Hostname: page-component-745bb68f8f-d8cs5 Total loading time: 0 Render date: 2025-02-12T01:37:52.997Z Has data issue: false hasContentIssue false

Dynamics of a stratified vortex under the complete Coriolis force: two-dimensional three-components evolution

Published online by Cambridge University Press:  24 October 2022

Iman Toghraei*
Affiliation:
LadHyX, CNRS, École polytechnique, Institut Polytechnique de Paris, 91120 Palaiseau, France
Paul Billant
Affiliation:
LadHyX, CNRS, École polytechnique, Institut Polytechnique de Paris, 91120 Palaiseau, France
*
Email address for correspondence: iman.toghraei@ladhyx.polytechnique.fr

Abstract

We study the dynamics of an initially axisymmetric and vertical Lamb–Oseen vortex in a stratified-rotating fluid under the complete Coriolis force on the $f$ plane, i.e. in the presence of a background rotation both along the vertical and horizontal directions. By a combination of direct numerical simulations and asymptotic analyses for small horizontal background rotation, we show that a critical layer appears at the radius where the angular velocity of the vortex is equal to the buoyancy frequency when the Froude number is larger than unity. This critical layer generates a vertical velocity that is invariant along the vertical and which first increases linearly with time and then saturates at an amplitude scaling like $Re^{1/3}/\widetilde{Ro}$, where $Re$ is the Reynolds number and $\widetilde {Ro}$ the non-traditional Rossby number based on the horizontal component of the background rotation. In turn, a quasi-axisymmetric anomaly of vertical vorticity is produced at the critical radius through the non-traditional Coriolis force. Below a critical $\widetilde {Ro}$ depending on $Re$, the Rayleigh's inflectional criterion is satisfied and a shear instability is subsequently triggered rendering the vertical vorticity fully non-axisymmetric. The decay of the angular velocity is then enhanced until it is everywhere lower than the buoyancy frequency. A theoretical criterion derived from the Rayleigh condition predicts well the instability. It shows that this phenomenon can occur even for a large non-traditional Rossby number $\widetilde {Ro}$ for large $Re$. Hence, the non-traditional Coriolis force might have much more impact on geophysical vortices than anticipated by considering the order of magnitude of $\widetilde {Ro}$.

Type
JFM Papers
Copyright
© The Author(s), 2022. Published by Cambridge University Press

1. Introduction

The Coriolis force due to the planetary rotation is an essential ingredient of geophysical fluid dynamics. When studying its effect on fluid motions, it is common to use the so-called traditional approximation that takes into account only the vertical component, $\varOmega _b \sin (\phi )$, of the planetary angular velocity vector $\boldsymbol {\varOmega _b}$ at a given latitude $\phi$. Its horizontal component $\varOmega _b \cos (\phi )$ is neglected mainly because the associated Coriolis force (called the non-traditional Coriolis force) involves vertical motions or appears in the vertical momentum equation, whereas geophysical flows are usually in hydrostatic balance with weak vertical motions compared with horizontal motions (Gerkema et al. Reference Gerkema, Zimmerman, Maas and Van Haren2008).

However, Gerkema et al. (Reference Gerkema, Zimmerman, Maas and Van Haren2008) have reviewed several circumstances where the effect of the non-traditional Coriolis force becomes non-negligible. This occurs, for example, when the vertical velocity is not small such as for the instability of Ekman layers (Wippermann Reference Wippermann1969; Etling Reference Etling1971) and for deep convection (Sheremet Reference Sheremet2004). In the latter case, its effect is particularly intuitive since convective cells become slanted along the axis of rotation instead of the direction of gravity. The non-traditional Coriolis force may have also many effects on equatorial flows (Hayashi & Itoh Reference Hayashi and Itoh2012; Igel & Biello Reference Igel and Biello2020) and on the propagation and frequency range of internal waves, especially when the stratification is weak (Gerkema et al. Reference Gerkema, Zimmerman, Maas and Van Haren2008). Recently, the non-traditional force has been shown to significantly modify several instabilities: the inertial instability (Tort, Ribstein & Zeitlin Reference Tort, Ribstein and Zeitlin2016; Kloosterziel, Carnevale & Orlandi Reference Kloosterziel, Carnevale and Orlandi2017), the symmetric instability (Zeitlin Reference Zeitlin2018) and the shear instability (Park et al. Reference Park, Prat, Mathis and Bugnet2021). Tort & Dubos (Reference Tort and Dubos2014) and Tort et al. (Reference Tort, Dubos, Bouchut and Zeitlin2014) have also derived shallow water models taking into account the complete Coriolis force.

In the case of vortices, Lavrovskii et al. (Reference Lavrovskii, Semenova, Slezkin and Fominykh2000) and Semenova & Slezkin (Reference Semenova and Slezkin2003) have shown analytically that the equilibrium shape of a meddy-like anticyclonic vortex in a stratified fluid is slightly tilted with respect to the horizontal in the presence of the full Coriolis force. However, they have assumed that the vortex has a uniform vorticity and is embedded within a fluid a rest. Hence, there exist both a discontinuity of vorticity and velocity at the vortex boundary. Here, we study numerically and theoretically the evolution of a vortex with a continuous distribution of vorticity under the complete Coriolis force. The vortex is initially axisymmetric and columnar with a vertical axis in a stratified-rotating fluid under the Boussinesq and $f$-plane approximations. Since there is a misalignment between the buoyancy force and the rotation vector, this configuration is somewhat similar to the tilted vortex in a stratified non-rotating fluid considered by Boulanger, Meunier & Le Dizes (Reference Boulanger, Meunier and Le Dizes2007, Reference Boulanger, Meunier and Le Dizes2008). They have shown that a critical layer develops at the radius where the angular velocity of the vortex is equal to the Brunt–Väisälä frequency. Near this critical layer, they observed an intense axial flow and strong density variations that are uniform along the vortex axis but that lead to a three-dimensional shear instability under certain circumstances. We will show that a similar critical layer develops in the present configuration when the Froude number is larger than unity. For some parameters, an instability will be also triggered but it will be two dimensional instead of being three dimensional and due to a different mechanism. In addition, we will show that the critical layer evolution contains two different regimes: first, an unsteady inviscid phase followed by a second viscous phase that can be steady or can evolve nonlinearly depending on the parameters. Such evolution is similar to that evidenced by Wang & Balmforth (Reference Wang and Balmforth2020, Reference Wang and Balmforth2021) in their studies of forced baroclinic critical layers. They have also reported the subsequent development of a two-dimensional shear instability and studied its nonlinear evolution by means of a reduced model. Our investigations are based on direct numerical simulations (DNS) coupled to asymptotic analyses of the critical layer in the limit of a small non-traditional Coriolis parameter following the lines of Boulanger et al. (Reference Boulanger, Meunier and Le Dizes2007, Reference Boulanger, Meunier and Le Dizes2008), Wang & Balmforth (Reference Wang and Balmforth2020, Reference Wang and Balmforth2021).

As a preliminary remark, we stress that the present study has been first carried out by means of three-dimensional numerical simulations. However, the flow turned out to remain independent of the vertical coordinate although the vertical velocity is non-zero. In other words, the dynamics were two dimensional but with three components of velocity (2D3C). As we will see later, this can be easily understood from the governing equations. For this reason, the subsequent simulations reported in this paper have been restricted to a two-dimensional configuration. However, in a future paper, we will show that the introduction of infinitely small three-dimensional perturbations may also lead to, for some parameters, a full three dimensionalization of the vortex, i.e. a 3D3C dynamics, via an axial shear instability similar to that reported by Boulanger et al. (Reference Boulanger, Meunier and Le Dizes2007, Reference Boulanger, Meunier and Le Dizes2008). We stress that the two-dimensional dynamics is still observed in this full three-dimensional configuration in a significant range of the parameters space.

The paper is organized as follows. The problem is first formulated in § 2. Direct numerical simulations are described in § 3. Asymptotic analyses are conducted for a small non-traditional Coriolis parameter in § 4. In § 5 the numerical and asymptotic results are compared. The origin of the full non-axisymmetric dynamics of the vortex will be investigated in § 6. Finally, the late evolution is discussed in § 7 and the conclusions are given in § 8.

2. Formulation of the problem

2.1. Governing equations

We use the incompressible Navier–Stokes equations under the Boussinesq approximation

(2.1)$$\begin{gather} \boldsymbol{\nabla}\boldsymbol{\cdot}\boldsymbol{u} = 0, \end{gather}$$
(2.2)$$\begin{gather}\frac{\partial \boldsymbol{u}}{\partial t} + \left( \boldsymbol{u}\boldsymbol{\cdot}\boldsymbol{\nabla} \right) \boldsymbol{u} ={-}\boldsymbol{\nabla} \left( \frac{p} {\rho_0} \right) + b \boldsymbol{e_z} - 2\boldsymbol{\varOmega_b} \times \boldsymbol{u} + \nu {\nabla}^{2}\boldsymbol{u}, \end{gather}$$
(2.3)$$\begin{gather} \frac{\partial b}{\partial t} + \boldsymbol{u}\boldsymbol{\cdot}\boldsymbol{\nabla} b + N^{2} u_z = \kappa \nabla^{2}b, \end{gather}$$

where $\boldsymbol {u}$ is the velocity field, $p$ is the pressure and $b = -g \rho /\rho _0$ is the buoyancy, $g$ is the gravity, $\rho$ is the density perturbation and $\rho _0$ is a constant reference density. Here $\boldsymbol {e_z}$ is the vertical unit vector and $\boldsymbol {\varOmega _b}$ is the background rotation vector. It is assumed to have not only a vertical component but also a horizontal component along the $y$ direction: $2\boldsymbol {\varOmega _b} = \tilde {f} \boldsymbol {e_y} + f \boldsymbol {e_z}$, where $f = 2\varOmega _b \sin {(\phi )}$ and $\tilde {f} = 2 \varOmega _b \cos {(\phi )}$ with $\phi$ the latitude or, equivalently, the angle between the background rotation vector and the unit vector in the $y$ direction, $\boldsymbol {e}_y$ (figure 1). Here $\nu$ and $\kappa$ are the viscosity of the fluid and diffusivity of the stratifying agent, respectively. The total density field $\rho _t$ reads $\rho _t(\boldsymbol {x},t)=\rho _0+\bar {\rho }(z)+\rho (x,t)$, where $\bar {\rho }$ is the mean density profile along the $z$ axis. The Brunt–Väisälä frequency

(2.4)\begin{equation} N = \sqrt{-\frac{g}{\rho_0}\frac{{\rm d}\bar{\rho}}{{\rm d}z}} \end{equation}

is assumed to be constant.

Figure 1. Sketch of the initial vortex in a stratified fluid and in the presence of a background rotation $\varOmega _b$ inclined with an angle $\phi$.

2.2. Initial conditions

A single vertical vortex with a Lamb–Oseen profile is taken as initial conditions. Its vorticity field reads

(2.5)\begin{equation} \boldsymbol{\omega} (\boldsymbol x,t=0) = \zeta \boldsymbol{e_z} = \frac{\varGamma}{{\rm \pi} a_0^{2}}\exp({-r^{2}/a_0^{2}}) \boldsymbol{e_z}, \end{equation}

where $\varGamma$ is the circulation, $a_0$ is the radius and $r$ is the radial coordinate. The geometry of the flow is sketched in figure 1.

2.3. Non-dimensionalization

Equations (2.1)–(2.3) are non-dimensionalized by using $2{\rm \pi} a_0^{2}/\varGamma$ and $a_0$ as time and length units,

(2.6)$$\begin{gather} \boldsymbol{\nabla}\boldsymbol{\cdot}{\boldsymbol{u}} = 0, \end{gather}$$
(2.7)$$\begin{gather}\frac{\partial \boldsymbol{u}}{\partial t} + \left( \boldsymbol{u}\boldsymbol{\cdot}\boldsymbol{\nabla} \right) \boldsymbol{u} ={-}\boldsymbol{\nabla}p + b \boldsymbol{e_z} - 2\left ( \frac{1}{Ro} \boldsymbol{e_z} + \frac{1}{\widetilde{Ro}} \boldsymbol{e_y} \right) \times \boldsymbol{u} + \frac{1}{Re} {\nabla}^{2} \boldsymbol{u}, \end{gather}$$
(2.8)$$\begin{gather}\frac{\partial b}{\partial t} + \boldsymbol{u}\boldsymbol{\cdot}\boldsymbol{\nabla} b + \frac{1}{F_h^{2}} u_z = \frac{1}{ReSc} \nabla^{2}b, \end{gather}$$

where the same notation has been kept for the non-dimensional variables. Note that $\rho _0$ has been eliminated by redefining the pressure $p$. The Reynolds, Froude, Rossby and Schmidt numbers are defined as

(2.9ae)\begin{equation} Re = \frac{\varGamma}{2{\rm \pi}\nu}, \quad F_h = \frac{\varGamma}{2{\rm \pi} a_0^{2} N}, \quad Ro = \frac{\varGamma}{{\rm \pi} a_0^{2}\,f}, \quad \widetilde{Ro} = \frac{\varGamma}{{\rm \pi} a_0^{2}\,\tilde{f}}, \quad Sc=\frac{\nu}{\kappa}. \end{equation}

Note that two Rossby numbers $Ro$ and $\widetilde {Ro}$ are defined based on the two components of the rotation vector. The Schmidt number is always set to unity. In the following, all results will be reported in non-dimensional form.

2.4. Numerical method

A pseudo-spectral method with periodic boundary conditions is used to integrate (2.1)–(2.3) in space (Deloncle, Billant & Chomaz Reference Deloncle, Billant and Chomaz2008). Time integration is performed with a fourth-order Runge–Kutta scheme. Most of the aliasing is removed by truncating the top one third of the modes along each direction. The viscous and diffusive terms are integrated exactly. The horizontal sizes of the computational domain have been set to $l_x=l_y=18$. As shown by Bonnici (Reference Bonnici2018) and Billant & Bonnici (Reference Billant and Bonnici2020), this is sufficiently large to minimize the effects of the periodic boundary conditions and to give results independent of the box sizes. In particular, the periodic boundary condition imposes that the net circulation over the domain is zero implying that the initial vorticity is not exactly (2.5) but $\zeta ^{\prime } = \zeta - \varGamma /(l_xl_y)$. However, with $l_x=l_y=18$, the artificial background vorticity $\varGamma /(l_xl_y)$ is weak and represents only $1\,\%$ of the maximum vorticity of the vortex. The horizontal resolution has been varied from $n_x=n_y=512$ for $Re=2000$ up to $n_x=n_y=1024$ for $Re=10\,000$. As mentioned in the introduction, preliminary simulations were fully three dimensional with a resolution $n_z$ and a vertical size $l_z$ similar to the horizontal ones. However, the flow is always observed to remain independent of the vertical coordinate. It can be seen indeed from (2.6)–(2.8) that if $\partial / \partial z = 0$ at $t=0$ then the flow will remain independent of the vertical coordinate for all time. Therefore, only two-dimensional simulations but with three components of velocity will be presented in the following. Several tests using different horizontal resolutions have been performed in order to verify the accuracy of the computations. For $Re=2000$, the velocity has been found to differ by less than $0.1\,\%$ when the resolution is increased from $512 \times 512$ to $1024 \times 1024$. Similarly, for $Re=10\,000$, the relative variation of the velocity is less than $0.5\,\%$ when the resolution is increased from $1024 \times 1024$ to $1536 \times 1536$.

3. Direct numerical simulations

3.1. Illustrative example of the vortex dynamic

To get an overview of the effect of the complete Coriolis force, we start by presenting the vortex evolution for the sample set of parameters $Re=2000$, $F_h=10$, $Ro=23.1$, $\phi =60^{\circ }$ ($\widetilde {Ro}=40$). Figures 2 and 3 show the evolution of the vertical velocity and vertical vorticity at six different times (a movie is available in the supplementary material available online at https://doi.org/10.1017/jfm.2022.812). Initially, the vortex is completely axisymmetric (figure 3a) and the vertical velocity is zero (figure 2a). As time goes on, a vertical velocity field with an azimuthal wavenumber $m=1$ develops (figure 2b). This structure tends to intensify and becomes more and more concentrated at a particular radius (figure 2c). Concomitantly, a ring of negative vertical vorticity appears and grows at the same radius (figure 3b,c). Later, the vertical velocity structure and the ring of anomalous vertical vorticity are no longer perfectly circular (figure 3df). Two negative vortices appear on the vertical vorticity ring and revolve around the vortex centre (figure 3ef). Simultaneously, the shape of the vertical velocity structure is deformed similarly (figure 2ef). As already mentioned, preliminary three-dimensional simulations with various vertical sizes $l_z$ of the computational domain and resolutions $n_z$ have shown that the velocity and vorticity fields remain always completely independent of the vertical coordinate, i.e. the same evolution is observed in any horizontal cross-section. It is also important to stress that this phenomenon occurs only in the presence of the complete Coriolis force. Indeed, if $\widetilde {Ro} = \infty$, the vertical velocity remains identically zero while the vertical vorticity simply decays by viscous diffusion.

Figure 2. Vertical velocity at different times: $(a)$ $t=1$ , $(b)$ $t=30$ , $(c)$ $t=60$, $(d)$ $t=75$, $(e)$ $t=80$ and $(f)$ $t=90$ for $Re=2000$, $F_h=10$, $Ro=23.1$, $\phi =60^{\circ }$ ($\widetilde {Ro}=40$).

Figure 3. Vertical vorticity at different times: $(a)$ $t=1$ , $(b)$ $t=30$ , $(c)$ $t=60$, $(d)$ $t=75$, $(e)$ $t=80$ and $(f)$ $t=90$ for $Re=2000$, $F_h=10$, $Ro=23.1$, $\phi =60^{\circ }$ ($\widetilde {Ro}=40$).

From figure 3 we can distinguish two phases in the evolution of the vortex. First a circular ring of anomalous vertical vorticity develops and then this ring becomes non-axisymmetric. A more in-depth analysis of these two phases will be discussed later. Let us first examine the effects of the control parameters on this phenomenon.

3.2. Effects of the stratification

When the Froude number is decreased from $F_h=10$ to $F_h=2$, the same evolution of the vertical velocity (figure 4ad) – see the movie in the supplementary material – and the vertical vorticity (figure 4eh) is observed but at a smaller radius. However, the anomaly of the vertical vorticity does not become negative this time and the maximum vertical velocity is also lower than for $F_h=10$. Figure 5(a) shows the evolution of the vertical velocity maximum $u_{zm}$. In fact, three phases can be distinguished. First, $u_{zm}$ increases linearly with time with some small oscillations superimposed that will be later attributed to inertia-gravity waves. The propagation of waves can be also seen in the vertical velocity field at the beginning of the movies. Then, $u_{zm}$ saturates and tends to slightly decrease. When $t\gtrsim 60$, large oscillations arise when the vortex becomes fully non-axisymmetric.

Figure 4. Vertical velocity (ad) and vertical vorticity (eh) at different times: (a,e) $t=10$, (bf) $t=40$, (c,g) $t=65$ and (d,h) $t=80$ for $Re=2000$, $F_h=2$, $Ro=23.1$, $\phi =60^{\circ }$ ($\widetilde {Ro}=40$).

Figure 5. Maximum vertical velocity as a function of time for $F_h=2$ and (a) $Re=2000$, $Ro=23.1$, $\phi =60^{\circ }$ ($\widetilde {Ro}=40$), (b) $Re=2000$, $Ro=20.3$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$) and (c) $Re=10\,000$, $Ro=20.3$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$).

If the Froude number is below unity, such evolution is no longer observed. In § 4 it will be shown that this phenomenon is due to the presence of a critical layer where the angular velocity of the vortex is equal to the non-dimensional Brunt–Väisälä frequency, i.e. $\varOmega = 1/F_h$, where $\varOmega$ is the non-dimensional angular velocity of the vortex corresponding to the vorticity field (2.5).

3.3. Effects of the Rossby numbers

When the traditional Rossby number $Ro$, which is based on the vertical component of the background rotation, is varied, while keeping the other numbers fixed ($Re$, $F_h$, $\widetilde {Ro}$), the vortex evolution remains strictly identical. This is consistent with the fact that the dynamics is independent of the vertical coordinate so that the Coriolis force associated with $Ro$ can be eliminated from (2.7) by redefining the pressure.

In contrast, varying the non-traditional Rossby number $\widetilde {Ro}$, which is based on the horizontal component of the background rotation, has important effects on the evolution of the vortex. Since the Rossby number $Ro$ has no effect, the effect of $\widetilde {Ro}$ has been studied by varying the latitude $\phi$ while keeping the background rotation rate $\varOmega _b$ constant. Hence, both $\widetilde {Ro}$ and $Ro$ varies. Figure 6 shows the evolution of the vertical velocity and vertical vorticity when the latitude is increased from $\phi = 60^{\circ }$ to $\phi =80^{\circ }$ ($\widetilde {Ro}$ is increased from $\widetilde {Ro}=40$ to $\widetilde {Ro}=115.2$) while keeping the other parameters fixed (a movie is available in the supplementary material). The initial evolution (figure 6a,b,ef) is similar to that in figure 4 but, later (figure 6c,d,g,h), the fields keep the same shape, i.e. no significant asymmetric deformations can be seen in contrast to figure 4(d,h). As seen in figure 5(b), only two phases are then present in the evolution of the maximum vertical velocity. First, a phase where $u_{zm}$ grows linearly with weak oscillations and, second, a phase where $u_{zm}$ remains approximately constant. In addition, the maximum vertical velocity is lower (figure 6b,c) and the anomalous vorticity ring is weaker (figure 6g) than in figure 4. At late time $t=200$ (figure 6d,h), we can see that the vertical velocity has decreased by viscous diffusion while the vertical velocity field has moved towards the centre of the vortex. This is consistent with the critical layer's interpretation since, as the angular velocity decays, the radius where $\varOmega = 1/F_h$ decreases.

Figure 6. Vertical velocity (ad) and vertical vorticity (eh) at different times: (a,e) $t=10$, (bf) $t=50$, (c,g) $t=150$ and (d,h) $t=200$ for $Re=2000$, $F_h=2$, $Ro=20.3$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$).

3.4. Effects of the Reynolds number

Figure 7 shows the evolution of the vertical velocity and vertical vorticity when the Reynolds number is increased from $Re=2000$ to $Re=10\,000$ while keeping the other parameters as in figure 6. A movie is also available in the supplementary material. The maximum vertical velocity is almost doubled and the vertical velocity field is much thinner and focused near a given radius (figure 7b) than in figure 6. Furthermore, the ring of anomalous vertical vorticity (figure 7f) is more intense. Later, asymmetric deformations of this ring and of the vertical velocity field are clearly visible (figure 7c,d,g,h) in contrast to figure 6. Oscillations are then visible in the evolution of $u_{zm}$ (figure 5c).

Figure 7. Vertical velocity (ad) and vertical vorticity (eh) at different times: (a,e) $t=10$, (bf) $t=100$, (c,g) $t=150$ and (d,h) $t=200$ for $Re=10\,000$, $F_h=2$, $Ro=20.3$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$).

3.5. Combined effects of $\widetilde {Ro}$ and $Re$

In the previous sections we have seen that the vertical vorticity becomes fully non-axisymmetric in a second stage if $\phi$ is sufficiently lower than $90^{\circ }$, i.e. if $\widetilde {Ro}$ is not too large (figure 4) or if the Reynolds number is large enough (figure 7), otherwise the vertical vorticity field remains quasi-axisymmetric (figure 6). Figure 8 summarizes several other simulations for various $\widetilde {Ro}$ and Reynolds numbers $Re$ keeping the Froude number equal to $F_h=2$. Yellow and blue symbols indicate simulations where the vertical vorticity remains quasi-axisymmetric or becomes non-axisymmetric, respectively. We can see that the critical Rossby number $\widetilde {Ro}_c$ above which the vortex remains quasi-axisymmetric increases with the Reynolds number from $\widetilde {Ro}_c \simeq 100$ for $Re=2000$ to $\widetilde {Ro}_c \simeq 400$ for $Re=10\,000$.

Figure 8. Map of the simulations in the parameter space $(Re,\widetilde {Ro})$ for $F_h=2$. The yellow and blue circles represent simulations where the vertical vorticity remains quasi-axisymmetric or not, respectively. The solid and dashed lines represent the criterion (6.14) for different values of (a,c): $(\infty,0)$ and $(\infty,0.4)$, respectively. The number near some points indicate the figure numbers where the corresponding simulation is shown.

4. Asymptotic analyses

In order to understand the vortex evolution observed in the DNS, it is interesting to perform an asymptotic analysis for a small horizontal component of the background rotation, i.e. for $\widetilde {Ro} \gg 1$, and for a large Reynolds number. To this end, it is first convenient to rewrite (2.6)–(2.8) in cylindrical coordinates $(r, \theta,z)$,

(4.1a)\begin{align} \frac{1}{r} \frac{\partial r u_r}{\partial r} + \frac{1}{r} \frac{\partial u_\theta}{\partial \theta} + \frac{\partial u_z}{\partial z} &= 0, \end{align}
(4.1b)\begin{align} \frac{\partial u_r}{\partial t} + u_r \frac{\partial u_r}{\partial r} + \frac{u_\theta}{r} \frac{\partial u_r}{\partial \theta} + u_z \frac{\partial u_r}{\partial z} - \frac{{u_\theta}^{2}}{r} &={-} \frac{\partial p}{\partial r} + \frac{2 u_\theta}{Ro} - \frac{2u_z}{\widetilde{Ro}} \cos\left(\theta\right) \nonumber\\ &\quad + \frac{1}{Re}\left ( \nabla^{2} u_r - \frac{2}{r^{2}}\,\frac{\partial u_\theta}{\partial\theta} \right), \end{align}
(4.1c)\begin{align} \frac{\partial u_\theta}{\partial t} + u_r \frac{\partial u_\theta}{\partial r} + \frac{u_\theta}{r} \frac{\partial u_\theta}{\partial\theta} + u_z \frac{\partial u_\theta}{\partial z} + \frac{u_r u_\theta}{r} &={-} \frac{1}{r} \frac{\partial p}{\partial \theta} - \frac{2 u_r}{Ro} + \frac {2u_z}{\widetilde{Ro}}\sin\left(\theta\right) \nonumber\\ &\quad + \frac{1}{Re}\left ( \nabla^{2} u_\theta + \frac{2}{r^{2}}\,\frac{\partial u_r}{\partial \theta} \right), \end{align}
(4.1d)\begin{align} \frac{\partial u_z}{\partial t} + u_r \frac{\partial u_z}{\partial r} + \frac{u_\theta}{r} \frac{\partial u_z}{\partial \theta} + u_z \frac{\partial u_z}{\partial z}&={-} \frac{\partial p}{\partial z} + b +\frac{2u_r}{\widetilde{Ro}} \cos\left(\theta\right) - \frac{2 u_\theta}{\widetilde{Ro}}\sin\left(\theta\right) \nonumber\\ &\quad +\frac{1}{Re} \nabla^{2} u_z, \end{align}
(4.1e)\begin{align} \frac{\partial b}{\partial t} + u_r \frac{\partial b}{\partial r} + \frac{u_\theta}{r} \frac{\partial b}{\partial\theta} + u_z \frac{\partial b}{\partial z} + \frac{u_z}{F_h^{2}} &= \frac{1}{ReSc}\nabla^{2} b. \end{align}

It is also convenient to consider the equation for the vertical vorticity $\zeta$,

(4.2)\begin{align} \frac{\partial\zeta}{\partial t} + \boldsymbol{u}\boldsymbol{\cdot}\boldsymbol{\nabla}\zeta &= \left(\boldsymbol{\omega}+\frac{2}{Ro} \boldsymbol{e_z} \right) \boldsymbol{\cdot}\boldsymbol{\nabla}u_z \nonumber\\ &\quad + \frac{2}{r\widetilde{Ro}}\left[ \frac{\partial }{\partial r} \left (r {u_z} \sin{(\theta)}\right) + \frac{\partial}{\partial \theta} \left({u_z} \cos{(\theta)} \right) \right] + \frac{1}{Re} \Delta \zeta, \end{align}

where $\boldsymbol {\omega } = \boldsymbol {\nabla }\times \boldsymbol {u}$.

The solution is expanded with the small parameter $\varepsilon = 2 / \widetilde {Ro} \ll 1$ in the form

(4.3)\begin{equation} (u_r,u_\theta,u_z,p,b) = (0, u_{\theta 0}, 0, p_0, 0) + \varepsilon ({u_r}_1,u_{\theta 1},u_{z1},{p}_1,{b}_1) + \dots, \end{equation}

where $u_{\theta 0} = r \varOmega$, with $\varOmega = (1-{\rm e}^{-r^{2}})/r^{2}$, is the non-dimensional angular velocity of the vortex corresponding to the vorticity field (2.5).

It is first instructive to consider a steady and non-diffusive flow, i.e. $\partial / \partial t =0$ and ${Re=\infty}$. Then, (4.1b) reduces at leading order to the cyclostrophic balance

(4.4)\begin{equation} -\frac{u_{\theta 0}^{2}}{r} ={-} \frac{\partial p_0}{\partial r} + \frac{2u_{\theta 0}}{Ro}, \end{equation}

whereas (4.1a), (4.1c)–(4.1e) are identically zero at leading order. At first order in $\varepsilon$, it is sufficient to consider only (4.1d) and (4.1e),

(4.5a)$$\begin{gather} \varOmega \frac{\partial u_{z1}}{\partial \theta} ={-} r\varOmega \sin{(\theta)} + b_{1} , \end{gather}$$
(4.5b)$$\begin{gather}\varOmega \frac{\partial b_{1}}{\partial\theta} ={-} \frac{u_{z1}}{F_h^{2}}. \end{gather}$$

The solution is

(4.6a)$$\begin{gather} u_{z1} = \frac{r\varOmega^{2}F_h^{2}}{F_h^{2}\varOmega^{2}-1} \cos{(\theta)}, \end{gather}$$
(4.6b)$$\begin{gather}{b}_{1} = \frac{-r\varOmega}{F_h^{2}\varOmega^{2}-1} \sin{(\theta)}, \end{gather}$$

showing that the Coriolis force due to the horizontal component of the background rotation (first term on the right-hand side of (4.5a)) forces a vertical velocity and buoyancy fields. These fields are independent of the vertical coordinate as observed in the DNS. However, we can remark that (4.6a)–(4.6b) are singular if there exists a radius $r_c$ where $\varOmega (r_c) = 1/F_h$. Such a critical radius exists wherever $F_h > 1$ since the non-dimensional angular velocity decreases from unity on the vortex axis to zero for $r \to \infty$.

A similar critical layer occurs in the case of a tilted vortex in a stratified fluid (Boulanger et al. Reference Boulanger, Meunier and Le Dizes2007) and in a stratified-rotating shear flow (Wang & Balmforth Reference Wang and Balmforth2020). This singularity can be smoothed if the flow is no longer assumed to be steady or inviscid. Although these two effects can a priori operate simultaneously, the unsteadiness turns out to be, first, the dominant effect while diffusive effects are negligible followed by a second phase where it is the opposite.

4.1. Unsteady inviscid analysis

Accordingly, we first consider (4.1a)–(4.1e) in the inviscid limit $Re = \infty$, but keep the time derivatives. At leading order, the equations reduce to

(4.7)$$\begin{gather} -\frac{u_{\theta 0}^{2}}{r} = \frac{\partial p_0}{\partial r} + \frac{2u_{\theta 0}}{Ro}, \end{gather}$$
(4.8)$$\begin{gather}\frac{\partial u_{\theta 0}}{\partial t} = 0, \end{gather}$$

so that $u_{\theta 0} = r \varOmega$ as before. At first order, (4.1d)–(4.1e) become

(4.9a)$$\begin{gather} \frac{\partial u_{z1}}{\partial t} + \varOmega \frac{\partial u_{z1}}{\partial \theta} = b_{1} - r\varOmega \sin{(\theta)}, \end{gather}$$
(4.9b)$$\begin{gather}\frac{\partial b_{1}}{\partial t} + \varOmega \frac{\partial b_{1}}{\partial \theta} ={-} \frac{u_{z1}}{F_h^{2}}. \end{gather}$$

The only difference with (4.6) is the presence of the time derivatives. By imposing ${u_{z1} = b_1 = 0}$ at $t = 0$, the solutions can be found in the form

(4.10a)$$\begin{gather} u_{z1} = {u_z}_{p} \mathrm{e}^{{\rm i}\theta} + {u^{*}_z}_{p} \mathrm{e}^{-{\rm i}\theta}, \end{gather}$$
(4.10b)$$\begin{gather}{b}_{1}= {b}_{p} \mathrm{e}^{{\rm i}\theta} + {b^{*}}_{p} \mathrm{e}^{-{\rm i}\theta}, \end{gather}$$

where the star denotes the complex conjugate and

(4.11a)$$\begin{gather} {u_z}_{p} = \frac{r \varOmega }{4} \left[- \frac{1}{\alpha} \left( 1 - \mathrm{e}^{{\rm i} \alpha t} \right) + \frac{1}{\beta} \left( 1 - \mathrm{e}^{-{\rm i} \beta t} \right) \right], \end{gather}$$
(4.11b)$$\begin{gather}{b}_{p} ={-} {\rm i} \frac{r \varOmega }{4 F_h} \left[ \frac{1}{\alpha} \left( 1 - \mathrm{e}^{{\rm i} \alpha t} \right)+ \frac{1}{\beta} \left( 1 - \mathrm{e}^{-{\rm i} \beta t} \right) \right], \end{gather}$$

with

(4.12a,b)\begin{equation} \alpha = \frac{1-F_h\varOmega}{F_h}, \quad \beta =\frac{1+F_h\varOmega}{F_h}. \end{equation}

Compared with the steady solution (4.6), the additional terms present in (4.11) correspond to waves generated at $t=0$ to satisfy the initial conditions. These waves oscillate at the frequencies $1/F_h - \varOmega$ and $-1/F_h - \varOmega$, i.e. the non-dimensional Brunt–Väisälä frequency with an additional Doppler shift coming from the azimuthal motion of the vortex. Hence, they correspond to inertia-gravity waves with a zero vertical wavenumber. In contrast to (4.6), we see now that the vertical velocity and buoyancy (4.10a)–(4.10b) are no longer singular at the radius $r_c$ where $\varOmega (r_c) = 1/F_h$. Indeed, we have $(1-e^{i\alpha t})/\alpha \simeq -i t$ when $\alpha \to 0$ so that (4.11a)–(4.11b) remain finite at $r=r_c$.

Following Wang & Balmforth (Reference Wang and Balmforth2020), the behaviour of these solutions in the vicinity of the critical radius can be studied more precisely by introducing the variable $\eta = r - r_c$. When $\eta \ll 1$, the vertical velocity approximates to

(4.13)\begin{align} u_{z1} &= \left[\! \left(\!\frac{r_c \varOmega_c}{2\eta \varOmega^{\prime}_c} + \frac{\varOmega_c}{2\varOmega^{\prime}_c} + \frac{r_c}{2} - \frac{r_c \varOmega^{\prime\prime}_c \varOmega_c }{{4\varOmega^{\prime}_c}^{2}} \!\right) \left(\!1\!-\cos{(\eta \varOmega^{\prime}_c t)}\!\right) + \frac{r_c}{4} \left(\!1\!-\cos{\left(\!\frac{2}{F_h}t\!\right)}\!\right) \!\right] \cos{(\theta)} \nonumber\\ &\quad - \left[ \left(\frac{r_c \varOmega_c}{2\eta \varOmega^{\prime}_c} + \frac{\varOmega_c}{2\varOmega^{\prime}_c} + \frac{r_c}{2} - \frac{r_c \varOmega^{\prime\prime}_c \varOmega_c }{4{\varOmega^{\prime}_c}^{2}} \right) \sin{(\eta \varOmega^{\prime}_c t)} + \frac{r_c}{4} \sin{\left(\frac{2}{F_h}t\right)} \right] \sin{(\theta)} \!+\! \textit{O} (\eta), \end{align}

where the subscript $c$ indicates a value taken at $r_c$. The terms involving $\eta \varOmega ^{\prime }_ct$ in the sin and cos functions have not been expanded in (4.13) since this quantity can be large when $t$ is large even for small $\eta$. The approximation (4.13) is therefore uniformly valid whatever $t$. Taking into account only the leading order, (4.13) can be simplified and rewritten as

(4.14)\begin{equation} u_{z1} = \frac{r_c \varOmega_c t}{2} \left[ \left(\frac{1-\cos{U}}{U}\right) \cos{(\theta)} - \frac{\sin{U}}{U}\sin{(\theta)} \right] + {O} (1), \end{equation}

where $U = \eta \varOmega ^{\prime }_c t$. This expression shows that the radial profile of the vertical velocity in the vicinity of $r_c$ depends only on the self-similar variable $U$. This implies that the vertical velocity will be more and more concentrated around $r_c$ as time increases. In addition, (4.14) shows that the amplitude of $u_{z1}$ will increase linearly with time for a fixed value of $U$. In practice, the approximation (4.14) will be accurate only when $t$ is large since the neglected terms are of order unity. Since $\varOmega _c = 1/F_h$, this will occur increasingly later when $F_h$ increases. We emphasize that the inertia-gravity wave oscillating at frequency $1/F_h+\varOmega _c=2/F_h$ is neglected in (4.14) unlike in (4.13).

4.2. Unsteady viscous analysis

Here, viscous and diffusive effects are taken into account in addition to the time evolution. Following Boulanger et al. (Reference Boulanger, Meunier and Le Dizes2007) and Wang & Balmforth (Reference Wang and Balmforth2020, Reference Wang and Balmforth2021), we assume that the Reynolds number is large $Re \gg 1$ and consider only the vicinity of the critical radius by introducing a rescaled radius such that $\tilde {r} = Re^{1/3}(r-r_c)$. We also assume that the evolution occurs over the slow time $T=Re^{-1/3}t$. Since $Re \gg 1$, the leading-order solution of (4.1a)–(4.1e) is still $u_{\theta 0} = r \varOmega$. At order $\varepsilon$, (4.1d)–(4.1e) become

(4.15a) \begin{align} & \frac{1}{Re^{1/3}}\frac{\partial u_{z1}}{\partial T}+ \left( \varOmega_c + \frac{\tilde{r} \varOmega_c^{\prime}}{Re^{1/3}} + {O} \left(\frac{1}{Re^{2/3}}\right) \right) \frac{\partial u_{z1}}{\partial \theta}\nonumber\\ &\quad = b_1 \!- \left( r_c \varOmega_c \!+ \frac{\tilde{r}(\varOmega_c+r_c\varOmega_c^{\prime})}{Re^{1/3}} + {O} \left(\frac{1}{Re^{2/3}}\right) \right) \sin{(\theta)} \!+ \frac{1}{Re^{1/3}} \frac{\partial^{2} u_{z1}}{\partial\tilde{r}^{2}} + {O} \left(\frac{1}{Re^{2/3}}\right) , \end{align}
(4.15b)\begin{align}& \frac{1}{Re^{1/3}} \frac{\partial b_{1}}{\partial T} \!+ \left(\!\varOmega_c + \frac{\tilde{r} \varOmega_c^{\prime}}{Re^{1/3}} + {O} \left(\!\frac{1}{Re^{2/3}}\!\right)\!\right) \frac{\partial b_1}{\partial\theta}={-} \frac{u_{z1}}{F_h^{2}} \!+ \frac{1}{Sc} \left[\!\frac{1}{Re^{1/3}} \frac{\partial^{2}{b}_1}{\partial\tilde{r}^{2}} + {O} \left(\!\frac{1}{Re^{2/3}}\!\right) \!\!\right] . \end{align}

These equations can be solved by expanding $u_{z1}$ and $b_1$ as

(4.16a)$$\begin{gather} {u_{z1}} = Re^{1/3} \left[ \tilde{u}_{z1}(\tilde{r},T) \mathrm{e}^{{\rm i}\theta} +{\rm c.c.} \right]+ \tilde{u}_{z2}(\tilde{r},T) \mathrm{e}^{{\rm i}\theta} +{\rm c.c.} + \dots, \end{gather}$$
(4.16b)$$\begin{gather}b_1 = Re^{1/3} \left[ \tilde{b}_1(\tilde{r},T) \mathrm{e}^{{\rm i}\theta} + {\rm c.c.} \right] + \tilde{b}_2(\tilde{r},T) \mathrm{e}^{{\rm i}\theta} +{\rm c.c.} + \dots\ . \end{gather}$$

By substituting (4.16a)–(4.16b) in (4.15a)–(4.15b), we get at order ${O}(Re^{1/3})$,

(4.17a)$$\begin{gather} {\rm i} \varOmega_c \tilde{u}_{z1} = \tilde{b}_1 , \end{gather}$$
(4.17b)$$\begin{gather}{\rm i} \varOmega_c \tilde{b}_1 ={-} \frac{ \tilde{u}_{z1}}{F_h^{2}}, \end{gather}$$

which both yield

(4.18)\begin{equation} \tilde{b}_1 = {\rm i} \varOmega_c \tilde{u}_{z1}. \end{equation}

The equations at order ${O} (1)$ are

(4.19a)$$\begin{gather} \frac{\partial\tilde{u}_{z1}}{\partial T} + {\rm i}\varOmega_c \tilde{u}_{z2} + {\rm i} \tilde{r}\varOmega_c^{\prime} \tilde{u}_{z1}= \tilde{b}_2 - \frac{r_c\varOmega_c}{2i} + \frac{{\rm d}^{2}\tilde{u}_{z1}}{{\rm d}\tilde{r}^{2}} , \end{gather}$$
(4.19b)$$\begin{gather}\frac{\partial\tilde{b}_{1}}{\partial T}+ {\rm i}\varOmega_c \tilde{b}_2 + {\rm i} \tilde{r}\varOmega_c^{\prime}\tilde{b}_1 ={-}\frac{1}{F_h^{2}} \tilde{u}_{z2} + \frac{1}{Sc} \frac{{\rm d}^{2}\tilde{b}_1}{{\rm d}\tilde{r}^{2}}. \end{gather}$$

By using (4.18), the solvability condition to find $\tilde {b}_2$ and $\tilde {u}_{z2}$ from (4.19a)–(4.19b) requires $\tilde {u}_{z1}$ to satisfy

(4.20)\begin{equation} \frac{\partial\tilde{u}_{z1}}{\partial T}+ {\rm i} \tilde{r}\varOmega_c^{\prime} \tilde{u}_{z1} = \frac{{\rm i}}{4} r_c\varOmega_c + \frac{1}{2} \left (1+\frac{1}{Sc}\right) \frac{{\rm d}^{2}\tilde{u}_{z1}}{{\rm d}\tilde{r}^{2}}. \end{equation}

As shown by Wang & Balmforth (Reference Wang and Balmforth2020, Reference Wang and Balmforth2021), the solution is

(4.21)\begin{equation} \tilde{u}_{z1} = {\rm i} A \frac{1}{\rm \pi} \int_0^{\left|{\varOmega_c^{\prime}}\right|T/ \gamma} \exp \left ( - \frac{z^{3}}{3} + {\rm i} \gamma \tilde{r} z \right) \mathrm{d}z, \end{equation}

where

(4.22a,b)\begin{equation} A=\frac{{\rm \pi} r_c \varOmega_c}{2 \left|{2\varOmega_c^{\prime}}\right|^{2/3}\left(1+\dfrac{1}{Sc}\right)^{1/3}}, \quad \gamma = \frac{ \left|{2\varOmega_c^{\prime}}\right|^{1/3}}{\left(1+\dfrac{1}{Sc}\right)^{1/3}}. \end{equation}

When $\left |{\varOmega _c^{\prime }}\right |T/ \gamma \gg 1$, the upper bound in the integral can be replaced by infinity so that (4.21) recovers the steady solution (Boulanger et al. Reference Boulanger, Meunier and Le Dizes2007)

(4.23)\begin{equation} \tilde{u}_{z1} = {\rm i} A \,{Hi}( {\rm i} \gamma \tilde{r}), \end{equation}

where ${Hi}$ is the Scorer's function (Abramowitz & Stegun Reference Abramowitz and Stegun1972). We will see that the time interval where both the unsteadiness and viscous effects are important is short so that (4.23) turns out to be reached quickly after the inviscid regime.

4.3. Effect on the vertical vorticity

We now turn to the study of the effect of the vertical velocity on the vertical vorticity. Since the flow is invariant along the vertical, the equation for the vertical vorticity (4.2) reduces to

(4.24)\begin{align} \frac{\partial\zeta}{\partial t} + \boldsymbol{u}\boldsymbol{\cdot} \boldsymbol{\nabla}\zeta = \frac{\varepsilon}{r}\left[ \frac{\partial }{\partial r} \left (r {u_z} \sin{(\theta)}\right) + \frac{\partial}{\partial\theta} \left({u_z} \cos{(\theta)} \right) \right] + \frac{1}{Re} \varDelta_h \zeta. \end{align}

This equation shows that the vertical velocity generated at order ${O}(\varepsilon )$ will in turn force a vertical vorticity field at order ${O}(\varepsilon ^{2})$. In order to compute this second-order horizontal flow, the vertical vorticity and streamfunction can be expanded as

(4.25)$$\begin{gather} \zeta = \zeta_0 + \varepsilon^{2} \zeta_2 + \dots, \end{gather}$$
(4.26)$$\begin{gather}\psi = \psi_0 + \varepsilon^{2} \psi_2 + \dots, \end{gather}$$

where $(\zeta _0,\psi _0)$ are the non-dimensional vertical vorticity and streamfunction of the base flow (2.5) and $\zeta _2 = \varDelta _h \psi _2$. The second-order vertical vorticity is governed by

(4.27)\begin{gather} \frac{\partial \zeta_{2}}{\partial t} + \varOmega \frac{\partial \zeta_2}{\partial \theta} - \frac{1}{r}\frac{\partial\psi_2}{\partial\theta} \frac{\partial\zeta_0}{\partial r} = \frac{1}{r}\left[ \frac{\partial}{\partial r} \left (r u_{z1} \sin{(\theta)}\right) + \frac{\partial}{\partial\theta} \left(u_{z1} \cos{(\theta)} \right) \right]+ \frac{1}{Re} \varDelta_h \zeta_2 . \end{gather}

Since the first-order vertical velocity $u_{z1}$ has an azimuthal wavenumber $m=1$, $\zeta _2$ and $\psi _2$ can be sought in the form

(4.28)$$\begin{gather} \zeta_2 = \zeta_{20} (r,t) + \left[ \zeta_{22}(r,t) {\rm e}^{2{\rm i}\theta} + {\rm c.c.} \right], \end{gather}$$
(4.29)$$\begin{gather}\psi_2 = \psi_{20} (r,t) + \left[ \psi_{22}(r,t) {\rm e}^{2{\rm i}\theta} + {\rm c.c.} \right]. \end{gather}$$

In the following, we will determine only the axisymmetric part $\zeta _{20}$ since we will show that the component $\zeta _{22}$ grows slower than $\zeta _{20}$.

4.3.1. Inviscid evolution of $\zeta _{20}$

When $u_{z1}$ follows the unsteady inviscid expression (4.10a), the axisymmetric vertical vorticity $\zeta _{20}$ is given by

(4.30)\begin{equation} \frac{\partial\zeta_{20}}{\partial t} ={-}\frac{{\rm i}}{2r} \frac{\partial}{\partial r} \left( r(u_{zp}^{*}-u_{zp}) \right). \end{equation}

By assuming that $\zeta _{20}=0$ at $t=0$, the solution reads

(4.31)\begin{align} {\zeta}_{20} &= \frac{ \zeta_0 }{4}\left( \frac{\cos{(\alpha t)}-1}{\alpha ^{2}} + \frac{\cos{(\beta t)}-1}{\beta^{2}} \right)\nonumber\\ &\quad + \frac{r\varOmega \varOmega^{\prime}}{4} \left[ \frac{\alpha t \sin{(\alpha t)} + 2(\cos{(\alpha t)}-1) }{\alpha^{3}} - \frac{ \beta t \sin{(\beta t)} + 2(\cos{(\beta t)}-1) }{\beta ^{3}} \right]. \end{align}

Considering the vicinity of the critical radius $\eta = r - r_c$, this solution approximates at leading order to

(4.32)\begin{equation} {\zeta}_{20} = \frac{r_c\varOmega_c}{{4\varOmega^{\prime}_c}^{2}\eta^{3}} \left[ 2 \left(1-\cos{(\eta\varOmega^{\prime}_c t)}\right) - \eta\varOmega_c^{\prime}t \sin{( \eta\varOmega_c^{\prime}t)} \right] + {O}\left(\frac{1}{\eta^{2}}\right), \end{equation}

where $\eta t$ has been considered finite as in (4.13) so that the approximation remains valid even for long time. In terms of the similarity variable $U$, (4.32) can be rewritten as

(4.33)\begin{equation} {\zeta}_{20} = \frac{r_c \varOmega_c \varOmega_c^{\prime} t^{3}}{4} \left[ \frac{2(1-\cos{U})-U\sin{U}}{U^{3}} \right] + {O} \left(\frac{1}{\eta^{2}}\right). \end{equation}

This expression shows that $\zeta _{20}$ remains finite and even vanishes when $U \to 0$, but $\zeta _{20}$ is more and more concentrated in the vicinity of the critical layer as time increases. Furthermore, its amplitude increases like $t^{3}$ at leading order.

Using the same approach for $\zeta _{22}$, it can be shown that its amplitude grows like $t^{2}$ instead of $t^{3}$. Indeed, the forcing term of $\zeta _{22}$ is proportional to $t^{2}$ like in the case of $\zeta _{20}$, but the left-hand side of (4.27) is dominated by the term $2 {\rm i} \varOmega _c \zeta _{22}$ instead of $\partial \zeta _{20} / \partial t$ for the axisymmetric component. Thus, $\zeta _{22}$ is proportional to $t^{2}$, at least initially. This explains why the vertical vorticity is observed in the DNS to remain quasi-axisymmetric during the first two phases.

4.3.2. Viscous evolution of $\zeta _{20}$

When the vertical velocity is given by the unsteady viscous solution (4.16a) and (4.21), it is possible to also obtain its effect on the second-order axisymmetric vertical vorticity $\zeta _{20}$. Expressing first (4.27) in terms of $\tilde {r}$ and $T$ gives at leading order when $Re \gg 1$,

(4.34)\begin{equation} \frac{1}{Re^{1/3}} \frac{\partial\zeta_{20}}{\partial T} = \frac{-{\rm i}}{2}Re^{2/3} \frac{\partial}{\partial\tilde{r}} \left(\tilde{u}_{z1}^{*}-\tilde{u}_{z1} \right) + {O} (Re^{1/3}) + \frac{1}{Re^{1/3}} \frac{\partial^{2}\zeta_{20}}{\partial\tilde{r}^{2}} + \dots\ . \end{equation}

As shown by Wang & Balmforth (Reference Wang and Balmforth2021), the exact solution of (4.34) can be found by means of a Fourier transform in $\tilde {r}$ using (4.21). This gives

(4.35)\begin{align} \zeta_{20} ={-}{\rm i} \frac{Re A}{2\gamma{\rm \pi}} \int_0^{\left|{\varOmega_c^{\prime}}\right|T} \exp \left( \frac{-q^{3}}{3\gamma^{3}} + {\rm i} q \tilde{r} \right) \left( \frac{1-\exp\left( q^{3}/\left|{\varOmega_c^{\prime}}\right| - q^{2}T \right)}{q} \right ) \mathrm{d}q + {\rm c.c.}.\end{align}

In the appendix an approximation of (4.35) valid for large time $T \gg 1$ is found in the form

(4.36)\begin{equation} \zeta_{20} = \zeta_{20}^{(1)} + \zeta_{20}^{(2)}, \end{equation}

where

(4.37a)$$\begin{gather} \zeta_{20}^{(1)} (\tilde{r}) = Re \frac{A}{2} \int_{0}^{\tilde{r}} \left( {{Hi}}^{*}( {\rm i} \gamma u) + {Hi}( {\rm i} \gamma u) \right) \mathrm{d}u, \end{gather}$$
(4.37b)$$\begin{gather}\zeta_{20}^{(2)} (\tilde{r}, t) ={-}Re \frac{A}{2\gamma} \mathrm{erf}\left( \frac{\tilde{r}}{\sqrt{4tRe^{{-}1/3}}} \right). \end{gather}$$

This approximation shows that $\zeta _{20}$ saturates and tends towards (4.37a) for $t Re^{-1/3} \rightarrow \infty$ for $\tilde {r} = {O}(1)$. However, when $|{\tilde {r}}| \rightarrow \infty$ and $tRe^{-1/3}$ is finite, $\zeta _{20}$ vanishes. The approximation (4.36) will be compared with numerical solutions of (4.34) as well as DNS in § 5.2.

To summarize, the axisymmetric vertical vorticity correction at order ${O}(\varepsilon ^{2})$, $\zeta _{20}$, follows two distinct regimes. First, it evolves purely inviscidly and grows like $t^{3}$ according to (4.31). Then, it tends to saturate and follows the approximation (4.36) for large times. This shows that $\zeta _{20}$ saturates towards the steady solution (4.37a) for finite $\tilde {r}$.

4.4. Nonlinear analysis of the critical layer

The viscous linear analysis of the critical layer in §§ 4.2 and 4.3 has shown that the vertical velocity scales as $u_z = {O}(\varepsilon Re^{1/3})$. This creates a vorticity correction of the order $\delta \zeta = {O}(\varepsilon ^{2}Re)$ (see (4.25), (4.28), (4.36)–(4.37)). Since $r = r_c + \tilde {r} / Re^{1/3}$ in the critical layer, the corresponding angular velocity correction $\delta \varOmega$ is given at leading order by $\delta \zeta \simeq r_c Re^{1/3} \partial \delta {\varOmega } / \partial {\tilde {r}}$ so that $\delta \varOmega = \textit {O}(\varepsilon ^{2} Re^{2/3})$. In turn, we see that this angular velocity correction would have the same order as the other terms of order ${O}(1/Re^{1/3})$ in (4.15) if $\varepsilon ^{2}Re^{2/3} = {O}(1/Re^{1/3})$, i.e. if

(4.38)\begin{equation} Re = \frac{\widetilde{Re}}{\varepsilon^{2}}, \end{equation}

where $\widetilde {Re}$ is of order unity. For this distinguished limit, there is therefore a nonlinear feedback of the horizontal flow on the evolutions of the vertical velocity and buoyancy, as considered by Wang & Balmforth (Reference Wang and Balmforth2020, Reference Wang and Balmforth2021). Using the scaling (4.38), the typical order of magnitudes of the different variables can be expressed in terms of $\varepsilon$ only: $u_z={O} ( \varepsilon ^{1/3} )$, $b= {O} ( \varepsilon ^{1/3} )$, $\delta \zeta = {O} ( 1 )$, $\delta \varOmega = {O} ( \varepsilon ^{2/3} )$, $u_r = {O} ( \varepsilon ^{4/3} )$ and the radius and slow time are $\tilde {r} = \varepsilon ^{-2/3}(r-r_c)$ and $T = \varepsilon ^{2/3}t$, respectively. We also assume that $\partial /\partial z=0$. Accordingly, we expand the variables as

(4.39a)$$\begin{gather} u_z = \varepsilon^{1/3} \left[ \tilde{u}_{z1} \mathrm{e}^{{\rm i}\theta} + {\rm c.c.} \right] +\varepsilon \left[ \tilde{u}_{z2} \mathrm{e}^{{\rm i}\theta} + {\rm c.c.} \right] + \dots, \end{gather}$$
(4.39b)$$\begin{gather}b = \varepsilon^{1/3} \left[ \tilde{b}_{1} \mathrm{e}^{{\rm i}\theta} + {\rm c.c.} \right] +\varepsilon \left[ \tilde{b}_{2} \mathrm{e}^{{\rm i}\theta} + {\rm c.c.} \right] + \dots, \end{gather}$$
(4.39c)$$\begin{gather}\varOmega = \varOmega_0 + \varepsilon^{2/3} \varOmega_1 +\dots, \end{gather}$$
(4.39d)$$\begin{gather}u_r = \varepsilon^{4/3} u_{r1} + \dots, \end{gather}$$
(4.39e)$$\begin{gather}\zeta = \zeta_0 + \zeta_1 + \varepsilon^{2/3} \zeta_2 + \dots, \end{gather}$$

where $\varOmega _0$ and $\zeta _0$ are the non-dimensional angular velocity and vorticity corresponding to (2.5). In the vicinity of $r_c$, they can be expanded as

(4.40a)$$\begin{gather} \varOmega_0 =\varOmega_c + \tilde{r} \varOmega_c^{\prime} \varepsilon^{2/3} + \dots, \end{gather}$$
(4.40b)$$\begin{gather}\zeta_0 =\zeta_c + \tilde{r} \zeta_c^{\prime} \varepsilon^{2/3} + \dots\ . \end{gather}$$

Note that there are only components of the form $\exp (\pm {\rm i}\theta )$ in (4.39a)–(4.39b) because the flow is invariant along the vertical and because third harmonics arise only at higher order. Then, (4.1d)–(4.1e) become at leading order

(4.41a)\begin{align} & \varepsilon \frac{\partial\tilde{u}_{z1}}{\partial T} + \varepsilon u_{r1} \frac{\partial \tilde{u}_{z1}}{\partial \tilde{r}} + {\rm i} \left( \varOmega_c + \varOmega_c^{\prime}\tilde{r}\varepsilon^{2/3} \right) \varepsilon^{1/3} \tilde{u}_{z1} + {\rm i} \varepsilon \varOmega_1 \tilde{u}_{z1} + {\rm i} \varepsilon \varOmega_c \tilde{u}_{z2}\nonumber\\ &\quad = \varepsilon^{1/3} \tilde{b}_{1} + \varepsilon \tilde{b}_{2} - \varepsilon \frac{r_c\varOmega_c}{2{\rm i}} + \frac{\varepsilon}{\widetilde{Re}} \frac{\partial^{2}\tilde{u}_{z1}}{\partial \tilde{r}^{2}} + {O} \left( \varepsilon^{5/3} \right), \end{align}
(4.41b)\begin{align} & \varepsilon \frac{\partial \tilde{b}_{1}}{\partial T} + \varepsilon u_{r1} \frac{\partial\tilde{b}_{1}}{\partial\tilde{r}} + {\rm i} \left( \varOmega_c + \varOmega_c^{\prime}\tilde{r}\varepsilon^{2/3} \right) \varepsilon^{1/3} \tilde{b}_{1} + {\rm i} \varepsilon \varOmega_1 \tilde{b}_{1} + {\rm i} \varepsilon \varOmega_c \tilde{b}_{2}\nonumber\\ &\quad ={-} \frac{\varepsilon^{1/3}}{F_h^{2}}\tilde{u}_{z1} - \frac{\varepsilon}{F_h^{2}}\tilde{u}_{z2} + \frac{\varepsilon}{\widetilde{Re}Sc} \frac{\partial^{2}\tilde{b}_{1}}{\partial\tilde{r}^{2}} + {O} \left( \varepsilon^{5/3} \right). \end{align}

Similarly, (4.24) becomes

(4.42)\begin{align} & \varepsilon^{2/3} \frac{\partial\zeta_{1}}{\partial T} + \left( \varOmega_c + \varOmega_c^{\prime}\tilde{r}\varepsilon^{2/3} \right)\frac{\partial\zeta_1}{\partial\theta} + \varepsilon^{2/3} \varOmega_1 \frac{\partial\zeta_1}{\partial\theta} + \varepsilon^{2/3} u_{r1} \frac{\partial\zeta_1}{\partial\tilde{r}} + \varepsilon^{2/3} \varOmega_c \frac{\partial\zeta_2}{\partial\theta} \nonumber\\ & \quad = \varepsilon^{2/3} \frac{\partial}{\partial\tilde{r}} \left( \left( \tilde{u}_{z1} \mathrm{e}^{{\rm i}\theta} + \tilde{u}_{z1}^{*} \mathrm{e}^{-{\rm i}\theta} \right) \sin{(\theta)} \right) + \frac{\varepsilon^{2/3}}{\widetilde{Re}} \frac{\partial^{2}\zeta_1}{\partial\tilde{r}^{2}} + {O} \left( \varepsilon^{4/3} \right). \end{align}

At leading order, (4.41)–(4.42) become

(4.43a)$$\begin{gather} \varepsilon^{1/3} {\rm i} \varOmega_c \tilde{u}_{z1} = \varepsilon^{1/3} \tilde{b}_{1}, \end{gather}$$
(4.43b)$$\begin{gather}\varepsilon^{1/3} {\rm i} \varOmega_c \tilde{b}_{1} ={-}\frac{\varepsilon^{1/3}}{F_h^{2}} \tilde{u}_{z1}, \end{gather}$$
(4.43c)$$\begin{gather}\varOmega_c \frac{\partial\zeta_1}{\partial\theta} = 0. \end{gather}$$

The first two equations are identical to (4.17) and the third one implies that $\zeta _1 \equiv \zeta _1(\tilde {r},T)$ and $u_{r1}=0$. At the next order, we have

(4.44a)$$\begin{gather} \frac{\partial \tilde{u}_{z1}}{\partial T} + {\rm i} \varOmega_c^{\prime}\tilde{r} \tilde{u}_{z1} + {\rm i} \varOmega_1 \tilde{u}_{z1} + {\rm i} \varOmega_c \tilde{u}_{z2} = \tilde{b}_{2} - \frac{r_c\varOmega_c}{2i} + \frac{1}{\widetilde{Re}}\,\frac{\partial^{2}\tilde{u}_{z1}}{\partial\tilde{r}^{2}}, \end{gather}$$
(4.44b)$$\begin{gather}\frac{\partial \tilde{b}_{1}}{\partial T} + {\rm i} \varOmega_c^{\prime}\tilde{r} \tilde{b}_{1} + {\rm i} \varOmega_1 \tilde{b}_{1} + {\rm i} \varOmega_c \tilde{b}_{2} ={-} \frac{\tilde{u}_{z2}}{F_h^{2}} + \frac{1}{\widetilde{Re}Sc}\,\frac{\partial^{2}\tilde{b}_{1}}{\partial \tilde{r}^{2}}, \end{gather}$$
(4.44c)$$\begin{gather}\frac{\partial \zeta_{1}}{\partial T} + \varOmega_c \frac{\partial \zeta_2}{\partial \theta}={-}\frac{{\rm i}}{2} \frac{\partial }{\partial \tilde{r}} \left( \tilde{u}_{z1}^{*} - \tilde{u}_{z1} + \tilde{u}_{z1} \mathrm{e}^{2{\rm i}\theta} - \tilde{u}_{z1}^{*} \mathrm{e}^{{-}2{\rm i}\theta} \right) + \frac{1}{\widetilde{Re}}\, \frac{\partial ^{2}\zeta_1}{\partial \tilde{r}^{2}} . \end{gather}$$

Equations (4.44a) and (4.44b) are identical to (4.19) except for the presence of the terms involving $\varOmega _1$. They can be combined to give

(4.45)\begin{equation} \frac{\partial \tilde{u}_{z1}}{\partial T} + {\rm i} \varOmega_c^{\prime}\tilde{r} \tilde{u}_{z1} + {\rm i} \varOmega_1 \tilde{u}_{z1} = \frac{i}{4}r_c \varOmega_c + \frac{1}{2\widetilde{Re}} \left(1 + \frac{1}{Sc} \right) \frac{\partial ^{2}\tilde{u}_{z1}}{\partial \tilde{r}^{2}}, \end{equation}

whereas (4.44c) splits into

(4.46a)$$\begin{gather} \frac{\partial \zeta_{1}}{\partial T} ={-}\frac{{\rm i}}{2} \frac{\partial }{\partial \tilde{r}} \left( \tilde{u}_{z1}^{*} - \tilde{u}_{z1} \right) + \frac{1}{\widetilde{Re}}\, \frac{\partial ^{2}\zeta_1}{\partial \tilde{r}^{2}}, \end{gather}$$
(4.46b)$$\begin{gather}\varOmega_c \frac{\partial \zeta_2}{\partial \theta} ={-}\frac{{\rm i}}{2} \frac{\partial }{\partial \tilde{r}} \left( \tilde{u}_{z1} \mathrm{e}^{2{\rm i}\theta} - \tilde{u}_{z1}^{*} \mathrm{e}^{{-}2{\rm i}\theta} \right). \end{gather}$$

Since $\zeta _1 = r_c \partial \varOmega _1 / \partial \tilde {r}$, (4.45)–(4.46a) form a closed system of equations. An equation for $\varOmega _1$ can be obtained by integrating (4.46a) as follows:

(4.47)\begin{equation} \frac{\partial \varOmega_{1}}{\partial T} ={-} \frac{{\rm i}}{2r_c} \left( \tilde{u}_{z1}^{*} - \tilde{u}_{z1} \right) + \frac{1}{\widetilde{Re}} \,\frac{\partial^{2}\varOmega_1}{\partial \tilde{r}^{2}}. \end{equation}

Equation (4.46b) shows that the azimuthal wavenumbers $m=\pm 2$ are generated in the vorticity only at higher order explaining again why the vorticity remains quasi-axisymmetric in the DNS during the first two phases. For this reason, components of the form $\exp (\pm 3 {\rm i}\theta )$ arise in the vertical velocity and density (4.39a)–(4.39b) only at the higher order $\varepsilon ^{5/3}$.

5. Comparison between the DNS and the asymptotic analyses

We now compare in detail the asymptotic and numerical results.

5.1. Vertical velocity

Figure 9 shows the evolution of the maximum vertical velocity $u_{zm}(\theta,t)$ (solid line) for $\theta =0$ (figure 9a) and $\theta ={\rm \pi} /2$ (figure 9b) for the set of parameters $Re=10\,000$, $F_h=2$, $Ro=20.3$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$). Two different angles are considered since the theoretical vertical velocity has the form $u_z = u_{zc} (r,t) \cos {(\theta )} + u_{zs} (r,t) \sin {(\theta )}$. Hence, the plots for $\theta =0$ and $\theta ={\rm \pi} /2$ allow us to check the predictions for $u_{zc}$ and $u_{zs}$, respectively.

Figure 9. Comparison between the maximum vertical velocity in the DNS (black solid line), predicted by the unsteady inviscid solution (4.10a) (red dashed line), by the unsteady viscous solution (4.16a), (4.21) (yellow dashed line), by the steady viscous solution (4.16a), (4.23) (green dashed line) and by the nonlinear equations (4.45), (4.47) (blue dashed line) for $(a)$ $\theta =0$ and $(b)$ $\theta ={\rm \pi} /2$ for $Re=10\,000$, $F_h=2$, $Ro=20.3$, $\phi =80^{\circ }$ and for $(c)$ $\theta =0$ and $(d)$ $\theta ={\rm \pi} /2$ for $Re=2000$, $F_h=2$, $Ro=20.3$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$).

We can see that the inviscid theoretical solution (4.10a) (red dashed line in figure 9a,b) predicts very well the initial linear increase of the maximum vertical velocity in the DNS for both angles. The unsteady viscous solution (4.16a), (4.21) (yellow dashed line) increases also linearly initially and is in good agreement with the DNS except that it lacks the small oscillations. They are indeed due to inertia-gravity waves oscillating at frequency $1/F_h + \varOmega = 2/F_h$ near $r_c$ (see the last term of (4.11a)) and of the two lines of (4.13) which are neglected in (4.14) and § 4.2.

When the growth of $u_{zm}$ is no longer linear, the time-dependent viscous solution (4.21) remains in very good agreement with the DNS and describes perfectly the transition towards the steady viscous solution (4.23) (green dashed line). The latter solution is close to the levels of saturation of $u_{zm}(\theta,t)$ for $\theta =0$ and $\theta = {\rm \pi}/2$ in the DNS, although the agreement is not as good as for the initial regime. However, the nonlinear equations (4.45), (4.47) (blue dashed lines) are in better agreement with the DNS indicating that nonlinear effects are also active in the saturation. In essence, none of the theoretical solutions can exhibit oscillations associated with the late non-axisymmetric evolution observed in the DNS.

Figure 9(c,d) shows a similar comparison when the Reynolds number is reduced to ${Re=2000}$, keeping the other parameters fixed. In this case, $u_{zm}(\theta =0,t)$ and $u_{zm}(\theta ={\rm \pi} /2,t)$ do not oscillate at late time in the DNS (black solid line). The agreement with (4.10a) and (4.21) or (4.23) are then excellent in the initial and saturation regimes, respectively. In addition, we see that the predictions of the nonlinear equations (4.45), (4.47) remain very close to the linear ones, i.e. (4.21), showing that nonlinear effects are weak in this case.

Figure 9 shows that the transition from the unsteady inviscid solution (4.10a) to the steady viscous one, (4.23), occurs in a short time range. Therefore, the unsteady viscous solution (4.21) can be well approximated by (4.10a) for $t \leqslant \mathcal {T}$ and (4.23) for $t \geqslant \mathcal {T}$, where

(5.1)\begin{equation} \mathcal{T} = 2 {\rm \pi}\frac{Re^{1/3}{Hi}(0)}{\left|{2 \varOmega_c^{\prime}}\right|^{2/3}(1+1/Sc)^{1/3}} \end{equation}

is the time when the overall maximum given by (4.10a) and (4.23) becomes equal. This time is independent of $\widetilde {Ro}$ and depends only on the Reynolds number, the Froude number via $\varOmega _c^{\prime }$ and the Schmidt number.

Figure 10 displays a detailed comparison of the radial profile of $u_z$ for $\theta =0$ and $\theta ={\rm \pi} /2$ predicted by (4.10a) (red dashed line) and observed in the DNS (black solid line) for different instants in the inviscid regime, i.e. $t \leqslant \mathcal {T}$ for the same parameters as figure 9(a,b). We see that the agreement is excellent even when $t \simeq \mathcal {T}$ (figure 10cf). The approximation (4.21) and the solution of the nonlinear equations (4.45), (4.47) for the vertical velocity are also represented by yellow and blue dashed lines, respectively, in figure 10(b,c,ef). The approximation (4.14) is almost identical to (4.21) in this time range and not represented. As expected, the agreement between the DNS and (4.21) or (4.45), (4.47) is very good near $r_c$ but deteriorates away from $r_c$. We can note that the blue and yellow dashed lines are superposed everywhere except close to the critical radius $r_c$ for $t=40$ (figure 10cf). In this region the blue dashed lines are in very good agreement with the DNS indicating that nonlinear effects are important there. However, away from $r_c$, it is (4.10a) (red dashed line) which better agrees with the DNS. For $t=5$ (figure 10a,d), the approximations (4.21) and (4.45), (4.47) are not accurate and not shown. This is because the profile of $u_z$ is not yet sufficiently localized around $r_c$ at this early time and, therefore, it cannot be well described by a local approximation near $r_c$. Indeed, we can see in figure 10(a,d) that the profiles of $u_z$ are quite different from those in figure 10(b,c,ef).

Figure 10. Comparison between the vertical velocity at $\theta =0$ (ac) and $\theta ={\rm \pi} /2$ (df) in the DNS (black solid line), predicted by the unsteady inviscid solution (4.10a) (red dashed line), by the unsteady viscous solution (4.21) (yellow dashed line) and by the nonlinear equations (4.45), (4.47) (blue dashed line) at (a,d) $t=5$, (b,e) $t=25$ and (cf) $t=40$ for $Re=10\,000$, $F_h=2$, $Ro=20.3$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$). The circle symbols represent the location of the critical radius in the unsteady inviscid solution (4.10a).

Figure 11 shows again the vertical velocity profiles observed in the DNS for the same parameters but, for $t \geqslant \mathcal {T}$, this time they are compared with the unsteady and steady viscous solutions (4.21) (yellow dashed lines) and (4.23) (green dashed lines) as well as the predictions of the nonlinear equations (4.45), (4.47) (blue dashed lines). At $t=40$ (figure 11a,d), the steady viscous solution (4.23) (green dashed lines) is already in good agreement with the DNS since $t=40$ is close to the time $\mathcal {T}$ where the transition from (4.10a) to (4.23) occurs (figure 9a,b). Nevertheless, it departs slightly from the DNS near $r_c$ unlike the unsteady viscous solution (4.21) and nonlinear predictions from (4.45), (4.47). At longer times, $t=60$ (figure 11b,e) and $t=80$ (figure 11cf), (4.21) and (4.23) become almost identical and remain in satisfactory agreement with the vertical velocity profiles of the DNS. Nevertheless, we can see a shift between the numerical and theoretical profiles. In contrast, the solution of the nonlinear equations (4.45), (4.47) remains in very good agreement with the DNS and does not exhibit such a shift. This indicates that the shift is due to the viscous and nonlinear variations of the angular velocity. For example, this makes the critical radius where $\varOmega (r_c) = 1/F_h$ to move towards the vortex centre, as seen in figure 11. This phenomenon is absent from the linear equations (4.21), (4.23) since they do not take into account any variation of the angular velocity.

Figure 11. Comparison between the vertical velocity at $\theta =0$ (ac) and $\theta ={\rm \pi} /2$ (df) in the DNS (black solid line), predicted by the unsteady viscous solution (4.16a), (4.21) (yellow dashed line), by the steady viscous solution (4.16a), (4.23) (green dashed line) and by the nonlinear equations (4.45), (4.47) (blue dashed line) at (a,d) $t=40$, (b,e) $t=60$, (cf) $t=80$ for $Re=10\,000$, $F_h=2$, $Ro=20.3$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$). The circle symbols represent the location of the critical radius in the unsteady viscous solution (4.16a).

5.2. Vertical vorticity

The asymptotic analyses have shown that the axisymmetric component of the vertical velocity is given by $\zeta = \zeta _0(r) + \varepsilon ^{2} \zeta _{20}(r,t) + \dots$, where $\zeta _{20}$ follows (4.31) and (4.35) in the inviscid and viscous regimes, i.e. $t \leqslant \mathcal {T}$ and $t\geqslant \mathcal {T}$, respectively. A global view of these two regimes and the associated approximations can be gained by plotting $\partial \zeta _{20} / \partial r$ at $r=r_c$ (figure 12). The black solid line shows the evolution of $\partial \zeta _{20} / \partial r (r_c,t)$ computed numerically from (4.35). Here $\partial \zeta _{20} / \partial r (r_c,t)$ increases initially like $t^{4}$ in agreement with the approximation (4.33) (red dashed line). Subsequently, for $t \gg \mathcal {T}$, $\partial \zeta _{20} / \partial r (r_c,t)$ increases more slowly and saturates towards $\partial \zeta _{20}^{(1)} / \partial r (r_c)$ (black dashed line) for $t \rightarrow \infty$. The approximation (4.36) (blue dashed line) is in good agreement with the solution (4.35) (black line) in this regime. It will allow us to derive a theoretical criteria for the onset of non-axisymmetry in § 6.4.

Figure 12. Comparison between the evolution of $\partial \zeta _{20}/\partial r(r_c,t)$ from the theoretical expressions: the unsteady viscous solution (4.35) (black solid line), the unsteady inviscid solution (4.31) (red dashed line), the viscous solutions (4.36) (blue dashed line) and (4.37a) (black dashed line) for $Re=10\,000$, $F_h=2$.

Figure 13 compares the radial profile of the theoretical vorticity $\zeta = \zeta _0(r) + \varepsilon ^{2} \zeta _{20}(r,t) + \dots$, with $\zeta _{20}$ given by the unsteady inviscid solution (4.31) (red dashed line) to the DNS when $t \leqslant \mathcal {T}$. The agreement is excellent and predicts very well the deformation of the vertical vorticity profile near $r_c \simeq 1.2$. A similar comparison is displayed in figure 14 for three different times such that $t \gtrsim \mathcal {T}$ and $\zeta _{20}$ is given by the unsteady viscous solution (4.35). The agreement continues to be very good even for $t=85$ (figure 14c), confirming the validity of the solution (4.35). A slight shift can however be noticed near $r_c$ and near the vortex axis. In contrast, the predictions of the nonlinear equations (4.45), (4.47) (blue dashed lines) are in very good agreement with the DNS near $r_c$. This shows again that nonlinear effects are not negligible in the critical layer.

Figure 13. Comparison between the vertical vorticity at $\theta ={\rm \pi} /2$ in the DNS (black solid line) and the asymptotic expressions $\zeta = \zeta _0 + \varepsilon ^{2}\zeta _{20}$ where $\zeta _{20}$ follows the unsteady inviscid solution (4.31) (red dashed line) and $\zeta = \zeta _0 + \zeta _1$ with $\zeta _1$ given by the nonlinear equations (4.45), (4.47) (blue dashed line) at $(a)$ $t=25$, $(b)$ $t=35$ and $(c)$ $t=40$ for $Re=10\,000$, $F_h=2$, $Ro=20.3$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$). The circle symbols represent the location of the critical radius in the unsteady inviscid solution (4.31).

Figure 14. Comparison between the vertical vorticity at $\theta ={\rm \pi} /2$ in the DNS (black solid line) and the asymptotic expressions $\zeta = \zeta _0 + \varepsilon ^{2}\zeta _{20}$ where $\zeta _{20}$ follows the unsteady viscous solution (4.35) (red dashed line) and $\zeta = \zeta _0 + \zeta _1$ with $\zeta _1$ given by the nonlinear equations (4.45), (4.47) (blue dashed line) at $(a)$ $t=50$, $(b)$ $t=65$ and $(c)$ $t=85$ for $Re=10\,000$, $F_h=2$, $Ro=20.3$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$). The circle symbols represent the location of the critical radius in the unsteady inviscid solution (4.35).

Finally, the prediction for late time $t \gg \mathcal {T}$ based on the steady viscous solution (4.36) has been compared with a DNS for a larger non-traditional Rossby number $\widetilde {Ro} = 500$, the other parameters being identical. The vortex remains then quasi-axisymmetric in the DNS (figure 8) allowing us to see the late evolution of the vorticity without the non-axisymmetric perturbations. Figure 15 shows that the deformation of the vorticity profile near $r_c$ is weak but well predicted by (4.36), although there is a shift for $t \geqslant 300$. In this case, the predictions from the nonlinear equations (4.45), (4.47) (not shown) are identical to those from (4.36) indicating that the nonlinear effects are weak. However, if we take also into account the global viscous decay of the leading-order vorticity $\zeta _0$ which is significant for these large times, then the agreement with the DNS is perfect (yellow dashed line).

Figure 15. Comparison between the vertical vorticity at $\theta ={\rm \pi} /2$ in the DNS (black solid line) and the asymptotic expressions $\zeta = \zeta _0 + \varepsilon ^{2}\zeta _{20}$ where $\zeta _{20}$ follows the viscous solution (4.36) (red dashed line) and $\zeta = \zeta _0 + \zeta _1$ with $\zeta _1$ given by the nonlinear equations (4.45), (4.47) and the viscous decay of $\zeta _0$ also taken into account (yellow dashed line) at $(a)$ $t=200$, $(b)$ $t=300$ and $(c)$ $t=400$ for $Re=10\,000$, $F_h=2$, $Ro=20.0$, $\phi =87.7^{\circ }$ ($\widetilde {Ro}=500$). The circle symbols represent the location of the critical radius in the viscous solution (4.36).

Besides, a feature of high interest is that the vertical vorticity profile exhibits two extrema near $r_c$, ${\rm d}\zeta / {\rm d}r = 0$, as soon as $t \geqslant 40$ (figures 13(c) and 14). According to Rayleigh's inflectional criterion (Rayleigh Reference Rayleigh1880), this is a necessary condition for the shear instability. However, only the first extremum, where $\zeta$ has a local minimum, satisfies the stiffer (Fjørtoft Reference Fjørtoft1950) instability condition $[\varOmega (r)-\varOmega (r_I)] \partial \zeta / \partial r < 0$, where $r_I$ is the extremum. This gives us hindsight on the possible origin of the late non-axisymmetric evolution of the vertical vorticity. The next section will investigate whether or not this hypothesis is correct.

6. Analysis of the non-axisymmetric evolution

As seen in the previous sections, a ring of anomalous vertical vorticity develops near the critical radius and, subsequently, this ring may become non-axisymmetric when the non-traditional Rossby number $\widetilde {Ro}$ is below a critical value depending on the Reynolds number (figure 8). In this section the main question is: what is the origin of this non-axisymmetric evolution? Is it due to a shear instability associated to the inflection point in the vertical vorticity profile (see figure 14) or is it an intrinsic behaviour of the vortex under the complete Coriolis force? Regarding the latter hypothesis, we have seen indeed from (4.27)–(4.29) that the vertical velocity field forces not only an axisymmetric vertical vorticity field but also a non-axisymmetric one with an azimuthal wavenumber $m=2$.

6.1. Azimuthal decomposition of $\zeta$ and $u_z$

In order to better understand the onset of non-axisymmetric vertical vorticity, we have first decomposed $\zeta$ thanks to an azimuthal Fourier transform

(6.1)\begin{equation} \hat{\zeta}_m (r,t) = \int_0^{2{\rm \pi}} \zeta (r,\theta,t) {\rm e}^{-{\rm i}m\theta} \,\mathrm{d}\theta, \end{equation}

where $\zeta (x,y,t)$ has been first interpolated on a grid of cylindrical coordinates $(r,\theta )$. The same transform has been applied to $u_z$ in order to obtain $\hat {u}_{z_m}$. The mean power in each azimuthal wavenumber is then defined as

(6.2)\begin{equation} E_{\zeta}(m,t)=\frac{ \displaystyle\int_0^{l_x/2}\hat{\zeta}_m^{2}(r,t) r\,\mathrm {d}r}{\displaystyle\int_0^{l_x/2} r\,\mathrm {d}r} = \frac{8}{l_x^{2}} \int_0^{l_x/2}\hat{\zeta}_m^{2}(r,t) r\,\mathrm {d}r,\end{equation}

and

(6.3)\begin{equation} E_{u_z}(m,t)= \frac{8}{l_x^{2}} \int_0^{l_x/2} \hat{u}_{z_m}^{2}(r,t) r\,\mathrm {d}r. \end{equation}

Figure 16(a) displays the evolution of the logarithm of the power of the first three azimuthal wavenumbers of the vertical vorticity. These are only even, i.e. $m=0$, $m=2$, $m=4$. Similarly, figure 16(b) shows the logarithm of the power of the first three azimuthal wavenumbers of $u_z$. They are odd in this case: $m=1$, $m=3$, $m=5$. It can be seen that $E_{\zeta }(0,t)$ and $E_{u_z}(1,t)$ (black solid lines) remain approximately constant except that $E_{u_z}(1,t)$ sustains large oscillations after $t \simeq 120$.

Figure 16. Evolution of the power $(a)$ $E_{\zeta }(m,t)$ for the azimuthal wavenumbers $m=0$ (black solid line), $m=2$ (red dashed line) and $m=4$ (green dashed line) and the power $(b)$ $E_{u_z}(m,t)$ for the azimuthal wavenumbers $m=1$ (black solid line), $m=3$ (red dashed line) and $m=5$ (green dashed line) for $Re=10\,000$, $F_h=2$, $Ro=20.3$ and $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$).

It is also worth pointing out that $E_{\zeta }(2,t)$ starts to grow at the beginning of the simulation (figure 16a) since $\zeta _{22}$ increases like $t^{2}$ due to the forcing by the vertical velocity as detailed in § 4.3.1. However, when $t \leqslant 80$, its power remains negligible compared with that of the axisymmetric mode, $E_\zeta (0,t)$. In contrast, after $t \simeq 80$, $E_{\zeta }(2,t)$ grows exponentially before saturating at $t \geqslant 120$. There is therefore a clear transition towards an exponential growth, a feature consistent with the instability hypothesis. Nevertheless, we can see that the azimuthal mode $m=3$ of $u_z$ (figure 16b) grows also exponentially at the same time. The higher modes, $m=4$ of $\zeta$ and $m=5$ of $u_z$, start to increase also exponentially but somewhat later. Thus, it is unclear if the growth of $E_{u_z}(3,t)$ is a consequence of the growth of $E_{\zeta }(2,t)$, if it is the opposite or if the exponential growth is due to a coupling between $E_{u_z}(3,t)$ and $E_{\zeta }(2,t)$.

6.2. Truncated model

To answer the latter question, we have derived a truncated model taking into account only the first azimuthal wavenumbers of each quantity. More precisely, the different variables have been written as

(6.4a)$$\begin{gather} u_r = \hat{u}_{r_{2c}} (r,t) \cos{(2\theta)} + \hat{u}_{r_{2s}} (r,t) \sin{(2\theta)}, \end{gather}$$
(6.4b)$$\begin{gather}u_\theta = \hat{u}_{\theta_0} (r,t) + \hat{u}_{\theta_{2c}} (r,t) \cos{(2\theta)} + \hat{u}_{\theta_{2s}} (r,t) \sin{(2\theta)}, \end{gather}$$
(6.4c)$$\begin{gather}p (r,\theta) = \hat{p}_0 (r,t) + \hat{p}_{2c} (r,t) \cos{(2\theta)} + \hat{p}_{2s} (r,t) \sin{(2\theta)}, \end{gather}$$
(6.4d)$$\begin{gather}\zeta = \hat{\zeta}_0 (r,t) + \hat{\zeta}_{2c} (r,t) \cos{(2\theta)} + \hat{\zeta}_{2s} (r,t) \sin{(2\theta)}, \end{gather}$$
(6.4e)$$\begin{gather}u_z = \hat{u}_{z_{1c}} (r,t) \cos{(\theta)} + \hat{u}_{z_{1s}} (r,t) \sin{(\theta)}, \end{gather}$$
(6.4f)$$\begin{gather}b = \hat{b}_{1c} (r,t) \cos{(\theta)} + \hat{b}_{1s} (r,t) \sin{(\theta)}. \end{gather}$$

These decompositions have been introduced in (4.1a)–(4.1e) and the following governing equations have been obtained for the vertical velocity and buoyancy by truncating all the higher modes:

(6.5a)\begin{align} \frac{\partial \hat{u}_{z_{1c}}}{\partial t} &={-}\frac{1}{2} \left ( \hat{u}_{r_{2c}} \frac{\partial \hat{u}_{z_{1c}}}{\partial r} + \hat{u}_{r_{2s}} \frac{\partial \hat{u}_{z_{1s}}}{\partial r} \right) - \frac{1}{r} \left (\hat{u}_{\theta_0} \hat{u}_{z_{1s}} + \frac{1}{2} ( \hat{u}_{\theta_{2c}} \hat{u}_{z_{1s}} - \hat{u}_{\theta_{2s}} \hat{u}_{z_{1c}} ) \right)\nonumber\\ &\quad + \hat{b}_{1c} + \frac{1}{Re} \nabla ^{2} \hat{u}_{z_{1c}}+ \frac{1}{\widetilde{Ro}}( \hat{u}_{r_{2c}} - \hat{u}_{\theta_{2s}} ), \end{align}
(6.5b)\begin{align} \frac{\partial \hat{u}_{z_{1s}}}{\partial t} &= \frac{1}{2} \left ( \hat{u}_{r_{2c}} \frac{\partial \hat{u}_{z_{1s}}}{\partial r} - \hat{u}_{r_{2s}} \frac{\partial \hat{u}_{z_{1c}}}{\partial r} \right) + \frac{1}{r} \left ( \hat{u}_{\theta_0} \hat{u}_{z_{1c}} - \frac{1}{2} ( \hat{u}_{\theta_{2c}} \hat{u}_{z_{1c}} + \hat{u}_{\theta_{2s}} \hat{u}_{z_{1s}} ) \right) \nonumber\\ &\quad + \hat{b}_{1s} + \frac{1}{Re} \nabla ^{2} \hat{u}_{z_{1s}} + \frac{1}{\widetilde{Ro}} ( \hat{u}_{r_{2s}} + \hat{u}_{\theta_{2c}} -2\hat{u}_{\theta_0} ) , \end{align}
(6.5c)\begin{align} \frac{\partial \hat{b}_{1c}}{\partial t} &={-}\frac{1}{2} \left ( \hat{u}_{r_{2c}} \frac{\partial \hat{b}_{1c}}{\partial r} + \hat{u}_{r_{2s}} \frac{\partial \hat{b}_{1s}}{\partial r} \right) - \frac{1}{r} \left( \hat{u}_{\theta_0} \hat{b}_{{1s}} + \frac{1}{2} ( \hat{u}_{\theta_{2c}} \hat{b}_{1s} - \hat{u}_{\theta_{2s}} \hat{b}_{1c} ) \right )\nonumber\\ &\quad - \frac{\hat{u}_{z_{1c}}}{F_h^{2}} + \frac{1}{ReSc} \nabla ^{2} \hat{b}_{{1c}}, \end{align}
(6.5d)\begin{align} \frac{\partial \hat{b}_{{1s}}}{\partial t} &= \frac{1}{2} \left ( \hat{u}_{r_{2c}} \frac{\partial \hat{b}_{1s}}{\partial r} - \hat{u}_{r_{2s}} \frac{\partial \hat{b}_{1c}}{\partial r} \right) + \frac{1}{r} \left ( \hat{u}_{\theta_0} \hat{b}_{{1c}} - \frac{1}{2} ( \hat{u}_{\theta_{2c}} \hat{b}_{1c} + \hat{u}_{\theta_{2s}} \hat{b}_{1s} ) \right) \nonumber\\ &\quad - \frac{\hat{u}_{z_{1s}}}{F_h^{2}} + \frac{1}{ReSc} \nabla ^{2} \hat{b}_{{1s}}. \end{align}

Applying the same truncation approach for the vertical vorticity gives

(6.6a)\begin{align} \frac{\partial \hat{\zeta}_0}{\partial t} &={-} \frac{1}{2} \left( \hat{u}_{r_{2c}} \frac{\partial \hat{\zeta}_{2c}}{\partial r} + \hat{u}_{r_{2s}} \frac{\partial \hat{\zeta}_{2s}}{\partial r} \right)- \frac{1}{r} \left( \hat{u}_{\theta_{2c}} \hat{\zeta}_{2s} - \hat{u}_{\theta_{2s}} \hat{\zeta}_{2c} \right) + \frac{1}{Re}\nabla ^{2} \hat{\zeta}_0 \nonumber\\ &\quad + \frac{1}{ \widetilde{Ro}} \left( \frac{\partial \hat{u}_{z_{1s}}}{\partial r} + \frac{\hat{u}_{z_{1s}}}{r} \right), \end{align}
(6.6b)$$\begin{align} \frac{\partial \hat{\zeta}_{2c}}{\partial t} &={-} \hat{u}_{r_{2c}} \frac{\partial \hat{\zeta}_0}{\partial r} - \frac{2 \hat{u}_{\theta_0}}{r} \hat{\zeta}_{2s} + \frac{1}{ Re} \nabla ^{2} \hat{\zeta}_{2c} - \frac{1}{\widetilde{Ro}} \left( \frac{\partial \hat{u}_{z_{1s}}}{\partial r} -\frac{\hat{u}_{z_{1s}}}{r} \right), \end{align}$$
(6.6c)$$\begin{align}\frac{\partial \hat{\zeta}_{2s}}{\partial t} &={-} \hat{u}_{r_{2s}} \frac{\partial\hat{\zeta}_0}{\partial r} + \frac{2 \hat{u}_{\theta_0}}{r} \hat{\zeta}_{2c} + \frac{1}{Re} \nabla ^{2} \hat \zeta_{2s} + \frac{1}{\widetilde{Ro}}\left( \frac{\partial \hat{u}_{z_{1c}}}{\partial r} -\frac{\hat{u}_{z_{1c}}}{r} \right), \end{align}$$

where

(6.7a)$$\begin{gather} \hat{\zeta}_0 = \frac{1}{r} \,\frac{\partial r \hat{u}_{\theta_0} }{\partial r}, \end{gather}$$
(6.7b)$$\begin{gather}\hat{\zeta}_{2c} = \frac{1}{r} \,\frac{\partial r \hat{u}_{\theta_{2c}} }{\partial r} - \frac{2}{r} \hat{u}_{r_{2s}}, \end{gather}$$
(6.7c)$$\begin{gather}\hat{\zeta}_{2s} = \frac{1}{r} \,\frac{\partial r \hat{u}_{\theta_{2s}} }{\partial r} + \frac{2}{r} \hat{u}_{r_{2c}}. \end{gather}$$

The divergence equation also implies that

(6.8a)$$\begin{gather} \frac{1}{r} \,\frac{\partial r \hat{u}_{r_{2c}} }{\partial r} + \frac{2}{r} \hat{u}_{\theta_{2s}} = 0, \end{gather}$$
(6.8b)$$\begin{gather}\frac{1}{r} \,\frac{\partial r \hat{u}_{r_{2s}} }{\partial r} - \frac{2}{r} \hat{u}_{\theta_{2c}} = 0. \end{gather}$$

Such a truncated model can be seen as a heuristic extension of the asymptotic analyses. Indeed, it takes into account both time dependence and diffusive effects in the evolution of the vertical velocity and buoyancy (6.5). Moreover, the modifications of the axisymmetric flow field ($\hat {u}_{\theta _0}$) and the generated $m=2$ mode ($\hat {u}_{r_{2c}}$, $\hat {u}_{r_{2s}}$, $\hat {u}_{\theta _{2c}}$, $\hat {u}_{\theta _{2s}}$) are also taken into account. The latter is governed by (6.6b)–(6.6c) and it appears also in the evolution of the axisymmetric flow field (6.6a). However, as in the asymptotic analyses, only the first azimuthal wavenumber of $u_z$ and $b$ are considered, i.e. the modes $m=3$, $m=5,\ldots$ are neglected. Similarly, the higher modes $m=4$, $m=6,\ldots$ are neglected in the evolution of the horizontal flow field (6.6).

Figure 17(a,b) compares the power $E_{\zeta } (2,t)$ and $E_{u_{z}} (1,t)$, respectively, obtained in the DNS and in the truncated model. There are some differences for $t \geqslant 100$ but, qualitatively, the same type of evolution as in the DNS is obtained with the truncated model. This is remarkable since the truncated model crudely neglects many azimuthal modes and, in particular, the azimuthal mode $m=3$ in the vertical velocity and buoyancy fields. Hence, this proves that the growth of $E_{u_z} (3,t)$ in figure 16(b) does not play a key role in the onset of non-axisymmetry in the vertical vorticity.

Figure 17. Evolution of the power $(a)$ $E_{\zeta }(2,t)$ and $(b)$ $E_{u_z}(1,t)$ in the DNS (black solid line) and truncated model (red dashed line) for $Re=10\,000$, $F_h=2$, $Ro=20.3$ and $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$).

We can take a step further in the understanding of this phenomenon by freezing the axisymmetric velocity field. In other words, the time evolution of $(\hat {u}_{\theta _0},\hat {\zeta }_0)$ is suppressed after a given time $t_f$. We also set the forcing terms due to the non-traditional Coriolis force to be zero in (6.6b) and (6.6c). Hence, the latter equations become

(6.9a)$$\begin{gather} \frac{\partial \hat{\zeta}_{2c}}{\partial t} ={-} \hat{u}_{r_{2c}} \frac{\partial \hat{\zeta}_0 (r,t_f)}{\partial r} - \frac{2 \hat{u}_{\theta_0}(r,t_f)}{r} \hat{\zeta}_{2s} + \frac{1}{Re} \nabla ^{2} \hat{\zeta}_{2c}, \end{gather}$$
(6.9b)$$\begin{gather}\frac{\partial \hat{\zeta}_{2s}}{\partial t} ={-} \hat{u}_{r_{2s}} \frac{\partial \hat{\zeta}_0 (r,t_f) }{\partial r} + \frac{2 \hat{u}_{\theta_0}(r,t_f)}{r} \hat{\zeta}_{2c} + \frac{1}{Re}\nabla ^{2} \hat{\zeta}_{2s}. \end{gather}$$

These equations describe simply the linear evolution of perturbations with azimuthal wavenumber $m=2$ on a steady axisymmetric vortex with azimuthal velocity $\hat {u}_{\theta _0}(r,t_f)$. The perturbations $(\hat {u}_{r_{2c}},\hat {u}_{\theta _{2c}})$ and $(\hat {u}_{r_{2s}},\hat {u}_{\theta _{2s}})$ are initialized by a white noise whose amplitude is adjusted so as to have a power of the same order as $E_{\zeta }(2,t_f)$. Figure 18 shows the evolution of the power $E_{\zeta }(2,t)$ for different freezing times $t_f$ ($t_f=40$, yellow dashed line; $t_f=50$, blue dashed line; $t_f=65$, red dashed line; $t_f=85$, green dashed line) compared with the evolution of $E_{\zeta }(2,t)$ in the truncated model (black solid line). Strikingly, we see that $E_{\zeta }(2,t)$ grows also exponentially regardless of the value of $t_f$ investigated. Furthermore, the growth rate, i.e. the slope, increases with $t_f$. Most interestingly, the growth rate for $t_f=85$ is close to that observed in the truncated model (black solid line). This demonstrates that the onset of non-axisymmetry in the vertical vorticity is due to an instability of the vortex profile. When the anomaly of vertical vorticity is sufficient to have an extremum, a shear instability with an azimuthal wavenumber $m=2$ develops. Subsequently, this triggers the growth of higher azimuthal modes through coupling with the non-traditional Coriolis force.

Figure 18. Evolution of the power $E_{\zeta }(2,t)$ in the truncated model (black solid line) or using (6.9) for different freezing times: $t_f=85$ (green dashed line), $t_f=65$ (red dashed line), $t_f=50$ (blue dashed line) and $t_f=40$ (yellow dashed line) for $Re=10\,000$, $F_h=2$, $Ro=20.3$ and $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$).

6.3. Equivalent vortex with piecewise uniform vorticity

A simple model of the instability can be obtained by considering the inviscid limit and by using a vortex with piecewise uniform vorticity with four concentric regions, as considered by Carton & Legras (Reference Carton and Legras1994) and Kossin, Schubert & Montgomery (Reference Kossin, Schubert and Montgomery2000). As shown by two examples in figure 19, the vorticity profile in the DNS can be crudely approximated by four levels of constant vorticity, i.e.

(6.10)\begin{equation} \zeta = \left\{ \begin{array}{ll} \zeta_1 = 2, & 0 < r < r_1, \\ \zeta_2 = \zeta_c - \delta_v/2, & r_1 < r < r_2, \\ \zeta_3 = \zeta_c + \delta_v/2, & r_2 < r < r_3, \\ \zeta_4 = 0, & r_3 < r, \end{array} \right. \end{equation}

where $r_1=r_c - \delta _h$, $r_2=r_c$ and $r_3= r_c + \delta _h$, with $\delta _v$ and $\delta _h$ the amplitude and size of the vorticity anomaly in the vicinity of the critical radius $r_c$. More explicitly, $\delta _v$ is the difference between the local maximum and minimum of the vorticity and $\delta _h$ is the distance between these two extrema. The corresponding angular velocity of the vortex is continuous and given by

(6.11)\begin{gather} \varOmega (r) = \tfrac{1}{2}\left\{ \begin{array}{ll} \zeta_1, & 0 < r < r_1, \\ \zeta_2 - (\zeta_2 -\zeta_1)(r_1/r)^{2}, & r_1 < r < r_2, \\ \zeta_3 - (\zeta_2 -\zeta_1)(r_1/r)^{2} - (\zeta_3 -\zeta_2)(r_2/r)^{2}, & r_2 < r < r_3, \\ - (\zeta_2 -\zeta_1)(r_1/r)^{2} - (\zeta_3 -\zeta_2)(r_2/r)^{2} + \zeta_3 (r_3/r)^{2}, & r_3 < r. \end{array} \right. \end{gather}

For a given Froude number $F_h$, the position of the critical radius $r_c$ and the value of $\zeta _c$ are fixed. Hence, the problem has only two control parameters: $\delta _v$ and $\delta _h$. The stability of such a vortex with respect to perturbations of the form $\psi \exp ({{\rm i}m\theta +\sigma t})$ is governed by the eigenvalue problem (Carton & Legras Reference Carton and Legras1994; Kossin et al. Reference Kossin, Schubert and Montgomery2000)

(6.12)\begin{align} &\begin{pmatrix} m \varOmega(r_1) + \frac{1}{2} (\zeta_2 - \zeta_1) & \frac{1}{2} (\zeta_2 - \zeta_1)(r_1/r_2)^{m} & \frac{1}{2} (\zeta_2 - \zeta_1)(r_1/r_3)^{m}\\ \frac{1}{2} (\zeta_3 - \zeta_2)(r_1/r_2)^{m} & m \varOmega(r_2) + \frac{1}{2} (\zeta_3 - \zeta_2) & \frac{1}{2} (\zeta_3 - \zeta_2)(r_2/r_3)^{m} \\ -\frac{1}{2} \zeta_3 (r_1/r_3)^{m} & -\frac{1}{2} \zeta_3 (r_2/r_3)^{m} & m \varOmega(r_3) - \frac{1}{2} \zeta_3 \end{pmatrix} \begin{pmatrix} \psi_1 \\ \psi_2 \\ \psi_3 \end{pmatrix} \nonumber\\ &\quad = \sigma \begin{pmatrix} \psi_1 \\ \psi_2 \\ \psi_3 \end{pmatrix} . \end{align}

Figure 19. Examples of the piecewise uniform vorticity (red line) fitting the continuous vertical vorticity profiles (black line) at $(a)$ $t=40$, $(b)$ $t=50$ for $Re=10\,000$, $F_h=2$, $Ro=20.3$ and $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$). The circle symbols represent the location of the critical radius.

Figure 20(a) shows the growth rate contours for the $m=2$ perturbations for $F_h=2$ as a function of $(\delta _v,\delta _h)$. We see that the growth rate is positive only when $\delta _v$ and $\delta _h$ are sufficiently away from zero in the ranges investigated. The symbols in figure 20(a) indicate the parameters $(\delta _v,\delta _h)$ estimated by fitting (6.10) to the vorticity field $\hat {\zeta }_0(r,t_f)$ at different times $t_f$ for $F_h=2$, for two different latitudes $\phi =80^{\circ }$ (red circles) and $\phi =75^{\circ }$ (black squares). For example, for $\phi =80^{\circ }$, the time $t_f$ varies from $t_f=45$ (leftmost point) to $t_f=85$ (rightmost point). The size of the vorticity anomaly $\delta _h$ does not vary very much and is around $\delta _h \simeq 0.2$ for both latitudes. In contrast, the amplitude of the anomaly $\delta _v$ increases with $t_f$ as expected. Figure 20(b) displays the growth rate as a function of $\delta _v$ (circle and square symbols). The dashed lines show the corresponding growth rate computed from (6.9), i.e. by considering the continuous vorticity profile $\hat {\zeta }_0(r,t_f)$ and with the perturbations initialized by white noise. This shows that the piecewise vortex model is able to predict quite well the growth rate of the instability observed in the truncated model, which is itself in good agreement with the DNS.

Figure 20. (a) Growth rate contours of the piecewise vortex model as a function of $\delta _h$ and $\delta _v$ for $F_h=2$. The contour interval is 0.03. The bold line indicates the growth rate $\sigma = 0$. The symbols correspond to the values of $\delta _h$ and $\delta _v$ estimated at different freezing times for $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$) (red circles) and $\phi =75^{\circ }$ ($\widetilde {Ro}=77.27$) (black squares) for $Re=10\,000$, $F_h=2$. (b) Growth rates as a function of $\delta _v$ obtained by the truncated model at different freezing times (symbols) and given by (6.12) for $m=2$ (dashed lines) for $\phi =75^{\circ }$ ($\widetilde {Ro}=77.27$) (black) and $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$) (red), $Re=10\,000$ and $F_h=2$.

Using (6.12), we have also computed the growth rate of higher azimuthal wavenumbers $m>2$. The results show that the most unstable wavenumber is not $m=2$ but is between $m=3$ and $m=5$ for the parameters indicated by the symbols in figure 20(a). Three reasons might explain the actual dominance of $m=2$. First, the velocity jumps in the piecewise vortex model could favour larger wavenumbers compared with a continuous vorticity profile. Second, we have seen from (4.27)–(4.28) that the non-traditional Coriolis force generates not only an axisymmetric vorticity at order $1/\widetilde {Ro}^{2}$ but also a vorticity field with an azimuthal wavenumber $m=2$. Figure 17(a) shows that the latter is weak before the onset of the instability. However, it is not zero and, therefore, this small amplitude could favour its dominance over more unstable higher wavenumbers whose initial amplitudes are much lower (see $m=4$ in figure 17a). Third, the vortex profile is continuously evolving with time while, in the stability problems (6.9) or (6.12), we have frozen this evolution. Hence, the $m=2$ wavenumber could be selected first when the vortex becomes slightly unstable. This early selection would then ensure its subsequent dominance even if it is no longer the most unstable wavenumber. Such an effect has been evidenced by Wang & Balmforth (Reference Wang and Balmforth2021) in their study of the evolution of the wavenumber selection as the critical layer becomes finer.

6.4. Theoretical criterion

Even if the Rayleigh–Fjørtoft criterion is only a necessary condition for the shear instability in inviscid fluids, we can try to use it to establish a theoretical criterion for the onset of the shear instability in the DNS for finite Reynolds numbers. Since the radial derivative of the vorticity is maximum at $r=r_c$ and is negative away from the critical radius, a necessary condition ensuring that there exists extrema, ${\rm d} \zeta / {\rm d} r = 0$, reads

(6.13)\begin{equation} \frac{{\rm d}\zeta (r_c,t_s)}{{\rm d}r} \geqslant c, \end{equation}

where ${c=0}$ is the minimum requirement for the existence of an inflection point, but we have explored also the consequences of larger values of $c$. Besides, the time $t_s$ will be set as $t_s=a\mathcal {T}$, where $a$ is a constant larger than unity. Indeed, the onset of the shear instability always occurs after the time $\mathcal {T}$. Therefore, the vorticity $\zeta$ will be taken as the asymptotic axisymmetric vorticity $\zeta = \zeta _0 + \varepsilon ^{2} \zeta _{20}$, where $\zeta _{20}$ is given by (4.36). As seen in figure 12, $\partial \zeta _{20}/\partial r(r_c,t)$ is indeed well predicted by (4.36) for large times. Then, (6.13) becomes

(6.14)\begin{equation} \frac{Re^{2/3}}{\widetilde{Ro}} \geqslant \left|{2 \varOmega_c^{\prime}}\right|^{1/3} \left (1+\frac{1}{Sc}\right)^{1/6} \sqrt{ \frac{\displaystyle { c - 3 \varOmega_c^{\prime} - r_c \varOmega_c^{\prime\prime}}}{\displaystyle { 2 {\rm \pi}r_c \varOmega_c \left( {Hi}(0) - \left(\frac{1+1/Sc}{8a{\rm \pi}^{2}{Hi}(0)}\right)^{1/2} \right)}}}. \end{equation}

Remarkably, the right-hand side depends only on the Froude number through $r_c$, the Schmidt number $S_c$ and the constants $a$ and $c$. The criterion (6.14) when ${c=0}$ and $a=\infty$ is represented by a solid line in figure 8. It delimits quite well the quasi-axisymmetric/non-axisymmetric domains observed in the DNS, except for the lowest Reynolds number investigated $Re=2000$. Such a difference for moderate Reynolds and Rossby numbers is not surprising since the asymptotics have been derived for a high Reynolds number and large Rossby number $\widetilde {Ro}$. Furthermore, viscous effects might damp the shear instability growth when the Reynolds number is moderate.

As seen from the piecewise vortex model, the shear instability for $m=2$ does not appear when $\delta _h \simeq 0.2$ as soon as $\delta _v >0$. It can be roughly estimated that the instability arises only when $\delta _v$ such that $\delta_v/\delta_h \gtrsim 0.4$ (figure 20a). Therefore, we can estimate $\partial \zeta /\partial r =$ $c \simeq \delta _v / \delta _h = 0.4$. The criterion (6.14) with this value of $c$ and $a = \infty$ is represented by a dashed line in figure 8. The agreement with the DNS is as good as the criterion (6.14) with ${c=0}$. The actual threshold is likely to be in between these two curves, i.e. in the hatched region (figure 8). Finally, we stress that the criterion (6.14) applies only to the shear instability due to an inflection point and not to other types of instability that may exist in viscous shear flows.

7. Late evolution of the vortex

Finally, figures 21 and 22 show the late evolution of the angular velocity profile (bottom row) when the instability develops or not, respectively. The DNS are the same as those already presented in § 3 for $Re=2000$, $F_h=2$ and different latitudes: $\phi =60^{\circ }$ ($\widetilde {Ro}=40$) (figures 4 and 21) and $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$) (figures 6 and 22). The corresponding vorticity fields are also shown again in the top row of figures 21 and 22 for convenience. In figure 21(c,d,g,h) we see that the instability ceases when the angular velocity is almost everywhere below $1/F_h$ (horizontal green dashed line), i.e. when a critical radius no longer exists. Due to the development of the critical layer and resulting instability, the decay of the angular velocity in the vortex core is accelerated compared with a pure viscous decay $\varOmega = ({1}/{r^{2}})(1- \exp ({-r^{2}}/{1+4t/Re}))$ (shown by red dashed lines in figure 21ef,g,h). When there is no instability (figure 22), the evolution of the angular velocity is slower and follows more closely a pure viscous diffusion law except in the vicinity of the critical radius where the decay is also slightly enhanced.

Figure 21. Vertical vorticity (ad) and angular velocity profile (eh) obtained from DNS (black line) at (a,e) $t=50$, (bf) $t=80$, (c,g) $t=120$ and (d,h) $t=250$ for $Re=2000$, $F_h=2$, $Ro=23.1$, $\phi =60^{\circ }$ ($\widetilde {Ro}=40$). The red dashed lines show the angular velocity profile if only viscous diffusion were active. The horizontal green dashed line represents the critical angular velocity value $1/F_{h}$.

Figure 22. Vertical vorticity (ad) and angular velocity profile (eh) obtained from DNS (black line) at (a,e) $t=50$, (bf) $t=80$, (c,g) $t=120$ and (d,h) $t=250$ for $Re=2000$, $F_h=2$, $Ro=23.1$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$). The red dashed lines show the angular velocity profile if only viscous diffusion were active. The horizontal green dashed line represents the critical angular velocity value $1/F_{h}$.

8. Conclusion

We have studied numerically and theoretically the evolution of a Lamb–Oseen vortex in a stratified-rotating fluid under the complete Coriolis force on the $f$ plane. The problem is governed mainly by the Froude number $F_h$, the Reynolds number $Re$ and the non-traditional Rossby number $\widetilde { Ro}$ based on the horizontal component of the background rotation.

Starting from a purely two-dimensional axisymmetric vortex, the DNS shows that a strong vertical velocity field with an azimuthal wavenumber $m=1$ is generated at a particular radius when the Froude number is larger than unity. This radius increases with the Froude number. Simultaneously, the vertical vorticity develops a quasi-axisymmetric anomaly near the same radius. Later, this anomalous ring may become fully non-axisymmetric with an azimuthal wavenumber $m=2$ when the Reynolds number is sufficiently large and the non-traditional Rossby number $\widetilde {Ro}$ not too large. At late time, the vorticity returns to a quasi-axisymmetric shape. Even if the vertical velocity is non-zero, all the fields remain independent of the vertical coordinate. In other words, the flow is 2D3C, i.e. two dimensional but with three velocity components. For this reason, the dynamics is independent of the traditional Rossby number $Ro$ based on the vertical component of the background rotation.

An asymptotic analysis for a large non-traditional Rossby number $\widetilde {Ro}$ has allowed us to unravel this evolution. First, it shows that the non-traditional Coriolis force generates a vertical velocity and buoyancy fields at order $1/\widetilde {Ro}$ that are invariant along the vertical. When the Froude number $F_h$ is larger than unity and in the absence of time dependence and viscous effects, these fields present a singularity at the radius where the angular velocity is equal to the Brunt–Väisälä frequency (i.e. the inverse of the Froude number in non-dimensional form). The asymptotic analyses show that this singularity is first regularized by the time dependence. This leads to a linear increase of the amplitude of the vertical velocity while the width of the critical layer shrinks at a rate inversely proportional to time. After a certain time, viscous effects saturate this evolution. The vertical velocity field is then steady with an amplitude proportional to $Re^{1/3}$ and a critical layer width scaling like $Re^{-1/3}$ as found by Boulanger et al. (Reference Boulanger, Meunier and Le Dizes2007) in the case of a tilted vortex in a stratified fluid. These asymptotic predictions are all in very good agreement with the DNS.

In turn, the non-traditional Coriolis force due to the vertical velocity modifies the vertical vorticity field at order $1/\widetilde {Ro}^{2}$. The dominant effect is the development of an axisymmetric ring of anomalous vorticity near the critical radius. This leads to the development of extrema in the vorticity profile. Again, the asymptotic predictions for the axisymmetric component of the vorticity are in good agreement with the DNS. Following Wang & Balmforth (Reference Wang and Balmforth2020, Reference Wang and Balmforth2021), we have further carried out a nonlinear asymptotic analysis that takes into account the effect of the anomaly of axisymmetric vorticity back on the evolution of the vertical velocity. The predictions of this nonlinear analysis are in better agreement with the DNS than those of the linear analysis indicating that both viscous and nonlinear effects operate in the critical layer.

In order to understand the origin of the subsequent non-axisymmetric evolution of the vorticity field, we have first decomposed the vertical velocity and vorticity in the DNS by an azimuthal Fourier transform. This analysis shows that several azimuthal modes grow exponentially during the onset of non-axisymmetry: the odd modes $m=3$, $m=5$, etc for the vertical velocity and the even modes $m=2$, $m=4$, etc for the vertical vorticity. We have then introduced a highly truncated model which keeps only the $m=1$ azimuthal wavenumber of the vertical velocity and the $m=0$ and $m=2$ wavenumbers of the vertical vorticity. Such a truncated model exhibits also an onset of non-axisymmetry like in the DNS demonstrating that this behaviour is not due to an unstable coupling between azimuthal modes. Furthermore, we have shown that if we freeze the profile of the axisymmetric component of the vertical vorticity at the time where the non-axisymmetry starts to appear and initializes the $m=2$ mode by white noise, the latter mode grows exponentially at a rate comparable to that observed in the DNS. In addition, the stability of an equivalent piecewise vortex model with four levels of vorticity has been investigated and found to give growth rates for $m=2$ in agreement with those of the truncated model. Altogether, this proves that the onset of non-axisymmetry comes from a two-dimensional shear instability related to the presence of a minimum in the vorticity profile. Finally, using the asymptotic expression of the axisymmetric component of the vorticity at late time at leading orders, the necessary condition for the shear instability has been converted into an instability condition in terms of $(Re,\widetilde {Ro})$. This condition delimits well the quasi-axisymmetric/non-axisymmetric domains in the parameter space $(Re,\widetilde {Ro})$.

The overall effect of the instability is to accelerate the decay of the angular velocity compared with a pure viscous diffusion. The instability ceases when the angular velocity is everywhere lower than the Brunt–Väisälä frequency (i.e. the inverse of the Froude number in non-dimensional form).

In summary, we have seen that the dynamics of a vortex for a large Reynolds number can be strongly affected by the non-traditional Coriolis force even if the non-traditional Rossby number $\widetilde {Ro}$ is large, i.e. even for a small value of the horizontal component of the background rotation. Since the typical Reynolds number of geophysical vortices is generally huge, this means that the non-traditional Coriolis force might have much more impact than expected by just considering its order of magnitude through the non-traditional Rossby number $\widetilde {Ro}$. It should be reminded however that another crucial condition is $F_h >1$ that ensures the presence of a critical layer. Hence, such process might affect intense but not too large vortices in geophysical flows.

In the future we will investigate the effect of three-dimensional perturbations on this phenomenon. Indeed, the vertical velocity is also responsible for an axial shear that might lead to another kind of shear instability if small three-dimensional perturbations are present as observed by Boulanger et al. (Reference Boulanger, Meunier and Le Dizes2007) for the case of a tilted vortex. It could be interesting also to study the configurations where the vortex is initially aligned with the background rotation vector or is not columnar.

Supplementary movies

Supplementary movies are available at https://doi.org/10.1017/jfm.2022.812.

Acknowledgements

We thank P. Meunier and T. Dubos for fruitful discussions and the referees for their helpful comments that have improved the manuscript.

Funding

This work was performed using HPC resources from GENCI-IDRIS (grant 2020-A0082A07419).

Declaration of interests

The authors report no conflict of interest.

Appendix A. Approximation of the solution of (4.34) for large time

The solution (4.35) can be simplified for large time $T \gg 1$, i.e. $t \gg Re^{1/3}$. It is first convenient to derive $\zeta _{20}$ with respect to $\tilde {r}$,

(A1)\begin{equation} \frac{\partial \zeta_{20}}{\partial \tilde{r}} = \frac{Re A}{2\gamma {\rm \pi}} \int_0^{\left|{\varOmega_c^{\prime}}\right|T} \exp \left( \frac{-q^{3}}{3\gamma^{3}} + {\rm i} q \tilde{r} \right) \left( 1-\exp\left( q^{3}/\left|{\varOmega_c^{\prime}}\right| - q^{2}T \right) \right ) \mathrm{d}q + {\rm c.c.} \end{equation}

By using the change of variable $z=q/\gamma$ for the first part of the integrand and $x=q\sqrt {T}$ for the second part, (A1) can be rewritten as

(A2)\begin{align} \frac{\partial\zeta_{20}}{\partial\tilde{r}} &= \frac{Re A}{2 {\rm \pi}} \int_0^{\left|{\varOmega_c^{\prime}}\right|T/\gamma} \exp \left( \frac{-z^{3}}{3} + {\rm i} \gamma z \tilde{r} \right) \mathrm{d}z\nonumber\\ &\quad -\frac{Re A}{2 \gamma {\rm \pi}\sqrt{T}} \int_0^{\left|{\varOmega_c^{\prime}}\right|T^{3/2}} \exp\left( \frac{-x^{3}}{3\gamma^{3}T^{3/2}} + {\rm i} \frac{x\tilde{r}}{\sqrt{T}} + \frac{x^{3}}{\left|{\varOmega_c^{\prime}}\right|T^{3/2}} - x^{2} \right) \mathrm{d} x + {\rm c.c.} \end{align}

When $T \gg 1$, the first integral tends to the Scorer's function (Abramowitz & Stegun Reference Abramowitz and Stegun1972) whereas the terms proportional to $1/T^{3/2}$ can be neglected compared with the other terms in the second integral. This yields

(A3)\begin{equation} \frac{\partial\zeta_{20}}{\partial \tilde{r}} \simeq \frac{Re A}{2} {Hi}({\rm i}\gamma \tilde{r}) - \frac{Re A}{2 \gamma {\rm \pi}\sqrt{T}} \int_0^{\left|{\varOmega_c^{\prime}}\right|T^{3/2}} \exp\left( - x^{2} + {\rm i} \frac{x\tilde{r}}{\sqrt{T}} \right) \mathrm{d} x + {\rm c.c.} \end{equation}

By introducing another change of variable

(A4)\begin{equation} U = x - \frac{{\rm i} \tilde{r}}{2\sqrt{T}}, \end{equation}

(A3) becomes

(A5)\begin{equation} \frac{\partial\zeta_{20}}{\partial\tilde{r}} \simeq \frac{Re A}{2} {Hi}({\rm i}\gamma \tilde{r}) - \frac{Re A}{2 \gamma {\rm \pi}\sqrt{T}} \exp \left( \frac{-\tilde{r}^{2}}{4T} \right) \int_{-{\rm i}\tilde{r}/(2\sqrt{T})}^{\left|{\varOmega_c^{\prime}}\right|T^{3/2}-{\rm i}\tilde{r}/(2\sqrt{T})} \mathrm{e}^{{-}U^{2}} \,\mathrm{d}U + {\rm c.c.} \end{equation}

The imaginary terms in the integral cancel with those of the complex conjugate giving

(A6)\begin{equation} \frac{\partial\zeta_{20}}{\partial\tilde{r}} = \frac{Re A}{2} \left[{Hi}({\rm i}\gamma \tilde{r}) + {Hi}^{*}({\rm i}\gamma \tilde{r}) \right] - \frac{Re A}{\gamma {\rm \pi}\sqrt{ T}} \exp \left( \frac{-\tilde{r}^{2}}{4T} \right) \int_{0}^{\left|{\varOmega_c^{\prime}}\right|T^{3/2}} \,\mathrm{e}^{{-}U^{2}} \,\mathrm{d}U. \end{equation}

The remaining integral can be approximated by $\sqrt {{\rm \pi} }/2$ since $T \gg 1$, leading finally to

(A7)\begin{equation} \frac{\partial\zeta_{20}}{\partial\tilde{r}} = \frac{Re A}{2} \left[{Hi}({\rm i}\gamma \tilde{r}) + {Hi}^{*}({\rm i}\gamma \tilde{r}) \right] - \frac{Re A}{2 \gamma \sqrt{{\rm \pi} T}} \exp \left( \frac{-\tilde{r}^{2}}{4T} \right). \end{equation}

Integrating back in $\tilde {r}$ gives the approximation (4.364.37).

References

REFERENCES

Abramowitz, M. & Stegun, I.A. (Ed.) 1972 Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables, 10th printing edn. US Government Printing Office.Google Scholar
Billant, P. & Bonnici, J. 2020 Evolution of a vortex in a strongly stratified shear flow. Part 2. Numerical simulations. J.Fluid Mech. 893, A18.CrossRefGoogle Scholar
Bonnici, J. 2018 Vertical decorrelation of a vortex by an external shear flow in a strongly stratified fluid. PhD thesis, Université Paris-Saclay.Google Scholar
Boulanger, N., Meunier, P. & Le Dizes, S. 2007 Structure of a stratified tilted vortex. J.Fluid Mech. 583, 443458.CrossRefGoogle Scholar
Boulanger, N., Meunier, P. & Le Dizes, S. 2008 Tilt-induced instability of a stratified vortex. J.Fluid Mech. 596, 120.CrossRefGoogle Scholar
Carton, X. & Legras, B. 1994 The life-cycle of tripoles in two-dimensional incompressible flows. J.Fluid Mech. 267, 5382.CrossRefGoogle Scholar
Deloncle, A., Billant, P. & Chomaz, J. 2008 Nonlinear evolution of the zigzag instability in stratified fluids: a shortcut on the route to dissipation. J.Fluid Mech. 599, 229239.CrossRefGoogle Scholar
Etling, D. 1971 The stability of an Ekman boundary layer flow as influenced by thermal stratification. Beitr. Phys. Atmos. 44, 168186.Google Scholar
Fjørtoft, R. 1950 Application of integral theorems in deriving criteria of stability for laminar flows and for the baroclinic circular vortex. Geophys. Publ. 17, 152.Google Scholar
Gerkema, T., Zimmerman, J.T.F., Maas, L.R.M. & Van Haren, H. 2008 Geophysical and astrophysical fluid dynamics beyond the traditional approximation. Rev. Geophys. 46, RG2004.CrossRefGoogle Scholar
Hayashi, M. & Itoh, H. 2012 The importance of the nontraditional Coriolis terms in large-scale motions in the tropics forced by prescribed cumulus heating. J.Atmos. Sci. 69 (9), 26992716.CrossRefGoogle Scholar
Igel, M.R. & Biello, J.A. 2020 The nontraditional Coriolis terms and tropical convective clouds. J.Atmos. Sci. 77 (12), 39853998.CrossRefGoogle Scholar
Kloosterziel, R.C., Carnevale, G.F. & Orlandi, P. 2017 Equatorial inertial instability with full Coriolis force. J.Fluid Mech. 825, 69108.CrossRefGoogle Scholar
Kossin, J.P., Schubert, W.H. & Montgomery, M.T. 2000 Unstable interactions between a hurricane's primary eyewall and a secondary ring of enhanced vorticity. J.Atmos. Sci. 57 (24), 38933917.2.0.CO;2>CrossRefGoogle Scholar
Lavrovskii, E.K., Semenova, I.P., Slezkin, L.N. & Fominykh, V.V. 2000 Mediterranean lenses as liquid gyroscopes in the ocean. Dokl. Phys. 45 (11), 606609.CrossRefGoogle Scholar
Park, J., Prat, V., Mathis, S. & Bugnet, L. 2021 Horizontal shear instabilities in rotating stellar radiation zones-II. Effects of the full Coriolis acceleration. Astron. Astrophys. 646, A64.CrossRefGoogle Scholar
Rayleigh, Lord 1880 On the stability, or instability, of certain fluid motions. Proc. Lond. Math. Soc. 9, 5770.Google Scholar
Semenova, I.P. & Slezkin, L.N. 2003 Dynamically equilibrium shape of intrusive vortex formations in the ocean. Fluid Dyn. 38 (5), 663669.CrossRefGoogle Scholar
Sheremet, V.A. 2004 Laboratory experiments with tilted convective plumes on a centrifuge: a finite angle between the buoyancy force and the axis of rotation. J.Fluid Mech. 506, 217244.CrossRefGoogle Scholar
Tort, M. & Dubos, T. 2014 Dynamically consistent shallow-atmosphere equations with a complete Coriolis force. Q. J. R. Meteorol. Soc. 140 (684), 23882392.CrossRefGoogle Scholar
Tort, M., Dubos, T., Bouchut, F. & Zeitlin, V. 2014 Consistent shallow-water equations on the rotating sphere with complete Coriolis force and topography. J.Fluid Mech. 748, 789821.CrossRefGoogle Scholar
Tort, M., Ribstein, B. & Zeitlin, V. 2016 Symmetric and asymmetric inertial instability of zonal jets on the-plane with complete Coriolis force. J.Fluid Mech. 788, 274302.CrossRefGoogle Scholar
Wang, C. & Balmforth, N.J. 2020 Nonlinear dynamics of forced baroclinic critical layers. J.Fluid Mech. 883, A12.CrossRefGoogle Scholar
Wang, C. & Balmforth, N.J. 2021 Nonlinear dynamics of forced baroclinic critical layers II. J.Fluid Mech. 917, A48.CrossRefGoogle Scholar
Wippermann, F. 1969 The orientation of vortices due to instability of the Ekman-boundary layer. Beitr. Phys. Atmos. 42 (4), 225244.Google Scholar
Zeitlin, V. 2018 Symmetric instability drastically changes upon inclusion of the full Coriolis force. Phys. Fluids 30 (6), 061701.CrossRefGoogle Scholar
Figure 0

Figure 1. Sketch of the initial vortex in a stratified fluid and in the presence of a background rotation $\varOmega _b$ inclined with an angle $\phi$.

Figure 1

Figure 2. Vertical velocity at different times: $(a)$$t=1$ , $(b)$$t=30$ , $(c)$$t=60$, $(d)$$t=75$, $(e)$$t=80$ and $(f)$$t=90$ for $Re=2000$, $F_h=10$, $Ro=23.1$, $\phi =60^{\circ }$ ($\widetilde {Ro}=40$).

Figure 2

Figure 3. Vertical vorticity at different times: $(a)$$t=1$ , $(b)$$t=30$ , $(c)$$t=60$, $(d)$$t=75$, $(e)$$t=80$ and $(f)$$t=90$ for $Re=2000$, $F_h=10$, $Ro=23.1$, $\phi =60^{\circ }$ ($\widetilde {Ro}=40$).

Figure 3

Figure 4. Vertical velocity (ad) and vertical vorticity (eh) at different times: (a,e) $t=10$, (bf) $t=40$, (c,g) $t=65$ and (d,h) $t=80$ for $Re=2000$, $F_h=2$, $Ro=23.1$, $\phi =60^{\circ }$ ($\widetilde {Ro}=40$).

Figure 4

Figure 5. Maximum vertical velocity as a function of time for $F_h=2$ and (a) $Re=2000$, $Ro=23.1$, $\phi =60^{\circ }$ ($\widetilde {Ro}=40$), (b) $Re=2000$, $Ro=20.3$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$) and (c) $Re=10\,000$, $Ro=20.3$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$).

Figure 5

Figure 6. Vertical velocity (ad) and vertical vorticity (eh) at different times: (a,e) $t=10$, (bf) $t=50$, (c,g) $t=150$ and (d,h) $t=200$ for $Re=2000$, $F_h=2$, $Ro=20.3$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$).

Figure 6

Figure 7. Vertical velocity (ad) and vertical vorticity (eh) at different times: (a,e) $t=10$, (bf) $t=100$, (c,g) $t=150$ and (d,h) $t=200$ for $Re=10\,000$, $F_h=2$, $Ro=20.3$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$).

Figure 7

Figure 8. Map of the simulations in the parameter space $(Re,\widetilde {Ro})$ for $F_h=2$. The yellow and blue circles represent simulations where the vertical vorticity remains quasi-axisymmetric or not, respectively. The solid and dashed lines represent the criterion (6.14) for different values of (a,c): $(\infty,0)$ and $(\infty,0.4)$, respectively. The number near some points indicate the figure numbers where the corresponding simulation is shown.

Figure 8

Figure 9. Comparison between the maximum vertical velocity in the DNS (black solid line), predicted by the unsteady inviscid solution (4.10a) (red dashed line), by the unsteady viscous solution (4.16a), (4.21) (yellow dashed line), by the steady viscous solution (4.16a), (4.23) (green dashed line) and by the nonlinear equations (4.45), (4.47) (blue dashed line) for $(a)$$\theta =0$ and $(b)$$\theta ={\rm \pi} /2$ for $Re=10\,000$, $F_h=2$, $Ro=20.3$, $\phi =80^{\circ }$ and for $(c)$$\theta =0$ and $(d)$$\theta ={\rm \pi} /2$ for $Re=2000$, $F_h=2$, $Ro=20.3$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$).

Figure 9

Figure 10. Comparison between the vertical velocity at $\theta =0$ (ac) and $\theta ={\rm \pi} /2$ (df) in the DNS (black solid line), predicted by the unsteady inviscid solution (4.10a) (red dashed line), by the unsteady viscous solution (4.21) (yellow dashed line) and by the nonlinear equations (4.45), (4.47) (blue dashed line) at (a,d) $t=5$, (b,e) $t=25$ and (cf) $t=40$ for $Re=10\,000$, $F_h=2$, $Ro=20.3$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$). The circle symbols represent the location of the critical radius in the unsteady inviscid solution (4.10a).

Figure 10

Figure 11. Comparison between the vertical velocity at $\theta =0$ (ac) and $\theta ={\rm \pi} /2$ (df) in the DNS (black solid line), predicted by the unsteady viscous solution (4.16a), (4.21) (yellow dashed line), by the steady viscous solution (4.16a), (4.23) (green dashed line) and by the nonlinear equations (4.45), (4.47) (blue dashed line) at (a,d) $t=40$, (b,e) $t=60$, (cf) $t=80$ for $Re=10\,000$, $F_h=2$, $Ro=20.3$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$). The circle symbols represent the location of the critical radius in the unsteady viscous solution (4.16a).

Figure 11

Figure 12. Comparison between the evolution of $\partial \zeta _{20}/\partial r(r_c,t)$ from the theoretical expressions: the unsteady viscous solution (4.35) (black solid line), the unsteady inviscid solution (4.31) (red dashed line), the viscous solutions (4.36) (blue dashed line) and (4.37a) (black dashed line) for $Re=10\,000$, $F_h=2$.

Figure 12

Figure 13. Comparison between the vertical vorticity at $\theta ={\rm \pi} /2$ in the DNS (black solid line) and the asymptotic expressions $\zeta = \zeta _0 + \varepsilon ^{2}\zeta _{20}$ where $\zeta _{20}$ follows the unsteady inviscid solution (4.31) (red dashed line) and $\zeta = \zeta _0 + \zeta _1$ with $\zeta _1$ given by the nonlinear equations (4.45), (4.47) (blue dashed line) at $(a)$$t=25$, $(b)$$t=35$ and $(c)$$t=40$ for $Re=10\,000$, $F_h=2$, $Ro=20.3$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$). The circle symbols represent the location of the critical radius in the unsteady inviscid solution (4.31).

Figure 13

Figure 14. Comparison between the vertical vorticity at $\theta ={\rm \pi} /2$ in the DNS (black solid line) and the asymptotic expressions $\zeta = \zeta _0 + \varepsilon ^{2}\zeta _{20}$ where $\zeta _{20}$ follows the unsteady viscous solution (4.35) (red dashed line) and $\zeta = \zeta _0 + \zeta _1$ with $\zeta _1$ given by the nonlinear equations (4.45), (4.47) (blue dashed line) at $(a)$$t=50$, $(b)$$t=65$ and $(c)$$t=85$ for $Re=10\,000$, $F_h=2$, $Ro=20.3$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$). The circle symbols represent the location of the critical radius in the unsteady inviscid solution (4.35).

Figure 14

Figure 15. Comparison between the vertical vorticity at $\theta ={\rm \pi} /2$ in the DNS (black solid line) and the asymptotic expressions $\zeta = \zeta _0 + \varepsilon ^{2}\zeta _{20}$ where $\zeta _{20}$ follows the viscous solution (4.36) (red dashed line) and $\zeta = \zeta _0 + \zeta _1$ with $\zeta _1$ given by the nonlinear equations (4.45), (4.47) and the viscous decay of $\zeta _0$ also taken into account (yellow dashed line) at $(a)$$t=200$, $(b)$$t=300$ and $(c)$$t=400$ for $Re=10\,000$, $F_h=2$, $Ro=20.0$, $\phi =87.7^{\circ }$ ($\widetilde {Ro}=500$). The circle symbols represent the location of the critical radius in the viscous solution (4.36).

Figure 15

Figure 16. Evolution of the power $(a)$$E_{\zeta }(m,t)$ for the azimuthal wavenumbers $m=0$ (black solid line), $m=2$ (red dashed line) and $m=4$ (green dashed line) and the power $(b)$$E_{u_z}(m,t)$ for the azimuthal wavenumbers $m=1$ (black solid line), $m=3$ (red dashed line) and $m=5$ (green dashed line) for $Re=10\,000$, $F_h=2$, $Ro=20.3$ and $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$).

Figure 16

Figure 17. Evolution of the power $(a)$$E_{\zeta }(2,t)$ and $(b)$$E_{u_z}(1,t)$ in the DNS (black solid line) and truncated model (red dashed line) for $Re=10\,000$, $F_h=2$, $Ro=20.3$ and $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$).

Figure 17

Figure 18. Evolution of the power $E_{\zeta }(2,t)$ in the truncated model (black solid line) or using (6.9) for different freezing times: $t_f=85$ (green dashed line), $t_f=65$ (red dashed line), $t_f=50$ (blue dashed line) and $t_f=40$ (yellow dashed line) for $Re=10\,000$, $F_h=2$, $Ro=20.3$ and $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$).

Figure 18

Figure 19. Examples of the piecewise uniform vorticity (red line) fitting the continuous vertical vorticity profiles (black line) at $(a)$$t=40$, $(b)$$t=50$ for $Re=10\,000$, $F_h=2$, $Ro=20.3$ and $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$). The circle symbols represent the location of the critical radius.

Figure 19

Figure 20. (a) Growth rate contours of the piecewise vortex model as a function of $\delta _h$ and $\delta _v$ for $F_h=2$. The contour interval is 0.03. The bold line indicates the growth rate $\sigma = 0$. The symbols correspond to the values of $\delta _h$ and $\delta _v$ estimated at different freezing times for $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$) (red circles) and $\phi =75^{\circ }$ ($\widetilde {Ro}=77.27$) (black squares) for $Re=10\,000$, $F_h=2$. (b) Growth rates as a function of $\delta _v$ obtained by the truncated model at different freezing times (symbols) and given by (6.12) for $m=2$ (dashed lines) for $\phi =75^{\circ }$ ($\widetilde {Ro}=77.27$) (black) and $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$) (red), $Re=10\,000$ and $F_h=2$.

Figure 20

Figure 21. Vertical vorticity (ad) and angular velocity profile (eh) obtained from DNS (black line) at (a,e) $t=50$, (bf) $t=80$, (c,g) $t=120$ and (d,h) $t=250$ for $Re=2000$, $F_h=2$, $Ro=23.1$, $\phi =60^{\circ }$ ($\widetilde {Ro}=40$). The red dashed lines show the angular velocity profile if only viscous diffusion were active. The horizontal green dashed line represents the critical angular velocity value $1/F_{h}$.

Figure 21

Figure 22. Vertical vorticity (ad) and angular velocity profile (eh) obtained from DNS (black line) at (a,e) $t=50$, (bf) $t=80$, (c,g) $t=120$ and (d,h) $t=250$ for $Re=2000$, $F_h=2$, $Ro=23.1$, $\phi =80^{\circ }$ ($\widetilde {Ro}=115.2$). The red dashed lines show the angular velocity profile if only viscous diffusion were active. The horizontal green dashed line represents the critical angular velocity value $1/F_{h}$.

Toghraei and Billant supplementary movie 1

Evolution of the vertical velocity (left) and vertical vorticity (right) for $Re=2000$, $F_h=10$, $Ro=23.1$, $\phi=60^{\circ}$ ($\widetilde{Ro}=40$).

Download Toghraei and Billant supplementary movie 1(Video)
Video 2.8 MB

Toghraei and Billant supplementary movie 2

Evolution of the vertical velocity (left) and vertical vorticity (right) for $Re=2000$, $F_h=2$, $Ro=23.1$, $\phi=60^{\circ}$ ($\widetilde{Ro}=40$).

Download Toghraei and Billant supplementary movie 2(Video)
Video 1.9 MB

Toghraei and Billant supplementary movie 3

Evolution of the vertical velocity (left) and vertical vorticity (right) for $Re=2000$, $F_h=2$, $Ro=20.3$, $\phi=80^{\circ}$ ($\widetilde{Ro}=115.2$).

Download Toghraei and Billant supplementary movie 3(Video)
Video 1.4 MB

Toghraei and Billant supplementary movie 4

Evolution of the vertical velocity (left) and vertical vorticity (right) for $Re=10000$, $F_h=2$, $Ro=20.3$, $\phi=80^{\circ}$ ($\widetilde{Ro}=115.5$).

Download Toghraei and Billant supplementary movie 4(Video)
Video 2.2 MB