. Li, y. Wang, and M. Peng, 2008: Tropical Cyclone En-ergy Dispersion in a Three-Dimensional Primitive Equation Model: Upper Tropospheric Influence. J. Atmos.Sci., 65 (7), 2272–2289.

Gray, W. M., 1968: Global view of the origin of tropical distur-bances and storms. Mon. Wea. Rev., 96, 669-700.

Holland, G. J., 1997: Maximum potential intensity of tropical cy-clones. J. Atmos. Sci., 54, 2519–2541.

Holton, J. R., 1992: An Introduction to Dynamic Meteorology. Academic Press, 511 pp.

Langland, R. H., and C.-S. Liou, 1996: Implementation of an E–e parameterization of vertical subgrid-scale mixing in a regional model. Mon. Wea. Rev., 124, 905–918.

Li, T., 2012: Synoptic and climatic aspects of tropical cyclogene-sis in western North Pacific. in Cyclones: Formation, triggers

Fig. 12. Radial-height cross section of symmetric component of tangential velocity (contour; ms-1) and vertical velocity (shading; ms-1) in (a) F05, (b) F10, (c) F15, and (d) F30.

Page 12: Dependence of tropical cyclone intensification on the ...iprc.soest.hawaii.edu/users/li/www/Lietal2012.pdfRossby deformation, the latter of which may determine the geostrophic adjustment

VoLuMe 212 TROPICAL CyCLONE RESEARCH AND REvIEW

and control, edited by K. Oouchi and H. Fudevasu, Noval Sci-ence Publishers, in press.

Li, T., B. Fu, X. Ge, B. Wang, M. Peng, 2003: Satellite data analy-sis and numerical simulation of tropical cyclone formation. Geophy. Res. Let., 2122-2126.

Li, T., X. Ge, B. Wang, and y. Zhu, 2006: Tropical cyclogenesis associated with Rossby wave energy dispersion of a pre-existing typhoon. Part II: Numerical simulations. J.Atmos.Sci., 63, 1390–1409.

Lin, y.-L., R. D. Rarley, and H. D. Orville, 1983: Bulk parameter-ization of the snow field in a cloud model. J. Appl. Meteor., 22, 1065-1092.

Merrill, RT. 1984 A comparison of large and small tropical cy-clones. Mon. Wea. Rev., 112, 1408-1418.

Nolan, David S., 2007: What is the trigger for tropical cyclogen-esis? Aust. Meteorol. Mag., 56 , 241-266.

Ooyama, K. v., 1969: Numerical simulation of the life cycle of tropical cyclones. J. atmos. Sci., 26, 3–40.

Schubert, W. H., and J. J. Hack, 1982: Inertial stability and tropi-cal cyclone development. J. Atmos. Sci., 39, 1687–1697.

Smith, R. K, and M. T. Montgomery, 2008: Balanced bound-ary layers in hurricane models. Q. J. R. Meteorol. Soc., 134, 1385–1395.

Smith, R. K, and S. vogl, 2008: A simple model of the hurricane boundary layer revisited. Q. J. R. Meteorol. Soc., 134, 337–351.

Smith, R. K., M. T. Montgomery, and C. Schmidt, 2010: Dynami-cal constraints on the intensity and size of tropical cyclones. Quart. J. Roy Met. Soc., in press.

vigh, J. L., and W. H. Schubert, 2009: Rapid development of the tropical cyclone warm core. J. Atmos. Sci., 66, 3335-3350.

Wang, B., and X. Zhou, 2008: Climate variability and predictabil-ity of rapid intensification in tropical cyclones in the western North Pacific, Meteor. Atmos. Phys., doi: 10.1007/s00703-006-0238-z, 1-16.

Wang y., 2001: An explicit simulation of tropical cyclones with a triply nested movable mesh primitive equation model: TCM3. Part I: Model description and control experiment. Mon. Wea. Rev., 129, 1370–1394.

Wang, y. 2002: vortex Rossby waves in a numerically simulated tropical cyclone. Part I: Overall structure, potential vorticity and kinetic energy budgets. J. Atmos. Sci.,59, 1213–1238.

Willoughby, H. E., 1988: The dynamics of the tropical cyclone core. Aust. Meteorol. Mag., 36, 183–191.

Willoughby, H. E., 1995: Mature structure and evolution. Global perspectives on tropical cyclones. WMO/TD-No.693, edited by Elsberry, R. L., World Meteorological Organization: Ge-neva, 21–62.

Xu, J., and y. Wang, 2010: Sensitivity of tropical cyclone inner core size and intensity to the radial distribution of surface en-tropy flux. J. Atmos. Sci., in press.