Hostname: page-component-68c7f8b79f-pksg9 Total loading time: 0 Render date: 2025-12-18T11:29:02.375Z Has data issue: false hasContentIssue false

Magneto-gravity-precessional instability

Published online by Cambridge University Press:  16 December 2025

Abdelaziz Salhi*
Affiliation:
Faculté des Sciences de Tunis, Université Tunis El Manar , El Manar 2092, Tunisia
Waleed Mouhali
Affiliation:
ECE Paris-Ecole d’ingénieurs, 10 Rue Sextius-Michel. CS 71520, 75725 Paris CEDEX 15, France
Dhaou Lassoued
Affiliation:
Département de Mathématique, Faculté des Sciences de Gabes, Université de Gabes, 6072 Gabes, Tunisia
Thierry Lehner
Affiliation:
Laboratory for the study of the Universe and eXtreme phenomena (LUX), Observatoire de Paris, Université PSL et Paris Cité CNRS (UMR-8102), 5 place Jules Janssen, Meudon 1F-92195 Meudon CEDEX, France
*
Corresponding author: Abdelaziz Salhi, salhidec55@gmail.com

Abstract

Magneto-gravity-precessional instability, which results from the excitation of resonant magneto-inertia-gravity (MIG) waves by a background shear generic to precessional flows, is addressed here. Two simple background precession flows, that of Kerswell (1993 Geophys. Astrophys. Fluid Dyn. vol 72, no. 1–4, pp. 107–144), and that of Mahalov (1993 Phys. Fluids A: Fluid Dyn. vol. 5, no. 4, pp. 891–900), are considered. We analytically perform an asymptotic analysis to order ${ O}(\varepsilon ),$ where $\varepsilon$ denotes the Poincaré number, i.e. the precession parameter, and determine the maximum growth rate of the destabilizing subharmonic resonances of MIG waves: that between two fast modes, that between two slow modes and that between a fast mode and a slow mode (mixed modes). The domains of the $(K_0 B_0/\varOmega _0, N/\varOmega _0)\hbox{-}$plane for which this instability operates are identified, where $1/K_0$ denotes a characteristic length scale, $B_0$ is the unperturbed Alfvén velocity, $\varOmega _0$ is the rotation rate and $N$ denotes the Brunt–Väisälä frequency. We demonstrate that the $N\rightarrow 0$ limit is, in fact, singular (discontinuous). At large $K_0B_0/\varOmega _0,$ stable stratification acts to suppress the destabilizing resonance between two fast modes as well as that between two slow modes, whereas it revives the destabilizing resonance between a fast mode and a slow mode provided $N\lt \varOmega _0,$ because, without stratification, the maximal growth rate of this instability approaches zero as $K_0B_0/\varOmega _0\rightarrow +\infty .$ This would be relevant for the generation of the mean electromotive force, and hence, the $\alpha \hbox{-}$effect in helical magnetized precessional flows under weak stable stratification. Diffusive effects on the instability is considered in the simple case where the magnetic and thermal Prandtl numbers are both equal to one.

Information

Type
JFM Papers
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press

1. Introduction

Many geophysical and astrophysical systems have magnetic fields and undergo rapid rotation; they are certainly likely to support magneto-inertia (or magneto-Coriolis (MC)) waves. The inertial waves, which are often observed in natural rotating fluid systems, arise from the intrinsic properties of a rotating fluid where the Coriolis force provides the restoring mechanism (see, e.g. Greenspan Reference Greenspan1969). Alfvén (or hydromagnetic) waves are found in plasma (such as the solar corona) or fluids with high electrical conductivity (such as the Earth’s core, (see, e.g. Finlay et al. Reference Finlay, Maus, Beggan, Bondar, Chambodut, Chernova and Zvereva2010)). They are incompressible transverse oscillations that propagate along field lines with magnetic tension as the restoring force (see, e.g. Tomczyk et al. Reference Tomczyk, McIntosh, Keil, Judge, Schad, Seeley and Edmondson2007). In the laboratory experiment of a magnetized turbulent Taylor–Couette flow of liquid metal by Nornberg et al. (Reference Nornberg, Ji, Schartman, Roach and Goodman2010), the combined fast and slow MC waves were clearly identified where the observed slow MC wave is damped. Magneto-inertia-gravity (MIG) waves, also called magnetic-Archimedean–Coriolis waves (see, e.g. Fearn   Proctor Reference Fearn and Proctor1983; Buffett, Knezek   Holme Reference Buffett, Knezek and Holme2016; Salhi et al. Reference Salhi, Baklouti, Godeferd, Lehner and Cambon2017, Reference Salhi, Khlifi, Marino, Feraco, Foldes and Cambon2024; Mouhali et al. Reference Mouhali, Salhi, Lehner and Cambon2024), are disturbances that propagate in rotating and stratified magnetized flows, where the combinations of the Coriolis force, the buoyancy force and the magnetic tension provide the restoring mechanism.

In the present study, we present analytical developments characterizing some parametric instabilities occurring in stratified precessing magnetized flows: the subharmonic instabilities resulting from the excitation of resonant MIG waves by a background shear generic to precessional flows (see, e.g. Kerswell Reference Kerswell1993; Mahalov Reference Mahalov1993; Mason   Kerswell Reference Mason and Kerswell2002). Indeed, the precession-driven dynamics represents a possible candidate for the development of both three-dimensional (3-D) waves with embedded two-dimensional (2-D) vortices that are present in several astrophysical and geophysical systems (see, e.g. Glampedakis, Andersson   Jones Reference Glampedakis, Andersson and Jones2008; Barker Reference Barker2016a ; Khlifi et al. Reference Khlifi, Salhi, Nasraoui, Godeferd and Cambon2018; Pizzi et al. Reference Pizzi, Mamatsashvili, Barker, Giesecke and Stefani2022; Kumar et al. Reference Kumar, Pizzi, Mamatsashvili, Giesecke, Stefani and Barker2024).

The problem of precessionally driven flow in a rotating container has been the subject of several experimental, theoretical and numerical studies for many decades considering spherical and spheroidal geometries (see, e.g. Poincaré Reference Poincaré1910; Busse Reference Busse1968; Kerswell Reference Kerswell1993; Zhang, Chan   Liao Reference Zhang, Chan and Liao2014; Cébron et al. Reference Cébron, Laguerre, Noir and Schaeffer2019; Lam, Kong   Zhang Reference Lam, Kong and Zhang2021) as well as cylindrical geometries (see, e.g. Gans Reference Gans1970; Vladimirov   Tarassov Reference Vladimirov and Tarassov1984; Mahalov Reference Mahalov1993; Lehner et al. Reference Lehner, Mouhali, Léorat and Mahalov2010; Mouhali et al. Reference Mouhali, Lehner, Léorat and Vitry2012; Gao et al. Reference Gao, Meunier, Le Dizès and Eloy2021; Giesecke et al. Reference Giesecke, Vogt, Pizzi, Kumar, Gonzalez, Gundrum and Stefani2024).

In a purely hydrodynamical context, flows within precessing containers result from the complex interplay between inertial waves, Ekman boundary layers and the base flow. They can eventually become unstable and create space-filling turbulence (see, e.g. Malkus Reference Malkus1968). Three kinds of instabilities have been suggested for precession-driven flows: boundary layer instabilities (see, e.g. Lorenzani   Tilgner Reference Lorenzani and Tilgner2001; Buffett Reference Buffett2021), bulk parametric instabilities (see, e.g. Kerswell Reference Kerswell1993; Lin, Noir   Jackson Reference Lin, Noir and Jackson2014; Burmann   Noir Reference Burmann and Noir2022)) and centrifugal instabilities (see, e.g. Giesecke et al. Reference Giesecke, Vogt, Gundrum and Stefani2019).

Directly relevant to the problem with precession and shear is the special case of Poincaré’s (Reference Poincaré1910) elegant constant-vorticity solution for a precessing, oblate spheroid, addressed by Kerswell (Reference Kerswell1993). As an idealization of a small-scale disturbance evolving at the centre of a precessing spheroid, Kerswell (Reference Kerswell1993) considered an unbounded basic flow, so that the effect of boundaries are ignored, in the precessing frame, of sheared circular streamlines,

(1.1) \begin{equation} {\boldsymbol{U}}=\varOmega _0\!\left ( -y\hat {\boldsymbol{x}}+\left (x-2\varepsilon z\right )\!\hat {\boldsymbol{y}}\right )\! ,\quad {\boldsymbol{\varOmega }}_{\!p}=\varOmega _{\!p}\hat {\boldsymbol{x}}=\varOmega _0\varepsilon \hat {\boldsymbol{x}}, \end{equation}

where $\boldsymbol{U}$ is the velocity field of the precessing flow, ${\boldsymbol{x}}=x\hat {{\boldsymbol{x}}}+ y\hat {{\boldsymbol{y}}}+z\hat {{\boldsymbol{z}}}$ denotes the vector position, $(\hat {{\boldsymbol{x}}}, \hat {{\boldsymbol{y}}},\hat {{\boldsymbol{z}}})$ is an orthonormal basis of the precessing frame, $\varOmega _0$ is the solid body rotation rate (a positive constant) and $\varepsilon =\varOmega _{\!p}/\varOmega _0$ denotes the Poincaré parameter (a constant). In their study of nonlinear dynamics in a precessing plane fluid layer, Mason   Kerswell (Reference Mason and Kerswell2002) used the background flow (1.1) with stress-free boundary conditions.

The base flow (1.1) is an exact incompressible solution of the Navier–Stokes equation. It can be split into two parts: one part corresponding to a solid body rotation around $\hat {\boldsymbol{z}}$ with circular streamlines and the other part corresponding to a plane linear shear ${\boldsymbol{U}}_{\!s}=-2\varOmega _0\varepsilon z\hat {{\boldsymbol{y}}}.$ A physical interpretation of the modified flow can be found here, as also in Salhi   Cambon (Reference Salhi and Cambon2009): the misalignment of the basic `rapid’ angular velocity ${\boldsymbol{W}}_0=\boldsymbol{\nabla }\times {\boldsymbol{U}}(\varepsilon =0)$ and the precessional one ${\boldsymbol{\varOmega }}_{\!p}$ , treated as an external Coriolis force, induces a gyroscopic torque. This torque can be exactly balanced by the plane shear ${\boldsymbol{U}}_{\!s}.$ We note that Wiener et al. (Reference Wiener, Hammer, Swanson and Donnelly1990), who studied experimentally the stability of Taylor–Couette flow subject to an external Coriolis force, have observed an axial component in the initial flow, and a transition to turbulence for larger rotation rates. Mahalov (Reference Mahalov1993) conjectured that this transition is indicative of parametric instabilities developing within the flow. Hereafter, the solution (1.1) is referred to as Kerswell’s base flow (KBF).

By using the formalism with projection of the disturbance fields onto Kelvin modes, Kerswell (Reference Kerswell1993) demonstrate the local (parametric) instability of these sheared circular streamlines to the straining field present, confirming that the effect of boundaries is largely secondary. The Kelvin modes are essentially 3-D Fourier modes, even if the wavevector ${\boldsymbol{k}}(t)$ can become time-dependent following the mean flow streamlines. The time dependency of the wavevector represents the advection of the plane wave $\exp (\textrm{i} {\boldsymbol{k}}(t){\boldsymbol{ \boldsymbol{\cdot }}} {\boldsymbol{x}})$ by the base flow (see, e.g. Cambon, Teissedre   Jeandel Reference Cambon, Teissedre and Jeandel1985; Bayly Reference Bayly1986; Craik   Criminale Reference Craik and Criminale1986). Note that the system of equations for disturbances in terms of Fourier (or Lagrangian) modes is recovered for small-scale disturbances travelling near any smooth base-flow trajectory, in the zonal asymptotic method of Lifschitz   Hameiri (Reference Lifschitz and Hameiri1991) with close connection with geometric optics: the velocity gradients of the base flow are treated as space-uniform in a domain of unspecified length scale, asymptotically small.

The mechanism of the parametric precessional instability is similar to that which underlies the elliptical instability (Bayly Reference Bayly1986; Pierrehumbert Reference Pierrehumbert1986): a resonance in which a pair of normal modes of oscillation on the undistorted circular flow become tuned to the underlying strain field. A review of different physical aspects of the elliptical instability in the theory of turbulence and in geophysics and astrophysics was done by Kerswell (Reference Kerswell2002) (also see the recent study by McKeown et al. Reference McKeown, Ostilla-Mónico, Pumir, Brenner and Rubinstein2020), and the study of the magneto-gravity-elliptic (MGE) instability by Salhi   Cambon (Reference Salhi and Cambon2023) (hereafter SC23).

Another simple example of a generic precessing flow is that considered by Mahalov (Reference Mahalov1993): an infinite column in which a tilted (sheared) streamline solution can exist under precession (see figure 1),

(1.2) \begin{equation} {\boldsymbol{U}}=\left (\varOmega _0r\right )\!\hat {\boldsymbol{\varphi }}-2\!\left (\varOmega _0\varepsilon r\sin \varphi \right )\!\hat {\boldsymbol{z}},\quad {\boldsymbol{\varOmega }}_{\!p}=\varOmega _0\varepsilon \left (\left (\cos \varphi \right )\!\hat {\boldsymbol{r}}-\left (\sin \varphi \right )\!\hat {\boldsymbol{\varphi }}\right )=\varOmega _0\varepsilon \hat {\boldsymbol{x}}, \end{equation}

where $(r,\varphi ,z)$ denotes a cylindrical coordinate system of an orthonormal basis $(\hat {\boldsymbol{r}},\hat {\boldsymbol{\varphi }},\hat {\boldsymbol{z}}).$ An associated boundary condition is that there is no flow through the boundaries. By means of a classical normal mode analysis of small disturbances (Drazin   Reid Reference Drazin and Reid2004), Mahalov (Reference Mahalov1993) showed that this distortion leads to 3-D parametric instabilities. In a Cartesian coordinate system of the orthonormal basis $(\hat {\boldsymbol{x}},\hat {\boldsymbol{y}},\hat {\boldsymbol{z}}),$ Mahalov’s solution can be rewritten as (see Salhi   Cambon Reference Salhi and Cambon2009),

(1.3) \begin{equation} {\boldsymbol{U}}=\varOmega _0\!\left ( -y\hat {\boldsymbol{x}}+x\!\hat {\boldsymbol{y}}-2\varepsilon y\hat {\boldsymbol{z}}\right )\! ,\quad {\boldsymbol{\varOmega }}_{\!p}=\varOmega _{\!p}\hat {\boldsymbol{x}}=\varOmega _0\varepsilon \hat {\boldsymbol{x}} ,\end{equation}

ignoring the boundary conditions. The latter unbounded base flow, which is an exact solution of the incompressible Navier–Stokes equation, can similarly model localized patches of precessing columns. Hereafter it is referred to as Mahalov’s base flow (MBF).

Figure 1. A slice of a vertical stratified precessing fluid column: a rotating fluid column seen in a frame rotating uniformly about the $\hat {\boldsymbol{x}}\hbox{-}$ axis with rate $\varOmega _{\!p}=\varepsilon \varOmega _0$ where $\varepsilon$ is the Poincaré number. Without precession $(\varepsilon =0),$ the unperturbed base flow, ${\boldsymbol{U}}=\varOmega _0r\hat \varphi ,$ has circular streamlines, and the isodensity planes are perpendicular to the vertical axis $\hat {\boldsymbol{z}},$ $\boldsymbol{\nabla }\varrho (z)=-(\rho _0/g)N^2\hat {\boldsymbol{z}},$ where $N^2=\text{const.}\gt 0$ is the square of the Brunt–Väisälä frequency. The effect of the Coriolis force induces a vertical mean shear that acts to balance the gyroscopic torque, so that the unperturbed velocity profile becomes ${\boldsymbol {U}} = \varOmega _0r (\hat {\boldsymbol{\varphi }}-2\varepsilon \sin \varphi \hat {\boldsymbol{z}})=\varOmega _0 (-y\hat {\boldsymbol{x}}+x\hat {\boldsymbol{y}}-2\varepsilon \hat {\boldsymbol{z}}).$ The trajectory of a fluid particle is in the plane perpendicular to the $z^*\hbox{-}$ axis and it is an ellipse (see § 2.2.1). The plane $(x^*,z^*)$ is obtained by a rotation, of angle $\gamma =-\tan ^{-1}(2\varepsilon ),$ of the plane $(x,z)$ around the $y\hbox{-}$ axis. Moreover, in the presence of the Coriolis force, the isodensity planes are perpendicular to the $z^*\hbox{-}$ axis, so that $\boldsymbol{\nabla }\varrho (z^*)=-(\rho _0/g)N^2\hat {\boldsymbol{z}}^*$ with ${\boldsymbol{n}}=\hat {\boldsymbol{z}}^*$ (see (2.17)). Here, the colour variation of the streamlines (from red to blue) represents the linear variation of the basic density, and ${\boldsymbol{B}}_0$ denotes the basic (constant) magnetic field.

The first direct numerical simulations of homogeneous hydrodynamical turbulence due to precession were performed in Barker (Reference Barker2016b ), and there were also earlier simulations with stress-free impermeable walls in Mason   Kerswell (Reference Mason and Kerswell2002). The direct numerical simulations studies indicate a saturation of the precessional instability due to the nonlinear interactions and the development of both 3-D waves with embedded 2-D vortices (see, e.g. Khlifi et al. Reference Khlifi, Salhi, Nasraoui, Godeferd and Cambon2018; Pizzi et al. Reference Pizzi, Mamatsashvili, Barker, Giesecke and Stefani2022). Note that the columnar vortices are commonly found in turbulence subjected to rapid rotation, and would be expected to correspond with zonal flows (Barker Reference Barker2016b ). However, in precessing magnetohydrodynamic (MHD) turbulence, the formation of the vortices is inhibited even for a weak background magnetic field (see, e.g. Barker, Reference Barker2016a ; Salhi, Khlifi   Cambon Reference Salhi, Khlifi and Cambon2019). It is worth noting that the latter study extends that of Salhi, Lehner   Cambon (Reference Salhi, Lehner and Cambon2010) which deals with a linear stability analysis of an unbounded precessing magnetized flow subjected to a background magnetic field, similarly to the Floquet analysis of Lebovitz   Zweibel (Reference Lebovitz and Zweibel2004) (hereafter LZ04) for unbounded elliptical magnetized flow.

In the context of the Earth’s magnetic field generation, i.e. the dynamo effect in the liquid outer core (see, e.g. Braginsky Reference Braginsky1991), Tilgner (Reference Tilgner1999) studied MHD flow in precessing spherical shells with and without an imposed magnetic dipole field. He showed that, when a magnetic dipole field with its dipole oriented along the rotation axis of the shell is applied, shear zones develop additional structure and change position and orientation (with respect to the case without a magnetic dipole). Note that several numerical simulations of precession driven dynamos have been performed in order to study the feasibility of dynamo action, as well as the existence of magnetic dipole reversals. In these studies, the incompressible MHD equations (see § 2) have been simulated with a non-zero initial magnetic energy (see, e.g. Cébron et al. Reference Cébron, Laguerre, Noir and Schaeffer2019; Etchevest, Fontana   Dmitruk Reference Etchevest, Fontana and Dmitruk2022).

As for the effects of stable density stratification on the precessing flow, Wei   Tilgner (Reference Wei and Tilgner2013) carried out a numerical calculation of the flow driven by both convection and precession in a rotating spherical shell. They found that a stable stratification suppresses possible precessional instabilities, whereas an unstable stratification can lead to both stabilization and destabilization, depending on the parameters such as Ekman, Nusselt, Rayleigh and Poincaré numbers (see, e.g. Wei Reference Wei2016; Vormann   Hansen Reference Vormann and Hansen2020).

The effect of an axial stable density stratification on Taylor columns has been previously investigated experimentally and theoretically (see, e.g. Davies Reference Davies1972; Caton, Janiaud   Hopfinger Reference Caton, Janiaud and Hopfinger2000; Hollerbach   Cally Reference Hollerbach and Cally2009). Benkacem et al. (Reference Benkacem, Salhi, Khlifi, Nasraoui and Cambon2022) extended Mahalov’s theoretical study for precessing ‘infinite’ Taylor columns (see the solution (1.2)) by including the effect of stratification: they implemented a correction to the basic buoyancy scalar field (see § 2). They showed that stable stratification acts in such a way as to make the subharmonic instability less efficient, so as it disappears for $N\geqslant 0.5\varOmega _0$ where $N$ is the Brunt–Väisälä frequency (defined in § 2). Similarly, in the case of unbounded strained-vortex flow with stable axial stratification studied by Miyazaki   Fukumoto (Reference Miyazaki and Fukumoto1992), the growth rates for the elliptical instability were invariably reduced: the subharmonic elliptical instability is completely suppressed when $N\geqslant \varOmega _0$ (see also Kerswell Reference Kerswell2002).

The outline of the paper is structured as follows. The Boussinesq MHD equations as well as the solution for the unperturbed fields (velocity, magnetic field and density) are presented in § 2. The perturbed system in physical space and its counterpart in Fourier space are presented in § 3. The use of the magnetic induction potential scalar (MIPS) (i.e. the inner product of the buoyancy gradient by the magnetic field, (see, Salhi et al. Reference Salhi, Lehner, Godeferd and Cambon2012)) which is an inviscid invariant, allows us to obtain a four-dimensional Floquet system. In § 4, the stability analysis of the Floquet system is addressed and asymptotic formulae for the maximal growth rate of the subharmonic instabilities are deduced. In § 5, a selection of numerical results is presented and compared with asymptotic formulae. The effect of diffusion is briefly addressed in the special case where the diffusion coefficients (kinetic, thermal and magnetic) are equal. Conclusions and perspectives are offered in § 6.

2. Mathematical formulation

2.1. The Boussinesq MHD equations

The Boussinesq approximation ignores density differences except where they appear in terms multiplied by the acceleration due to the buoyancy force (see, e.g. Spiegel   Veronis Reference Spiegel and Veronis1960; Salhi et al. Reference Salhi, Lehner, Godeferd and Cambon2013). Accordingly, the continuity equation reduces to $\boldsymbol{\nabla }{\boldsymbol{\cdot }} \tilde {\boldsymbol{u}}=0$ as for an incompressible fluid.

In a frame rotating uniformly about a fixed axis at the rate ${{\boldsymbol{\varOmega }}}_{\!p}$ , the Boussinesq MHD equations for the instantaneous velocity field $\tilde {\boldsymbol{u}}({\boldsymbol{x}},t),$ magnetic field $\tilde {\boldsymbol{b}}({\boldsymbol{x}},t)$ and buoyancy scalar $\tilde \vartheta ({\boldsymbol{x}},t)$ can be written as (see, e.g. Davidson Reference Davidson2013)

(2.1a) \begin{align} \left (\partial _t +\tilde {{\boldsymbol{u}}}{\boldsymbol{\cdot }}{\boldsymbol{\nabla }}\right )\!{\tilde {{\boldsymbol{u}}}}&= -\boldsymbol{\nabla }{\tilde p}-2{{\boldsymbol{\varOmega }}}_{\!p}\times \tilde {{\boldsymbol{u}}} +\big (\tilde {{\boldsymbol{b}}}\,{\boldsymbol{\cdot }} {\boldsymbol{\nabla }}\big )\tilde {{\boldsymbol{b}}} +\tilde {\vartheta }{\boldsymbol{n}}+\nu {\nabla} ^2 \tilde {{\boldsymbol{u}}}, \end{align}
(2.1b) \begin{align} \left (\partial _t +\tilde {{\boldsymbol{u}}}{\boldsymbol{\cdot }}{\boldsymbol{\nabla }}\right ){\tilde {{\boldsymbol{b}}}}&= \big (\tilde {{\boldsymbol{b}}}\,{\boldsymbol{\cdot }}{\boldsymbol{\nabla }}\big )\tilde {{\boldsymbol{u}}}+\eta {\nabla} ^2\tilde {{\boldsymbol{b}}}, \end{align}
(2.1c) \begin{align} \left (\partial _t +\tilde {{\boldsymbol{u}}}{\boldsymbol{\cdot }}{\boldsymbol{\nabla }}\right )\!{\tilde {\vartheta }}&=\kappa {\nabla} ^2\tilde {\vartheta }, \end{align}
(2.1d) \begin{align} \boldsymbol{\nabla }{\boldsymbol{\cdot }}\, {\tilde {{\boldsymbol{u}}}}&=0, \end{align}
(2.1e) \begin{align} \boldsymbol{\nabla }{\boldsymbol{\cdot }}\, {\tilde {{\boldsymbol{b}}}}&=0. \end{align}

Here, the magnetic field is scaled using Alfvén velocity units: it is divided by $\sqrt {\rho _0\mu _0}$ where $\rho _0$ and $\mu _0$ are the constant density and the magnetic permeability of the fluid, $\tilde p$ being the total pressure (including the centrifugal potential and the magnetic part) divided by the constant density $\rho _0,$ and ${\boldsymbol{\varOmega }}_{\!p}$ is the rotation vector of the reference frame which is constant and $\nu ,$ $\eta$ and $\kappa$ are positive constants that denote the kinematic viscosity, magnetic diffusivity and thermal diffusivity, respectively, and $\boldsymbol{n}$ is a unit vector that aligns with the basic buoyancy scalar gradient. As indicated by an anonymous referee, the natural systems possess a non-negligible adiabatic gradient, sometimes very strong (unlike in experimental situations), and hence, $\tilde \vartheta$ should be really defined as an excess over the adiabatic profile (influencing also the basic pressure field), so that the equations indeed look like (2.1).

The equation for the absolute vorticity vector, $\tilde {{\boldsymbol{w}}}= \boldsymbol{\nabla }\times \tilde {{\boldsymbol{u}}}+2{\boldsymbol{\varOmega }}_{\!p},$ is obtained by taking the curl of the momentum equation

(2.2) \begin{equation} \partial _t {\tilde {{\boldsymbol{w}}}}= \boldsymbol{\nabla }\times \big (\tilde {{\boldsymbol{u}}}\times \tilde {{\boldsymbol{w}}}+\tilde {{\boldsymbol{j}}}\times \tilde {{\boldsymbol{b}}}\big )+\boldsymbol{\nabla }\tilde \vartheta \times {\boldsymbol{n}}+\nu {\nabla} ^2\tilde {{\boldsymbol{w}}}, \end{equation}

where $\tilde {{\boldsymbol{j}}}=\boldsymbol{\nabla }\times \tilde {{\boldsymbol{b}}}$ is the normalized current density. In addition, we report here the equation for the gradient of the buoyancy scalar,

(2.3) \begin{equation} \left (\partial _t +\tilde {{\boldsymbol{u}}}{\boldsymbol{\cdot }}{\boldsymbol{\nabla }}\right )\big (\boldsymbol{\nabla }{\tilde {\vartheta }}\big )=-\big (\boldsymbol{\nabla }\tilde {{\boldsymbol{u}}}\big )^T{\boldsymbol{\cdot }} \big (\boldsymbol{\nabla }\tilde {\vartheta }\big )+\kappa {\nabla} ^2\big (\boldsymbol{\nabla }\tilde \vartheta \big ), \end{equation}

where $T$ denotes the transpose. Accordingly, from (2.2) and (2.3) we deduce the equation for the potential vorticity, $\tilde \varPi _\kappa = \tilde {{\boldsymbol{w}}}{\boldsymbol{\cdot }}{\boldsymbol{\nabla }} \tilde {\vartheta },$ which is very useful as an invariant in stratified geophysical flows (see, e.g. Pedlosky Reference Pedlosky2013),

(2.4) \begin{equation} \left (\partial _t +\tilde {{\boldsymbol{u}}}{\boldsymbol{\cdot }}{\boldsymbol{\nabla }}\right )\!{\tilde \varPi _\kappa }= \big (\boldsymbol{\nabla }\tilde \vartheta \big ){\boldsymbol{\cdot }} \big ( \boldsymbol{\nabla }\times \big (\tilde {{\boldsymbol{j}}}\times \tilde {{\boldsymbol{b}}}\big )\big )+ \nu \big (\boldsymbol{\nabla }\tilde \vartheta \big ){\boldsymbol{\cdot }} {\nabla} ^2\tilde {\boldsymbol{w}}+ \kappa \tilde {\boldsymbol{w}}{\boldsymbol{\cdot }}{\nabla} ^2\big (\boldsymbol{\nabla }\tilde \vartheta \big ). \end{equation}

It follows that, in the case of a magnetized Boussinesq ideal fluid, $\tilde \varPi _\kappa$ is not a Lagrangian invariant: while it removes the baroclinic torque, it does not remove the Lorentz torque. In contrast, the so-called MIPS, $\tilde \varPi _m=\tilde {\boldsymbol{b}}{\boldsymbol{\cdot }} \boldsymbol{\nabla }\tilde \vartheta$ (see, Salhi et al. Reference Salhi, Lehner, Godeferd and Cambon2012), which is a solution of the following transport equation:

(2.5) \begin{equation} \left (\partial _t +\tilde {{\boldsymbol{u}}}{\boldsymbol{\cdot }}{\boldsymbol{\nabla }}\right )\!{\tilde \varPi _m}= \eta \big (\boldsymbol{\nabla }\tilde \vartheta \big ){\boldsymbol{\cdot }} {\nabla} ^2\tilde {\boldsymbol{b}}+ \kappa \tilde {\boldsymbol{b}}{\boldsymbol{\cdot }}{\nabla} ^2\big (\boldsymbol{\nabla }\tilde \vartheta \big ), \end{equation}

deduced from (2.1b ) and (2.3), is a Lagrangian invariant for a magnetized Boussinesq ideal fluid. As will be shown later, the use of the magnetic invariant (i.e. MIPS) makes it possible to reduce the linear system of ordinary differential equations, which represents a Floquet problem of the inviscid stability problem, from a five-dimensional system to a four-dimensional one only.

2.2. The unperturbed system

The solutions of system (2.1) and (2.2) are conveniently decomposed into a ‘basic flow’ ( ${\boldsymbol{U}},P,{\boldsymbol{W}}, {\boldsymbol{B}},\varTheta )$ and a `disturbance’ $({\boldsymbol{u}},p,{\boldsymbol{w}}, {\boldsymbol{b}},\vartheta )$ as

(2.6) \begin{equation} \tilde {\boldsymbol{u}}={\boldsymbol{U}}+{\boldsymbol{u}},\quad \tilde p=P+ p,\quad \tilde {\boldsymbol{w}}= {{\boldsymbol{W}}}+{\boldsymbol{w}},\quad \tilde {\boldsymbol{b}}={\boldsymbol{B}}_0+{\boldsymbol{b}},\quad \tilde \vartheta =\varTheta +\vartheta , \end{equation}

where the basic flow is a solution of system (2.1). As indicated in § 1, we consider the unbounded basic flows KBF and MBF given by (1.1) and (1.3), respectively, that can be thought to represent a small patch of a stratified precessing magnetized flow subject to a background magnetic field.

To discuss the admissibility conditions, i.e. the basic flow must be a solution of the system (2.1) (see, Craik Reference Craik1989), it is convenient to consider the equation for the basic absolute vorticity (2.2), the induction (2.1b ) and the equation for the basic buoyancy scalar gradient (2.3) supplemented by (2.1d , e ),

(2.7a) \begin{align} \left (\partial _t+{\boldsymbol{U}}{\boldsymbol{\cdot }} \boldsymbol{\nabla }\right )\!{\boldsymbol{W}}&={\boldsymbol{W}}{\boldsymbol{\cdot }}\boldsymbol{\nabla }{\boldsymbol{U}}+ \boldsymbol{\nabla }\times \left (\left ({\boldsymbol{B}}_0{\boldsymbol{\cdot }}\boldsymbol{\nabla }\right )\!{\boldsymbol{B}}_0\right )+\nu {\nabla} ^2{\boldsymbol{W}}, \end{align}
(2.7b) \begin{align} \left (\partial _t +{{\boldsymbol{U}}}{\boldsymbol{\cdot }}{\boldsymbol{\nabla }}\right )\!{{{\boldsymbol{B}}_0}}&={\boldsymbol{B}}_0{\boldsymbol{\cdot }}\boldsymbol{\nabla }{\boldsymbol{U}}+\eta {\nabla} ^2\!{{\boldsymbol{B}}_0}, \end{align}
(2.7c) \begin{align} \left (\partial _t +{{\boldsymbol{U}}}{\boldsymbol{\cdot }}{\boldsymbol{\nabla }}\right )\!\left (\boldsymbol{\nabla }{{\varTheta }}\right )&=- \!\left (\boldsymbol{\nabla }{\varTheta }\right )\!{\boldsymbol{\cdot }} \boldsymbol{\nabla }\!\left ({\boldsymbol{U}}\right )^T+\kappa {\nabla} ^2\!\left (\boldsymbol{\nabla }\varTheta \right )\! , \end{align}

with $\boldsymbol{\nabla }{\boldsymbol{\cdot }} {\boldsymbol{U}}=0$ and $\boldsymbol{\nabla }{\boldsymbol{\cdot }} {\boldsymbol{B}}_0=0.$ Recall that the unit vector $\boldsymbol{n}$ aligns with the basic buoyancy scalar gradient, so that, $\boldsymbol{\nabla }\varTheta \times {\boldsymbol{n}}={{\boldsymbol{0}}}.$

2.2.1. Basic velocity

Recall that the KBF, i.e. the velocity field of the precessing flow described by (1.1), approximates the laminar Poincaré flow far from the boundaries of a spheroid if $\varepsilon \ll 1,$ and the tidal deformation is neglected (see, Poincaré Reference Poincaré1910; Kerswell Reference Kerswell1993; Salhi   Cambon Reference Salhi and Cambon2009). In this case, the basic absolute vorticity vector, ${\boldsymbol{W}}=\boldsymbol{\nabla }\times {\boldsymbol{U}}+2\varOmega _{\!p}=2\varOmega _0 (2\varepsilon \hat {\boldsymbol{x}}+\hat {\boldsymbol{z}} ),$ does not align with the solid body rotation. In counterpart, the MBF, i.e. the velocity field of the precessing flow described by (1.3), characterizes an infinite column in which a tilted (sheared) streamline solution can exist under precession (see, Mahalov Reference Mahalov1993; Salhi   Cambon Reference Salhi and Cambon2009). In this case, ${\boldsymbol{W}}=2\varOmega _0\hat {\boldsymbol{z}}$ aligns with the solid body rotation.

Accordingly, in the pure hydrodynamic case, so that ${\boldsymbol{B}}_0$ and $\boldsymbol{\nabla }\varTheta$ are zero, both KBF and MBF are exact solutions of (2.7a ). Note that, the main difference between these two base flows is the cross-gradient direction of the plane shear: for the KBF it is along $\hat {\boldsymbol{y}}$ and varies linearly with respect to the vertical space coordinate $z,$ while for the MBF it is along the vertical axis $\hat {\boldsymbol{z}}$ and varies linearly with respect to the horizontal space coordinate  $y.$ Note that the trajectory of a fluid particle can be determined by resolving the equation $d_t{\boldsymbol{x}}=\boldsymbol{\nabla }{\boldsymbol{U}}{\boldsymbol{\cdot }}{\boldsymbol{x}}.$ Accordingly, for the KBF, for example, we find

(2.8) \begin{equation} x(\tau )=x_{0}\cos \tau -y_{0}\sin \tau ,\ y(\tau )=x_{0}\sin \tau +y_{0}\cos \tau ,\ z(\tau )=z_0+2\varepsilon (x-x_0), \end{equation}

where $\tau =\varOmega _0t$ is a dimensionless time and ${\boldsymbol{x}}_0=(x_0,y_0,z_0)^T$ is the initial (at $t=0)$ vector position. It follows that the fluid particle describes an ellipse in the $(x^*,y^*)\hbox{-}$ plane,

(2.9a) \begin{align}&\qquad\qquad {\boldsymbol{x}}=x\hat {\boldsymbol{x}}+y\hat {\boldsymbol{y}}+z\hat {\boldsymbol{z}}=x^*\hat {\boldsymbol{x}}^*+y^*\hat {\boldsymbol{y}}^*+z^*\hat {\boldsymbol{z}}^*, \end{align}
(2.9b) \begin{align} x^*&=\frac {1}{\sqrt {1+4\varepsilon ^2}}\left (x+2\varepsilon z\right )\! ,\quad y^*=y,\quad z^*=\frac {1}{\sqrt {1+4\varepsilon ^2}}\left (-2\varepsilon x+z\right )\! . \end{align}

Indeed, one easily verifies that $z^*$ (given by (2.9b )) is time-independent and

(2.10) \begin{equation} \frac {1}{\sqrt {1+4\varepsilon ^2}}\left (x^*-2\varepsilon z^*\right )^2+y^{*2}=\text{const}. \end{equation}

On the other hand, we also note that in some previous studies (see, Mason   Kerswell Reference Mason and Kerswell2002; Barker Reference Barker2016b ; Pizzi et al. Reference Pizzi, Mamatsashvili, Barker, Giesecke and Stefani2022) the `mantle frame’ has been used: this consists in applying the following transformation to the solution (1.1):

(2.11) \begin{equation} x_1=x \cos t+y\sin t,\quad x_2=-x\sin t+y\cos t,\quad x_3=z. \end{equation}

2.2.2. Basic magnetic field and buoyancy scalar

Given the solutions (1.1) and (1.3), a constant basic magnetic field ${\boldsymbol{B}}_0,$ which is a trivial solution of (2.7a ), must also satisfy (2.7b ), so that, ${\boldsymbol{B}}_0{\boldsymbol{\cdot }} \boldsymbol{\nabla }{\boldsymbol{U}}=0,$ and hence, ${\boldsymbol{B}}_0$ aligns with the basic absolute vorticity:

(2.12a) \begin{align} {\boldsymbol{B}}_0&=B_0\!\left (2\varepsilon \hat {\boldsymbol{x}}+\hat {\boldsymbol{z}}\right )\quad \text{for the KBF}, \end{align}
(2.12b) \begin{align} {\boldsymbol{B}}_0&=B_0\hat {\boldsymbol{z}}\quad \text{for the MBF}. \end{align}

Here, $B_0$ is a positive constant.

Given the above solutions for ( ${\boldsymbol{U}}, {\boldsymbol{B}}_0),$ in a stratified MHD case under the Boussinesq approximation the basic buoyancy scalar must be a solution of the (2.7c ).

For the KBF (given by (1.1)), a vertical constant basic density gradient, $\boldsymbol{\nabla }\varrho =({\rm d}\varrho /{\rm d}z)\hat {\boldsymbol{z}},$ so that

(2.13) \begin{equation} \varTheta =-\frac {g}{\rho _0}\left (\frac {{\rm d}\varrho }{{\rm d}z}\right ) z=N^2z,\quad N^2=-\frac {g}{\rho _0}\frac {{\rm d}\varrho }{{\rm d}z},\quad {\boldsymbol{n}}=\frac {\boldsymbol{\nabla }\varTheta }{\Vert {\boldsymbol{\nabla }\varTheta }\Vert }=\hat {\boldsymbol{z}} \end{equation}

is an admissible solution. Here $N$ being the Brunt–Väisälä frequency ( $N^2\gt 0$ for a stable stratification).

For the MBF (given by (1.3)), we consider that, a priori, the basic buoyancy scalar varies linearly with the space coordinates (see, e.g. Craik Reference Craik1989), $\varTheta =(\partial _x\varTheta )x+ (\partial _y\varTheta )y+N^2 {\boldsymbol{z}},$ so that, without precession, it reduces to $\varTheta =N^2 {\boldsymbol{z}}.$ By substituting the above form into (2.7c ), we obtain, $ (\boldsymbol{\nabla }{\boldsymbol{U}} )^T{\boldsymbol{\cdot }} (\boldsymbol{\nabla }\varTheta )={{\boldsymbol{0}}},$

(2.14) \begin{equation} \left (\!\!\begin{array}{ccc} 0&1&0\\ -1&0&-2\varepsilon \\ 0&0&0\end{array}\!\!\right )\!{\boldsymbol{\cdot }}\left ( \!\!\begin{array}{c} \partial _x\varTheta \\ \partial _y\varTheta \\ N^2\end{array}\!\!\right )=\left (\!\!\begin{array}{c} 0\\0\\0 \end{array}\!\!\right ) \end{equation}

or equivalently,

(2.15) \begin{equation} \partial _y\varTheta =0\quad \text{and}\quad \partial _x\varTheta =-2\varepsilon N^2. \end{equation}

Accordingly, the basic buoyancy scalar takes the form

(2.16) \begin{equation} \varTheta ({\boldsymbol{x}})={N^2}\! \left (-2\varepsilon x+z\right )\! ,\quad {\boldsymbol{n}}=\frac {\boldsymbol{\nabla }\varTheta }{\Vert {\boldsymbol{\nabla }\varTheta }\Vert }=\frac {1}{\sqrt {1+4\varepsilon ^2}}\left (-2\varepsilon \hat {\boldsymbol{x}}+\hat {\boldsymbol{z}}\right )\! . \end{equation}

The presence of the additional horizontal component of the basic buoyancy scalar, which vanishes in the case without precession $(\varepsilon =0),$ is thereby due to the gyroscopic torque as for the additional mean shear. In a cylindrical coordinate system $(r,\varphi ,z)$ the solution (2.16) can be rewritten as

(2.17) \begin{equation} \varTheta ({\boldsymbol{x}})={N^2}\! \left (-2\varepsilon \left (\cos \varphi \right )\! r+z\right )\! ,\quad {\boldsymbol{n}}=\frac {1}{\sqrt {1+4\varepsilon ^2}}\left [-2\varepsilon \left ((\cos \varphi )\hat {\boldsymbol{r}}-(\sin \varphi )\hat {\boldsymbol{\varphi }}\right )+\hat {\boldsymbol{z}}\right ]\!. \end{equation}

Thus, we recover the solution proposed by Benkacem et al. (Reference Benkacem, Salhi, Khlifi, Nasraoui and Cambon2022) who extended Mahalov’s theoretical study (Mahalov Reference Mahalov1993) by taking into account the effect of stable stratification on precessing columns: they proposed a first-order correction to the profile of the basic buoyancy scalar because the Coriolis force alters the base flow (see their Appendix A).

For the sake of clarity, we repeat here the base flows considered in the present study,

(2.18) \begin{equation} \boldsymbol{\nabla }{\boldsymbol{U}}=\varOmega _0\!\left (\!\!\begin{array}{ccc} 0&-1&0\\ 1&0&-2\varepsilon \\ 0&0&0 \end{array}\!\!\right )\! , \ {\boldsymbol{\varOmega }}_{\!p}= \varOmega _0\!\left (\!\!\begin{array}{c} \varepsilon \\ 0\\ 0 \end{array}\!\!\right )\! , \ {\boldsymbol{B}}_0= {B_0}\left (\!\!\begin{array}{c} 2\varepsilon \\ 0\\ 1 \end{array}\!\!\right )\! ,\ \boldsymbol{\nabla }\varTheta = N^2\!\left (\!\!\begin{array}{c} 0\\ 0\\ 1 \end{array}\!\!\right )\! ,\end{equation}

for the KBF and

(2.19) \begin{equation} \boldsymbol{\nabla }{\boldsymbol{U}}=\varOmega _0\!\left (\!\!\begin{array}{ccc} 0&-1&0\\ 1&0&0\\ 0&-2\varepsilon &0 \end{array}\!\!\right )\! , \ {\boldsymbol{\varOmega }}_{\!p}= \varOmega _0\!\left (\!\!\begin{array}{c} \varepsilon \\ 0\\ 0 \end{array}\!\!\right )\! , \ {\boldsymbol{B}}_0= B_0\!\left (\!\!\begin{array}{c} 0\\ 0\\ 1 \end{array}\!\!\right )\! ,\ \boldsymbol{\nabla }\varTheta = {N^2}\left (\!\!\begin{array}{c} -2\varepsilon \\ 0\\ 1 \end{array}\!\!\right )\! ,\end{equation}

for the MBF.

3. Perturbed system

In the following, we consider the case of a non-diffusive fluid. Diffusivity effects with the assumption that the diffusion coefficients are equal $(\nu = \eta = \kappa )$ or, equivalently, the magnetic and thermal Prandtl numbers are equal to one are briefly discussed at the end of § 5.

3.1. Linearized system in physical space

We substitute the solutions (2.6) into the system (2.1) and linearize. The resulting perturbed equations are

(3.1a) \begin{align} \left (\partial _t +{{\boldsymbol{U}}}{\boldsymbol{\cdot }}{\boldsymbol{\nabla }}\right )\!{{\boldsymbol{u}}}&= -\boldsymbol{\nabla }{p}-{\boldsymbol{u}}{\boldsymbol{\cdot }}\boldsymbol{\nabla }{\boldsymbol{U}}-2\varepsilon \varOmega _0\hat {\boldsymbol{x}}\times {{\boldsymbol{u}}} +\left ({\boldsymbol{B}}_0{\boldsymbol{\cdot }} {\boldsymbol{\nabla }}\right )\!{{\boldsymbol{b}}} +{\vartheta }{\boldsymbol{n}}, \end{align}
(3.1b) \begin{align} \left (\partial _t +{{\boldsymbol{U}}}{\boldsymbol{\cdot }}{\boldsymbol{\nabla }}\right )\!{{\boldsymbol{b}}}&={\boldsymbol{b}}{\boldsymbol{\cdot }}\boldsymbol{\nabla }{\boldsymbol{U}}+ \left ({\boldsymbol{B}}_0{\boldsymbol{\cdot }}{\boldsymbol{\nabla }}\right )\!{{\boldsymbol{u}}}, \end{align}
(3.1c) \begin{align} \left (\partial _t +{{\boldsymbol{U}}}{\boldsymbol{\cdot }}{\boldsymbol{\nabla }}\right )\!{\vartheta }&=-{\boldsymbol{u}}{\boldsymbol{\cdot }}\boldsymbol{\nabla }\varTheta, \end{align}
(3.1d) \begin{align} \boldsymbol{\nabla }{\boldsymbol{\cdot }}\, {{{\boldsymbol{u}}}}&=0, \end{align}
(3.1e) \begin{align} \boldsymbol{\nabla }{\boldsymbol{\cdot }}\, {{{\boldsymbol{b}}}}&=0. \end{align}

For an inviscid and non-diffusive Boussinesq fluid, the potential induction magnetic scalar (MIPS) $\tilde \varPi _m=\varPi _m+\varpi _m+\varpi _m^{(nl)}$ is a Lagrangian invariant, as already indicated (see (2.5)), and its linear part takes the form

(3.2) \begin{equation} \varpi _m={\boldsymbol{b}}{\boldsymbol{\cdot }}\boldsymbol{\nabla }\varTheta +{\boldsymbol{B}}_0{\boldsymbol{\cdot }} \boldsymbol{\nabla }\vartheta \, \end{equation}

while $\varpi _m^{(nl)}={\boldsymbol{b}}{\boldsymbol{\cdot }} \boldsymbol{\nabla }\vartheta$ denotes its nonlinear part which is omitted in the present study.

3.2. Floquet system in wave space

The disturbances are expressed in terms of plane waves, for which the direction and the speed of propagation depend on time (see, Bayly Reference Bayly1986; Craik   Criminale Reference Craik and Criminale1986),

(3.3) \begin{equation} \big ({\boldsymbol{u}},p,{\boldsymbol{b}},N^{-1}\vartheta \big )({\boldsymbol{x}},t)=\big (\hat {\boldsymbol{u}},\hat p,\hat {\boldsymbol{b}},\hat \vartheta \big )({\boldsymbol{k}},t)\exp \!\left (\textrm{i}{\boldsymbol{k}}(t){\boldsymbol{\cdot }}{\boldsymbol{x}}\right ) \end{equation}

where $\textrm{i}^2=-1,$ ${\boldsymbol{k}}(t)$ is the wavevector and $\hat \vartheta ({\boldsymbol{k}})$ has a dimension of a velocity. Accordingly, the material derivative of the fluctuating velocity $\boldsymbol{u}$ can be rewritten as

(3.4) \begin{equation} \left (\partial _t +{{\boldsymbol{U}}}{\boldsymbol{\cdot }}{\boldsymbol{\nabla }}\right )\!{{\boldsymbol{u}}}({\boldsymbol{x}},t)=\left [\partial _t\hat {\boldsymbol{u}}+\textrm{i}\!\left (\left (d_t{\boldsymbol{k}}+\left (\boldsymbol{\nabla }{\boldsymbol{U}}\right )^T{\boldsymbol{\cdot }}\,{\boldsymbol{k}}\right )\!{\boldsymbol{\cdot }}{\boldsymbol{x}}\right )\!\hat {\boldsymbol{u}}\right ]\exp \!\left (\textrm{i}{\boldsymbol{k}}(t){\boldsymbol{\cdot }}{\boldsymbol{x}}\right )\! . \end{equation}

To remove the explicit dependence on $\boldsymbol{x}$ in the resulting equations for the Fourier amplitudes $\hat {\boldsymbol{u}},$ $\hat {\boldsymbol{b}}$ and $\hat \vartheta$ one has to ensure that ${\boldsymbol{k}}(t )$ varies in time according to the eikonal equation (see, Bayly Reference Bayly1986; Craik   Criminale Reference Craik and Criminale1986),

(3.5) \begin{equation} d_t{\boldsymbol{k}}=-\!\left (\boldsymbol{\nabla }{\boldsymbol{U}}\right )^T{\boldsymbol{\cdot }} {\boldsymbol{k}} \end{equation}

where $d_t ({\boldsymbol{\cdot }})\equiv d ({\boldsymbol{\cdot }})/{\rm d}t.$

Substituting the plane waves solution (3.3) into the system (3.1) and taking into account the eikonal (3.5), we obtain

(3.6a) \begin{align} d_t{\hat {\boldsymbol{u}}}&=-\textrm{i} \hat p {\boldsymbol{k}} -\hat {\boldsymbol{u}}{\boldsymbol{\cdot }} \boldsymbol{\nabla }{\boldsymbol{U}}-2\varepsilon \varOmega _0\hat {\boldsymbol{x}}\times {\hat {\boldsymbol{u}}} +\textrm{i} \!\left ({\boldsymbol{k}}{\boldsymbol{\cdot }}{\boldsymbol{B}}_0\right )\!{\hat {\boldsymbol{b}}} +N\hat \vartheta {\boldsymbol{n}}, \end{align}
(3.6b) \begin{align} d_t{\hat {\boldsymbol{b}}}&=\hat {\boldsymbol{b}}{\boldsymbol{\cdot }} \boldsymbol{\nabla }{\boldsymbol{U}}+ \textrm{i} \!\left ({\boldsymbol{k}}{\boldsymbol{\cdot }}{\boldsymbol{B}}_0\right )\!{\hat {\boldsymbol{u}}}, \end{align}
(3.6c) \begin{align} d_t{\hat \vartheta }&=-N^{-1}\hat {\boldsymbol{u}}{\boldsymbol{\cdot }}\boldsymbol{\nabla }\varTheta, \end{align}

together with $ {\boldsymbol{k}}{\boldsymbol{\cdot }} {{\hat {\boldsymbol{u}}}}=0$ and ${\boldsymbol{k}} {\boldsymbol{\cdot }} {{\hat {\boldsymbol{b}}}}=0.$ The use of the latter conditions allows one to eliminate the Fourier amplitude of fluctuating pressure,

(3.7) \begin{equation} \hat p=\textrm{i} k^{-2}\left (\left (\hat {\boldsymbol{u}}{\boldsymbol{\cdot }} \boldsymbol{\nabla }{\boldsymbol{U}}\right )\!{\boldsymbol{\cdot }}\,{\boldsymbol{k}}+\left (\left (\boldsymbol{\nabla }{\boldsymbol{U}}\right )^T{\boldsymbol{\cdot }}\,{\boldsymbol{k}}\right )\!{\boldsymbol{\cdot }}\,\hat {\boldsymbol{u}}+2\varepsilon \varOmega _0 \left (\hat {\boldsymbol{x}}\times \hat {\boldsymbol{u}}\right )\!{\boldsymbol{\cdot }}\,{\boldsymbol{k}}-N\left ({\boldsymbol{n}}{\boldsymbol{\cdot }}\,{\boldsymbol{k}}\right )\!\hat \vartheta \right ) \end{equation}

and to reduce the above seven-component Floquet system to a five-component version, where $k=\Vert {\boldsymbol{k}}\Vert .$ The spectral counterpart of the linear part of MIPS given by (3.2) reads

(3.8) \begin{equation} \hat \varpi _m=\hat {\boldsymbol{b}}{\boldsymbol{\cdot }}\boldsymbol{\nabla }\varTheta +\textrm{i} \!\left ({\boldsymbol{k}}{\boldsymbol{\cdot }}{\boldsymbol{B}}_0\right )\!\hat \vartheta =\text{const.} \end{equation}

It will be used to reduce the five-dimensional Floquet system to a inhomogeneous four-dimensional Floquet system.

3.3. The KBF case

The developments presented in this section as well as those presented in § 3.4 concern the KBF. Those which concern the MBF are reported in Appendix A for greater clarity for the reader.

3.3.1. Time-periodic wavevector

For the KBF (described by (2.18)), (3.5) reduces to

(3.9) \begin{equation} d_tk_1=-\varOmega _0k_2,\quad d_tk_2=\varOmega _0k_1,\quad d_tk_3=2\varepsilon \varOmega _0 k_2=-2\varepsilon d_tk_1, \end{equation}

so that

(3.10) \begin{equation} d_tk^2=d_t\big(k_1^2+k_2^2+k_3^2\big)= 4\varOmega _0\varepsilon k_2k_3, \end{equation}

where $k=\Vert {\boldsymbol{k}}\Vert .$ Equation (3.9) can be solved to give

(3.11) \begin{equation} k_1=k_{\!p}\cos (\tau -\phi ),\quad k_2=k_{\!p}\sin (\tau -\phi ),\quad k_3=k_0-2\varepsilon k_1, \end{equation}

where $\tau =\varOmega _0t$ is a dimensionless time, $\tan \phi =-k_{20}/k_{10}$ and

(3.12) \begin{equation} k_{\!p}=\sqrt {k_1^2+k_2^2}=\sqrt {k_{10}^2+k_{20}^2},\quad k_0=k_{30}+2\varepsilon k_{10}=k_3+2\varepsilon k_1 ,\end{equation}

in which ${\boldsymbol{K}}_0=(k_{10},k_{20},k_{30})^T$ is the initial wavevector. Both $k_{\!p}$ and $k_0$ are time-independent.

For purposes of studying stability, we may set $\phi =0.$ This is easily seen by making the substitution $\varOmega _0t'= \varOmega _0 t - \phi ,$ which eliminates $\phi$ from the equation.

3.3.2. One-dimensional two component and 2-D three component flow cases

For $k_{\!p}=0,$ so that $k_1=k_2=0$ : the wavevector aligns with the vertical unit vector $\hat {\boldsymbol{z}},$ ${\boldsymbol{k}}=k_{3}\hat {\boldsymbol{z}}=k_{30}\hat {\boldsymbol{z}}.$ In this case, the constraints ${\boldsymbol{k}}{\boldsymbol{\cdot }}\,\hat {\boldsymbol{u}}=0$ and ${\boldsymbol{k}}{\boldsymbol{\cdot }}\,\hat {\boldsymbol{b}}=0$ give $\hat u_3=0$ and $\hat b_3=0,$ respectively. Thus, this flow configuration corresponds to one-dimensional two component (1D2C) flows. System (3.6) reduces to a four-dimensional system for the velocity and magnetic field horizontal fluctuations which are not affected by the effect of stratification (at least for the linear regime, i.e. the nonlinear interactions are omitted),

(3.13) \begin{equation} d_t \left (\!\!\begin{array}{c} \hat u_1\\ \hat u_2\\ \hat b_1\\ \hat b_2 \end{array}\!\!\right )= \left (\!\! \begin{array}{cccc} 0&\varOmega _0&\textrm{i} k_3B_0&0\\ -\varOmega _0&0&0&\textrm{i} k_3B_0\\ \textrm{i} k_3B_0&0&0&-\varOmega _0\\ 0&\textrm{i} k_3B_0&\varOmega _0&0\\ \end{array}\!\!\right ) {\boldsymbol{\cdot }} \left (\!\!\begin{array}{c} \hat u_1\\ \hat u_2\\ \hat b_1\\ \hat b_2 \end{array}\!\!\right ) \end{equation}

with $\hat \vartheta =\text{const.},$ in agreement with (3.8). It is not difficult to show that there can be no instability associated with the solution of the above differential system.

For the case where $k_0=0,$ so that $k_3=-2\varepsilon k_1$ and ${\boldsymbol{k}}=k_1 (\hat {\boldsymbol{x}}-2\varepsilon \hat {\boldsymbol{z}} )+k_2\hat {\boldsymbol{y}},$ and the constraints ${\boldsymbol{k}}{\boldsymbol{\cdot }}\,\hat {\boldsymbol{u}}=0$ and ${\boldsymbol{k}}{\boldsymbol{\cdot }}\, \hat {\boldsymbol{b}}=0$ reduce to

(3.14) \begin{equation} k_1\big(\hat u_1+2\varepsilon \hat u_3\big)+k_2\hat u_2=0,\quad k_1\big (\hat b_1+2\varepsilon \hat b_3\big )+k_2\hat b_2=0. \end{equation}

Moreover, in the linear regime, there is no effect of the basic magnetic field on the velocity and density fluctuations. The magnetic field fluctuations behave like a passive field in a precessing stratified flow, since the system (3.1) reduces to

(3.15a) \begin{align} d_t{\hat {\boldsymbol{u}}}&=-\textrm{i} \hat p {\boldsymbol{k}} -\hat {\boldsymbol{u}}{\boldsymbol{\cdot }} \boldsymbol{\nabla }{\boldsymbol{U}}-2\varepsilon \varOmega _0\hat {\boldsymbol{x}}\times {\hat {\boldsymbol{u}}} +N\hat \vartheta {\boldsymbol{n}}, \end{align}
(3.15b) \begin{align} d_t{\hat {\boldsymbol{b}}}&=\hat {\boldsymbol{b}}{\boldsymbol{\cdot }}\boldsymbol{\nabla }{\boldsymbol{U}}, \end{align}
(3.15c) \begin{align} d_t{\hat \vartheta }&=-N^{-1}\hat {\boldsymbol{u}}{\boldsymbol{\cdot }}\boldsymbol{\nabla }\varTheta , \end{align}

or equivalently,

(3.16a) \begin{align} \hat \omega _{3}&=\text{const.},\quad \hat b_3=\text{const.},\quad k_1\hat b_2-k_2\hat b_1 +2\varepsilon k_2\hat b_3= \text{const.}, \end{align}
(3.16b) \begin{align} d_\tau \hat \psi &=-2\textrm{i}\varepsilon k_1\hat \omega _3+\left (N/\varOmega _0\right )k_{\!p}^2\hat \vartheta , \end{align}
(3.16c) \begin{align} d_\tau \vartheta &=-\!\left (N/\varOmega _0\right )\!\hat u_3, \end{align}

where $\hat \psi =k^2\hat u_3,$ $k^2=k_{\!p}^2+k_3^2=k_{\!p}^2+4\varepsilon ^2k_1^2= k_{\!p}^2 (1+4\varepsilon ^2\cos ^2\tau )$ and $\hat \omega _3\equiv \textrm{i}(k_1\hat u_2-k_2\hat u_1)$ denotes the Fourier amplitude of the vertical vorticity component ( ${\boldsymbol{\omega }} =\boldsymbol{\nabla }\times {\boldsymbol{u}})$ . Accordingly, we deduce the following inhomogeneous second-order differential equation (i.e. a Hill’s equation):

(3.17) \begin{equation} d_{\tau \tau }\hat \psi +{{\left (N/\varOmega _0\right )}^2}\big (1+2\varepsilon ^2\!\left (1+2\cos 2\tau \right )\big )\hat \psi =-2\varepsilon k_{\!p}\hat \omega _3\sin \tau . \end{equation}

The stability analysis of (3.17) indicates that there is no instability at order ${ O}(\varepsilon ).$ In counterpart, on times cales of order $\varepsilon ^{-2},$ the solutions are unstable at and near ${N/\varOmega _0}=1$ with a maximal growth rate, $\sigma _{m}={\varepsilon ^2}/{2}$ (more details are given in the Supplementary material).

3.4. Floquet system for the 3-D three component flow case

For perturbations with wavenumber $k_0\ne 0,$ we transform the resulting Floquet system in terms of the following variables to facilitate subsequent calculations (see, Salhi et al. Reference Salhi, Lehner and Cambon2010):

(3.18a) \begin{align} \varOmega _0 \hat c_1&=\left (k_1\hat u_2-k_2\hat u_1\right )+2\varepsilon k_2\hat u_3, \quad \varOmega _0 \hat c_2=-k_0\hat u_3, \end{align}
(3.18b) \begin{align} \varOmega _0\hat c_3&=\big (k_1\hat b_2-k_2\hat b_1\big )+2\varepsilon k_2\hat b_3, \quad \varOmega _0 \hat c_4=-k_0\hat b_3,\quad \varOmega _0 \hat c_5=-k_0\hat \vartheta . \end{align}

Given (3.7), which reduces to

(3.19) \begin{equation} \hat p=-\textrm{i} k^{-2}\big (2\varOmega _0\!\left (k_1\hat u_2-k_2\hat u_1\right )+6\varepsilon \varOmega _0k_2\hat u_3-2\varepsilon \varOmega _0 k_3\hat u_2+k_3N\hat \vartheta \big ), \end{equation}

we transform the system (3.6) according to these new variables,

(3.20a) \begin{align} d_\tau \hat c_1&=-4\varepsilon \left (\frac {k_2k_3}{k^2}+\varepsilon \frac {k_1k_2}{k^2}\right )\!\hat c_1-2\!\left (1+\varepsilon \frac {k_1}{k_0}+4\varepsilon ^3\frac {k_1k_2^2}{k_0k^2}\right )\!\hat c_2+\textrm{i} {\mathcal B}\mu \hat c_3-2\varepsilon {\mathcal N} \frac {k_2k_{\!p}^2}{k_0k^2}\hat c_5, \end{align}
(3.20b) \begin{align} d_\tau \hat c_2&=2\frac {k_0\!\left (k_0-\varepsilon k_1\right )}{k^2}\hat c_1+4\varepsilon ^2\frac {k_1k_2}{k^2}\hat c_2+ \textrm{i} {\mathcal B}\mu \hat c_4+{\mathcal N} \frac {k_{\!p}^2}{k^2}\hat c_5, \end{align}
(3.20c) \begin{align} d_\tau \hat c_3&=\textrm{i} {\mathcal B}\mu \hat c_1, \end{align}
(3.20d) \begin{align} d_\tau \hat c_4&=\textrm{i} {\mathcal B}\mu \hat c_2, \end{align}
(3.20e) \begin{align} d_\tau \hat c_5&=-{\mathcal N}\hat c_2 \end{align}

where $d_\tau ({\boldsymbol{\cdot }})=\varOmega _0^{-1}d_t({\boldsymbol{\cdot }})$ and

(3.21) \begin{equation} -\!1\leqslant \mu =\frac {k_0}{\sqrt {k_{\!p}^2+k_0^2}}\leqslant 1,\quad {\mathcal B}=\frac {B_0\sqrt {k_{\!p}^2+k_0^2}}{\varOmega _0},\quad {\mathcal N}=\frac {N}{\varOmega _0}, \end{equation}

which are the parameters of the Floquet system in addition to the pression parameter $\varepsilon .$ We note that, at small scales, i.e. $\sqrt {k_{\!p}^2+k_0^2}\gg 1,$ $\mathcal B$ is large even if $B_0/\varOmega _0\sim 1.$

On the other hand, combining (3.20d ) and (3.20e ), we deduce the following relation:

(3.22) \begin{equation} {\mathcal N}\hat c_4+\textrm{i} {\mathcal B}\mu \hat c_5= -\varOmega _0^{-1}k_0\!\left (N\hat b_3+\textrm{i} B_0k_0\hat \vartheta \right )=-\varOmega _0^{-1}k_0\hat \varpi = \text{const.,} \end{equation}

which represents the spectral counterpart of the linear part of MIPS (see (3.8)). Accordingly, system (3.20) can be further reduced to a fourth-order inhomogeneous Floquet system for $\hat {\boldsymbol{c}}= (\hat c_1,\hat c_2,\hat c_3,\hat c_4 )^T$ ,

(3.23) \begin{equation} d_\tau \hat {\boldsymbol{c}}=\boldsymbol{D}{\boldsymbol{\cdot }}\, \hat {\boldsymbol{c}} +\hat {\boldsymbol{\varphi }}, \end{equation}

where the only non-zero elements of $\boldsymbol{D}(\tau )$ are

(3.24a) \begin{align} D_{11}&=-4\varepsilon \left (\frac {k_2k_3}{k^2}+\varepsilon \frac {k_1k_2}{k^2}\right )\! , \end{align}
(3.24b) \begin{align} D_{12}&=-2\!\left (1+\varepsilon \frac {k_1}{k_0}+4\varepsilon ^3\frac {k_1k_2^2}{k_0k^2}\right )\! , \end{align}
(3.24c) \begin{align} D_{13}&=D_{31}=D_{42}=\textrm{i} {\mathcal B}\mu , \end{align}
(3.24d) \begin{align} D_{14}&=-2\textrm{i}\varepsilon \frac {{\mathcal N}}{{\mathcal B}\mu ^2}\frac {k_2k_{\!p}^2}{k^3}, \end{align}
(3.24e) \begin{align} D_{21}&=2\frac {k_0\!\left (k_0-\varepsilon k_1\right )}{k^2}, \end{align}
(3.24f) \begin{align} D_{22}&= 4\varepsilon ^2\frac {k_1k_2}{k^2}, \end{align}
(3.24g) \begin{align} D_{24}&=\textrm{i} {\mathcal B}\mu +\textrm{i} \frac {{\mathcal N}}{{\mathcal B}\mu }\frac {k_{\!p}^2}{k^2}. \end{align}

Therefore, the homogeneous system is

(3.25) \begin{equation} d_\tau \hat {\boldsymbol{c}}=\boldsymbol{D}{\boldsymbol{\cdot }}\,\hat {\boldsymbol{c}}, \end{equation}

while the non-zero components of the inhomogeneous term in (3.23) take the form

(3.26) \begin{equation} \hat \varphi _1=-2\textrm{i}\varepsilon \frac {N\varOmega _0}{B_0} \frac {k_2k_{\!p}^2}{k_0k^2}\hat \varpi _m,\quad \hat \varphi _2= \textrm{i}\frac {N\varOmega _0}{B_0} \frac {k_{\!p}^2}{k^2}\hat \varpi _m, \end{equation}

and it can be seen as a time-varying forcing excitation. The linear system (3.23) has the properties $\boldsymbol{D}(\tau +T) =\boldsymbol{D}(\tau )$ and $\hat {\boldsymbol{\varphi }}(\tau +T)=\hat {\boldsymbol{\varphi }}(\tau )$ where $T = 2\pi$ is the period common to both the matrix $\boldsymbol{D}(\tau )$ and the vector $\hat {\boldsymbol{\varphi }}(\tau ).$ Floquet theory does not address stability of the inhomogeneous system described by (3.20) where the ‘forcing excitation’ $\hat {\boldsymbol{\varphi }}(\tau )$ is present. However, the $T\hbox{-}$ periodic nature of $\hat {\boldsymbol{\varphi }}(\tau )$ allows an extension to the theory (see, Slane   Tragesser Reference Slane and Tragesser2011). By adding a forcing excitation, the general behaviour predicted by Floquet theory for the homogeneous system changed only for the case of semisimple Floquet multipliers that are identically equal to one, i.e. one is the only multiplier on the unit circle and is semisimple. In this case, the presence of the forcing excitation caused the solution to diverge (see, Slane   Tragesser Reference Slane and Tragesser2011). Incidentally, for several values of the parameters $({\mathcal B},{\mathcal N},\varepsilon , \mu ),$ we have numerically determined the stability and instability regions by integrating the $5\times 5$ Floquet system (without involving MIPS, i.e. system (3.20)) as well as the resulting $4\times 4$ Floquet system with $\hat \varpi _m=0$ (i.e. system (3.25)). We obtained the same results for both systems (up to numerical errors). We note that the determinant of the fundamental matrix solution is unity at $\tau =2\pi$ for both Floquet systems (see (4.3)). This indicates that the neutral case of system (3.25) could be characterized by the fact that unity is not the only multiplier on the unit circle.

3.5. Resonant cases of MIG waves

In this section, we start from the case without precession, so that both the background shear and the Coriolis force vanish. In that case, there are MIG waves. The resonant cases of MIG waves have been characterized by SC23 in their study of the MGE instability . Because some resonant cases of MIG waves become destabilizing in stratified magnetized precessing flows $(\varepsilon \ne 0),$ we briefly recall here the analysis made by SC23.

We denote by $\boldsymbol{D}_0$ the matrix $\boldsymbol{D}$ for $\varepsilon =0$ (i.e. circular streamlines)

(3.27) \begin{equation} \boldsymbol{D}_0=\varOmega _0\!\left (\!\!\begin{array}{cccc} 0&-2\varOmega _0&\textrm{i} \omega _a&0\\ (2\varOmega _0)^{-1}\omega _r^2&0&0&\textrm{i} \big (\omega _a^2+\omega _g^2\big )\omega _a^{-1}\\ \textrm{i} \omega _a&0&0&0\\ 0&\textrm{i} \omega _a&0&0 \end{array}\!\!\right ) \end{equation}

where $\omega _r,$ $\omega_a$ and $\omega _g$ are the frequencies of inertial, Alfvén and gravity waves, respectively,

(3.28a) \begin{align} \omega _r&=\frac {2\vert {\boldsymbol{\varOmega }}_0{\boldsymbol{\cdot }} {\boldsymbol{K}}_0\vert }{K_0}=2\varOmega _0\vert \mu \vert , \end{align}
(3.28b) \begin{align} \omega _a&={\boldsymbol{B}}_0{\boldsymbol{\cdot }} {\boldsymbol{K}}_0=B_0k_0=B_0 K_0\vert \mu \vert , \end{align}
(3.28c) \begin{align} \omega _g&=N\frac {\Vert \hat {\boldsymbol{z}}\times {\boldsymbol{K}}_0\Vert }{K_0}=N\sqrt {1-\mu ^2}. \end{align}

The eigenvalues $\sigma _j$ $( j = 1, 2, 3, 4)$ of the constant matrix $\boldsymbol{D}_0$ such that

(3.29) \begin{equation} \left \vert \boldsymbol{D}_0-\sigma \boldsymbol{I}_4\right \vert =\left \vert \boldsymbol{D}_0-\textrm{i} \omega \boldsymbol{I}_4\right \vert =0, \end{equation}

where $\boldsymbol{I}_4$ is the four-dimensional unit matrix, take the form

(3.30a) \begin{align} -\varOmega _0^2\sigma _{1,2}^2&=\omega ^2_{1,2}=\frac {1}{2}\left [2\omega _a^2+\omega _r^2+\omega _g^2+\sqrt {\left (\omega _r^2+\omega _g^2\right )^2+4\omega _r^2\omega _a^2}\ \right ]\!, \end{align}
(3.30b) \begin{align} -\varOmega _0^2\sigma _{3,4}^2&=\omega _{3,4}^2=\frac {1}{2}\left [2\omega _a^2+\omega _r^2+\omega _g^2-\sqrt {\left (\omega _r^2+\omega _g^2\right )^2+4\omega _r^2\omega _a^2}\ \right ]\!, \end{align}

or equivalently,

(3.31a) \begin{align} \omega _1=-\omega _{2}&=\frac {\varOmega _0}{\sqrt {2}}\sqrt {\left (4+2{\mathcal B}^2-{\mathcal N}^2\right )\mu ^2+{\mathcal N}^2+\sqrt {\left [\left (4-{\mathcal N}^2\right )\mu ^2+{\mathcal N}^2\right ]^2+16{\mathcal B}^2\mu ^4}}, \end{align}
(3.31b) \begin{align} \omega _3=-\omega _{4}&=\frac {\varOmega _0}{\sqrt {2}}\sqrt {\left (4+2{\mathcal B}^2-{\mathcal N}^2\right )\mu ^2+{\mathcal N}^2-\sqrt {\left [\left (4-{\mathcal N}^2\right )\mu ^2+{\mathcal N}^2\right ]^2+16{\mathcal B}^2\mu ^4}}. \end{align}

Because $\omega _1\geqslant \omega _3,$ fast waves have frequency $\omega _1,$ while slow waves have frequency $\omega _3.$ In the non-magnetized case, fast modes reduce to the purely hydrodynamic modes (i.e. inertia-gravity waves). In counterpart, slow waves are purely magnetic modes because $\omega _3$ reduces to zero for ${\mathcal B}=0$ (see LZ04). The expression of the eigenvectors associated with the eigenvalues $\sigma _1,\sigma _2,\sigma _3$ and $\sigma _4$ is reported in Appendix B.

The resonant cases of MIG waves are those parameter values $(\mu , {\mathcal N},{\mathcal B},)$ such that

(3.32) \begin{equation} \omega _i-\omega _j= n \varOmega _0,\quad (i\ne j). \end{equation}

As will be shown later, some resonant cases of order $n=1$ are destabilizing, whereas in the case of an elliptical flow the only resonant cases that can lead to instability are those for which the integer $n$ is even (see SC23). Indeed, in the precessing flow case, the components of the wavevector vary in time as $\sim \exp (\pm \textrm{i} \varOmega _{0} t)$ , reflecting the $\varOmega _{0}$ -frequency of the strain field, while in the elliptical flow case they vary as $\sim \exp (\pm 2 \textrm{i} \varOmega _{0} t)$ , corresponding to a doubled frequency. This difference directly accounts for the fact that precessing flows can exhibit destabilizing resonances already for $n = 1$ , whereas elliptical flows only admit even- $n$ resonances.

Note that the condition (3.32) for resonance readily extends the one by Bayly (Reference Bayly1986) for basic elliptical flow instability ( ${\mathcal B}=0$ and ${\mathcal N}=0),$ so that, $4 \varOmega \mu = n \varOmega _0,$ that immediately yielded $\mu = n/4=0.5$ for the elliptical flow ( $n=2)$ and $n/4=0.25$ for the precessing flow ( $n=1)$ (see, Kerswell Reference Kerswell1993; Salhi   Cambon Reference Salhi and Cambon2009) giving the origin of time-dependent instability tongues at vanishing $\varepsilon$ . We also note that the analysis of triad instability principle by Waleffe (Reference Waleffe1992) shows that the elliptical instability corresponds to a forward-interaction: the two modes with eigenfrequency $\omega _1=\omega _r$ and $\omega _2=-\omega _r$ have opposite polarities and are coupled with the mean flow, which is associated with a zero frequency, the unstable modes are thereby two resonant inertial waves associated with the uniform background rotation.

4. Asymptotic analysis of the Floquet system

Following the study of the MGE instability by SC23, we extend the asymptotic method of LZ04 to determine, at leading order in $\varepsilon ,$ the maximal growth rate of the solution of the four-component Floquet system (3.25). We note that the analytical developments presented in the present study are somewhat more complex than those performed by SC23 for the elliptic instability. This is mainly due to the fact that the unperturbed basic state for the precessional flow (see (2.18) and (2.19)) is somewhat more complex than that for the elliptic flow.

We indicate that in the remainder of this article the frequencies $\omega _1,\omega _2,\omega _3$ and $\omega _4$ are normalized by $\varOmega _0,$ except where they appear in the figures and in (5.3).

4.1. Expansion in Taylor series of the Floquet multiplier matrix

We denote by ${\boldsymbol \varPhi }(\tau ,\varepsilon ,\mu ,{\mathcal B}, {\mathcal N})$ the fundamental matrix solution,

(4.1) \begin{equation} d_\tau {\boldsymbol \varPhi }=\boldsymbol{D}{\boldsymbol{\cdot }} {\boldsymbol \varPhi },\quad {\boldsymbol \varPhi }(\tau =0)=\boldsymbol{I}_4 \end{equation}

and by $\boldsymbol{M}= {\boldsymbol \varPhi }(2\pi ,\varepsilon ,\mu ,{\mathcal B},{\mathcal N})$ the Floquet multiplier matrix. According to Floquet–Lyapunov theorem, $\boldsymbol \varPhi$ is expressible in the form (see, Kuchment Reference Kuchment1993),

(4.2) \begin{equation} \boldsymbol \varPhi (\tau )=\boldsymbol{F}\!(\tau )\exp \!\left (\boldsymbol{K}\!\tau \right )\! , \end{equation}

where $\boldsymbol{F}\!(\tau )$ is a non-singular continuous $2\pi \hbox{-}$ periodic $4\times 4$ matrix-function (whose derivative is an integrable piecewise-continuous function) and $\boldsymbol{K}$ is a constant matrix. The determinant of $\boldsymbol \varPhi$ is unity at $\tau =2\pi ,$ $\vert {\boldsymbol \varPhi (2\pi )}\vert =1$ since

(4.3) \begin{equation} \text{trace}\ \boldsymbol{D}=\sum _{j=1}^4 D_{\textit{jj}}=-4\varepsilon \frac {k_2k_3}{k^2}=-\frac {1}{k^2}{d_\tau k^2}. \end{equation}

It follows that whenever $\lambda$ is an eigenvalue of the monodromy matrix, $\boldsymbol{M}=\boldsymbol \varPhi (2\pi ),$ so also are its inverse $\lambda ^{-1}$ and its complex conjugate $\overline {\lambda }$ (see LZ04). Consequently, in the stable case, eigenvalues of $\boldsymbol{M}$ lie on the unit circle. If any eigenvalue $\lambda$ of $\boldsymbol{M}$ has modulus exceeding one, this implies that there is indeed an exponentially growing solution. The growth rates are then given by

(4.4) \begin{equation} \sigma =\frac {1}{2\pi }\log \left (\lambda \right )\! . \end{equation}

It follows that whenever $\lambda$ is an eigenvalue of $\boldsymbol{M},$ so also are its inverse $\lambda ^{-1}$ and its complex conjugate $\overline {\lambda }.$ We denote by

(4.5) \begin{equation} {p}(\lambda ,\varepsilon )=\vert \boldsymbol{M}-\lambda \boldsymbol{I}_{\!4}\vert \end{equation}

the characteristic polynomial of the Floquet multiplier matrix $\boldsymbol{M}$ and by $\varLambda _1,\varLambda _2,\varLambda _3$ and $\varLambda _4$ its roots. A necessary condition for stability is that each root lie on the unit circle (see LZ04).

We expand the Floquet multiplier matrix $\boldsymbol{M}\!(\varepsilon ,\mu , {\mathcal B}, {\mathcal N})$ in Taylor series in the neighbourhood of $(\varepsilon ,\mu )=(0,\mu _0),$ holding $({\mathcal B},{\mathcal N})$ constant,

(4.6) \begin{equation} \boldsymbol{M}=\boldsymbol{M}_0(\mu _0,0)+\varepsilon \boldsymbol{M}_\varepsilon (\mu _0,0)+ (\mu -\mu _0)\boldsymbol{M}_{\!\mu} (\mu _0,0)+{\cdots} \end{equation}

where $\boldsymbol{M}_\varepsilon = (\partial \boldsymbol{M}/\partial \varepsilon ),$ $\boldsymbol{M}_{\!\mu} = (\partial \boldsymbol{M}/\partial \mu )$ and the dots indicate higher-order terms in $\varepsilon$ and $\mu -\mu _0.$ Generally, at sufficiently small $\varepsilon ,$ the region in the $(\varepsilon , \mu )\hbox{-}$ plane where instability occurs is typically a wedge with apex at a point $(\varepsilon , \mu )) = (0, \mu _0)$ and boundaries

(4.7) \begin{equation} \mu =\mu _0+{\eta _\pm }\varepsilon, \end{equation}

where the slopes $\eta _+$ and $\eta _{-}$ are to be found. The instability (if it exists) has a bandwidth $ (\eta _+-\eta _- )\varepsilon$ that is, for given $\varepsilon ,$ $\mathcal B$ and ${\mathcal N},$ the length of the $\mu \hbox{-}$ interval for which the wavenumbers are unstable. Therefore, (4.6) can be rewritten as

(4.8a) \begin{align} \boldsymbol{M}&=\boldsymbol{M}_0+\varepsilon \boldsymbol{M}_1+{ O}(\varepsilon ^2), \end{align}
(4.8b) \begin{align} \boldsymbol{M}_1&=\boldsymbol{M}_\varepsilon +{\eta } \boldsymbol{M}\mu . \end{align}

Accordingly, we no longer need the designation $\mu _0$ and, hereinafter, use the symbol $\mu$ in its place.

To determine matrices $\boldsymbol{M}_0,$ $\boldsymbol{M}_\varepsilon$ and $\boldsymbol{M}_{\!\mu} ,$ we expand, for a given $\tau \in [0, 2\pi ],$ $\boldsymbol \varPhi$ and $\boldsymbol{D},$

(4.9a) \begin{align} {\boldsymbol \varPhi }(\tau ,\varepsilon )&={\boldsymbol \varPhi }_0(\tau ,\mu ,0)+\varepsilon {\boldsymbol \varPhi }_1(\tau ,\mu ,0)+{ O}(\varepsilon ^2), \end{align}
(4.9b) \begin{align} \boldsymbol{D}(\tau ,\varepsilon )&=\boldsymbol{D}_0+\varepsilon \boldsymbol{D}_\varepsilon (\tau ,0)+{ O}(\varepsilon ^2), \end{align}

where ${\boldsymbol \varPhi }_0(\tau =0)=\boldsymbol{I}_4$ and ${\boldsymbol \varPhi }_1(\tau =0)=\boldsymbol{0}.$ We notice that we find the same expression for the matrix $\boldsymbol{D}_\varepsilon (\tau , \varepsilon )$ for both KBF and MBF cases (see Appendix B2). Thus, to leading order in $\varepsilon ,$ the stability problem for these two basic flows is the same.

Substituting (4.9) into (4.1), we obtain

(4.10) \begin{align} d_\tau {\boldsymbol \varPhi }_0=\boldsymbol{D}_0{\boldsymbol{\cdot }} {\boldsymbol \varPhi }_0,\quad d_\tau {\boldsymbol \varPhi }_1=\boldsymbol{D}_0{\boldsymbol{\cdot }} {\boldsymbol \varPhi }_1+\boldsymbol{D}_\varepsilon {\boldsymbol{\cdot }} {\boldsymbol \varPhi }_0, \end{align}

with solution ${\boldsymbol \varPhi }_0(\tau )=\text{e}^{\tau \boldsymbol{D}_0}$ and

(4.11) \begin{equation} {\boldsymbol \varPhi }_1(\tau )={\boldsymbol \varPhi }_0(\tau ){\boldsymbol{\cdot }}\left (\int _0^\tau {\boldsymbol \varPhi }_0^{-1}(s){\boldsymbol{\cdot }}\boldsymbol{D}_\varepsilon (s) {\boldsymbol{\cdot }} {\boldsymbol \varPhi }_0(s){\rm d}s\right )\! . \end{equation}

Because the characteristic polynomial $\text{p}(\lambda ,\varepsilon )$ is the same in any coordinate system and the four eigenvalues of the matrix $\boldsymbol{D}_0,$ given by (3.30) are distinct as long as ${\mathcal B}\ne 0$ and $0\lt \mu ^2\lt 1,$ we transform the solution in the basis diagonalizing $\boldsymbol{D}_0,$ so that

(4.12a) \begin{align} \tilde {\boldsymbol{M}}_0&={\boldsymbol{T}}^{-1}{\boldsymbol{\cdot }} {\boldsymbol{M}}_0{\boldsymbol{\cdot }} \boldsymbol{T}=\textbf{diag}\big ({\text{e}}^{2\pi \sigma _1}, {\text{e}}^{-2\pi \sigma _1}, {\text{e}}^{2\pi \sigma _3}, \text{e}^{-2\pi \sigma _3}\big )= \textbf{diag}\left (\lambda _1, \lambda _2, \lambda _3, \lambda _4\right )\! , \end{align}
(4.12b) \begin{align} \tilde {\boldsymbol{M}}_\varepsilon &={\boldsymbol{T}}^{-1}{\boldsymbol{\cdot }} {\boldsymbol{M}}_\varepsilon {\boldsymbol{\cdot }} \boldsymbol{T}=\tilde {\boldsymbol{M}}_0{\boldsymbol{\cdot }} \tilde {\boldsymbol{J}}, \end{align}
(4.12c) \begin{align} \tilde {J}_{\textit{ij}}&=\left ({\boldsymbol{T}}^{-1}\right )_{im}{ T}_{lj}\int _0^{2\pi } {\text{e}}^{(\sigma _j-\sigma _i)\tau }\left ({\boldsymbol{D}}_\varepsilon \right )_{ml}(\tau ){\rm d}\tau , \end{align}

where the columns of the matrix $\boldsymbol{T}$ are the eigenvectors associated with the eigenvalues $\sigma _j$ ( $j=1,2,3,4)$ given by (B1) and $\boldsymbol{T}^{-1}$ is its inverse given by (B2).

To complete the construction of the matrix $\tilde {\boldsymbol{M}}_1,$ which appears in (4.8), we need the derivative of $\tilde {\boldsymbol{M}}_0(\mu )=\tilde {\boldsymbol{M}}(\mu ,0)$ with respect to $\mu ,$

(4.13) \begin{equation} {\eta }\tilde {\boldsymbol{M}}_\mu ={\eta }\frac {\partial \tilde {\boldsymbol{M}}_0}{\partial \mu }=\frac {2\textrm{i}\pi {\eta }}{\varOmega }\ \textbf{diag}\left ( \frac {\partial \omega _1}{\partial \mu }\lambda _1, \frac {\partial \omega _2}{\partial \mu }\lambda _2, \frac {\partial \omega _3}{\partial \mu }\lambda _3, \frac {\partial \omega _4}{\partial \mu }\lambda _4 \right )\! , \end{equation}

where $\omega _1,$ $\omega _2,$ $\omega _3$ and $\omega _4$ are given by (3.31).

When there is a subharmonic instability associated with the resonant case $\omega _\ell -\omega _j=\pm 1$ ( $j\ne \ell ),$ its maximum growth rate $\sigma _{m}$ is given by

(4.14) \begin{equation} \sigma _{m}=\frac {\varepsilon }{2\pi }\Re \sqrt {\tilde J_{\ell j}\tilde J_{j\ell }}\quad (\text{with}\ \ell \ne j), \end{equation}

recalling that the frequencies $\omega _j$ ( $j=1,2,3,4)$ are normalized by $\varOmega _0.$ In the following sections, we present the analytical results of the maximum growth rates associated with the three unstable cases, namely the fast–fast, slow–slow and fast–slow destabilizing resonances (calculation details are given in the Supplementary material).

4.2. Case of unstable resonance between two fast modes or between two slow modes

4.2.1. Domains of the resonant cases in the $({\mathcal B},{\mathcal N})\hbox{-}$ plane

The resonance condition between two fast modes is obtained by substituting the expression of $\omega _1$ given by (3.31) into the following identity:

(4.15) \begin{equation} \omega _1-\omega _2=2\omega _1=1\ \implies \ 16\omega _1^4=1, \end{equation}

or equivalently,

(4.16) \begin{equation} 4{\mathcal B}^2\big ({\mathcal B}^2-{\mathcal N}^2\big)\mu ^4+\big (4{\mathcal B}^2{\mathcal N}^2-2{\mathcal B}^2-4+{\mathcal N}^2\big )\mu ^2+\left (\frac {1}{4}-{\mathcal N}^2\right )=0, \end{equation}

with $0\lt \mu ^2\lt 1.$ Because the replacement $\mu \rightarrow -\mu$ in (4.16) result in the same condition, we may therefore assume without loss of generality that $\mu \gt 0.$ We note that the resonant condition between two slow modes, $\omega _3-\omega _4=1,$ is also described by (4.16), and the real roots of (4.16), such that $0\lt \mu ^2\lt 1,$ if they exist, are ordered. The smaller of the two roots characterizes the resonant cases between two fast modes where

(4.17) \begin{equation} \left ({\mathcal B},{\mathcal N}\right )\in \left [0,+\infty \right )\times \left [0,{1}/{2}\right ]\cup {\mathcal D}_f,\quad \lim _{\substack {{\mathcal B} \to +\infty \\ 0\leqslant {\mathcal N}\leqslant \frac {1}{2}}} {\mathcal B}^2\mu ^2=\frac {1}{4}-{\mathcal N}^2,\quad \lim _{\substack {{\mathcal N} \to +\infty \\ \frac {1}{2}\lt {\mathcal B}\leqslant \frac {\sqrt {5}}{2}}} \mu ^2=\frac {1}{4{\mathcal B}^2}. \end{equation}

The other root of (4.16), such that $0\lt \mu ^2\lt 1,$ characterizes the resonant cases between two slow modes where

(4.18) \begin{equation} \left ({\mathcal B},{\mathcal N}\right )\in \big [{\sqrt {5}}/{2},+\infty \big )\times \left [0,+\infty \right )\cup {\mathcal D}_f,\quad \lim _{\substack {{\mathcal B} \to +\infty \\ 0\leqslant {\mathcal N}\lt +\infty }} {\mathcal B}^2\mu ^2=\frac {1}{4},\quad \lim _{\substack {{\mathcal N} \to +\infty \\ \frac {1}{2}\lt {\mathcal B}\lt +\infty }} \mu ^2=1. \end{equation}

Here, the domain ${\mathcal D}_f$ is defined as follows (see also figure 2):

(4.19) \begin{equation} \forall \left ({\mathcal B},{\mathcal N}\right )\in {\mathcal D}_f \Leftrightarrow \ \text{for given}\ {\mathcal N}\in \big ({{\sqrt {6}}/{2}},+\infty \big )\quad \implies \quad \frac {1}{2}\lt f({\mathcal N})\leqslant {\mathcal B}\lt \frac {\sqrt {5}}{2}, \end{equation}

where $f : (\sqrt {6}/2, +\infty )\rightarrow (1/2,\sqrt {5}/2 )$ is a continuous decreasing function

(4.20) \begin{align} \forall {\mathcal N}\!\in \big ({\sqrt {6}}/{2},+\infty \big ), f({\mathcal N})=\frac {1}{2{\mathcal N}^2} \sqrt {\left ({\mathcal N}^2+2\right )^2-6+2\sqrt {(4{\mathcal N}^2-1)({\mathcal N}^4-1)}}\!\underset {{\mathcal N}\rightarrow +\infty }{\longrightarrow } \!\frac {1}{2}. \end{align}

Figure 2. Resonant cases of MIG waves. (a) Resonant cases (of order $n=1$ ) occur for $({\mathcal B},{\mathcal N})$ belonging to the coloured domains. (b) Resonance between two fast modes ( $\omega _1-\omega _2=\varOmega _0)$ where (I) $\equiv [0,+\infty )\times [0,1]$ and (II) $\equiv {\mathcal D}_f$ (see (4.19)). (c) Resonance between two slow modes ( $\omega _3-\omega _4=\varOmega _0)$ where (III) $\equiv [\sqrt {5}/2,+\infty )\times [0,+\infty ).$ (d) Resonance between a fast mode and a slow mode ( $\omega _1-\omega _4=\varOmega _0$ or $\omega _1-\omega _3=\varOmega _0)$ where (IV) $\equiv [0,+\infty )\times [0,1]$ and (V) $\equiv {\mathcal D}_m$ (see (4.30)).

For the pure hydrodynamic case ( ${\mathcal B}=0,{\mathcal N}=0),$ one has $\mu = {1}/{4}.$ For the resonant cases between to fast modes and for given ${{\mathcal N}\in [0,{1}/{2}]},$ the parameter $\mu$ decreases from $\mu =(\sqrt {1-4{\mathcal N}^2})/(2\sqrt {4-{\mathcal N}^2})$ as $\mathcal B$ increases, approaching zero when ${\mathcal B}\rightarrow +\infty .$ Likewise, for given ${1}/{2}\lt {\mathcal B}={\mathcal B}_f\leqslant {\sqrt {5}}/{2}$ such that $ ({\mathcal N},{\mathcal B} )\in {\mathcal D}_f,$ $\mu$ increases from $\mu (f^{-1}({\mathcal B}_f),{\mathcal B}_f )$ approaching $1$ when ${\mathcal N}\rightarrow +\infty$ (see also figure 3 a).

Figure 3. Resonant cases of MIG waves. (a) Variation of the parameter $\mu =k_0/K_0$ versus ${\mathcal N}=N/\varOmega _0$ for ${\mathcal B}= B_0K_0/\varOmega _0=0.75,\ 1.0,\ \sqrt {5}/2$ such that $({\mathcal B},{\mathcal N})\in {\mathcal D}_f$ (see (4.19)) associated with the resonant cases between two fast modes ( $\omega _1-\omega _2=\varOmega _0)$ or between two slow modes ( $\omega _3-\omega _4=\varOmega _0).$ (b) Variation of the parameter $\mu =k_0/K_0$ versus $\mathcal B$ for ${\mathcal N}=1.2,\ 2.0,\ 3.0$ such that $({\mathcal B},{\mathcal N})\in {\mathcal D}_m$ (see (4.30)) associated with the resonant cases between a fast mode and a slow mode ( $\omega _1-\omega _4=\varOmega _0$ or $\omega _1-\omega _3=\varOmega _0).$ Here, $K_0=\sqrt {k_0^2+k_{\!p}^2}.$

For the resonant cases between two slow modes, which exist only in the magnetized case, and for given ${\mathcal B}\in [{\sqrt {5}}/{2},+\infty ),$ $\mu$ decreases from $\mu =1/(2(\sqrt {1+{\mathcal B}^2}-1))$ as $\mathcal N$ increases approaching $1/(2{\mathcal B})$ when ${\mathcal N}\rightarrow +\infty .$ Likewise, for given ${1}/{2}\lt {\mathcal B}\leqslant {\sqrt {5}}/{2},$ so that $ ({\mathcal N},{\mathcal B} )\in {\mathcal D}_f,$ $\mu$ decreases from $\mu (f^{-1}({\mathcal B}),{\mathcal B} )$ approaching $1/(2{\mathcal B})$ when ${\mathcal N}\rightarrow +\infty$ (see also figure 3 a).

4.2.2. Maximal growth rate

If a subharmonic instability resulting from the resonance between two fast (respectively, slow) modes exists, its maximal growth rate, denoted by $\sigma _{\textit{mf}}$ ( $\sigma _{\textit{ms}}),$ is given by

(4.21) \begin{equation} \sigma _{\textit{mf}}=\frac {\varepsilon }{2\pi }\Re \sqrt {\tilde J_{12}\tilde J_{21}}, \quad \sigma _{\textit{ms}}=\frac {\varepsilon }{2\pi }\Re \sqrt {\tilde J_{34}\tilde J_{43}}. \end{equation}

After some analytical developments given in the Supplementary material, we find

(4.22) \begin{align}& \frac {\sigma _{\textit{mf}}}{\varepsilon },\ \frac {\sigma _{\textit{ms}}}{\varepsilon }=\left \vert \frac {\left (4{\mathcal B}^2\mu ^2+1\right )\mu \sqrt {1-\mu ^2}} {8\left [2\!\left (4+2{\mathcal B}^2-{\mathcal N}^2\right )\mu ^2+2{\mathcal N}^2-1\right ]}\right \vert \nonumber \\&\quad \times \big \vert \big (\big (4{\mathcal B}^2\mu ^2-1\big )\big ({\mathcal B}^2\big (1-4{\mathcal N}^2-4\big ({\mathcal B}^2-{\mathcal N}^2\big )\mu ^2\big )+4\big )+2\big (4{\mathcal B}^2\mu ^2-3\big )\big )\big \vert , \end{align}

in which $0\lt \mu ^2\lt 1$ is the smaller of the two roots of (4.16) when it is a resonance between two fast modes or the other root when it is a resonance between two slow modes.

Some results reported in previous studies of stratified or magnetized precessing flows can be recovered from (4.22),

(4.23a) \begin{align} \text{for}&\ {\mathcal B}={\mathcal N}=0,\quad \mu =\frac {1}{4},\quad \frac {\sigma _{\textit{mf}}}{\varepsilon }=\frac {5\sqrt {15}}{32}, \end{align}
(4.23b) \begin{align} \text{for}&\ {\mathcal B}=0,\quad \mu =\frac {1}{2}\sqrt {\frac {1-4{\mathcal N}^2}{4-{\mathcal N}^2}},\quad \frac {\sigma _{\textit{mf}}}{\varepsilon }=\frac {5\sqrt {15}}{8}\frac {\sqrt {1-4{\mathcal N}^2}}{\left (4-{\mathcal N}^2\right )}, \end{align}
(4.23c) \begin{align} \text{for}&\ {\mathcal N}=0,\ \mu =\frac {1}{2\big (\sqrt {1+{\mathcal B}^2}+1\big ) },\ \frac {\sigma _{\textit{mf}}}{\varepsilon }= \frac {1}{8}\frac {\big (2\sqrt {{\mathcal B}^2+1}+3\big )\sqrt {4 \big (1+\sqrt {1+{\mathcal B}^2}\big )^2\!-1}}{\big (1+\sqrt {{\mathcal B}^2+1}\big )^2}. \end{align}

Equation (4.23b ) indicates that the maximal growth rate is zero at ${\mathcal N}={1}/{2}.$ Therefore, in the non-magnetized stratified precessing case ( ${\mathcal B}=0),$ the subharmonic instability is completely suppressed by stratification when $\mathcal N$ reaches $ {1}/{2}$ (see, Benkacem et al. Reference Benkacem, Salhi, Khlifi, Nasraoui and Cambon2022). For the non-stratified magnetized precessing case ( ${\mathcal N}=0),$ (4.23c ) indicates that $\sigma _{\textit{mf}}/\varepsilon$ decreases as $\mathcal B$ increases so as

(4.24) \begin{equation} \lim _{{\mathcal N}=0,\ {\mathcal B}\rightarrow +\infty } \frac {\sigma _{\textit{mf}}}{\varepsilon }=\frac {1}{2}. \end{equation}

Recall that, at small scales, i.e. $\sqrt {k_{\!p}^2+k_0^2}\gg 1,$ $\mathcal B$ is large even if $B_0/\varOmega _0\sim 1.$

In the stratified magnetized precessing case, the analysis of (4.22) is more subtle as shown in the following.

For $0\lt {\mathcal N}\lt 0.5,$ we verify using the second identity in (4.17) that $\sigma _{\textit{mf}}$ approaches zero for large ${\mathcal B}.$ Given (4.24), this clearly proves that the ${\mathcal N}\rightarrow 0$ limit is, in fact, singular (discontinuous).

For ${\mathcal N}=0.5,$ (4.16) gives $\mu =0$ (i.e. the smaller of the two roots) and hence, $\sigma _{\textit{mf}}=0,$ signifying that there is no parametric instability associated with the resonant case of order $n=1$ between two fast modes. It is worth mentioning that the case $\mu =0,$ so that $k_0=0$ corresponds to a 2-D three component (2D3C) flow (see the end of § 3.3.2).

For $ {1}/{2}\lt {\mathcal B}\leqslant {\sqrt {5}}/{2},$ i.e. $({\mathcal B},{\mathcal N})\in {\mathcal D}_f,$ we verify using the third identity in (4.17) that $\sigma _{\textit{mf}}$ also approaches zero for large ${\mathcal N}.$

Figure 4(a) shows $\sigma _{\textit{mf}}/\varepsilon$ versus $\mathcal B$ for ${\mathcal N}=0,\ 0.2,\ 0.3,\ 0.35$ and ${\mathcal N}=0.45,$ so that $({\mathcal B},{\mathcal N})\in [0,+\infty )\times [0,1/2],$ while figure 4(b) illustrates the variation of $\sigma _{\textit{mf}}/\varepsilon$ versus $\mathcal B$ for ${\mathcal N}=2,$ so that $({\mathcal B},{\mathcal N})\in {\mathcal D}_f.$ The numerical results for $\varepsilon =0.05,$ which will be discussed in § 5, are also reported .

Figure 4. Maximal growth rate of destabilizing resonant cases given by the asymptotic formulae. (a) Resonant cases between two fast modes for some values of $({\mathcal B},{\mathcal N})\in [0,+\infty )\times [0,1/2].$ (b) Resonant cases between two fast modes or between two slow modes for some values of $({\mathcal B},{\mathcal N})\in {\mathcal D}_f$ (see (4.19)). (c) Resonant cases between two slow modes for some values of $({\mathcal B},{\mathcal N})\in [\sqrt {5}/2,+\infty )\times [0,+\infty ).$ (d) Resonant cases between a fast mode and a slow mode for some values of $({\mathcal B},{\mathcal N})\in [0,+\infty )\times [0,1].$ The numerical results obtained for $\varepsilon =0.05$ (symbol) are also reported.

For the non-stratified magnetized case ( ${\mathcal N}=0),$ the expression of the maximal growth rate $\sigma _{\textit{ms}}$ of the destabilizing resonance between two slow mode deduced from (4.22) is of the form

(4.25) \begin{equation} \frac {\sigma _{\textit{ms}}}{\varepsilon }= \frac {\big(2+2{\mathcal B}^2-3\sqrt {{\mathcal B}^2+1}\big)\sqrt {4 \big(\sqrt {1+{\mathcal B}^2}-1\big)^2-1}}{8\sqrt {1+{\mathcal B}^2}\big(\sqrt {{\mathcal B}^2+1}-1\big)^2}\underset {{\mathcal B}\rightarrow +\infty }{\longrightarrow }\frac {1}{2} \end{equation}

where ${\sqrt {5}}/{2}\lt {\mathcal B}\lt +\infty ,$ in agreement with the previous results by Salhi et al. (Reference Salhi, Lehner and Cambon2010). In this case, ${\sigma _{\textit{ms}}}/{\varepsilon }$ increases from $0$ at ${\mathcal B}={\sqrt {5}}/{2}$ to ${1}/{2}$ as ${\mathcal B}\rightarrow +\infty .$

For the stratified magnetized case such that $ ({\mathcal B},{\mathcal N} ) \in [ {\sqrt {5}}/{2},+\infty )\times [0,+\infty )\cup {\mathcal D}_f,$ we use the second and third identities in (4.18) to deduce from (4.22) that

(4.26) \begin{equation} \lim _{{\mathcal N}\gt 0,\ {\mathcal B}\rightarrow +\infty } \frac {\sigma _{\textit{ms}}}{\varepsilon }=0, \quad \lim _{{\mathcal B}\gt \frac {1}{2},\ {\mathcal N}\rightarrow +\infty } \frac {\sigma _{\textit{ms}}}{\varepsilon }=0. \end{equation}

Figure 4(c) shows $\sigma _{\textit{ms}}/\varepsilon$ versus $\mathcal B$ for ${\mathcal N}=0,\ 0.5,\ 1$ and ${\mathcal N}=2,$ so that $({\mathcal B},{\mathcal N})\in [0,+\infty )\times [\sqrt {5}/2,+\infty ),$ while figure 4(b) illustrates the variation of $\sigma _{\textit{ms}}/\varepsilon$ versus $\mathcal B$ for ${\mathcal N}=2,$ so that ${({\mathcal B},{\mathcal N})\in {\mathcal D}_f}.$ These figures show the expected agreement between the numerical results and the asymptotic formulae.

Accordingly, we may conclude that, unlike the non-stratified magnetized case for which $\lim _{{\mathcal B}\rightarrow +\infty }\sigma _{\textit{ms}}/\varepsilon ={1}/{2},$ the destabilizing resonance between two slow modes is weakened in the presence of stable stratification in the sense that its growth rate is reduced and approaches zero for strong stratification. Given (4.25), this clearly proves that the ${\mathcal N}\rightarrow 0$ limit is, in fact, singular (discontinuous).

4.3. Destabilizing resonance between a slow mode and a fast mode

In this section, we will show that the resonant case associated with $\omega _1-\omega _4=1$ is destabilizing. In contrast, the resonant case associated with $\omega _1-\omega _3=1$ is stable.

By substituting the expression of $\omega _1^2$ and $\omega _3^2$ given by (3.31) into the following identity:

(4.27) \begin{equation} \omega _1-\omega _4=\omega _1+\omega _3=1\implies \omega _1^4+\omega _3^4-2\omega _1^2\omega _3^2-2\big(\omega _1^2+\omega _3^2\big)+1=0, \end{equation}

we obtain

(4.28) \begin{equation} \big (\big (4-{\mathcal N}^2\big )^2+16{\mathcal B}^2\big )\mu ^4-2\!\big (2{\mathcal B}^2+\big ({\mathcal N}^2-4\big )\big ({\mathcal N}^2-1\big )\big )\mu ^2+\big ({\mathcal N}^2-1\big )^2=0. \end{equation}

We note that the resonant condition $\omega _1-\omega _3=1,$ is also described by (4.16), and the real roots of (4.16), such that $0\lt \mu ^2\lt 1,$ if they exist, are ordered. The smaller of the two roots characterizes the resonant cases characterized by $\omega _1-\omega _4=1$ where

(4.29) \begin{equation} \left ({\mathcal B},{\mathcal N}\right )\in (0,+\infty )\times \left [0,1\right ]\cup {\mathcal D}_m,\quad \lim _{\substack {{\mathcal B} \to +\infty {\mathcal N}\in [0,+\infty [}} {\mathcal B}^2\mu ^2=\frac {\left (1-{\mathcal N}^2\right )^2}{4}. \end{equation}

Here, the domain ${\mathcal D}_m$ is defined as follows:

(4.30) \begin{equation} \forall \left ({\mathcal B},{\mathcal N}\right )\in {\mathcal D}_m \Leftrightarrow \ \text{for given}\ {\mathcal N}\in \left ]1,+\infty \right [\ \implies \ 0\lt \sqrt {{3{\mathcal N}^2\!\left ({\mathcal N}^2-1\right )}}\leqslant {\mathcal B}\ . \end{equation}

Accordingly, for given ${\mathcal N}\in [0,1[,$ $\mu ({\mathcal B},{\mathcal N})$ decreases from $\mu (0,{\mathcal N})=\sqrt {(1-{\mathcal N}^2)/(4-{\mathcal N}^2)}$ as $\mathcal B$ increases approaching zero when ${\mathcal B}\rightarrow +\infty .$ For ${\mathcal N}=1,$ one has $\mu =0,$ and hence there is no instability for this resonant case (see the end of § 3.3.2). For given ${\mathcal N}\gt 1,$ such that $ ({\mathcal B},{\mathcal N} )\in {\mathcal D}_m,$ $\mu ({\mathcal B},{\mathcal N} )$ decreases from $\mu ({\mathcal B}_m,{\mathcal N} )=\sqrt { ({\mathcal N}^2-1 )/ (7{\mathcal N}^2-4 )}$ , approaching zero when ${\mathcal B}\rightarrow +\infty ,$ where ${\mathcal B}_m^2={{3{\mathcal N}^2 ({\mathcal N}^2-1 )}}$ (see also figure 2 c).

The maximal growth rate of the destabilizing resonant case $\omega _1-\omega _4=1,$ is denoted by $\sigma _{\textit{mm}}$ ( $m$ for mixed) is then described by (4.14) which, in this case, reduces to

(4.31) \begin{equation} \sigma _{\textit{mm}}=\frac {\varepsilon }{2\pi }\left (\Re \alpha \right )_{\textit{max}}=\frac {\varepsilon }{2\pi }\Re \sqrt {\tilde J_{14}\tilde J_{41}} \end{equation}

where

(4.32) \begin{align} \frac {{\tilde J}_{14}}{2\textrm{i} \pi }&= \frac {\mu \sqrt {1-\mu ^2}}{\left (\omega _3^2-\omega _1^2\right )}\left [ \frac {\left (\mu ^2{\mathcal B}^2-\omega _3^2\right )\!\omega _1\omega _3}{\mu ^2{\mathcal B}^2}+\frac {\left (\mu ^2{\mathcal B}^2-\omega _3^2\right )^2}{4\mu ^4{\mathcal B}^2}\frac {\omega _1}{\omega _3}\big (\omega _3-\big (1-\mu ^2\big ){\mathcal N}^2\big )\right . \nonumber \\&\quad \left . +\big (4\mu ^2-1\big )\omega _3- {\mathcal N}^2\big (1-\mu ^2\big)\frac {\left (\mu ^2{\mathcal B}^2-\omega _3^2\right )}{\omega _3} \right ]\!, \end{align}
(4.33) \begin{align} \frac {{\tilde J}_{41}}{2\textrm{i} \pi }&= \frac {\mu \sqrt {1-\mu ^2}}{\left (\omega _3^2-\omega _1^2\right )}\left [ \frac {\left (\mu ^2{\mathcal B}^2-\omega _1^2\right )\!\omega _1\omega _3}{\mu ^2{\mathcal B}^2}+\frac {\left (\mu ^2{\mathcal B}^2-\omega _1^2\right )^2}{4\mu ^4{\mathcal B}^2}\frac {\omega _3}{\omega _1}\big (\omega _1-\big (1-\mu ^2\big ){\mathcal N}^2\big )\right . \nonumber \\&\quad \left . +\big (4\mu ^2-1\big )\omega _1- {\mathcal N}^2\big (1-\mu ^2\big )\frac {\left (\mu ^2{\mathcal B}^2-\omega _1^2\right )}{\omega _1} \right ]\!, \end{align}

where detail calculations are given in the Supplementary material.

In the non-magnetized case ( ${\mathcal N}=0)$ the resulting expression of $\sigma _{\textit{mm}}$ is given by

(4.34) \begin{equation} \frac {\sigma _{\textit{mm}}}{\varepsilon }=\frac {{\mathcal B}^2\sqrt {3+4{\mathcal B}^2}}{8\left (1+{\mathcal B}^2\right )^2}\underset {{\mathcal B}\rightarrow +\infty }{\longrightarrow } 0. \end{equation}

However, in the magnetized case, analytically obtaining a compact form for the product ${\tilde J}_{14}{\tilde J}_{41}$ that only involves the parameters $\mu ,$ $\mathcal N$ and $\mathcal B$ is quite complex. Nevertheless, some results characterizing the behaviour of $\sigma _{\textit{mm}}$ can be drawn by analysing the behaviour of ${\tilde J}_{14}{\tilde J}_{41}$ for large ${\mathcal B}.$ Recall that, at small scales, i.e. $\sqrt {k_{\!p}^2+k_0^2}\gg 1,$ $\mathcal B$ is large even if $B_0/\varOmega _0\sim 1.$

Using the second identity in (4.29) we deduce the following equivalent forms for the normalized frequencies $\omega _1$ and $\omega _3$ given by (3.31),

(4.35) \begin{equation} \omega _1\underset {{\mathcal B}\rightarrow +\infty }{\large {\sim }}\frac {1+{\mathcal N}^2}{2},\quad \omega _3\underset {{\mathcal B}\rightarrow +\infty }{\sim }\mu {\mathcal B}\underset {{\mathcal B}\rightarrow +\infty }{\sim }\frac {1-{\mathcal N}^2}{2}, \end{equation}

so that $\omega _1+\omega _3=\omega _1-\omega _4=1.$ The resulting equivalent form for the product ${\tilde J}_{14}{\tilde J}_{41}$ is given by

(4.36) \begin{equation} \frac {{\tilde J}_{14}{\tilde J}_{41}} {4\pi ^2} \underset {{\mathcal B}\rightarrow +\infty }{\sim } \frac {\omega _3^2}{4\mu ^2{\mathcal B}^2} \frac {\left (\mu ^2{\mathcal B}^2-\omega _1^2\right )^2}{\left (\omega _3^2-\omega _1^2\right )^2}\left (1-\frac {{\mathcal N}^2}{\omega _1}\right ) \underset {{\mathcal B}\rightarrow +\infty }{\sim }\frac {1}{4}\frac {\left (1-{\mathcal N}^2\right )}{\left (1+{\mathcal N}^2\right )}. \end{equation}

Thus, the maximal growth rate of the destabilizing resonance between a fast mode and a slow mode such that $\omega _1-\omega _4=1$ has the following behaviour:

(4.37) \begin{equation} \frac {\sigma _{\textit{mm}}}{\varepsilon }\underset {{\mathcal B}\rightarrow +\infty }{\sim } \frac {1}{2}\sqrt {\frac {1-{\mathcal N}^2} {1+{\mathcal N}^2}}. \end{equation}

Figure 4(d) shows the variation of $\sigma _{\textit{mm}}/\varepsilon$ versus $\mathcal N$ for ${\mathcal B}=10^2,\ 10^3,\ 10^4$ and ${\mathcal B}\rightarrow +\infty .$ The numerical results obtained for $\varepsilon =0.05$ and ${\mathcal B}=10^2$ are also displayed in figure 4(d). As it can be seen, there is a good agreement between the numerical results (see § 5) and the asymptotic formulae.

Because in the non-magnetized case ( ${\mathcal N}=0),$ $\sigma _{\textit{mm}}$ tends to zero as ${\mathcal B}\rightarrow +\infty$ (see 4.34), while in the magnetized case (4.37) gives

(4.38) \begin{equation} \lim _{\substack {{\mathcal B} \to +\infty {\mathcal N}\rightarrow 0^+}}\frac {\sigma _{\textit{mm}}}{\varepsilon }=\frac {1}{2}, \end{equation}

we can conclude that for the destabilizing resonance associated with $\omega _1-\omega _4=1,$ the limit ${\mathcal N}\rightarrow 0$ is also, in fact, singular (discontinuous). This signifies that, at large ${\mathcal B},$ the presence of stable stratification enhances the destabilizing resonance between a fast (hydrodynamic) mode and a slow (magnetic) mode, provided $0\lt {\mathcal N}\lt 1.$ This would be relevant for the study of the generation of mean electromotive force ( ${\mathcal E}=\langle {\boldsymbol{u}}\times {\boldsymbol{b}}\rangle ,$ (see, Moffatt Reference Moffatt1978; Moffatt   Dormy Reference Moffatt and Dormy2019) in strongly magnetized precessing flows under a weak (stable) axial stratification, especially since in the present analysis the electromotive force generation mechanism does not require diffusivity to operate (see, Mizerski, Bajer   Moffatt Reference Mizerski, Bajer and Moffatt2012) for the case of the magneto-elliptic instability of rotating systems).

5. Numerical results

In this section we present a selection of numerical results. These are obtained by integrating the system given by (3.25) to obtain the fundamental matrix solution $\boldsymbol{\varPhi }(\tau ,\mu ,\varepsilon ,{\mathcal B},{\mathcal N})$ (see (4.1)) and evaluating it at $\tau =2\pi$ to get the Floquet matrix $\boldsymbol{M}(\mu ,\varepsilon ,{\mathcal B},{\mathcal N}).$

5.1. Identification of the instability tongues

For fixed $({\mathcal B},{\mathcal N}),$ we use the resonance conditions (4.16), and (4.28) for the identification of instabilities of order $n=1$ (if they exist). Note that the (3.32) also allows us to identify the high-order destabilizing resonances ( $n\geqslant 2)$ which are excluded by the procedure leading to the asymptotic formulae. On the other hand, for the same value of the couple $({\mathcal B} , {\mathcal N} ),$ we consider several values of $\varepsilon$ uniformly distributed in the interval $ (0, 0.25 )$ and, for each of these values of $\varepsilon ,$ we integrate numerically the Floquet system (4.1) and determine the growth rate $\Re \sigma$ (corresponding to the maximum modulus eigenvalue) for $5000$ values of $\mu$ (evenly distributed) in $ (0, 1 ).$ Obviously, $\sigma (\mu )=0$ if there is no instability. Thus, in the plane $(\mu , \varepsilon +\Re \sigma ) ,$ the region of instability that emanates from the point of abscissa $\mu$ characterizes a subharmonic instability as illustrated by figure 5 obtained for ${\mathcal B}=0.75$ and ${\mathcal N}=0.25.$ For these values, the destabilizing resonances detected by the computations are that between two fast modes which correspond to the tongue emanating from the point ( $\mu = 0.196, \varepsilon +\Re \sigma =0),$ and that between a fast mode and a slow mode which corresponds to the tongue emanating from the point ( $\mu = 0.380, \varepsilon +\Re \sigma =0$ ). The third instability tongue detected by the calculations emanates from the point ( $\mu = 0.434,\varepsilon +\Re \sigma =0$ ) and characterizes the destabilizing resonance of order $n=2$ between two fast modes. We note that this instability is more developed in the MBF case than in the KBF case, recalling that the matrix $\boldsymbol{D}$ in (3.23) does not have the same expression for these two base flows (see (3.24) and (A13)), except at order ${ O}(\varepsilon )$ (see (B4)).

Figure 5. Magneto-gravity-precessional instabilities. The figure shows, for fixed $\varepsilon ,$ $\varepsilon +\Re \sigma$ versus $\mu$ for ${\mathcal B}=0.75$ and ${\mathcal N}=0.25$ and $100$ values of $\varepsilon$ evenly distributed in the interval $[0,0.25].$ (a) The case of KBF; (b) the case of MBF. The instability region emanating from the point ( $\mu = 0.196, \varepsilon +\Re \sigma =0)$ is associated with the destabilizing resonance (of order $n=1)$ between two fast modes, while that emanating from the point ( $\mu = 0.380, \varepsilon +\Re \sigma =0$ ) is associated with the destabilizing resonance (of order $n=1)$ between a fast mode and a slow mode. The third region, which emanates from the point ( $\mu = 0.434,\varepsilon +\Re \sigma =0$ ), characterizes the destabilizing resonance of order $n=2$ between two fast modes.

For both background flows (KBF and MBF), the agreement between the asymptotic formulae and the numerical results is quite good for sufficiently small $\varepsilon (\leqslant 0.05),$ as illustrated in figures 4, 6 and 7. On the other hand, when $\varepsilon$ is not small enough, say $0.05\lt \varepsilon \lt 0.25$ , the subharmonic instabilities are strongest only in the KBF case, so the agreement between the asymptotic formulae and the numerical results is also quite good (see figures 7 a and 8 a).

Figure 6. Magneto-gravity-precessional instabilities. Maximal growth rate of dominant instability normalized by $\sigma _0=(5\sqrt {15}/32)\varepsilon$ plotted as a function of $0\leqslant {\mathcal B}=K_0B_0/\varOmega _0\leqslant 8$ and $0\leqslant {\mathcal N}=N/\varOmega _0\leqslant 2$ : (a) numerical results for $\varepsilon =0.05$ ; (b) asymptotic analysis results.

Figure 7. Magneto-gravity-precessional instabilities. Maximal growth rate of dominant instability normalized by $\sigma _0=(5\sqrt {15}/32)\varepsilon$ plotted as a function of $0\leqslant {\mathcal B}=K_0B_0/\varOmega _0\leqslant 8$ and $0\leqslant {\mathcal N}=N/\varOmega _0\leqslant 2$ for $\varepsilon =0.2.$ Here (a) KBF; (b) MBF.

Figure 8. Magneto-gravity-precessional instabilities. Maximal growth rate of dominant instability normalized by $\sigma _0=(5\sqrt {15}/32)\varepsilon$ plotted as a function of ${\mathcal N}=N/\varOmega _0$ for ${\mathcal B}=k_0B_0/\varOmega _0=5$ and $\varepsilon =0.05,\ 0.10,\ 0.20.$ Here (a) KBF; (b) MBF.

In figures 6 and 7, we show the continuous variation of the maximal growth rate $\sigma _m$ (maximum $\sigma$ over $0\leqslant \mu \leqslant 1)$ of the dominant instability, normalized by $\sigma _0=(5\sqrt {15}/32)\varepsilon ,$ plotted as a function of $0\leqslant {\mathcal B}\leqslant 8$ and $0\leqslant {\mathcal N}\leqslant 2.$ Figure 6(a) displays the numerical results obtained for $\varepsilon =0.05$ in the KBF case, while figure 6(b) shows the analytical results $\sigma _m/\sigma _0$ where

(5.1) \begin{equation} \sigma _m=\max (\sigma _{\textit{mf}},\sigma _{\textit{ms}},\sigma _{\textit{mm}}), \end{equation}

recalling that $\sigma _{\textit{mf}},$ $\sigma _{\textit{ms}}$ and $\sigma _{\textit{mm}}$ are given by (4.22) and (4.31), respectively. The numerical results for $\varepsilon =0.2$ are shown in figure 7 (figure 7 a, KBF case; figure 7 b, MBF case). Figure 8 shows $\sigma _m/\varepsilon$ as a function of $\mathcal N$ for ${\mathcal B}=5,$ $\varepsilon =0.05,\ 0.1$ and $\varepsilon =0.2.$ The analytical results are also reported in figure 8.

In the case of MBF and when $\varepsilon$ is not sufficiently small, it is rather the destabilizing resonance of order $n=2$ between two slow modes which is the strongest provided that ${\mathcal N}\gt 1$ and ${\mathcal B}\gt 2,$ as illustrated by figures 7(b) and 8(b). Recall that high-order destabilizing resonances ( $n\geqslant 2)$ are excluded by the procedure leading to the asymptotic formulae.

5.2. Diffusive effects in the case where $\nu =\eta =\kappa$

In this section, we consider the case where $\nu =\eta =\kappa ,$ so that the magnetic and thermal numbers are equal to one. In this case and given the definition (3.29), we show that the frequency of MIG waves is given by

(5.2) \begin{equation} \omega _j^{(v)}=\omega _j+\textrm{i} \nu k^2\quad (j=1,2,3,4), \end{equation}

where the superscript $(v)$ characterizes the viscous case and $\omega _j$ is the frequency of MIG waves for a non-diffusive fluid which is given by (3.30). Therefore, the condition of resonance

(5.3) \begin{equation} \omega _j^{(v)}-\omega _\ell ^{(v)}=\omega _j-\omega _\ell =n\varOmega _0,\quad (\ell \ne j) \end{equation}

is not modified when $\nu =\eta =\kappa .$ As for the destabilizing resonances (if they exist), they are affected by viscosity as shown in the following.

We formally rewrite the system (3.20) as

(5.4) \begin{equation} d_\tau \hat c_j={\mathcal L}_{\textit{jm}}\hat c_m \quad (j,m=1,{\cdots} ,5). \end{equation}

By taking into account diffusive terms such that $\nu =\eta =\kappa ,$ the above system is transformed as

(5.5) \begin{equation} d_\tau \hat c_j^{(v)}={\mathcal L}_{\textit{jm}}\hat c_m^{(v)} -\varOmega _0^{-1}\nu k^2\hat c_j^{(v)}\quad (j,m=1,{\cdots} ,5), \end{equation}

where its solution can be expressed in terms of that of (5.4) as

(5.6) \begin{equation} \hat c_j^{(v)}(\tau )=\hat c_j(\tau )\exp \!\left (-\frac {\nu }{\varOmega _0}\int _0^\tau k^2(s){\rm d}s\right )\! . \end{equation}

Thus, the linear part of the spectral counterpart of MIPS is zero because, for purposes of studying stability, we have set $\hat \varpi _m=0$ (see (3.22)),

(5.7) \begin{equation} k_0\varOmega _0^{-1}\hat \varpi ^{(v)}_m(\tau )=-\big ({\mathcal N}\hat c_4^{(v)}+\textrm{i} {\mathcal B}\hat \vartheta ^{(v)}\big )=k_0\varOmega _0^{-1}\hat \varpi _m \exp \!\left (-\frac {\nu }{\varOmega _0}\int _0^\tau k^2(s){\rm d}s\right )=0. \end{equation}

Accordingly, for a diffusive fluid such that $\nu =\eta =\kappa ,$ the growth rate of the destabilizing resonance (if it exists) takes the form (see (4.1) and (4.4))

(5.8) \begin{equation} \sigma ^{(v)}=\sigma -\frac {\nu }{2\pi \varOmega _0}\int _0^{2\pi } k^2(s){\rm d}s=\sigma -Re^{-1}L_0^2\!\left (k_0^2+k_{\!p}^2\right )\left (1+2\varepsilon ^2\!\left (1-\mu ^2\right )\right )\! , \end{equation}

where $Re=\varOmega _0L_0^2/\nu$ is the Reynolds number and $L_0$ a characteristic length scale. For the case of MBF (see (2.19) and Appendix A), we obtain

(5.9) \begin{equation} \sigma ^{(v)}=\sigma -Re^{-1}L_0^2\!\left (k_0^2+k_{\!p}^2\right )\left (1+4\varepsilon ^2\mu ^2\right )\! . \end{equation}

Thus, at order ${ O}(\varepsilon ),$ the effect of viscosity is the same for both basic flows. At ${L_0\sqrt {k_0^2+k_{\!p}^2}\sim 1},$ the dominant subharmonic instability survives the diffusive effects if $Re \gt Re_c=\sigma _m^{-1}.$

6. Concluding remarks

We have analysed here the joint influence of a stable stratification and an external uniform magnetic field on the stability of an unbounded flow with sheared circular streamlines of a perfectly conducting Boussinesq fluid. Two simple background flows permitting a study of parametric instabilities resulting from the excitation of MIG waves by the shear generic to precessing flows were adopted. The first basic state, which is described by (2.18), extends that of Kerswell (Reference Kerswell1993) and Mason   Kerswell (Reference Mason and Kerswell2002), and represents an idealization of a small-scale disturbance evolving at the centre of a precessing spheroid. The second basic state (see (2.19)) extends that of Mahalov (Reference Mahalov1993) that represents an infinite column in which a tilted (sheared) streamline solution can exist under precession. These simple models allows us to formulate the stability problem as a system of equations for disturbances in terms of Lagrangian Fourier modes which is universal for wavelengths of the perturbation sufficiently small with respect to the scale of variation of the mean velocity gradients.

Without precession ( $\varepsilon =0,$ i.e. the case with circular streamlines), both the basic magnetic field (of intensity $B_0)$ and stable stratification (of strength $N$ ), via the mean buoyancy gradient, are in the axial direction which aligns with the solid body rotation with rate $\varOmega _0.$ In this case and in the non-diffusive linear limit, there are MIG waves with fast and slow waves of dispersion relation given by (3.31). The domains in the plane $ ({\mathcal B},{\mathcal N} )$ which characterize the resonant cases of order $n=1$ of these waves have been identified. (see figure 2), recalling that ${\mathcal B}=B_0K_0/\varOmega _0$ and ${\mathcal N}=N/\varOmega _0$ where $K_0=\sqrt {k_{\!p}^2+k_0^2},$ and $k_0$ and $k_{\!p}$ are defined by (3.12) for KBF and by (A3) for MBF. Some of these resonant cases become destabilizing when excited by the background shear (of rate $2\varepsilon \varOmega _0)$ generic to precessing flows.

For infinite wavelength in the direction that aligns with the unperturbed magnetic field, so that $k_0=0,$ the precessing flow is 2D3C. In this case, the stability problem is governed by a Hill equation (i.e. (3.17) for KBF or (A9) for MBF). Stability analysis of the Hill equation indicates that there is a parametric instability of order ${O} (\varepsilon ^2 )$ at and near ${\mathcal N}=1$ with a maximum growth rate of approximately $\sigma _m=\varepsilon ^2/2.$

For $k_0\ne 0,$ thanks to MIPS, which is invariant for a non-diffusive system, it was possible to reduce the five-dimensional Floquet system to a four-dimensional inhomogeneous Floquet system with parameters $\varepsilon ,$ $\mu =k_0/K_0,$ $B_0K_0/\varOmega _0$ and $N/\varOmega _0.$ At order ${O}(\varepsilon ),$ the resulting Floquet system is the same for KBF and MBF. It has been analysed by extending the analytical technique developed by Lebovitz   Zweibel (Reference Lebovitz and Zweibel2004); which allowed us to determine the maximal growth rate of the destabilizing resonances (of order $n=1)$ between two fast modes ( $\sigma _{\textit{mf}})$ between two slow modes ( $\sigma _{\textit{ms}})$ or between a fast mode and a slow mode ( $\sigma _{\textit{mm}}).$

Without stratification ( ${\mathcal N}=0),$ the maximum growth rate of the destabilizing resonance between two fast modes, which takes the value $\sigma _{\textit{mf}}=(5\sqrt {15}/32)\varepsilon$ at ${\mathcal B}=0,$ decreases as $\mathcal B$ increases approaching $\sigma _{\textit{mf}}=\varepsilon /2$ as ${\mathcal B}\rightarrow +\infty .$ In the presence of stable stratification, this parametric instability is weakened when $({\mathcal B},{\mathcal N})$ belongs to the domain $ ([0,+\infty [\times [0,0.5[ )\cup {\mathcal D}_f$ where ${\mathcal D}_f$ is defined by (4.19), and it does not exist for $({\mathcal B},{\mathcal N})$ outside this domain. In addition, $\sigma _{\textit{mf}}$ approaches zero as ${\mathcal B}\rightarrow +\infty$ (see (4.22)), recalling that the large $\mathcal B$ corresponds either to a large unperturbed magnetic field, or to a large $K_0$ (i.e. small scales).

As for the instability associated with the resonant case (or order $n=1)$ between two slow modes, in the absence of stratification, it exists only for ${\mathcal B}\geqslant \sqrt {5}/2$ and its maximal growth rate increases as $\mathcal B$ increases approaching $\sigma _{\textit{ms}}=\varepsilon /2$ as ${\mathcal B}\rightarrow +\infty .$ This instability is also weakened by stable stratification, so that as $\sigma _{\textit{ms}}$ tends to zero as ${\mathcal B}\rightarrow +\infty$ . We note that the case of an elliptical flow, stable stratification enhanced the instability associated with the resonant case ( $n=2)$ between two slow modes, so as $\sigma _{\textit{ms}}$ approaches $\varepsilon /2$ as ${\mathcal B}\rightarrow +\infty .$

For the stratified magnetized precessing flow, it is rather the destabilizing resonance between a fast mode and a slow mode which is enhanced by stratification provided ${\mathcal N}\lt 1,$ $\sigma _{\textit{mm}}=(\varepsilon /2)\sqrt {(1-{\mathcal N}^2)/ (1+{\mathcal N}^2)}$ as ${\mathcal B}\rightarrow +\infty .$ In fact, in the case without stratification, it found that $\sigma _{\textit{mm}}$ approaches zero as ${\mathcal B}\rightarrow +\infty .$

Our analytical results clearly prove that, for destabilizing resonances (of order $n=1)$ between two fast modes, between two slow modes or between a fast mode and a slow mode, the limit ${\mathcal N}\rightarrow 0$ is in fact singular (discontinuous).

The asymptotic formulae, which characterize only the resonant cases of order $n=1$ and not those of order $n\geqslant 2,$ are in good agreement with the numerical results at least for sufficiently small $\varepsilon$ . For KBF, the destabilizing resonances of order $n=1$ are the strongest, so that there is a good agreement between the asymptotic formulae and the numerical results even if $\varepsilon (\leqslant 0.2)$ is not small. In return, for MBF, the destabilizing resonances of order $n\geqslant 2$ can be the strongest when $\varepsilon$ is not sufficiently small, notably for ${\mathcal N}\gt 1$ and $\mathcal B$ large (see figures 7 b and 8 b).

The simple case considering diffusive effects where $\nu =\eta =\kappa ,$ i.e. the magnetic and thermal Prandtl numbers are both equal to one, was briefly discussed. It follows that the instability survives viscous effects as long as the Reynolds number remains greater than the inverse of the maximum growth rate obtained in the inviscid case.

We note that the study of nonlinear effects on the magneto-gravitational-precessional instability in the case of (statistically) homogeneous turbulence constitutes the motivation for further studies for which the present analytical developments can serve as a parametric study.

Supplementary material

Supplementary material is available at https://doi.org/10.1017/jfm.2025.10881.

Declaration of interests

The authors report no conflict of interest.

Appendix A. The case of MBF

In Appendix A we consider the MBF, which is described by (2.19), and derive the resulting Floquet system governing the stability problem in this basic flow case.

A.1. Floquet system in wave space

A.1.1. Time-dependent wavevector

For the MBF (described by (2.19)), (3.5),

(A1) \begin{equation} d_tk_1=-\varOmega _0k_2,\quad d_tk_2=\varOmega _0\!\left (k_1+2\varepsilon k_3\right )\! ,\quad d_tk_3=0, \end{equation}

can be solved to give

(A2) \begin{equation} k_1=k_{p}\cos \tau -2\varepsilon k_3,\quad k_2=k_{\!p}\sin \tau ,\quad k_3=k_{30}, \end{equation}

where

(A3) \begin{equation} k_{p}= \sqrt { \left (k_{10}+2\varepsilon k_{30}\right )^2+k_{20}^2}=\sqrt { \left (k_{1}+2\varepsilon k_{3}\right )^2+k_{2}^2}, \quad k_0\equiv k_3. \end{equation}

Both $k_{\!p}$ and $k_0$ are time-independent, so that $\mu =k_0/\sqrt {k_0^2+k_{\!p}^2}$ (see (3.21). The wavevector trajectories are circles with sheared centres.

A.1.2. Floquet system

Substituting the form (2.19) and the plane wave solution (3.3) into the system (3.6), we obtain the following seven-dimensional system:

(A4a) \begin{align} d_t\hat u_1&= -\textrm{i} k_1\hat p+\varOmega _0 \hat u_2+ \textrm{i}\!\left (k_3B_0\right )\!\hat b_1-\frac {2\varepsilon N}{\sqrt {1+4\varepsilon ^2}}\hat \vartheta , \end{align}
(A4b) \begin{align} d_t\hat u_2&= -\textrm{i} k_2\hat p-\varOmega _0 \hat u_1+2\varepsilon \varOmega _0 \hat u_3+ \textrm{i}\!\left (k_3B_0\right )\!\hat b_2, \end{align}
(A4c) \begin{align} d_t\hat u_3&= -\textrm{i} k_3\hat p+ \textrm{i}\!\left (k_3B_0\right )\!\hat b_3+\frac {N}{\sqrt {1+4\varepsilon ^2}}\hat \vartheta , \end{align}
(A4d) \begin{align} d_t\hat b_1&=-\varOmega _0 \hat b_2+\textrm{i}\!\left (k_3B_0\right )\!\hat u_1, \end{align}
(A4e) \begin{align} d_t\hat b_2&=\varOmega _0 \hat b_1+\textrm{i}\!\left (k_3B_0\right )\!\hat u_2, \end{align}
(A4f) \begin{align} d_t\hat b_3&=-2\varepsilon \varOmega _0 \hat b_2+\textrm{i}\!\left (k_3B_0\right )\!\hat u_3, \end{align}
(A4g) \begin{align} d_t{\hat {\vartheta }}&=N \!\left (2\varepsilon \hat u_1-\hat u_3\right )\! , \end{align}

together with the conditions ${\boldsymbol{k}}{\boldsymbol{\cdot }}\hat {{\boldsymbol{u}}}=0$ and ${\boldsymbol{k}}{\boldsymbol{\cdot }}\hat {{\boldsymbol{b}}}=0.$ The use of these two conditions allows one to deduce the Fourier amplitude of the pressure disturbance,

(A5) \begin{equation} \hat p=-{\textrm{i}}k^{-2}\left [2\varOmega _0\!\left (k_1\hat u_2-k_2\hat u_1\right )+2\varepsilon \varOmega _0\!\left (k_2\hat u_3+k_3 \hat u_2\right )+\frac {(k_3-2\varepsilon k_1)N}{\sqrt {1+4\varepsilon ^2}}\hat \vartheta \right ]\!. \end{equation}

A.1.3. The case where $k_{\!p}=0$

For $k_{\!p}=0,$ so that $k_2=0$ and $k_1+2\varepsilon k_3=0,$ the flow is 1D2C since the constraints ${\boldsymbol{k}}{\boldsymbol{\cdot }}\,\hat {\boldsymbol{u}}=0$ and ${\boldsymbol{k}}{\boldsymbol{\cdot }}\,\hat {\boldsymbol{b}}=0$ give

(A6) \begin{equation} k_3\!\left (\hat u_3-2\varepsilon \hat u_1\right )=0,\quad k_3\big (\hat b_3-2\varepsilon \hat b_1\big )=0, \end{equation}

respectively. In this case, system (A4) reduces to a four-dimensional system for the velocity and magnetic field horizontal fluctuations which are not affected by the effect of stratification (at least for the linear regime),

(A7) \begin{equation} d_t\left (\!\!\begin{array}{c} \hat u_1\\ \hat u_2\\ \hat b_1\\ \hat b_2 \end{array}\!\!\right )=\left (\!\! \begin{array}{cccc} 0&\frac {(1-4\varepsilon ^2)} {(1+4\varepsilon ^2)}\varOmega _0&\textrm{i} (k_3B_0)&0\\ -(1-4\varepsilon ^2)\varOmega _0&0&0&\textrm{i} (k_3B_0)\\ \textrm{i} (k_3B_0)&0&0&-\varOmega _0\\ 0&\textrm{i} (k_3B_0)&\varOmega _0&0 \end{array}\!\!\right )\!{\boldsymbol{\cdot }} \left (\!\!\begin{array}{c} \hat u_1\\ \hat u_2\\ \hat b_1\\ \hat b_2\end{array}\!\!\right ) \end{equation}

with $\hat \vartheta = \text{const}.$ It is not difficult to show that there can be no instability associated with the solution of the above differential system.

A.1.4. The case where $k_3=0$

For $k_3=0,$ which corresponds to an infinite wavelength along the solid body rotation axis, the frequency of Alfvén waves vanishes ( $\omega _a=k_3B_0=0),$ and hence, in the linear regime, the flow dynamics is not affected by the Lorentz force. In this case, one easily verifies that

(A8) \begin{equation} \hat \omega _3-2\varepsilon \hat \omega _1= \textrm{i}\!\left (k_1\hat u_2-k_2\hat u_1\right )-2\textrm{i}\varepsilon k_2\hat u_3=\text{const.,} \end{equation}

is a constant of motion. Accordingly, the resulting stability problem for $k_3=0$ is governed by the following Hill equation:

(A9) \begin{equation} d_{\tau \tau }\hat u_3+\frac {{\mathcal N}^2}{\sqrt {1+4\varepsilon ^2}}\big (1+2\varepsilon ^2\big (1-\sin 2\tau \big )\big )\hat u_3=0. \end{equation}

The analysis of the solution of this last equation can be done in a similar way to that of the solution of the (3.17). Therefore, at order ${O}(\varepsilon ^2)$ , the solution is unstable at and near ${\mathcal N}=1$ with a maximal growth rate, $\sigma _{m}=({1}/{2})\varepsilon ^2.$

A.2. Reduced Floquet system

A.2.1. Change of variables

For $k_3\ne 0,$ we introduce the following variables to facilitate subsequent calculations (see, Salhi et al. Reference Salhi, Lehner and Cambon2010):

(A10a) \begin{align} \varOmega _0\hat c_1&=\left (k_1+2\varepsilon k_3\right )\!\hat u_2-k_2\hat u_1,\quad \varOmega _0\hat c_2=k_3\!\left (2\varepsilon \hat u_1 -\hat u_3\right )\! , \end{align}
(A10b) \begin{align} \varOmega _0\hat c_3&=\left (k_1+2\varepsilon k_3\right )\!\hat b_2-k_2\hat b_1,\quad \varOmega _0\hat c_4=k_3\big (2\varepsilon \hat b_1 -\hat b_3\big ). \end{align}

In terms of these variables, the linear part of the spectral counterpart of MIPS (see (3.2)) reads

(A11) \begin{equation} \frac {\hat {\varpi }_m}{\varOmega _0}= \frac {{\mathcal N}}{\sqrt {1+4\varepsilon ^2}}\big (-2\varepsilon \hat b_1+\hat b_3\big )+\textrm{i} {\mathcal B}\hat \vartheta =- \frac {{\mathcal N}}{k_3\sqrt {1+4\varepsilon ^2}}\hat c_4+\textrm{i} {\mathcal B}\hat \vartheta , \end{equation}

recalling that the parameters $\mathcal B$ and $\mathcal N$ are defined by (3.21). For purposes of studying stability, we may set $\hat {\varpi }_m=0,$ as indicated previously.

A.2.2. Reduced homogeneous Floquet system

Accordingly, the resulting four-dimensional homogeneous Floquet system for $\hat {\boldsymbol{c}}=(\hat c_1,\hat c_2,\hat c_3,\hat c_4)^T$ takes the form

(A12) \begin{equation} d_\tau \hat {\boldsymbol{c}}=\boldsymbol{D}{\boldsymbol{\cdot }}\, \hat {\boldsymbol{c}} \end{equation}

where the non-zero elements of the periodic matrix $\boldsymbol{D}$ are

(A13a) \begin{align} D_{11}&=-4\varepsilon \frac {k_2k_3}{k^2}-4\varepsilon ^2\frac {k_1k_2k_\perp ^2}{k^2k_{p}^2}-8\varepsilon ^3\frac {k_2k_3k_1^2}{k^2k_{p}^2}, \end{align}
(A13b) \begin{align} D_{12}&=-2-2\varepsilon \frac {k_1}{k_3}+8\varepsilon ^3\frac {k_1k_3k_2^2}{k^2k_{p}^2}, \end{align}
(A13c) \begin{align} D_{13}&=D_{31}=D_{42}={\textrm{i}}{\mathcal B}, \end{align}
(A13d) \begin{align} D_{14}&=-\frac {2\textrm{i}\varepsilon {\mathcal N}}{\left ({1+4\varepsilon ^2}\right ){\mathcal B}}\frac {k_2}{k_3}\frac {\left (k_\perp ^2+2\varepsilon k_1k_3\right )}{k^2}, \end{align}
(A13e) \begin{align} D_{21}&=2\frac {k_3^2}{k^2}-2\varepsilon \frac {k_1k_3}{k^2}+8\varepsilon ^3\frac {k_1k_3k_2^2}{k^2k_{p}^2}, \end{align}
(A13f) \begin{align} D_{22}&=4\varepsilon ^2\frac {k_1k_2k_\perp ^2}{k^2k_{p}^2}+8\varepsilon ^3\frac {k_2k_3k_1^2}{k^2k_{p}^2}, \end{align}
(A13g) \begin{align} D_{24}&={\textrm{i}}{\mathcal B}+\frac {\textrm{i}{\mathcal N}}{{\left ({1+4\varepsilon ^2}\right )}{\mathcal B}}\frac {\left (k_\perp ^2+4\varepsilon k_1k_3+4\varepsilon ^2\!\left (k_2^2+k_3^2\right )\right )}{k^2}, \end{align}

where $k_\perp =\sqrt {k_1^2+k_2^2}.$

Appendix B. Eigenvectors of the matrix $\boldsymbol{D}_0$ and the expression of $\boldsymbol{D}_\varepsilon$

In this appendix we report the eigenvectors of the matrix ${\boldsymbol{D}}_0$ (see (3.27)) that are associated with the eigenvalues $\sigma _1,\sigma _2=-\sigma _1,\sigma _3$ and $\sigma _4=-\sigma _3$ given by (3.31). Likewise, we report in this appendix the expression of the derivative of the matrix $\boldsymbol{D}$ with respect to the parameter $\varepsilon$ to order ${\mathcal O}(\varepsilon ),$ say $\boldsymbol{D}_\varepsilon .$

B.1. Eigenvectors associated with the eigenvalues of the matrix $\boldsymbol{D}_0$

The eigenvectors are the columns of the following matrix:

(B1) \begin{equation} \boldsymbol{T}=\left (\!\!\!\begin{array}{cccc} \sigma _1&-\sigma _1&\sigma _3&-\sigma _3\\[3pt] -\dfrac {1}{2}\left (\sigma _1^2+m^2\right )&-\dfrac {1}{2}\left (\sigma _1^2+m^2\right )&-\dfrac {1}{2}\left (\sigma _3^2+m^2\right )&-\dfrac {1}{2}\left (\sigma _3^2+m^2\right )\\[3pt] \textrm{i} m&\textrm{i} m&\textrm{i} m&\textrm{i} m\\[3pt] -\dfrac {\textrm{i}}{2}\frac {m}{\sigma _1}\left (\sigma _1^2+m^2\right )&\dfrac {\textrm{i}}{2}\dfrac {m}{\sigma _1}\left (\sigma _1^2+m^2\right )&-\dfrac {\textrm{i}}{2}\dfrac {m}{\sigma _3}\left (\sigma _3^2+m^2\right )&\dfrac {\textrm{i}}{2}\dfrac {m}{\sigma _3}\left (\sigma _3^2+m^2\right ) \end{array}\!\!\right ) \end{equation}

with $m={\mathcal B}\mu .$ The matrix $\boldsymbol{T}$ is inversible,

(B2) \begin{equation} \boldsymbol{T}^{-1}=\dfrac {1}{\left (\sigma _1^2-\sigma _3^2\right )}\left (\!\!\! \begin{array}{cccc} \dfrac {1}{2}\dfrac {\sigma _1}{m^2} \left (\sigma _3^2+m^2\right )&-1&\dfrac {\textrm{i}}{2m} \left (\sigma _3^2+m^2\right )&- \dfrac {\textrm{i}\sigma _1\sigma _3^2}{m^3}\\[8pt] -\dfrac {1}{2}\dfrac {\sigma _1}{m^2} \left (\sigma _3^2+m^2\right )&-1&\dfrac {\textrm{i}}{2m} \left (\sigma _3^2+m^2\right )& \dfrac {\textrm{i}\sigma _1\sigma _3^2}{m^3}\\[8pt] -\dfrac {1}{2}\dfrac {\sigma _3}{m^2} \left (\sigma _1^2+m^2\right )&1&-\dfrac {\textrm{i}}{2m} \left (\sigma _1^2+m^2\right )& \dfrac {\textrm{i}\sigma _3\sigma _1^2}{m^3}\\[8pt] \dfrac {1}{2}\dfrac {\sigma _3}{m^2} \left (\sigma _1^2+m^2\right )&1&-\dfrac {\textrm{i}}{2m} \left (\sigma _1^2+m^2\right )&- \dfrac {\textrm{i}\sigma _3\sigma _1^2}{m^3} \end{array}\!\!\right )\! , \end{equation}

so that,

(B3) \begin{equation} \tilde {\boldsymbol{D}}_0=\boldsymbol{T}^{-1}{\boldsymbol{\cdot }} \boldsymbol{D}_0{\boldsymbol{\cdot }} \boldsymbol{T}=\text{diag}\left (\sigma _1,-\sigma _1,\sigma _3,-\sigma _3\right )\! , \end{equation}

in the base diagonalizing $\boldsymbol{D}_0.$ Such a basis is used to facilitate the calculations leading to the expression of the amplification rates of destabilizing resonances of order $n=1.$

B.2. Expression of the matrix $\boldsymbol{D}_\varepsilon$

From (3.24) and (A13) giving the non-zero elements of the matrix $\boldsymbol{D}$ for the KBF and MBF cases, respectively, we deduce that the matrix $\boldsymbol{D}_\varepsilon$ (appearing in the expansion (4.9)) is the same for these two basic flows. The non-zero elements of $\boldsymbol{D}_\varepsilon$ are

(B4a) \begin{align} \left (\boldsymbol{D}_\varepsilon \right )_{11}&= 2\textrm{i} \mu \sqrt {1-\mu ^2}\big (\text{e}^{{i}\tau }-\text{e}^{-{i}\tau }\big ), \end{align}
(B4b) \begin{align} \left (\boldsymbol{D}_\varepsilon \right )_{12}&=-\frac {\sqrt {1-\mu ^2}}{\mu }\big (\text{e}^{{i}\tau }+\text{e}^{-{i}\tau }\big ), \end{align}
(B4c) \begin{align} \left (\boldsymbol{D}_\varepsilon \right )_{14}&=- \frac {{\mathcal N}^2}{{\mathcal B}}\frac {\left (1-\mu ^2\right )\sqrt {1-\mu ^2}}{\mu ^2}\big (\text{e}^{{i}\tau }-\text{e}^{-{i}\tau }\big ), \end{align}
(B4d) \begin{align} \left (\boldsymbol{D}_\varepsilon \right )_{21}&= \mu \sqrt {1-\mu ^2}\big (4\mu ^2-1\big )\big (\text{e}^{{i}\tau }+\text{e}^{-{i}\tau }\big ), \end{align}
(B4e) \begin{align} \left (\boldsymbol{D}_\varepsilon \right )_{24}&=-2\textrm{i} \frac {{\mathcal N}^2}{{\mathcal B}}\frac {\left (1-\mu ^2\right )\sqrt {1-\mu ^2}}{\mu ^2}\big (\text{e}^{{i}\tau }+\text{e}^{-{i}\tau }\big ). \end{align}

Thus, the asymptotic analysis to order ${\mathcal O}(\varepsilon )$ is the same for both basic flows (KBF and MBF).

References

Barker, A.J. 2016 a On turbulence driven by axial precession and tidal evolution of the spin–orbit angle of close-in giant planets. Mon. Not. R. Astron. Soc. 460 (3), 23392350.CrossRefGoogle Scholar
Barker, A.J. 2016 b Non-linear tides in a homogeneous rotating planet or star: global simulations of the elliptical instability. Mon. Not. R. Astron. Soc. 459 (1), 939956.CrossRefGoogle Scholar
Bayly, B.J. 1986 Three-dimensional instability of elliptical flow. Phys. Rev. Lett. 57 (17), 2160.CrossRefGoogle ScholarPubMed
Benkacem, N., Salhi, A., Khlifi, A., Nasraoui, S.   Cambon, C. 2022 Destabilizing resonances of precessing inertia-gravity waves. Phys. Rev. E 105 (3), 035107.CrossRefGoogle ScholarPubMed
Braginsky, S.I. 1991 Towards a realistic theory of the geodynamo. Geophys. Astrophys. Fluid Dyn. 60 (1–4), 89134.CrossRefGoogle Scholar
Buffett, B., Knezek, N.   Holme, R. 2016 Evidence for MAC waves at the top of Earths core and implications for variations in length of day. Geophys. J. Intl 204 (3), 17891800.CrossRefGoogle Scholar
Buffett, B.A. 2021 Conditions for turbulent Ekman layers in precessionally driven flow. Geophys. J. Intl 226 (1), 5665.CrossRefGoogle Scholar
Burmann, F.   Noir, J. 2022 Experimental study of the flows in a non-axisymmetric ellipsoid under precession. J. Fluid Mech. 932, A24.CrossRefGoogle Scholar
Busse, F.H. 1968 Steady fluid flow in a precessing spheroidal shell. J. Fluid Mech. 33 (4), 739751.CrossRefGoogle Scholar
Cambon, C., Teissedre, C.   Jeandel, D. 1985 Etude d’effets couplés de déformation et de rotation sur une turbulence homogène. Journal de mécanique théorique et appliquée 4 (5), 629657.Google Scholar
Caton, F., Janiaud, B.   Hopfinger, E.J. 2000 Stability and bifurcations in stratified Taylor–Couette flow. J. Fluid Mech. 419, 93124.CrossRefGoogle Scholar
Cébron, D., Laguerre, R., Noir, J.   Schaeffer, N. 2019 Precessing spherical shells: flows, dissipation, dynamo and the lunar core. Geophys. J. Intl 219 (Supplement 1), S34S57.CrossRefGoogle Scholar
Craik, A.D.D.   Criminale, W.O. 1986 Evolution of wavelike disturbances in shear flows: a class of exact solutions of the Navier–Stokes equations. Proc. R. Soc. Lond. A. Math. Phys. Sci. 406 (1830), 1326.Google Scholar
Craik, A.D.D. 1989 The stability of unbounded two-and three-dimensional flows subject to body forces: some exact solutions. J. Fluid Mech. 198, 275292.CrossRefGoogle Scholar
Davidson, P.A. 2013 Turbulence in Rotating, Stratified and Electrically Conducting Fluids. Cambridge University Press.CrossRefGoogle Scholar
Davies, P.A. 1972 Experiments on Taylor columns in rotating stratified fluids. J. Fluid Mech. 54 (4), 691717.CrossRefGoogle Scholar
Drazin, P.G.   Reid, W.H. 2004 Hydrodynamic Stability. Cambridge University Press.CrossRefGoogle Scholar
Etchevest, M., Fontana, M.   Dmitruk, P. 2022 Behavior of hydrodynamic and magnetohydrodynamic turbulence in a rotating sphere with precession and dynamo action. Phys. Rev. Fluids 7 (10), 103801.CrossRefGoogle Scholar
Fearn, D.R.   Proctor, M.R.E. 1983 Hydromagnetic waves in a differentially rotating sphere. J. Fluid Mech. 128, 120.CrossRefGoogle Scholar
Finlay, C.C., Maus, S., Beggan, C.D., Bondar, T.N., Chambodut, A., Chernova, T.A.   Zvereva, T.I. 2010 International geomagnetic reference field: the eleventh generation. Geophys. J. Intl 183 (3), 12161230.Google Scholar
Gans, R.F. 1970 On the precession of a resonant cylinder. J. Fluid Mech. 41 (4), 865872.CrossRefGoogle Scholar
Gao, D., Meunier, P., Le Dizès, S.   Eloy, C. 2021 Zonal flow in a resonant precessing cylinder. J. Fluid Mech. 923, A29.CrossRefGoogle Scholar
Giesecke, A., Vogt, T., Gundrum, T.   Stefani, F. 2019 Kinematic dynamo action of a precession-driven flow based on the results of water experiments and hydrodynamic simulations. Geophys. Astrophys. Fluid Dyn. 113 (1–2), 235255.CrossRefGoogle Scholar
Giesecke, A., Vogt, T., Pizzi, F., Kumar, V., Gonzalez, F.G., Gundrum, T.   Stefani, F. 2024 The global flow state in a precessing cylinder. J. Fluid Mech. 998, A30.CrossRefGoogle Scholar
Glampedakis, K., Andersson, N.   Jones, D.I. 2008 Stability of precessing superfluid neutron stars. Phys. Rev. Lett. 100 (8), 081101.CrossRefGoogle ScholarPubMed
Greenspan, H.P. 1969 The Theory of Rotating Fluids. Cambridge University Press.Google Scholar
Hollerbach, R.   Cally, P.S. 2009 Nonlinear evolution of axisymmetric twisted flux tubes in the solar tachocline. Solar Phys. 260, 251260.CrossRefGoogle Scholar
Kerswell, R.R. 1993 The instability of precessing flow. Geophys. Astrophys. Fluid Dyn. 72 (1-4), 107144.CrossRefGoogle Scholar
Kerswell, R.R. 2002 Elliptical instability. Annu. Rev. Fluid Mech. 34 (1), 83113.CrossRefGoogle Scholar
Khlifi, A., Salhi, A., Nasraoui, S., Godeferd, F.   Cambon, C. 2018 Spectral energy scaling in precessing turbulence. Phys. Rev. E 98 (1), 011102.CrossRefGoogle ScholarPubMed
Kuchment, P.A. 1993 Floquet Theory for Partial Differential Equations, vol. 60. Springer.CrossRefGoogle Scholar
Kumar, V., Pizzi, F., Mamatsashvili, G., Giesecke, A., Stefani, F.   Barker, A.J. 2024 Dynamo action driven by precessional turbulence. Phys. Rev. E 109 (6), 065101.CrossRefGoogle ScholarPubMed
Lam, K., Kong, D.   Zhang, K. 2021 Localised thermal convection in rotating spheres that undergo weak precession. Geophys. Astrophys. Fluid Dyn. 115 (3), 280296.CrossRefGoogle Scholar
Lebovitz, N.R.   Zweibel, E. 2004 Magnetoelliptic instabilities. Astrophys. J. 609 (1), 301.CrossRefGoogle Scholar
Lehner, T., Mouhali, W., Léorat, J.   Mahalov, A. 2010 Mode coupling analysis and differential rotation in a flow driven by a precessing cylindrical container. Geophys. Astrophys. Fluid Dyn. 104 (4), 369401.CrossRefGoogle Scholar
Lifschitz, A.   Hameiri, E. 1991 Local stability conditions in fluid dynamics. Phys. Fluids A: Fluid Dyn. 3 (11), 26442651.CrossRefGoogle Scholar
Lin, Y., Noir, J.   Jackson, A. 2014 Experimental study of fluid flows in a precessing cylindrical annulus. Phys. Fluids 26, 046604.CrossRefGoogle Scholar
Lorenzani, S.   Tilgner, A. 2001 Fluid instabilities in precessing spheroidal cavities. J. Fluid Mech. 447, 111128.CrossRefGoogle Scholar
Mahalov, A. 1993 The instability of rotating fluid columns subjected to a weak external Coriolis force. Phys. Fluids A: Fluid Dyn. 5 (4), 891900.CrossRefGoogle Scholar
Malkus, W.V.R. 1968 Precession of the Earth as the cause of geomagnetism: experiments lend support to the proposal that precessional torques drive the earth’s dynamo. Science 160 (3825), 259264.CrossRefGoogle Scholar
Mason, R.M.   Kerswell, R.R. 2002 Chaotic dynamics in a strained rotating flow: a precessing plane fluid layer. J. Fluid Mech. 471, 71106.CrossRefGoogle Scholar
McKeown, R., Ostilla-Mónico, R., Pumir, A., Brenner, M.P.   Rubinstein, S.M. 2020 Turbulence generation through an iterative cascade of the elliptical instability. Sci. Adv. 6 (9), eaaz2717.CrossRefGoogle ScholarPubMed
Miyazaki, T.   Fukumoto, Y. 1992 Three-dimensional instability of strained vortices in a stably stratified fluid. Phys. Fluids A: Fluid Dyn. 4 (11), 25152522.CrossRefGoogle Scholar
Mizerski, K.A., Bajer, K.   Moffatt, H.K. 2012 The mean electromotive force generated by elliptic instability. J. Fluid Mech. 707, 111128.CrossRefGoogle Scholar
Moffatt, H.K. 1978 Magnetic field generation in electrically conducting fluids. In Cambridge Monographs on Mechanics and Applied Mathematics.Google Scholar
Moffatt, K.   Dormy, E. 2019 Self-Exciting Fluid Dynamos, vol. 59. Cambridge University Press.CrossRefGoogle Scholar
Mouhali, W., Lehner, T., Léorat, J.   Vitry, R. 2012 Evidence for a cyclonic regime in a precessing cylindrical container. Exp. Fluids 53, 16931700.CrossRefGoogle Scholar
Mouhali, W., Salhi, A., Lehner, T.   Cambon, C. 2024 Waves and non-propagating mode in stratified and rotating magnetohydrodynamic turbulence. Phys. Fluids 36, 125194.CrossRefGoogle Scholar
Nornberg, M.D., Ji, H., Schartman, E., Roach, A.   Goodman, J. 2010 Observation of magnetocoriolis waves in a liquid metal Taylor–Couette experiment. Phys. Rev. Lett. 104 (7), 074501.CrossRefGoogle Scholar
Pedlosky, J. 2013 Geophysical Fluid Dynamics. Springer Science   Business Media.Google Scholar
Pierrehumbert, R.T. 1986 Universal short-wave instability of two-dimensional eddies in an inviscid fluid. Phys. Rev. Lett. 57 (17), 2157.CrossRefGoogle Scholar
Pizzi, F., Mamatsashvili, G., Barker, A.J., Giesecke, A.   Stefani, F. 2022 Interplay between geostrophic vortices and inertial waves in precession-driven turbulence. Phys. Fluids 34, 125135.CrossRefGoogle Scholar
Poincaré, H. 1910 Sur la précession des corps déformables. Bulletin astronomique, Observatoire de Paris 27 (1), 321356.CrossRefGoogle Scholar
Salhi, A.   Cambon, C. 2009 Precessing rotating flows with additional shear: stability analysis. Phys. Rev. E 79 (3), 036303.CrossRefGoogle ScholarPubMed
Salhi, A., Lehner, T.   Cambon, C. 2010 Magnetohydrodynamic instabilities in rotating and precessing sheared flows: an asymptotic analysis. Phys. Rev. E 82 (1), 016315.CrossRefGoogle ScholarPubMed
Salhi, A., Lehner, T., Godeferd, F.   Cambon, C. 2012 Magnetized stratified rotating shear waves. Phys. Rev. E 85 (2), 026301.CrossRefGoogle ScholarPubMed
Salhi, A., Lehner, T., Godeferd, F.   Cambon, C. 2013 Wave–vortex mode coupling in astrophysical accretion disks under combined radial and vertical stratification. Astrophys. J. 771 (2), 103.CrossRefGoogle Scholar
Salhi, A., Baklouti, F.S., Godeferd, F., Lehner, T.   Cambon, C. 2017 Energy partition, scale by scale, in magnetic Archimedes Coriolis weak wave turbulence. Phys. Rev. E 95 (2), 023112.CrossRefGoogle ScholarPubMed
Salhi, A., Khlifi, A.   Cambon, C. 2019 Nonlinear effects on the precessional instability in magnetized turbulence. Atmosphere-BASEL 11 (1), 14.CrossRefGoogle Scholar
Salhi, A.   Cambon, C. 2023 Magneto-gravity-elliptic instability. J. Fluid Mech. 963, A9.CrossRefGoogle Scholar
Salhi, A., Khlifi, A., Marino, R., Feraco, F., Foldes, R.   Cambon, C. 2024 Waves and non-propagating modes in stratified MHD turbulence subject to a weak mean magnetic field. J. Fluid Mech. 1001, A28.CrossRefGoogle Scholar
Slane, J.   Tragesser, S. 2011 Analysis of periodic nonautonomous inhomogeneous systems. Nonlinear Dyn. Syst. Theory 11 (2), 183198.Google Scholar
Spiegel, E.A.   Veronis, G. 1960 On the Boussinesq approximation for a compressible fluid. Astrophys. J. 131, 442.CrossRefGoogle Scholar
Tilgner, A. 1999 Magnetohydrodynamic flow in precessing spherical shells. J. Fluid Mech. 379, 303318.CrossRefGoogle Scholar
Tomczyk, S., McIntosh, S.W., Keil, S.L., Judge, P.G., Schad, T., Seeley, D.H.   Edmondson, J. 2007 Alfvén waves in the solar corona. Science 317 (5842), 11921196.CrossRefGoogle ScholarPubMed
Vladimirov, V.A.   Tarassov, V.F. 1984 Forced regimes of motion of a rotating fluid. Sov. Phys. D 30, 654.Google Scholar
Vormann, J.   Hansen, U. 2020 Characteristics of a precessing flow under the influence of a convecting temperature field in a spheroidal shell. J. Fluid Mech. 891, A15.CrossRefGoogle Scholar
Waleffe, F. 1992 The nature of triad interactions in homogeneous turbulence. Phys. Fluids A: Fluid Dyn. 4 (2), 350363.CrossRefGoogle Scholar
Wei, X.   Tilgner, A. 2013 Stratified precessional flow in spherical geometry. J. Fluid Mech. 718, R2.CrossRefGoogle Scholar
Wei, X. 2016 The combined effect of precession and convection on the dynamo action. Astrophys. J. 827 (2), 123.CrossRefGoogle Scholar
Wiener, R.J., Hammer, P.W., Swanson, C.E.   Donnelly, R.J. 1990 Stability of Taylor–Couette flow subject to an external Coriolis force. Phys. Rev. Lett. 64 (10), 1115.CrossRefGoogle Scholar
Zhang, K., Chan, K.H.   Liao, X. 2014 On precessing flow in an oblate spheroid of arbitrary eccentricity. J. Fluid Mech. 743, 358384.CrossRefGoogle Scholar
Figure 0

Figure 1. A slice of a vertical stratified precessing fluid column: a rotating fluid column seen in a frame rotating uniformly about the $\hat {\boldsymbol{x}}\hbox{-}$axis with rate $\varOmega _{\!p}=\varepsilon \varOmega _0$ where $\varepsilon$ is the Poincaré number. Without precession $(\varepsilon =0),$ the unperturbed base flow, ${\boldsymbol{U}}=\varOmega _0r\hat \varphi ,$ has circular streamlines, and the isodensity planes are perpendicular to the vertical axis $\hat {\boldsymbol{z}},$$\boldsymbol{\nabla }\varrho (z)=-(\rho _0/g)N^2\hat {\boldsymbol{z}},$ where $N^2=\text{const.}\gt 0$ is the square of the Brunt–Väisälä frequency. The effect of the Coriolis force induces a vertical mean shear that acts to balance the gyroscopic torque, so that the unperturbed velocity profile becomes ${\boldsymbol {U}} = \varOmega _0r (\hat {\boldsymbol{\varphi }}-2\varepsilon \sin \varphi \hat {\boldsymbol{z}})=\varOmega _0 (-y\hat {\boldsymbol{x}}+x\hat {\boldsymbol{y}}-2\varepsilon \hat {\boldsymbol{z}}).$ The trajectory of a fluid particle is in the plane perpendicular to the $z^*\hbox{-}$axis and it is an ellipse (see § 2.2.1). The plane $(x^*,z^*)$ is obtained by a rotation, of angle $\gamma =-\tan ^{-1}(2\varepsilon ),$ of the plane $(x,z)$ around the $y\hbox{-}$axis. Moreover, in the presence of the Coriolis force, the isodensity planes are perpendicular to the $z^*\hbox{-}$axis, so that $\boldsymbol{\nabla }\varrho (z^*)=-(\rho _0/g)N^2\hat {\boldsymbol{z}}^*$ with ${\boldsymbol{n}}=\hat {\boldsymbol{z}}^*$ (see (2.17)). Here, the colour variation of the streamlines (from red to blue) represents the linear variation of the basic density, and ${\boldsymbol{B}}_0$ denotes the basic (constant) magnetic field.

Figure 1

Figure 2. Resonant cases of MIG waves. (a) Resonant cases (of order $n=1$) occur for $({\mathcal B},{\mathcal N})$ belonging to the coloured domains. (b) Resonance between two fast modes ($\omega _1-\omega _2=\varOmega _0)$ where (I) $\equiv [0,+\infty )\times [0,1]$ and (II) $\equiv {\mathcal D}_f$ (see (4.19)). (c) Resonance between two slow modes ($\omega _3-\omega _4=\varOmega _0)$ where (III) $\equiv [\sqrt {5}/2,+\infty )\times [0,+\infty ).$ (d) Resonance between a fast mode and a slow mode ($\omega _1-\omega _4=\varOmega _0$ or $\omega _1-\omega _3=\varOmega _0)$ where (IV) $\equiv [0,+\infty )\times [0,1]$ and (V) $\equiv {\mathcal D}_m$ (see (4.30)).

Figure 2

Figure 3. Resonant cases of MIG waves. (a) Variation of the parameter $\mu =k_0/K_0$ versus ${\mathcal N}=N/\varOmega _0$ for ${\mathcal B}= B_0K_0/\varOmega _0=0.75,\ 1.0,\ \sqrt {5}/2$ such that $({\mathcal B},{\mathcal N})\in {\mathcal D}_f$ (see (4.19)) associated with the resonant cases between two fast modes ($\omega _1-\omega _2=\varOmega _0)$ or between two slow modes ($\omega _3-\omega _4=\varOmega _0).$ (b) Variation of the parameter $\mu =k_0/K_0$ versus $\mathcal B$ for ${\mathcal N}=1.2,\ 2.0,\ 3.0$ such that $({\mathcal B},{\mathcal N})\in {\mathcal D}_m$ (see (4.30)) associated with the resonant cases between a fast mode and a slow mode ($\omega _1-\omega _4=\varOmega _0$ or $\omega _1-\omega _3=\varOmega _0).$ Here, $K_0=\sqrt {k_0^2+k_{\!p}^2}.$

Figure 3

Figure 4. Maximal growth rate of destabilizing resonant cases given by the asymptotic formulae. (a) Resonant cases between two fast modes for some values of $({\mathcal B},{\mathcal N})\in [0,+\infty )\times [0,1/2].$ (b) Resonant cases between two fast modes or between two slow modes for some values of $({\mathcal B},{\mathcal N})\in {\mathcal D}_f$ (see (4.19)). (c) Resonant cases between two slow modes for some values of $({\mathcal B},{\mathcal N})\in [\sqrt {5}/2,+\infty )\times [0,+\infty ).$ (d) Resonant cases between a fast mode and a slow mode for some values of $({\mathcal B},{\mathcal N})\in [0,+\infty )\times [0,1].$ The numerical results obtained for $\varepsilon =0.05$ (symbol) are also reported.

Figure 4

Figure 5. Magneto-gravity-precessional instabilities. The figure shows, for fixed $\varepsilon ,$$\varepsilon +\Re \sigma$ versus $\mu$ for ${\mathcal B}=0.75$ and ${\mathcal N}=0.25$ and $100$ values of $\varepsilon$ evenly distributed in the interval $[0,0.25].$ (a) The case of KBF; (b) the case of MBF. The instability region emanating from the point ($\mu = 0.196, \varepsilon +\Re \sigma =0)$ is associated with the destabilizing resonance (of order $n=1)$ between two fast modes, while that emanating from the point ($\mu = 0.380, \varepsilon +\Re \sigma =0$) is associated with the destabilizing resonance (of order $n=1)$ between a fast mode and a slow mode. The third region, which emanates from the point ($\mu = 0.434,\varepsilon +\Re \sigma =0$), characterizes the destabilizing resonance of order $n=2$ between two fast modes.

Figure 5

Figure 6. Magneto-gravity-precessional instabilities. Maximal growth rate of dominant instability normalized by $\sigma _0=(5\sqrt {15}/32)\varepsilon$ plotted as a function of $0\leqslant {\mathcal B}=K_0B_0/\varOmega _0\leqslant 8$ and $0\leqslant {\mathcal N}=N/\varOmega _0\leqslant 2$: (a) numerical results for $\varepsilon =0.05$; (b) asymptotic analysis results.

Figure 6

Figure 7. Magneto-gravity-precessional instabilities. Maximal growth rate of dominant instability normalized by $\sigma _0=(5\sqrt {15}/32)\varepsilon$ plotted as a function of $0\leqslant {\mathcal B}=K_0B_0/\varOmega _0\leqslant 8$ and $0\leqslant {\mathcal N}=N/\varOmega _0\leqslant 2$ for $\varepsilon =0.2.$ Here (a) KBF; (b) MBF.

Figure 7

Figure 8. Magneto-gravity-precessional instabilities. Maximal growth rate of dominant instability normalized by $\sigma _0=(5\sqrt {15}/32)\varepsilon$ plotted as a function of ${\mathcal N}=N/\varOmega _0$ for ${\mathcal B}=k_0B_0/\varOmega _0=5$ and $\varepsilon =0.05,\ 0.10,\ 0.20.$ Here (a) KBF; (b) MBF.

Supplementary material: File

Salhi et al. supplementary material

Salhi et al. supplementary material
Download Salhi et al. supplementary material(File)
File 173 KB