1. Introduction
The potential vorticity conservation principle provides a basis for understanding midlatitude weather systems, both through balanced models and primitive equation models. In many tropical weather systems, such as the ITCZ and tropical cyclones, the release of latent heat plays a crucial role, so that potential vorticity is not materially conserved. However, even in these cases, the relevant dynamics is “slow manifold dynamics,” with adjusted wind and mass fields intimately related to the evolving potential vorticity field. Thus, the analysis of tropical flows in terms of potential vorticity dynamics yields insights into such processes as ITCZ breakdown, the formation of easterly waves, and into the extreme inner core structures of tropical cyclones. The primary purpose of the present paper is to study the inner core potential vorticity structure of tropical cyclones simulated with a high resolution, axisymmetric, nonhydrostatic tropical cyclone model whose thermodynamic/dynamic foundations and discretizations are based on the work of Ooyama (1990, 2001, 2002, hereafter O90, O1, O2, respectively). The model has several unique features: 1) a very accurate treatment of moist processes within the context of equilibrium thermodynamics, with a simple switch to include or exclude the effects of ice; 2) an associated potential vorticity (PV) principle and an invertibility principle, both exactly derivable from the original model equations; 3) the inclusion of the thermodynamic and dynamic effects of precipitation, in particular the vertical transport of entropy and momentum by precipitation; 4) spatial numerics based on the cubic spline transform (CST) method, which results in small computational dispersion errors and noise-free nesting.
We begin in section 2 by presenting a concise, self-contained description of the model. In section 3 and appendix B we derive the potential vorticity and invertibility principles associated with our cloudy, precipitating model atmosphere. The moist generalization of the dry Ertel potential vorticity turns out to be P = ρ−1ζ · ∇θρ, where ρ is the total density of moist air, ζ is the absolute vorticity vector, and θρ is the virtual potential temperature. In the axisymmetric case, the vorticity vector has components in the vertical, radial, and azimuthal directions, but ∇θρ has components in the vertical and radial directions only. As a consequence, the azimuthal component of ζ is lost in performing the product ζ · ∇θρ, and the potential vorticity simplifies to P = ρ−1 {(−∂υ/∂z) (∂θρ/∂r) + [ f + ∂(rυ)/r∂r] (∂θρ/∂z)}, where υ is the azimuthal component of the wind. Since the evolving hurricane remains close to a state of gradient and hydrostatic balance, υ and ρ can be expressed in terms of the pressure p. Then, since θρ is a function of ρ and p only, the potential vorticity can be expressed solely in terms of the pressure. In other words, the P field contains all the required information about the balanced part of the wind and mass fields, and the invertibility principle, which determines the pressure from the potential vorticity, is an elliptic partial differential equation in the radial-vertical plane. Understanding the evolution of the P field is thus a crucial part of understanding the whole intensification problem.
In section 4 we present the results of the control experiment, paying particular attention to the potential vorticity dynamics. With 500-m horizontal and vertical resolution in the inner core, a remarkable, hollow tower potential vorticity structure emerges in the quasi–steady state, with values of potential vorticity as high as 275 PV units (PVU; where 1 PVU = 1.0 × 10−6 m2 s−1 K kg−1) in the eyewall. This is in sharp contrast to the “zero PV picture” that emerges when equivalent potential temperature or saturation equivalent potential temperature is used as the thermodynamic variable in the definition of potential vorticity.
Section 5 contains a discussion of selected sensitivity experiments, including the important effects of ice and the effects of precipitation terms in the entropy and momentum budgets. Concluding remarks are given in section 6.
2. Model
Consider atmospheric matter to consist of dry air, airborne moisture, and precipitation, with respective densities ρa, ρm, ρr, so that the total density is given by ρ = ρa + ρm + ρr.1 The mass density of airborne moisture is the sum of the densities of water vapor and airborne condensate, so that ρm = ρυ + ρc. However, the partition of ρm into ρυ and ρc is not considered in the prognostic stage, but only later in the diagnostic stage. In cylindrical coordinates, with the assumption of axisymmetry, the mass conservation law for dry air is ∂ρa/∂t + ∂(ρaru)/r∂r + ∂(ρaw)/∂z = 0, the advective form of which is (2.1). Similarly, the conservation law for ρr is ∂ρr/∂t + ∂(ρrru)/r∂r + ∂[ρr(w + W)]/∂z = Qr, where w + W is the vertical velocity of the precipitation (so that W is the fall velocity of the precipitation relative to the dry air and airborne moisture) and Qr is the rate of conversion from airborne moisture to precipitation. Defining the precipitation mixing ratio as μr = ρr/ρa, the conservation law for ρr can be combined with (2.1) and written in the advective form (2.3). Finally, the conservation law for the airborne moisture is ∂ρm/∂t + ∂(ρmru)/r∂r + ∂(ρmw + Fm)/∂z = −Qr, where Fm is the boundary layer turbulent flux of water vapor. Adding this conservation law for ρm to the conservation law for ρr and converting the result into an advective form for the total water mixing ratio μ = (ρm + ρr)/ρa, we obtain (2.2). Note that, although there are four types of matter (with densities ρa, ρυ, ρc, ρr), there are only three prognostic equations, (2.1)–(2.3), for the distribution of mass. As we shall see, the separation of the airborne moisture density ρm into the vapor density ρυ and the airborne condensate density ρc will be accomplished diagnostically in the two alternatives of (2.14).
The total entropy density is σ = σa + σm + σr, consisting of the sum of the entropy densities of dry air, airborne moisture, and precipitation. Since the vertical entropy flux is given by σaw + σmw + σr(w + W) + ρaFs = σw + σrW + ρaFs, we can write the flux form of the entropy conservation principle as ∂σ/∂t + ∂(σru)/r∂r + ∂(σw + σrW + ρaFs)/∂z = 0, where Fs denotes the turbulent eddy flux in the atmospheric boundary layer, and where radiative effects have been neglected. Defining s = σ/ρa as the “dry-air-specific” entropy of moist air, we can combine (2.1) with the above entropy conservation principle to obtain (2.4).
We next consider the momentum equations. Defining the absolute angular momentum per unit mass as m = rυ + ½fr 2, we can write the absolute angular momentum budget as ∂(ρm)/∂t + ∂(ρmru)/r∂r + ∂[(ρw + ρrW + Fm)m]/∂z = −∂(ρarFυ)/∂z, where Fυ denotes the turbulent eddy flux of υ in the boundary layer. With the aid of the continuity equation for total density, this absolute angular momentum budget can be written as (2.6). In a similar fashion we can derive the radial wind Eq. (2.5), where p = pa + pυ is the sum of the partial pressures of dry air and water vapor. The derivation of the vertical equation of motion is somewhat more complicated. While all the matter moves with the same horizontal velocity, the vertical velocity of dry air and airborne moisture is w, while the vertical velocity of precipitation is w + W. The prediction of both w and w + W is equivalent to predicting W, which is inconsistent with the diagnostic treatment of W through parameterized cloud microphysics. A solution to this problem, proposed in O1, is to write a single budget equation for the total vertical momentum ρw + ρrW, and then approximate this equation by neglecting the material derivative of W along the precipitation path, that is, by neglecting ∂W/∂t + u(∂W/∂r) + (w + W)(∂W/∂z). With this approximation our vertical momentum equation becomes (2.7). The neglect of the material derivative of W along the precipitation path is consistent with the parameterization Eq. (2.18), which gives a slow variation of W because of the small fractional power of ρr and the inverse square root of ρa. The largest errors due to this assumption are expected in a small region near the melting level, where the ice factor fice results in an increase of the fall velocity of the precipitation.
























































The domain extends from the sea surface to a height of 24 km and from the vortex center to a radius of 1536 km. The outer boundary is open, and the solutions are assumed to decay exponentially with an e-folding distance of 1400 km. The top and bottom boundaries are assumed rigid, with the boundary condition w = 0. To reduce the effect of the reflection of gravity waves off the top boundary, we have also included Rayleigh-type damping terms in (2.4)–(2.7). These terms damp the winds to zero and the specific entropy to its background value. The damping coefficient vanishes for z ≤ 18 km and increases linearly with height from 18 km to a maximum value of 0.015 s−1 at 24 km. Thus, the region between 18 km and the lid at 24 km is a sponge layer that effectively damps vertically propagating gravity waves that would otherwise reenter the region below 18 km after falsely reflecting off the lid. In all the cross sections of the present paper, only the region below 18 km is displayed.
The model spatial numerics are based on the CST method described in O2. As the name implies, the CST method uses the cubic B spline as the basis function. Because the first two derivatives of the B spline are continuous, the CST method has small computational dispersion errors, similar to the Fourier spectral method. Yet, because the B spline is locally defined, the CST method allows flexibility with regard to boundary conditions. With reduced dispersion errors and flexible boundary conditions, the CST method provides for noise-free nesting. To take advantage of this, the domain is discretized into a series of nested grids. As illustrated in Fig. 1, the horizontal domain consists of six grids. The horizontal grid spacing Δr within each grid increases by a factor of 2 from the finest grid at 0.5 km to the coarsest grid at 16.0 km. The vertical domain, in contrast, consists of a single grid with 48 grid intervals and a grid spacing Δz of 0.5 km. Time integration is accomplished in a two-stage process. In the first stage all the prognostic variables are advanced explicitly using the leapfrog scheme with a small enough time step for stability of gravity waves. In the second stage these explicit predictions are implicitly adjusted via a second-order elliptic equation. This particular implementation of the semi-implicit method allows a tenfold increase of the time step. The time steps are 2.5 s on the finest grid, 5 s on the next two grids, and 10 s on the coarsest three grids.
3. Potential vorticity equation and invertibility principle


The variable P has two important properties that contribute to its fundamental importance (Schubert 2004): (i) it reduces to the classical Ertel PV in the limit of a completely dry atmosphere; (ii) it is invertible, that is, it carries all the dynamical information about the balanced part of the wind and mass fields. Other choices for the scalar field in the definition of PV are lacking in this regard. For example, if saturation equivalent potential temperature is used in place of θρ, property (i) is lost, while if equivalent potential temperature is used in place of θρ, property (ii) is lost.








In this paper we will not be concerned with actually solving the invertibility principle (3.2)–(3.5). Rather, we simply use the existence of the principle as justification for studying model output fields of P. In section 4 we will use a transformed version of (3.1) to understand the PV dynamics of a quasi-steady tropical cyclone. When attempting to obtain physical understanding of certain phenomena, it has been said (Stommel 1995) that complex models such as (2.1)–(2.27) are “not much help: like vegetable soup, they have too many ingredients to reveal which one imparts the flavor.” In the case of a tropical cyclone, it is definitely the PV that imparts the flavor.
4. Control experiment
The initial condition for all experiments is a purely azimuthal vortex in gradient and hydrostatic balance, with a horizontally uniform relative humidity field. At the surface the radial distribution of initial azimuthal velocity is 2υm(r/rm)/[1 + (r/rm)2], with the maximum wind specified as υm = 12 m s−1 and the radius of maximum wind as rm = 100 km. The initial azimuthal velocity is assumed to decrease linearly with height to zero at z = 18 km, and to be zero between 18 km and the model top at 24 km. Figure 1 shows this initial azimuthal wind field. The associated pressure anomaly field has a minimum of −7.1 hPa at the surface in the vortex core. The far-field temperature and humidity is taken from Jordan’s (1958) mean hurricane season sounding and is shown in Fig. 2. At the initial time this sounding is approximately valid at all radii, since the initial horizontal temperature gradient is weak, with the lower tropospheric vortex core approximately 1.1 K warmer than the far field.
To summarize the temporal and spatial evolution of this control (CNTL) experiment, Fig. 3 depicts the 3-h running mean surface tangential wind speed as a function of radius and time. Between 40 and 50 h, a very small, intense central vortex develops. As discussed in appendix C, this small, central vortex is not the result of a problem with CST numerics. In fact such small, central vortices also appear in previous high resolution axisymmetric tropical cyclone simulations (e.g., Yamasaki 1983; Willoughby et al. 1984) that use entirely different numerics. Although strong cumulus convection (referred to as a hub cloud) is sometimes observed in the center of a hurricane eye, the convection appearing near r = 0 in our CNTL experiment is unrealistically deep and intense. Beyond 80 h, however, this central vortex weakens substantially, while a secondary tangential wind maximum forms at about 30 km and propagates inward, intensifying along the way. This wind maximum is collocated with a ring of convection that eventually contracts to become the new tropical cyclone eyewall. From 100 to 180 h, the tropical cyclone gradually intensifies. As shown in Fig. 3, this overall increase of intensity is not steady but highly variable. For instance, during the 12-h period following 118 h, the maximum tangential wind oscillates over 20 m s−1. This variability, which is very similar to that obtained by Willoughby et al. (1984), is caused by disturbances of the boundary layer inflow and the formation of secondary rings of convection that propagate inward. The secondary circulation of a new, inward-propagating ring and the restriction of high θe inflow into the core, results in the dissipation of the primary ring, and its ultimate replacement by the secondary ring.
Beyond 180 h, the vortex settles into a quasi steady state. Figures 4a–c depict the 180–240-h average primary and secondary circulations. The model steady-state cyclone is similar in structure to observed storms but is more intense. The tangential wind has a maximum of over 90 m s−1, with the peak located beneath the eyewall at a radius of 14 km. This intense vortex is produced by the advection of high angular momentum air toward the center. This same air is later advected away from the center within the outflow branch of the secondary circulation, producing an anticyclonic vortex with a wind speed of −30 m s−1 at a radius of 1250 km and a height of 14 km. The horizontal branches of the secondary circulation are concentrated into shallow layers at the surface and near the tropopause. As shown in Fig. 4b, the inflow branch is confined3 below 2 km, with a maximum inward radial flow of 31.8 m s−1 at r = 21 km. The maximum inflow lies just outside the radius of maximum tangential wind. The most intense outflow is confined to a layer between 12 and 18 km. The maximum outflow of 29.5 m s−1 is located at a radius of about 75 km, which is more than 50 km away from the eyewall. The two vertical branches of the secondary circulation are very different in terms of intensity and horizontal scale. Immediately above and sloping away from the region of maximum surface convergence is the ascending branch of the secondary circulation, embedded within the eyewall cloud. In the ascending branch, vertical velocities reach 5.5 m s−1 just above the boundary layer and in the upper troposphere. The width of the ascending branch increases with height. In contrast, the compensating subsidence in the descending branch of the secondary circulation is very weak and extends into the far field of the domain. The transverse circulation depicted in Fig. 4 provides a deformation field in the (r, z) plane, with especially large deformation in the lower troposphere at radii between 10 and 20 km. This frontogenetic effect (Eliassen 1959; Emanuel 1997) acts to crowd together the absolute angular momentum surfaces, resulting in large vorticity.
Figure 5 shows the 180–240-h mean cross sections of T and the water vapor mixing ratio μυ. The maximum T ′, disregarding the effect of the very small central vortex, is approximately 17 K and is located at a height of 12 km. The axis of maximum T ′ extends both outward into the stratiform precipitation and downward along the inner edge of the eyewall. The horizontal extent of these temperature anomalies depends on the local radius of deformation, defined by (gH)1/2{[ f + ∂(rυ)/r∂r]( f + 2υ/r)}−1/2, where H is the equivalent depth and where the factor in the denominator is a measure of the inertial stability or “stiffness” of the vortex. When the radius of deformation is small, the vortex is stiffened such that the horizontal extent of the secondary circulation is restricted. Assuming (gH)1/2 ≈ 60 m s−1, the radius of deformation is less than 5 km along the inner edge of the eyewall; however, within the outflow, it is larger than 300 km. Therefore, we observe a narrow secondary circulation and adiabatic warming on the inner edge of the eyewall, as compared to a broad secondary circulation outside the eyewall.
The dynamics of the eye and eyewall also have a distinct influence on the distribution of water vapor, as shown in Fig. 5b. Radially, the water vapor mixing ratio μυ is a maximum in the eyewall, because of vertical advection within the updraft. Because of subsidence within the eye, μυ is a minimum along the inside of the eyewall. At the surface, the large flux of water vapor increases μυ outside 25 km to 23.9 g kg−1, which is 5.5 g kg−1 greater than the initial value. The corresponding surface relative humidity is nearly 100%. Beneath the eyewall, downdrafts produce a local surface minimum of 21.6 g kg−1 at a radius of 14 km. Above the 0°C isotherm, the terminal velocity of precipitation is approximately 1.5–2 m s−1, whereas below this level, it increases to as much as 8 m s−1 within the eyewall. Because of the relatively small terminal velocity aloft and the intense outflow, precipitation is advected far from the eyewall, resulting in surface precipitation rates of about 10–40 mm h−1. However, beneath the eyewall, W and μr are both large, producing a precipitation rate of over 200 mm h−1.
Figure 6 shows 180–240-h mean cross sections of the vertical component of the absolute vorticity ζ, the virtual potential temperature θρ, and the potential vorticity anomaly P′ = P −












Since the P field determines the primary circulation and the θ̇ρ field, along with frictional effects, determine the secondary circulation, (4.6) can be interpreted as the fundamental relation between the primary and secondary parts of the steady-state circulation.
In the steady state, a parcel of eyewall air erupting from the boundary layer on a given angular momentum surface stays on this same outward-sloping surface as it spirals upward, crossing θρ surfaces and changing its PV at the rate P(∂θ̇ρ/∂θρ), which is positive below the level of maximum θ̇ρ and negative above this level. Note that the level of maximum θ̇ρ is also the level of maximum P through the coupling described by (4.6). All parcels erupting from the boundary layer on the same angular momentum surface have the same Lagrangian history, but the Lagrangian histories are generally distinct on different angular momentum surfaces. Unfortunately, when applied to the CNTL experiment, the above argument is compromised by the unsteadiness illustrated in Fig. 3. However, as we shall see, the argument more accurately holds for the “no ice experiment” discussed in section 5.
It is interesting to note that the generalized moist PV, defined by (3.5), is only a modest modification of the dry Ertel PV, which is obtained by replacing ρ with ρa and θρ with θ in (3.5). This modest modification is in sharp contrast to other proposed PV generalizations that involve use of the equivalent potential temperature or the saturation equivalent potential temperature, both of which result in drastic differences with the dry Ertel PV. To confirm that (3.5) yields PV fields that are very similar to dry Ertel PV fields, we have produced a 180–240-h mean cross section of the dry Ertel PV. This dry PV cross section (not shown) is nearly identical to Fig. 6c. Thus, the dry Ertel PV is useful for diagnostic analysis of moist models (e.g., Yau et al. 2004; Braun et al. 2006). In addition, although a hurricane is definitely a phenomenon involving moist physics, a reasonable approximation to the balanced dynamics can be constructed (e.g., Schubert and Alworth 1987; Möller and Smith 1994) using the dry PV invertibility principle and the dry PV evolution equation, as long as diabatic and frictional effects are properly included in the latter equation.
5. Sensitivity experiments
a. Sensitivity to ice microphysics
In our representation of thermodynamics and microphysics there are two effects of ice: (i) the latent heat effects of ice are incorporated by computing E(T), from which L(T), C(T), D(T), ρ*υ(T) are determined, as a temperature-dependent interpolation of the saturation vapor pressure over a plane surface of water and a plane surface of ice; (ii) the reduced settling speed of geometrically complex ice particles is included in the model through the temperature-dependent ice factor defined by (2.16). Remembering that our CNTL experiment included both these effects of ice, we now perform a sensitivity experiment with no latent heat of fusion and with the ice factor set to unity.4
Figure 7 shows the 3-h-averaged surface tangential wind as a function of r, t for this “no ice” (NICE) experiment, while Figs. 8 –10 show the quasi-steady-state structure in the (r, z) plane. The formats of Figs. 8 –10 are identical to Figs. 4 –6 except that the fields are averaged from 120 to 240 h. Comparing Fig. 7 with Fig. 3, we see that in the NICE experiment the simulated tropical cyclone develops more rapidly and attains a more intense steady state than in the CNTL experiment. In addition the NICE experiment is distinctly less variable. Comparison of Figs. 8 –10 with Figs. 4 –6 indicates that the NICE experiment has stronger tangential winds, a smaller radius of maximum tangential winds, a low-level inflow penetrating further inward, a narrower and more intense updraft, a warmer and dryer central core, a thinner annular ring of eyewall cloud, and a narrower and more intense outward tilting region of maximum ζ and P. In particular, note that the peak values of P in the NICE experiment are approximately 400 PVU, which is even more extreme than those found in the CNTL experiment. Note also that our 0.5 km by 0.5 km resolution on the inner grid is required to accurately capture these extreme features of the ζ and P fields. These differences between the CNTL and NICE experiments are caused by the development and contraction of secondary eyewalls in the CNTL experiment, features that do not develop in the NICE experiment. The reason that secondary eyewalls do not form in the NICE experiment is that extensive stratiform precipitation does not develop. Above 5 km the downward terminal velocity of precipitation reaches 10 m s−1 in the NICE experiment and only 2 m s−1 in the CNTL experiment. Even though the CNTL experiment has a weaker time-averaged secondary circulation than the NICE experiment, the small terminal velocities in the CNTL experiment allow the precipitation to be lofted into the outflow and advected far from the eyewall. However, in the NICE experiment, much of the precipitation falls from the sloping updraft without being ejected into the outflow. As a result, the surface precipitation extends to almost 100 km in the CNTL experiment but only about 45 km in the NICE experiment. Furthermore, since the precipitation is distributed over a smaller area in the NICE experiment, the precipitation rate is a factor of four greater than in the CNTL experiment. Without the stratiform precipitation to provide the mesoscale downdrafts that induce surface convergence and ascent, secondary eyewalls do not form in the NICE experiment.
These differences between the CNTL and NICE experiments are consistent with the results reported by Willoughby et al. (1984), Lord et al. (1984), and Lord and Lord (1988) using a nonhydrostatic moist model with 1-km vertical grid spacing and 2-km radial grid spacing in the inner core, but with a more elaborate microphysical parameterization. To compare our model with the Willoughby–Lord model, it is useful to measure the complexity of a nonhydrostatic moist model by the number of prognostic equations over and above the five required for a dry model. For the nonhydrostatic moist model (2.1)–(2.25) the count is two, that is, (2.2) for the total water mixing ratio μ and (2.3) for the precipitation mixing ratio μr. In fact, this nonhydrostatic moist model can be considered the model of maximum simplicity among the class of models that explicitly calculate the movement of solid and liquid precipitation from its formation to its impact with the earth’s surface. Contributing to the simplicity of this model is the fact that the distribution of precipitation is described by a single scalar field μr, even though the precipitation can be solid or liquid. In contrast, the count for the Willoughby–Lord model is six, one for each of the mixing ratios of water vapor, cloud water, cloud ice, rain, snow, and graupel. With the increased number of prognostic equations comes an increased number of conversion processes. For example, with the model of maximum simplicity, there are only three parameterized conversions, Qauto, Qcol, Qevap, as given by (2.19)–(2.21). In the Willoughby–Lord model the number of parameterized conversions increases to twenty-four, which requires a more extensive theoretical and observational basis for the microphysical parameterization. Thus, it is interesting that very similar results to those of obtained by the Willoughby–Lord model can be obtained with a microphysical model of maximum simplicity.
In a second sensitivity experiment (not shown) we included the latent heat effect of ice but set fice to unity. This yields results qualitatively similar to the NICE experiment. In both of these experiments with fice = 1, the precipitation is not advected away from the eyewall but rapidly falls from the sloping updraft. Without extensive stratiform precipitation and the resulting horizontal dipole of freezing and melting, secondary eyewalls do not form; thus, the latent effects of ice alone cannot explain the significant differences between the NICE and CNTL experiments. In a third sensitivity experiment (not shown) we excluded the latent heat effect of ice but left the ice factor as defined by (2.16). This experiment yields results qualitatively similar to the CNTL experiment. Thus, our results indicate that the reduced terminal fall velocity associated with frozen precipitation is the primary factor responsible for the qualitative differences between the CNTL and NICE experiments. The latent effects of ice are a secondary effect. Thus, using a simple microphysical scheme, we have obtained results consistent with those obtained using more sophisticated parameterizations (Willoughby et al. 1984; Lord et al. 1984; Lord and Lord 1988).
b. Sensitivity to precipitation effects
One distinctive feature of our model is the inclusion in (2.4)–(2.7) of terms representing the vertical fluxes of entropy and momentum by precipitation. To evaluate the sensitivity of simulated tropical cyclones to these vertical fluxes by precipitation, we have performed two additional experiments, the first of which includes the W term in (2.4) but excludes the W terms in (2.5)–(2.7). The results of this experiment (not shown) demonstrate that the precipitation terms in the momentum equations have relatively little impact on tropical cyclone development. This result is not unexpected considering that the momentum of the precipitation, ρrW, is typically much smaller than ρw.
The final sensitivity experiment excludes the W term in (2.4) but includes the W terms in (2.5)–(2.7). The results of this experiment (not shown) differ substantially from the CNTL. For instance, the tropical cyclone intensifies more rapidly and attains a more intense steady state than the CNTL experiment. These differences result from changes in the boundary layer entropy, with the θe of inflowing air being 10 K warmer beneath the eyewall compared to the CNTL experiment. With higher θe in the boundary layer, the development of secondary eyewalls is inhibited. We conclude that the vertical flux of entropy by precipitation is an important effect that should be included in tropical cyclone models. Perhaps this is not surprising, since it is well known that accurate simulation of the boundary layer moist entropy budget is an important part of the tropical cyclone forecasting problem.
6. Concluding remarks
We have presented results from an axisymmetric, nonhydrostatic, full-physics model of the tropical cyclone and have used the moist potential vorticity principle as a diagnostic for interpreting the quasi-steady-state structure. This analysis demonstrates how the P and θ̇ρ fields become locked together in a thin leaning tower on the inner edge of the eyewall cloud. In addition, our sensitivity experiments indicate that accurate simulations of tropical cyclones require 1) the inclusion of ice and, in particular, the slow terminal velocities associated with frozen precipitation above the melting level; 2) the inclusion of the vertical transport of moist entropy by precipitation.
Since the numerical model used in this study is based on the equation set (2.1)–(2.25), it is interesting to comment on the usefulness of these equations as a physical model of a tropical cyclone. We certainly have the most confidence in the prognostic Eqs. (2.1)–(2.7) and the equilibrium thermodynamics (2.8)–(2.15) as being part of an accurate description of nature. Although precipitation microphysics has been parameterized as a bulk process, (2.16)–(2.18) and (2.21) have considerable observational and laboratory support (e.g., see Kessler 1969). However, (2.19) and (2.20) must be regarded as a crude parameterization of the intricate process of precipitation formation. As for the air–sea interaction parameterization (2.23)–(2.25), the most uncertain aspect is (2.25), especially the values of CD and CH at high wind speeds. Finally, a major limitation of the model (2.1)–(2.25) is the assumption of axisymmetry.
Consistent with the results of Persing and Montgomery (2003), our results show that the quasi-steady-state intensity for an axisymmetric storm may be much greater than predicted by energetically based, steady-state maximum potential intensity (MPI) theories. To explain this discrepancy, Persing and Montgomery have used the nonhydrostatic model developed by Rotunno and Emanuel (1987) to show that past model simulations supporting such MPI theories may have been limited by their coarse resolution and large diffusion—problems that have been overcome in both the Persing–Montgomery simulations and those presented here. At the same time it should be realized that high resolution, low diffusion, axisymmetric models probably produce overly intense vortices because of the suppression of variations in the azimuthal direction. In fact, the PV distributions produced in such models would be unstable in a combined baroclinic–barotropic sense if the models allowed variations in the azimuthal direction. Over their life cycle such instabilities would radially mix the PV, thereby reducing the maximum tangential wind (Schubert et al. 1999). In addition, although difficult to observe and thus inadequately documented, Kelvin–Helmholtz instability along the sloping inner edge of the eyewall may also play a role in limiting the maximum tangential winds.
Although the PV mixing process can be crudely parameterized in an axisymmetric model by using radial diffusion or hyperdiffusion, there are unrealistic aspects associated with this type of parameterization (Kossin and Schubert 2003). In this regard it should be recalled that our model has no explicit frictional effects above the boundary layer, although the CST method does include a third-order derivative constraint in the least squares minimization that defines the transform of spatial fields to nodal amplitudes. This third order derivative constraint is equivalent, in wavenumber space, to a sharp, sixth-order, low-pass filter that effectively eliminates small-scale errors at the resolution limit. This filter is weak compared to the usual diffusion or hyperdiffusion used in most finite difference models, so that the model flow above the boundary layer should be regarded as quite inviscid. Thus, the assumption of axisymmetry, in conjunction with the nearly inviscid nature of the flow above the boundary layer, probably leads to the overly intense vortices produced by the model.
Even with the above limitations in mind, our results help answer the question, “What is a quasi-steady-state hurricane?” It is no doubt a complicated structure involving all three flow components and many moist thermodynamic fields. But, at its core, it is an extreme structure in which the P field and the θ̇ρ field have become intimately coupled in such a way that they vary in a similar fashion along a tightly packed group of absolute angular momentum surfaces. To capture the formation and asymmetric evolution of such extreme PV structures in full-physics 3D models is obviously a very challenging problem, but it may be necessary for accurate intensity forecasts. Perhaps this is part of a sobering realization that hurricane intensity forecasting is fundamentally much more difficult than hurricane track forecasting.
Acknowledgments
The authors thank Paul Ciesielski, William Cotton, Rob Fleishauer, Matthew Garcia, Richard Johnson, James Kossin, John McGinley, Brian McNoldy, Michael Montgomery, John Persing, Richard Taft, Jonathan Vigh, and three anonymous reviewers for helpful comments. This work was supported by NASA/CAMEX Grant NAG5-11010, NASA/TCSP Grant 04-0007-0031, NSF Grant ATM-0332197, and NOAA Grant NA17RJ1228.
REFERENCES
Braun, S. A., 2002: A cloud-resolving simulation of Hurricane Bob (1991): Storm structure and eyewall buoyancy. Mon. Wea. Rev., 130 , 1573–1592.
Braun, S. A., and W-K. Tao, 2000: Sensitivity of high-resolution simulations of Hurricane Bob (1991) to planetary boundary layer parameterizations. Mon. Wea. Rev., 128 , 3941–3961.
Braun, S. A., M. T. Montgomery, and Z. Pu, 2006: High-resolution simulation of Hurricane Bonnie (1998). Part I: The organization of eyewall vertical motion. J. Atmos. Sci., 63 , 19–42.
Eliassen, A., 1959: On the formation of fronts in the atmosphere. The Rossby Memorial Volume, Rockefeller Institute and Oxford University, 277–287.
Emanuel, K. A., 1995: Sensitivity of tropical cyclones to surface exchange coefficients and a revised steady-state model incorporating eye dynamics. J. Atmos. Sci., 52 , 3969–3976.
Emanuel, K. A., 1997: Some aspects of hurricane inner-core dynamics and energetics. J. Atmos. Sci., 54 , 1014–1026.
Emanuel, K. A., 2003: A similarity hypothesis for air–sea exchange at extreme wind speeds. J. Atmos. Sci., 60 , 1420–1428.
Hausman, S. A., 2001: Formulation and sensitivity analysis of a nonhydrostatic, axisymmetric tropical cyclone model. Atmospheric Science Paper 701, Colorado State University, 210 pp.
Jordan, C. L., 1958: Mean soundings for the West Indies area. J. Meteor., 15 , 91–97.
Kessler, E., 1969: On the Distribution and Continuity of Water Substance in Atmospheric Circulations. Meteor. Monogr., No. 32, Amer. Meteor. Soc., 84 pp.
Klemp, J. B., and R. Wilhelmson, 1978: The simulation of three-dimensional convective storm dynamics. J. Atmos. Sci., 35 , 1070–1096.
Kossin, J. P., and W. H. Schubert, 2003: Diffusion versus advective rearrangement of a circular vortex sheet. J. Atmos. Sci., 60 , 586–589.
Lighthill, J., 1999: Ocean spray and the thermodynamics of tropical cyclones. J. Eng. Math., 35 , 11–42.
Lord, S. J., and J. M. Lord, 1988: Vertical velocity in an axisymmetric, nonhydrostatic tropical cyclone model. J. Atmos. Sci., 45 , 1453–1461.
Lord, S. J., H. E. Willoughby, and J. M. Piotrowicz, 1984: Role of a parameterized ice phase microphysics in an axisymmetric, nonhydrostatic tropical cyclone model. J. Atmos. Sci., 41 , 2836–2848.
Malkus, J. S., and H. Riehl, 1960: On the dynamics and energy transformations in steady state hurricanes. Tellus, 12 , 1–20.
Manzello, S. L., and J. C. Yang, 2002: An experimental study of a water droplet impinging on a liquid surface. Exp. Fluids, 32 , 580–589.
Möller, J. D., and R. K. Smith, 1994: The development of potential vorticity in a hurricane-like vortex. Quart. J. Roy. Meteor. Soc., 120 , 1255–1265.
Ooyama, K. V., 1969: Numerical simulation of the life cycle of tropical cyclones. J. Atmos. Sci., 26 , 3–40.
Ooyama, K. V., 1990: A thermodynamic foundation for modeling the moist atmosphere. J. Atmos. Sci., 47 , 2580–2593.
Ooyama, K. V., 2001: A dynamic and thermodynamic foundation for modeling the moist atmosphere with parameterized microphysics. J. Atmos. Sci., 58 , 2073–2102.
Ooyama, K. V., 2002: The cubic-spline transform method: Basic definitions and tests in a 1D single domain. Mon. Wea. Rev., 130 , 2392–2415.
Persing, J., and M. T. Montgomery, 2003: Hurricane superintensity. J. Atmos. Sci., 60 , 2349–2371.
Rein, M., 1996: The transitional regime between coalescing and splashing drops. J. Fluid Mech., 306 , 145–165.
Rogers, R., S. Chen, J. Tenerelli, and H. Willoughby, 2003: A numerical study of the impact of vertical shear on the distribution of rainfall in Hurricane Bonnie (1998). Mon. Wea. Rev., 131 , 1577–1599.
Rotunno, R., and K. A. Emanuel, 1987: An air–sea interaction theory for tropical cyclones. Part II: Evolutionary study using a nonhydrostatic axisymmetric numerical model. J. Atmos. Sci., 44 , 542–561.
Schubert, W. H., 2004: A generalization of Ertel’s potential vorticity to a cloudy, precipitating atmosphere. Meteor. Z., 13 , 465–471.
Schubert, W. H., and B. T. Alworth, 1987: Evolution of potential vorticity in tropical cyclones. Quart. J. Roy. Meteor. Soc., 113 , 147–162.
Schubert, W. H., M. T. Montgomery, R. K. Taft, T. A. Guinn, S. R. Fulton, J. P. Kossin, and J. P. Edwards, 1999: Polygonal eyewalls, asymmetric eye contraction, and potential vorticity mixing in hurricanes. J. Atmos. Sci., 56 , 1197–1223.
Schubert, W. H., S. A. Hausman, M. Garcia, K. V. Ooyama, and H-C. Kuo, 2001: Potential vorticity in a moist atmosphere. J. Atmos. Sci., 58 , 3148–3157.
Stommel, H. M., 1995: Autobiography. Collected Works of Henry M. Stommel, N. G. Hogg and R. N. Huang, Eds., Vol. 1, Amer. Meteor. Soc. 380 pp.
Willoughby, H. E., H-L. Jin, S. J. Lord, and J. M. Piotrowicz, 1984: Hurricane structure and evolution as simulated by an axisymmetric nonhydrostatic numerical model. J. Atmos. Sci., 41 , 1169–1186.
World Meteorological Organization, 1979: General. Vol. I, Technical Regulations, WMO 49, 3 pp.
Yamasaki, M., 1983: A further study of the tropical cyclone without parameterizing the effects of cumulus convection. Pap. Meteor. Geophys., 34 , 221–260.
Yau, M. K., Y. Liu, D-L. Zhang, and Y. Chen, 2004: A multiscale numerical study of Hurricane Andrew (1992). Part VI: Small-scale inner-core structures and wind streaks. Mon. Wea. Rev., 132 , 1410–1433.
Zhang, D-L., and X. Wang, 2003: Dependence of hurricane intensity and structures on vertical resolution and time-step size. Adv. Atmos. Sci., 20 , 711–725.
Zhang, D-L., Y. Liu, and M. K. Yau, 2000: A multiscale numerical study of Hurricane Andrew (1992). Part III: Dynamically induced vertical motion. Mon. Wea. Rev., 128 , 3772–3788.
APPENDIX A
List of Symbols
Mass densities, mixing ratios, temperatures, pressures, velocities
ρa mass density of dry air
ρυ mass density of water vapor
ρc mass density of airborne condensate
ρr mass density of precipitating water substance
ρm = ρυ + ρc mass density of airborne moisture
ρ = ρa + ρm + ρr total mass density
μm = ρm/ρa mixing ratio of airborne moisture
μr = ρr/ρa mixing ratio of precipitating water substance
μ = μm + μr mixing ratio of total water substance
T1 temperature for thermodynamic state 1
T2 temperature for thermodynamic state 2
T = max(T1, T2) temperature
Tρ = p/(ρRa) virtual temperature
θρ = Tρ(p0/p)κ virtual potential temperature
θ = T(p0/pa)κ dry potential temperature
θe = T0 exp[σ/(ρacpa)] equivalent potential temperature
pa partial pressure of dry air
pυ partial pressure of water vapor
p = pa + pυ total pressure of moist air
u, υ, w radial, azimuthal, and vertical components of the velocity of dry air and airborne moisture
W vertical velocity of precipitation (relative to air)
w = w + ρ−1 (ρrW + Fm) density-weighted-mean vertical velocity
Specific entropies (J kg−1 K−1) and entropy densities (J m−3 K−1)
sa(ρa, T) specific entropy of dry air, defined by sa(ρa, T) = cυa ln(T/T0) − Ra ln(ρa/ρa0)
s(1)m(ρm, T) specific entropy of airborne moisture in state 1: s(1)m(ρm, T) = cυυ ln(T/T0) − Rυ ln(ρm/ρ*υ0) + L(T0)/T0
s(2)m(ρm, T) specific entropy of airborne moisture in state 2: s(2)m(ρm, T) = C(T) + D(T)/ρm
sr = C(T2) specific entropy of condensed water
s = σ/ρa dry-air-specific entropy of moist air
σa = ρasa entropy density of dry air
σm = ρmsm entropy density of airborne water substance
σr = ρrsr entropy density of precipitating water substance
σ = σa + σm + σr total entropy density
S1(ρa, ρm, T) entropy density function for state 1, defined by S1(ρa, ρm, T) = ρasa(ρa, T) + ρms(1)m (ρm, T)
S2(ρa, ρm, T) entropy density function for state 2, defined by S2(ρa, ρm, T) = ρasa(ρa, T) + ρms(2)m (ρm, T)
Constants and defined functions of temperature
f = 5.0 × 10−5 s−1 Coriolis parameter
g = 9.80665 m s2 acceleration of gravity
Ra = 287.05 J kg−1 K−1 gas constant of dry air
Rυ = 461.51 J kg−1 K−1 gas constant of water vapor
cpa = 1004.675 J kg−1 K−1 specific heat of dry air at constant pressure
cpυ = 1850.0 J kg−1 K−1 specific heat of water vapor at constant pressure
cυa = cpa − Ra specific heat of dry air at constant volume
cυυ = cpυ − Rυ specific heat of water vapor at constant volume
κ = Ra/cpa
p0 = 100 kPa Reference pressure
T0 = 273.15 K Reference temperature
ρa0 = p0/(RaT0) reference density for dry air
ρ*υ0 = ρ*υ(T0) mass density of saturated vapor at T0
ρr0 = 1.0 × 10−3kg m−3 reference density for precipitation
W0 = 5.5206 m s−1 reference fall velocity
τauto = 1000 s Autoconversion time scale
τcol = 455 s Collection time scale
τevap = 763 s Evaporation time scale
E(T) saturation vapor pressure; E(T) is synthesized from the saturation vapor pressures over water and ice
ρ*υ(T) = E(T)/(RυT) mass density of saturated vapor
L(T) = RυT 2[d ln E(T)/dT] specific latent heat for vaporizing condensate at T
C(T) entropy of a unit mass of condensate at T, given by C(T) = cυυ ln(T/T0) − Rυ ln[ρ*υ(T)/ρ*υ0] + L(T0)/T0 − L(T)/T
D(T) = L(T)ρ*υ(T)/T gain of entropy per unit volume by evaporating a sufficient amount of water, ρ*υ(T), to saturate the volume at T
Others
Fm, Fs, Fu, Fυ boundary layer turbulent fluxes of water vapor, entropy, radial, and azimuthal momentum
Qr conversion rate of ρm to ρr
(ξ, ζ) = (−∂m/r∂z, ∂m/r∂r) radial and vertical components of vorticity
D/DT operator ∂/∂t + u∂/∂r + w∂/∂z
D /Dt operator ∂/∂t + u∂/∂r +w ∂/∂zm = rυ + ½fr 2 absolute angular momentum per unit mass
ṁ = Dm/Dt source term for absolute angular momentum
=D m/Dt source term for absolute angular momentum: = ṁ + ρ−1(ρrW + Fm)(∂m/∂z)θ̇ρ = Dθρ/Dt source term for virtual potential temperature
ρ =D θρ/Dt source term for virtual potential temperature: ρ = θ̇ρ + ρ−1(ρrW + Fm)(∂θρ/∂z)P potential vorticity, as defined as (3.5), or by P = ρ−1∂(m, θρ)/r∂(r, z)
APPENDIX B
Potential Vorticity Equation




















APPENDIX C
Axisymmetric Shallow Water Equations




In the top panel of Fig. C1, the thick solid line (labeled Axi.Nonlin) shows the time history of the surface height h at r = 0 for the axisymmetric, nonlinear case, that is, the case when the flow is governed by (C.1) and (C.2). The thin solid line (labeled Axi.Linear) shows the corresponding time history for the axisymmetric, linear case, that is, the case when the flow is governed by the linearized versions of (C.1) and (C.2). Corresponding results for the cases in which axisymmetry is replaced by slab symmetry are shown by the curves labeled Slab.Nonlin and Slab.Linear. The bottom panel of Fig. 11 shows the radial profiles of h at the time of maximum h at r = 0. A prominent recoil column occurs only in the axisymmetric, nonlinear case (with the derivative constraint filter parameter, described in O2, chosen as lc = 2). Although there is no “truth solution” with which to compare Fig. 11, the recoil column produced in the axisymmetric nonlinear case is qualitatively similar to such columns produced in laboratory experiments (e.g., Rein 1996).
We have run similar experiments with different resolutions, with different values of lc, and with different initial conditions. An examination of all these results does not reveal anything peculiar or unexpected in the work of the derivative constraint filter under the axisymmetric assumption. We have also run similar experiments with the Coriolis force and with an azimuthal component υ, but the only effect of rotation is a progressive reduction of the recoil height with increasing f. From these experiments we conclude that there is no fundamental problem with the CST numerics applied to axisymmetric flows.

Vertical cross section of the initial tangential wind. Headings above this and the following figures indicate (left) the experiment, (center) the contoured variable, including units and contour interval Δ, and (right) the time in units of hh:mm:ss.s. Perturbation variables are identified by the (-bg) to the right of the variable name, indicating that a background state has been subtracted. The vertically oriented dashed lines mark the interfaces between nested grids, while the distances straddling these lines indicate the horizontal grid spacing to either side.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Vertical cross section of the initial tangential wind. Headings above this and the following figures indicate (left) the experiment, (center) the contoured variable, including units and contour interval Δ, and (right) the time in units of hh:mm:ss.s. Perturbation variables are identified by the (-bg) to the right of the variable name, indicating that a background state has been subtracted. The vertically oriented dashed lines mark the interfaces between nested grids, while the distances straddling these lines indicate the horizontal grid spacing to either side.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1
Vertical cross section of the initial tangential wind. Headings above this and the following figures indicate (left) the experiment, (center) the contoured variable, including units and contour interval Δ, and (right) the time in units of hh:mm:ss.s. Perturbation variables are identified by the (-bg) to the right of the variable name, indicating that a background state has been subtracted. The vertically oriented dashed lines mark the interfaces between nested grids, while the distances straddling these lines indicate the horizontal grid spacing to either side.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Initial far-field soundings of temperature (solid) and dewpoint temperature (dashed).
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Initial far-field soundings of temperature (solid) and dewpoint temperature (dashed).
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1
Initial far-field soundings of temperature (solid) and dewpoint temperature (dashed).
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Isolines of the 3-h-averaged surface tangential wind speed as a function of (r, t) from 0 to 100 km and 0 to 240 h for the CNTL experiment. The contour interval is 15 m s−1, starting from the 10 m s−1 contour.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Isolines of the 3-h-averaged surface tangential wind speed as a function of (r, t) from 0 to 100 km and 0 to 240 h for the CNTL experiment. The contour interval is 15 m s−1, starting from the 10 m s−1 contour.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1
Isolines of the 3-h-averaged surface tangential wind speed as a function of (r, t) from 0 to 100 km and 0 to 240 h for the CNTL experiment. The contour interval is 15 m s−1, starting from the 10 m s−1 contour.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Time-averaged (180–240 h) cross sections of the (a) tangential, (b) radial, and (c) vertical wind speeds (m s−1) for the CNTL experiment. Solid curves indicate positive values of u, υ, or w, while the dotted curves indicate zero or negative values. Shading in this figure and the following two figures denotes the region where the airborne condensate mixing ratio is greater than 0.1 g kg−1.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Time-averaged (180–240 h) cross sections of the (a) tangential, (b) radial, and (c) vertical wind speeds (m s−1) for the CNTL experiment. Solid curves indicate positive values of u, υ, or w, while the dotted curves indicate zero or negative values. Shading in this figure and the following two figures denotes the region where the airborne condensate mixing ratio is greater than 0.1 g kg−1.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1
Time-averaged (180–240 h) cross sections of the (a) tangential, (b) radial, and (c) vertical wind speeds (m s−1) for the CNTL experiment. Solid curves indicate positive values of u, υ, or w, while the dotted curves indicate zero or negative values. Shading in this figure and the following two figures denotes the region where the airborne condensate mixing ratio is greater than 0.1 g kg−1.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Time-averaged (180–240 h) cross section of the (a) temperature deviation (temperature minus height-dependent background temperature) and (b) water vapor mixing ratio for the CNTL experiment.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Time-averaged (180–240 h) cross section of the (a) temperature deviation (temperature minus height-dependent background temperature) and (b) water vapor mixing ratio for the CNTL experiment.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1
Time-averaged (180–240 h) cross section of the (a) temperature deviation (temperature minus height-dependent background temperature) and (b) water vapor mixing ratio for the CNTL experiment.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Time-averaged (180–240 h) cross sections of (a) vertical component of absolute vorticity, (b) virtual potential temperature, and (c) potential vorticity anomaly for the CNTL experiment.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Time-averaged (180–240 h) cross sections of (a) vertical component of absolute vorticity, (b) virtual potential temperature, and (c) potential vorticity anomaly for the CNTL experiment.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1
Time-averaged (180–240 h) cross sections of (a) vertical component of absolute vorticity, (b) virtual potential temperature, and (c) potential vorticity anomaly for the CNTL experiment.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Same as Fig. 3, but for the NICE experiment.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Same as Fig. 3, but for the NICE experiment.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1
Same as Fig. 3, but for the NICE experiment.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Same as Fig. 4 but for 120–240-h time average of the NICE experiment.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Same as Fig. 4 but for 120–240-h time average of the NICE experiment.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1
Same as Fig. 4 but for 120–240-h time average of the NICE experiment.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Same as Fig. 5 but for 120–240-h time average of the NICE experiment.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Same as Fig. 5 but for 120–240-h time average of the NICE experiment.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1
Same as Fig. 5 but for 120–240-h time average of the NICE experiment.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Same as Fig. 6 but for 120–240-h time average of the NICE experiment.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Same as Fig. 6 but for 120–240-h time average of the NICE experiment.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1
Same as Fig. 6 but for 120–240-h time average of the NICE experiment.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Fig. C1. Results of shallow water experiments on the Rayleigh recoil problem under four different dynamics: axisymmetric, nonlinear; axisymmetric, linear; slab symmetric, nonlinear; slab symmetric, linear. (upper) The fluid depth h (m) at r = 0 as a function of time (in seconds), and (lower) h as function of r (in kilometers) at the time of maximum fluid depth at r = 0.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1

Fig. C1. Results of shallow water experiments on the Rayleigh recoil problem under four different dynamics: axisymmetric, nonlinear; axisymmetric, linear; slab symmetric, nonlinear; slab symmetric, linear. (upper) The fluid depth h (m) at r = 0 as a function of time (in seconds), and (lower) h as function of r (in kilometers) at the time of maximum fluid depth at r = 0.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1
Fig. C1. Results of shallow water experiments on the Rayleigh recoil problem under four different dynamics: axisymmetric, nonlinear; axisymmetric, linear; slab symmetric, nonlinear; slab symmetric, linear. (upper) The fluid depth h (m) at r = 0 as a function of time (in seconds), and (lower) h as function of r (in kilometers) at the time of maximum fluid depth at r = 0.
Citation: Journal of the Atmospheric Sciences 63, 1; 10.1175/JAS3601.1
A list of symbols is provided in appendix A.
Although the assumed boundary layer depth of 1000 m is consistent with many observed tropical flows, the inner core hurricane boundary layer may be shallower than 1000 m, which requires high vertical resolution in the lower troposphere. For a detailed study of the effect of lower tropospheric vertical resolution on hurricane intensity, see Zhang and Wang (2003).
The vertical confinement of the boundary layer radial inflow is even more pronounced in models with high vertical resolution in the lowest 2 km (Zhang et al. 2000; Braun 2002; Rogers et al. 2003).
More extensive discussions of these and other numerical experiments can be found in Hausman (2001).