1. Introduction
Strong tropical cyclones (TCs) often experience secondary eyewall formation (SEF; Willoughby et al. 1982), in which a new eyewall forms outside the original eyewall. It is usually followed by an eyewall replacement cycle, and hence, predicting SEF is important for accurate TC forecasts. However, there is no widely accepted theory for SEF despite the significant effort by many researches, especially in the last decade. We here present a new theory focusing on flow interaction between the boundary layer and the free atmosphere.
SEF often occurs at a radius of 2–4 times the radius of maximum wind (RMW) in recently intensified TCs (Willoughby et al. 1982; Black and Willoughby 1992; Dodge et al. 1999; Houze et al. 2007; Kossin and Sitkowski 2009; Kuo et al. 2009; Wu et al. 2016). The key parameters and processes for SEF proposed by previous studies are inertial instability (Willoughby et al. 1982), terrain effect (Hawkins 1983), ice microphysics (Willoughby et al. 1984), inflow surge (Molinari and Skubis 1985; Molinari and Vollaro 1989), vortex Rossby waves (Montgomery and Kallenbach 1997; Terwey and Montgomery 2003), potential-vorticity intrusion from upper layer (Nong and Emanuel 2003), axisymmetrization of vorticity disturbance (Kuo et al. 2004, 2008), the rectification of eddies in the β skirt outside the primary eyewall (Terwey and Montgomery 2008), supergradient wind (Huang et al. 2012; Wu et al. 2012), wind field expansion (Rozoff et al. 2012; Sun et al. 2013), radial vorticity gradient outside the RMW where relative vorticity is low (Kepert 2013), top-down process associated with interaction with upper-level trough (Leroux et al. 2013), rainband heating (Zhu and Zhu 2014; Zhang et al. 2016), and upper-level outflow interactions (Dai et al. 2017).
Kepert (2013) showed the importance of the radial gradient of relative vorticity in a low-vorticity environment using both linear and nonlinear boundary layer models. According to the Ekman theory for a rotating fluid, Ekman pumping velocity is proportional to the radial vorticity gradient and inversely proportional to the square of the absolute vorticity itself. Kepert (2013) and Kepert and Nolan (2014) proposed a positive feedback that can lead to SEF: when the radial gradient of vertical relative vorticity is steep, Ekman pumping velocity is large, the upward velocity is enhanced by convection and causes further convergence, and the vorticity gradient steepens. However, the previous studies consider vorticity profiles after SEF and secondary wind maximum, in which a secondary peak of tangential velocity is already present outside the RMW. From ensembles of numerical simulations, Zhang et al. (2016) showed that the diabatic heating outside the RMW in the upper layer proceeds downward, changing the radial vorticity profile, and the secondary eyewall develops through the interaction between convection and boundary layer processes. Although the interaction might work once the vorticity gradient is steep, the proposed mechanism is incomplete because it offers no explanation of conditions under which outer rainbands produce such a strong vorticity gradient (otherwise, rainbands would always cause SEF). Therefore, the underlying mechanism of SEF, especially for the initial process (i.e., when and where SEF occurs), is unclear.
This study proposes a linear instability via an interaction between Ekman pumping and the flow in the free atmosphere, which is not associated with moist convection, unlike the previous theories, and hypothesizes that very early processes of SEF are attributed to this instability. Once the initial processes of SEF proposed in this study create a local maximum of vertical velocity or tangential velocity outside the primary eyewall, the heating–convergence feedback might work to develop the secondary eyewall, followed by an eyewall replacement cycle (Kepert 2013; Kepert and Nolan 2014).
The rest of the paper is organized as follows. We develop a linear theory for a 1.5-layer shallow-water system in Cartesian coordinates and eigenvalue analyses and numerical integrations for a nonlinear equation set are conducted in section 2. The theory is extended to cylindrical coordinates and a condition for the instability to occur is derived and linear stability analyses of an axisymmetric 1.5-layer shallow-water system are performed in section 3. The unstable condition is tested by applying it to simulation results by two nonhydrostatic full-physics models in section 4. The results are discussed by comparing them with previous studies in section 5. This study is concluded in section 6.
2. Extension of the Ekman theory in Cartesian coordinates to high–wind speed conditions
a. Classic Ekman theory


Schematic of the one-dimensional, 1.5-layer fluid system on the f plane. The system consists of free atmosphere with a depth of h and boundary layer with
Citation: Journal of the Atmospheric Sciences 75, 11; 10.1175/JAS-D-18-0042.1
The Ekman damping process for velocity perturbations can be interpreted as follows: once a positive
Schematic of the spatial profiles of velocity disturbances that respond to (top) the initial perturbation of
Citation: Journal of the Atmospheric Sciences 75, 11; 10.1175/JAS-D-18-0042.1
The classic Ekman theory is widely accepted and has provided many insights into geophysical fluid dynamics. For example, the effects of Ekman pumping on Rossby waves are modeled by adding vertical velocity at the top of the boundary layer that is proportional to vertical relative vorticity (e.g., James 1995; Vallis 2017). Although the theory can capture fundamental processes for various phenomena, the assumptions may limit its application. In particular, it is assumed that the background flow is not fast and its spatial gradient (relative vorticity) is not large. Hence, the theory may not be applicable to some circumstances such as jets or tropical cyclones in which the horizontal velocity is fast in the lower free atmosphere.
b. Extension of Ekman theory to fast velocity conditions
We here hypothesize that the relationship in (1) can possibly be reversed when we include the advection in the boundary layer. If this were to occur,


















A series of numerical calculations for the equations for the free atmosphere [(3)–(5)] and vertical velocity [(8)] have been performed to test the hypothesized instability by artificially setting the coefficients
c. The condition for instability













d. Eigenvalue analysis









The eigenvalue equation [(16)] was discretized on a staggered grid system and eigenvalues and eigenvectors were computed with MATLAB. As in the test calculations in appendix C, the spatial derivatives in (15) were discretized by the fourth-order central difference scheme, and periodic boundary conditions were used. We have tested the second-order difference scheme and found that the results are insensitive to the discretization scheme. The nondimensional domain width was 8 with 201 grid points and uniform grid spacing of 0.04. The initial velocity field was a jetlike profile
Figures 3a and 3b show spatial profiles of
Spatial profiles of (a)
Citation: Journal of the Atmospheric Sciences 75, 11; 10.1175/JAS-D-18-0042.1
The eigenmodes with the maximum eigenvalue are shown in Figs. 3c and 3d. In this case, the real part of the maximum eigenvalue (i.e., growth rate) is 18.73 and corresponds to an e-folding time of 1.33 × 104 s (≈3.7 h). The vertical velocity
We conducted a series of eigenvalue analyses for wide ranges of Ro and the spatial parameter
List of parameters in the eigenvalue analysis for Cartesian coordinates.
Figure 4 displays a phase diagram in space of two parameters: Ro and the domain maximum of
Phase diagram in the parameter space (domain maximum of
Citation: Journal of the Atmospheric Sciences 75, 11; 10.1175/JAS-D-18-0042.1
e. Numerical integrations of nonlinear equation set


List of parameters that are constant in this study.
The equations were solved numerically using a finite difference method. As in the eigenvalue analysis, the quantities were allocated on a staggered grid system. The time and space derivatives in the equations were discretized by the third-order Runge–Kutta method and by the fourth-order central difference scheme, respectively. At the first time step, w was estimated from υ, and then the equations for the free atmosphere were stepped forward. The integration was conducted up to
Figure 5 shows spatial distributions of velocities at t = 0.02 in the control case, in which Ro = 0.80 and
Spatial profiles of (a) υ (black) and its deviation from the initial value
Citation: Journal of the Atmospheric Sciences 75, 11; 10.1175/JAS-D-18-0042.1
Figure 6a displays a space–time cross section of υ in the control case and time series of domain-maximum w. The maximum w amplifies with time. The maximum w keeps growing until the end of the simulation, and the locations of velocity peaks seen in Fig. 5 are nearly fixed in space. The spatial profiles of velocity during the integration (Fig. 5) strongly suggest that the proposed feedback works to amplify w. Since no explicit diffusion is incorporated into the discretized equation set, the linearly unstable modes continue to grow.
(a) Time–space cross section of
Citation: Journal of the Atmospheric Sciences 75, 11; 10.1175/JAS-D-18-0042.1
Figure 7 depicts a phase diagram in Ro and the domain-maximum
As in Fig. 4, but for the results of nonlinear integrations. The circles indicate experiments in which the peak w exceeds 5.0 during the integration period. The gray circles represent experiments in which the minimum absolute vorticity in the domain is negative at the initial time. The gray × marks show the experiments in which velocities decay as in the classic Ekman theory.
Citation: Journal of the Atmospheric Sciences 75, 11; 10.1175/JAS-D-18-0042.1
3. Theory in cylindrical coordinates
























Eigenvalue analysis in a cylindrical coordinate














The radial dimension was discretized on a staggered grid system with uniform grid spacing of 0.04. The domain size was 80.0 (i.e., 80 times the RMW) with 2001 grid points. The radial derivatives were discretized by a fourth-order central difference scheme. At the boundaries (r = 0 and 80),
Figures 8a and 8b show radial profiles of basic-state
Radial profiles of basic-state (a) tangential velocity
Citation: Journal of the Atmospheric Sciences 75, 11; 10.1175/JAS-D-18-0042.1
The eigenmodes of
We examine the sensitivity of the solution to Ro and vortex parameters
List of parameters in the eigenvalue analysis for cylindrical coordinates.
Figure 9 shows the growth rate as a function of experimental parameters. The growth rate increases with Ro,
Growth rate as a function of experimental parameters: (a) Ro; (b) radius of secondary jump of vorticity relative to the RMW,
Citation: Journal of the Atmospheric Sciences 75, 11; 10.1175/JAS-D-18-0042.1














Growth rate as a function of
Citation: Journal of the Atmospheric Sciences 75, 11; 10.1175/JAS-D-18-0042.1
4. SEF simulated in full-physics models
a. Axisymmetric model
A long-term simulation was performed using an axisymmetric version of Cloud Model 1 (CM1; version 17; Bryan and Fritsch 2002), which solves fully compressible, nonhydrostatic equations. The experimental setting was basically the same as Hakim (2013) in which a number of eyewall replacement cycle events were simulated during his long-term simulation for 500 days. The differences from Hakim’s (2013) simulation are listed below. The initial mass, temperature, and water vapor mixing ratio was obtained from Jordan’s (1958) tropical mean sounding for hurricane season. The Coriolis parameter was 5 × 10−5 s−1, and the sea surface temperature was 301.15 K, both of which were fixed during integration. The radiative effects were incorporated by the NASA Goddard scheme, which was called every 5 min. The integration period was 60 days (=1440 h) with an output interval of 3 h for prognostic quantities. The grid spacing for the radial direction was 2 km, whereas the spacing for the vertical direction was stretched from the bottom to top. The minimum grid spacing for the vertical direction was 0.25 km at the lowest model layer, stretching to 1.0 km near the top.
Figurea 11a and 11b show time series of TC intensity defined as the maximum tangential velocity at z = 2 km and its radius. The TC achieves maximum intensity around t = 340 h after an intensification phase, decreases to about 60 m s−1, and then is approximately steady afterward, which is qualitatively consistent with the long-term simulations in previous studies (Hakim 2011, 2013). We shall call the period after t = 500 h a quasi-steady state. During the quasi-steady state, rapid changes in intensity and RMW are observed several times. As pointed out by Hakim (2011), the intensity oscillation appears to be due to the eyewall replacement cycle associated with SEF.
Time series of (a) TC intensity and (b) RMW of the simulated TC in an axisymmetric full-physics model, (c) radius–height cross section of tangential velocity (gray contours), radial velocity (black contours), and mixing ratio of hydrometeors (shaded) averaged during the quasi-steady state, and (d) radius–time cross section of tangential velocity at z = 2 km. The contour intervals of tangential velocity and radial velocity in (c) are 10 and 5 m s−1, respectively. The contour of 50 m s−1 is highlighted by the white line in (d).
Citation: Journal of the Atmospheric Sciences 75, 11; 10.1175/JAS-D-18-0042.1
Figure 11c depicts a radius–height cross section of tangential and radial velocities and mixing ratio of hydrometeors averaged during the quasi-steady state. Both the velocity and cloud fields are consistent with the observational studies (e.g., Hawkins and Imbembo 1976): the tangential velocity has a peak at the top of the boundary layer (~1.5 km) and decays in both the radial and vertical directions, the inflow and outflow regions appear at the surface and around the 15-km level, and the hydrometeor mixing ratio is large around the RMW from the 1.5- to 15-km altitudes.
Figure 11d depicts a radius–time cross section of tangential velocity at z = 2 km. The RMW is about 60 km during the quasi-steady state as seen in Fig. 11b. Interestingly, a strong velocity region periodically forms about every 80 h, and the region proceeds inward after formation. The mixing ratio of hydrometeors and vorticity fields shows similar behavior to the tangential velocity (not shown). It is suggested that this cyclic process is the eyewall replacement cycle and formation of strong tangential velocity outside the RMW is SEF.
The SEF events were extracted from the simulation by applying the following conditions: the RMW is greater than that at the previous and next output time, the increase from the previous step is greater than 10 km, and the new RMW is more than 20 km beyond its average value during the quasi-steady state (~60 km in this case). Applying the conditions, six SEF events were detected in the quasi-steady period, and the six events were combined to make composites.
Figure 12a displays a radius–time cross section of tangential velocity at z = 2 km subtracted from the temporal average from
(a) Radius–time cross section of composited tangential velocity subtracted from the temporal average from t = −40 to 40 h (shaded) and
Citation: Journal of the Atmospheric Sciences 75, 11; 10.1175/JAS-D-18-0042.1
Figure 12b shows the time series of
b. Three-dimensional model
We conducted a numerical simulation using a three-dimensional full-physics numerical model, Weather Research and Forecasting (WRF) Model, version 3.4.1 (Skamarock et al. 2008). The experimental setting is described in appendix E.
Figures 13a and 13b show time series of TC intensity and RMW of the simulated TC. The TC experiences two different intensification phases and becomes quasi steady after t = 45 h. The intensity weakens from t = 65 to 76 h and reintensifies afterward. The RMW gradually decreases, jumps outward at t = 76 h, and then decreases again. Structural evolution of the simulated TC (shown below) indicates that the sudden changes in TC intensity and RMW are associated with SEF and the eyewall replacement process. We define t = 76 h as the time of SEF.
Time series of (a) TC intensity and (b) RMW of the simulated TC in the WRF Model; (c) radius–height cross section of azimuthally averaged tangential velocity (gray contours), radial velocity (black contours), and mixing ratio of hydrometeors (shaded) averaged from t = 48 to 54 h; and (d) radius–time cross section of azimuthally averaged tangential velocity at z = 2 km (shaded), diabatic heating rate vertically averaged from z = 1 to 10 km (gray contours), and RMW (gray dashed lines). The contour intervals of tangential velocity, radial velocity, and hydrometeor mixing ratio in (c) are 10 and 5 m s−1 and 5 g kg−1, respectively. The contour level of 5 K h−1 is shown in (d).
Citation: Journal of the Atmospheric Sciences 75, 11; 10.1175/JAS-D-18-0042.1
Figure 13c depicts a radius–height cross section of the azimuthally averaged hydrometeor mixing ratio, tangential velocity, and radial velocity, which are temporally averaged from t = 48 to 54 h. Figure 13d depicts a radius–time cross section of azimuthally averaged tangential velocity at z = 2 km and diabatic heating rate that is averaged vertically from z = 2 to 10 km. The tangential velocity intensifies with decreasing RMW to t = 45 h, and then the RMW is approximately constant. A notable point is that the tangential velocity field expands with time after the TC achieves the maximum intensity. The temporal changes in structure and intensify during the SEF event are consistent with observations.
Figure 14a displays a radius–time cross section of the deviation of tangential velocity from the average between t = 48 and 96 h, diabatic heating averaged vertically from z = 1 to 10 km, and
(a) Radius–time cross section of azimuthally averaged tangential velocity at z = 2 km subtracted by the temporal average from t = 45 to 55 h of
Citation: Journal of the Atmospheric Sciences 75, 11; 10.1175/JAS-D-18-0042.1
Figure 14b shows a times series of
5. Discussion
The theory indicates that a small gradient of vorticity can produce large
The initial amplification of the radial gradient of vorticity to satisfy the unstable condition might be caused by some processes such as diabatic heating in rainbands (Rozoff et al. 2012; Zhang et al. 2016). The present 3D simulation shows little convective organization before the secondary peak of tangential velocity forms (Fig. 14a). Rather, many isolated convective cells are present, but the accumulated effect of these can produce the needed vorticity gradient. Once a secondary peak of tangential velocity forms, a symmetric convective region is organized around that radius. Subsequently, the secondary peak and organized convection may amplify together through the feedback.
Figure 15 illustrates how the proposed feedback mechanism works, with a comparison to the classic Ekman feedback (and damping). Starting from perturbation of vertical velocity with a positive peak, the radial and tangential velocities in the free atmosphere will be produced (cf. Fig. 2). This change will result in faster velocity in the x-direction in the boundary layer to the left of the peak of the initial perturbation of vertical velocity and slower velocity to the right. Thus, the responding flow in the boundary layer tends to cause divergence and decrease w. On the other hand, when the unstable condition is satisfied, the change in flow fields in the free atmosphere rather produces slower velocity to the left and faster to the right, which increases in w. This would happen if the decrease in
Schematic of proposed mechanism. (a) Divergent flow caused by w increases
Citation: Journal of the Atmospheric Sciences 75, 11; 10.1175/JAS-D-18-0042.1
6. Conclusions
We demonstrate an instability via interaction between the free atmosphere and the Ekman layer based on the classic Ekman pumping theory and hypothesize that the very early process of SEF in TCs can be attributed to this instability. The classic Ekman theory indicates that vertical motion at the top of the boundary layer is proportional to the spatial gradient of velocity (relative vorticity), and this results in a decay of velocity perturbations in the free atmosphere. Here, it was hypothesized that once the Ekman pumping velocity is negatively proportional to the velocity gradient under some special conditions, disturbances can grow exponentially.
The instability happens when velocity is fast and changes rapidly in the horizontal direction. Using a 1.5-layer shallow-water model on the f plane [(3)–(7)], we analytically derived an unstable condition (13), in which the proportionality coefficient of relative vorticity perturbation in the
Eigenvalue analyses were performed for the linearized equation set. A positive eigenvalue (growth rate) was obtained, and the eigenvector for the maximum eigenvalue showed a velocity field consistent with the hypothesis. In the parameter space of Ro and
We extended the theory to cylindrical coordinates. By linearizing the nondimensionalized equations, it was shown that the instability occurs where Ro
The hypothesis was tested by computing the dimensional values of
Another SEF was simulated using WRF in an idealized environment with nonzero shear. A secondary peak of tangential velocity formed around 5.0–5.5 times the RMW and intensified while proceeding inward, while accompanied by strong convective heating. The local value of
The two modeling results suggest that the initial process of SEF, the formation of a secondary peak of a tangential velocity, may be attributed to the instability, and thus
Acknowledgments
Y. Miyamoto was supported by JSPS Scientific Research 26–358 for the JSPS fellowship program for overseas researchers. D. S. Nolan was supported by the NSF through Grant AGS-1654831. This work was supported by Keio University Academic Development Funds for Joint Research. The authors thank Dr. Jeff Kepert and two anonymous reviewers for their careful reviews and Drs. Hiroshi Taniguchi and Shigenori Otsuka for fruitful discussions.
APPENDIX A
Equations for the One-Dimensional 1.5-Layer Shallow-Water System in Cartesian Coordinates



















APPENDIX B
Condition for the Instability in Cartesian Coordinates




















APPENDIX C
Test Calculations for the Instability
To verify the hypothesized sensitivity of solutions to the coefficient
The initial perturbation horizontal velocities were zero everywhere at the initial time. White noise perturbations with amplitude of 1.0 were added to the vertical velocity. Given
The present hypothesis indicates that an instability occurs when
Time series of domain-maximum
Citation: Journal of the Atmospheric Sciences 75, 11; 10.1175/JAS-D-18-0042.1
APPENDIX D
Equations for the Rapidly Rotating, Axisymmetric, 1.5-Layer Shallow-Water System

















APPENDIX E
Details of the WRF Simulation
The three-dimensional, full-physics simulation used in section 4b was produced using the WRF Model, version 3.4.1. The simulation depicts the early development and then rapid intensification of a tropical cyclone in a large, zonally periodic channel using the idealized modeling framework of Nolan (2011), with wind shear balanced by a meridional temperature gradient. However, in many other aspects, the simulation is modeled after the hurricane nature run simulations described by Nolan et al. (2013) and Nolan and Mattocks (2014). The simulation was originally intended to be one of a larger set of idealized simulations used by Koltz and Nolan (2018, manuscript submitted to Wea. Forecasting) for observing system tests, but its robust SEF and ERC made it useful for this paper.
The outer domain has 240 × 180 grid points in the zonal and meridional directions with 27-km grid spacing. Three nested, vortex-following grids are used with 9-, 3-, and 1-km grid spacings, with 180 × 180, 360 × 360, and 480 × 480 grid points, respectively. There are 60 vertical levels between the surface and 20-km altitude. The grid spacings in the WRF vertical coordinate are the same as used in the nature runs (see Fig. 2 of Nolan et al. 2013), and the physical parameterizations are also all identical, including the use of the one-dimensional mixed-layer cooling model of Pollard et al. (1972) available in WRF, version 3.4.1.
The initial vortex and background flow are very similar to the simulations described in Nolan (2011) and Nolan and McGauley (2012). The outer domain is initialized with mean easterly flow of 5 m s−1 at the surface, which smoothly increases to 5 m s−1 of westerly flow between 850 and 20 hPa. An axisymmetric vortex, balanced by temperature and pressure anomalies, is introduced into the eastern end of the domain. The symmetric initial vortex has a maximum tangential wind of 15 m s−1 at r = 135 km and z = 1.5 km. The radial profile of the tangential wind varies as a modified Rankine vortex with decay parameter a = 1/3 and decreases with height using the same analytical function as in (4.2) of Stern and Nolan (2011).
The Coriolis parameter is constant across the channel domain with value
Finally, as the simulation proceeds, the wind, temperature, and moisture fields on the outer domain are relaxed back to their initial values with a 24-h relaxation time scale. This relaxation does not apply on the nested grids. The relaxation keeps the environmental sounding and wind shear profile around the tropical cyclone roughly constant as the storm develops.
REFERENCES
Black, M. L., and H. E. Willoughby, 1992: The concentric eyewall cycle of Hurricane Gilbert. Mon. Wea. Rev., 120, 947–957, https://doi.org/10.1175/1520-0493(1992)120<0947:TCECOH>2.0.CO;2.
Bryan, G. H., and J. M. Fritsch, 2002: A benchmark simulation for moist nonhydrostatic numerical models. Mon. Wea. Rev., 130, 2917–2928, https://doi.org/10.1175/1520-0493(2002)130<2917:ABSFMN>2.0.CO;2.
Dai, Y., S. Majumdar, and D. S. Nolan, 2017: Secondary eyewall formation in tropical cyclones by outflow–jet interaction. J. Atmos. Sci., 74, 1941–1958, https://doi.org/10.1175/JAS-D-16-0322.1.
Dodge, P., R. W. Burpee, and F. D. Marks Jr., 1999: The kinematic structure of a hurricane with sea level pressure less than 900 mb. Mon. Wea. Rev., 127, 987–1004, https://doi.org/10.1175/1520-0493(1999)127<0987:TKSOAH>2.0.CO;2.
Ekman, V. W., 1905: On the influence of the Earth’s rotation on ocean currents. Arch. Math. Astron. Phys., 2, 1–52.
Eliassen, A., and M. Lystad, 1977: The Ekman layer of a circular vortex: A numerical and theoretical study. Geophys. Norv., 31, 1–16.
Hakim, G. J., 2011: The mean state of axisymmetric hurricanes in statistical equilibrium. J. Atmos. Sci., 68, 1364–1376, https://doi.org/10.1175/2010JAS3644.1.
Hakim, G. J., 2013: The variability and predictability of axisymmetric hurricanes in statistical equilibrium. J. Atmos. Sci., 70, 993–1005, https://doi.org/10.1175/JAS-D-12-0188.1.
Hawkins, H. F., 1983: Hurricane Allen and island obstacles. J. Atmos. Sci., 40, 1360–1361, https://doi.org/10.1175/1520-0469(1983)040<1360:HAAIO>2.0.CO;2.
Hawkins, H. F., and S. M. Imbembo, 1976: The structure of a small, intense Hurricane Inez 1966. Mon. Wea. Rev., 104, 418–442, https://doi.org/10.1175/1520-0493(1976)104<0418:TSOASI>2.0.CO;2.
Holton, J. R., 1992: An Introduction to Dynamic Meteorology. 3rd ed. Academic Press, 511 pp.
Houze, R. A., Jr., S. S. Chen, B. F. Smull, W.-C. Lee, and M. M. Bell, 2007: Hurricane intensity and eyewall replacement. Science, 315, 1235–1239, https://doi.org/10.1126/science.1135650.
Huang, Y.-H., M. T. Montgomery, and C.-C. Wu, 2012: Concentric eyewall formation in Typhoon Sinlaku (2008). Part II: Axisymmetric dynamical processes. J. Atmos. Sci., 69, 662–674, https://doi.org/10.1175/JAS-D-11-0114.1.
James, I. N., 1995: Introduction to Circulating Atmospheres. Cambridge University Press, 448 pp.
Jordan, C. L., 1958: Mean soundings for the West Indies area. J. Meteor., 15, 91–97, https://doi.org/10.1175/1520-0469(1958)015<0091:MSFTWI>2.0.CO;2.
Kepert, J. D., 2013: How does the boundary layer contribute to eyewall replacement cycles in axisymmetric tropical cyclones? J. Atmos. Sci., 70, 2808–2829, https://doi.org/10.1175/JAS-D-13-046.1.
Kepert, J. D., 2017: Time and space scales in the tropical cyclone boundary layer, and the location of the eyewall updraft. J. Atmos. Sci., 74, 3305–3323, https://doi.org/10.1175/JAS-D-17-0077.1.
Kepert, J. D., and D. S. Nolan, 2014: Reply to “Comments on ‘How does the boundary layer contribute to eyewall replacement cycles in axisymmetric tropical cyclones?’” J. Atmos. Sci., 71, 4692–4704, https://doi.org/10.1175/JAS-D-14-0014.1.
Kossin, J. P., and M. Sitkowski, 2009: An objective model for identifying secondary eyewall formation in hurricanes. Mon. Wea. Rev., 137, 876–892, https://doi.org/10.1175/2008MWR2701.1.
Kuo, H.-C., L.-Y. Lin, C.-P. Chang, and R. T. Williams, 2004: The formation of concentric vorticity structures in typhoons. J. Atmos. Sci., 61, 2722–2734, https://doi.org/10.1175/JAS3286.1.
Kuo, H.-C., W. H. Schubert, C.-L. Tsai, and Y.-F. Kuo, 2008: Vortex interactions and the barotropic aspects of concentric eyewall formation. Mon. Wea. Rev., 136, 5183–5198, https://doi.org/10.1175/2008MWR2378.1.
Kuo, H.-C., C.-P. Chang, Y.-T. Yang, and H. J. Jiang, 2009: Western North Pacific typhoons with concentric eyewalls. Mon. Wea. Rev., 137, 3758–3770, https://doi.org/10.1175/2009MWR2850.1.
Leroux, M.-D., M. Plu, D. Barbary, F. Roux, and P. Arbogast, 2013: Dynamical and physical processes leading to tropical cyclone intensification under upper-level trough forcing. J. Atmos. Sci., 70, 2547–2565, https://doi.org/10.1175/JAS-D-12-0293.1.
Molinari, J., and S. Skubis, 1985: Evolution of the surface wind field in an intensifying tropical cyclone. J. Atmos. Sci., 42, 2865–2879, https://doi.org/10.1175/1520-0469(1985)042<2865:EOTSWF>2.0.CO;2.
Molinari, J., and D. Vollaro, 1989: External influences on hurricane intensity. Part I: Outflow layer eddy angular momentum fluxes. J. Atmos. Sci., 46, 1093–1105, https://doi.org/10.1175/1520-0469(1989)046<1093:EIOHIP>2.0.CO;2.
Montgomery, M. T., and R. J. Kallenbach, 1997: A theory for vortex Rossby-waves and its application to spiral bands and intensity changes in hurricanes. Quart. J. Roy. Meteor. Soc., 123, 435–465, https://doi.org/10.1002/qj.49712353810.
Nolan, D. S., 2011: Evaluating environmental favorableness for tropical cyclone development with the method of point-downscaling. J. Adv. Model. Earth Syst., 3, M08001, https://doi.org/10.1029/2011MS000063.
Nolan, D. S., and M. G. McGauley, 2012: Tropical cyclogenesis in wind shear: Climatological relationships and physical processes. Cyclones: Formation, Triggers and Control, K. Oouchi and H. Fudeyasu, Eds., Nova Science Publishers, 1–36.
Nolan, D. S., and C. A. Mattocks, 2014: Development and evaluation of the second hurricane nature run using the joint OSSE nature run and the WRF Model. 31st Conf. on Hurricane and Tropical Meteorology, San Diego, CA, Amer. Meteor. Soc., 91, https://ams.confex.com/ams/31Hurr/webprogram/Paper244751.html.
Nolan, D. S., R. Atlas, K. T. Bhatia, and L. R. Bucci, 2013: Development and validation of a hurricane nature run using the joint OSSE nature run and the WRF Model. J. Adv. Model. Earth Syst., 5, 382–405, https://doi.org/10.1002/jame.20031.
Nong, S., and K. Emanuel, 2003: A numerical study of the genesis of concentric eyewalls in hurricanes. Quart. J. Roy. Meteor. Soc., 129, 3323–3338, https://doi.org/10.1256/qj.01.132.
Ooyama, K., 1969: Numerical simulation of the life cycle of tropical cyclones. J. Atmos. Sci., 26, 3–40, https://doi.org/10.1175/1520-0469(1969)026<0003:NSOTLC>2.0.CO;2.
Pollard, R. T., P. B. Rhines, and R. O. R. Y. Thompson, 1972: The deepening of the wind-mixed layer. Geophys. Fluid Dyn., 4, 381–404, https://doi.org/10.1080/03091927208236105.
Rozoff, C. M., D. S. Nolan, J. P. Kossin, F. Zhang, and J. Fang, 2012: The roles of an expanding wind field and inertial stability in tropical cyclone secondary eyewall formation. J. Atmos. Soc., 69, 2621–2643, https://doi.org/10.1175/JAS-D-11-0326.1.
Skamarock, W. C., and Coauthors, 2008: A description of the Advanced Research WRF version 3. NCAR Tech. Note NCAR/TN-475+STR, 123, 113 pp., https://doi.org/10.5065/D68S4MVH.
Stern, D. P., and D. S. Nolan, 2011: On the vertical decay rate of the maximum tangential winds in tropical cyclones. J. Atmos. Sci., 68, 2073–2094, https://doi.org/10.1175/2011JAS3682.1.
Sugimoto, N., K. Ishioka, and S. Yoden, 2007: Balance regimes for the stability of a jet in an f-plane shallow water system. Fluid Dyn. Res., 39, 353–377, https://doi.org/10.1016/j.fluiddyn.2006.07.004.
Sun, Y. Q., Y. Jiang, B. Tan, and F. Zhang, 2013: The governing dynamics of the secondary eyewall formation of Typhoon Sinlaku (2008). J. Atmos. Sci., 70, 3818–3837, https://doi.org/10.1175/JAS-D-13-044.1.
Terwey, W. D., and M. T. Montgomery, 2003: Vortex waves and evolution in sharp vorticity gradient vortices. Colorado State University Bluebook 734, 97 pp.
Terwey, W. D., and M. T. Montgomery, 2008: Secondary eyewall formation in two idealized, full-physics modeled hurricanes. J. Geophys. Res., 113, D12112, https://doi.org/10.1029/2007JD008897.
Vallis, G. K., 2017: Atmospheric and Oceanic Fluid Dynamics. Cambridge University Press, 964 pp.
Willoughby, H. E., J. A. Clos, and M. B. Shoreibah, 1982: Concentric eyewalls, secondary wind maxima, and the development of the hurricane vortex. J. Atmos. Sci., 39, 395–411, https://doi.org/10.1175/1520-0469(1982)039<0395:CEWSWM>2.0.CO;2.
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, https://doi.org/10.1175/1520-0469(1984)041<1169:HSAEAS>2.0.CO;2.
Willoughby, H. E., R. W. R. Darling, and M. E. Rahn, 2006: Parametric representation of the primary hurricane vortex. Part II: A new family of sectionally continuous profiles. Mon. Wea. Rev., 134, 1102–1120, https://doi.org/10.1175/MWR3106.1.
Wu, C.-C., Y.-H. Huang, and G. Lien, 2012: Concentric eyewall formation in Typhoon Sinlaku (2008). Part I: Assimilation of T-PARC data based on the ensemble Kalman filter (EnKF). J. Atmos. Sci., 140, 506–527, https://doi.org/10.1175/MWR-D-11-00057.1.
Wu, C.-C., Y.-H. Huang, and Z. Tan, 2016: Secondary eyewall formation in tropical cyclones. Dynamics and Predictability of Large-Scale, High-Impact Weather and Climate Events, Cambridge University Press, 168–175.
Zhang, F., D. Tao, Y. Q. Sun, and J. D. Kepert, 2016: Dynamics and predictability of secondary eyewall formation in sheared tropical cyclones. J. Adv. Model. Earth Syst., 9, 89–112, https://doi.org/10.1002/2016MS000729.
Zhu, Z., and P. Zhu, 2014: The role of outer rainband convection in governing the eyewall replacement cycle in numerical simulations of tropical cyclones. J. Geophys. Res. Atmos., 119, 8049–8072, https://doi.org/10.1002/2014JD021899.
We here consider the atmosphere in which the interior layer is located above the boundary layer. The basic concept of the present theory can be applied to the ocean and other flows.