Hostname: page-component-857557d7f7-bkbbk Total loading time: 0 Render date: 2025-11-21T01:19:02.334Z Has data issue: false hasContentIssue false

Subcritical transition and multistability in liquid-metal magnetoconvection with sidewalls

Published online by Cambridge University Press:  19 November 2025

Matthew McCormack
Affiliation:
School of Mathematics and Maxwell Institute for Mathematical Sciences, University of Edinburgh, UK
Andrei Teimurazov
Affiliation:
Max Planck Institute for Dynamics and Self-Organization, Göttingen 37077, Germany
Olga Shishkina
Affiliation:
Max Planck Institute for Dynamics and Self-Organization, Göttingen 37077, Germany
Moritz Linkmann*
Affiliation:
School of Mathematics and Maxwell Institute for Mathematical Sciences, University of Edinburgh, UK
*
Corresponding author: Moritz Linkmann, moritz.linkmann@ed.ac.uk

Abstract

The motionless conducting state of liquid-metal convection with an applied vertical magnetic field confined in a vessel with insulating sidewalls becomes linearly unstable to wall modes through a supercritical pitchfork bifurcation. Nevertheless, we show that the transition proceeds subcritically, with stable finite-amplitude solutions with different symmetries existing at parameter values beneath this linear stability threshold. Under increased thermal driving, the branch born from the linear instability becomes unstable and solutions are attracted to the most subcritical branch, which follows a quasiperiodic route to chaos. Thus, we show that the transition to turbulence is controlled by this subcritical branch and hence turbulent solutions have no connection to the initial linear instability. This is further quantified by observing that the subcritical equilibrium solution sets the spatial symmetry of the turbulent mean flow and thus organises large-scale structures in the turbulent regime.

Information

Type
JFM Rapids
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

Liquid-metal magnetoconvection (MC), where an electrically conducting fluid is subjected to a time-independent magnetic field and an imposed temperature gradient, is relevant to engineering applications such as liquid-metal cooling systems for nuclear fusion reactors (Davidson Reference Davidson1999), geoastrophysical systems such as flows in liquid-metal planetary cores (Jones Reference Jones2011) and associated laboratory experiments (Stefani Reference Stefani2024). Despite its relevance, however, a number of uncertainties persist regarding the transition to turbulence in this system, especially in confined containers relevant to experiments and engineering applications. Here, we focus on the case where the uniform applied magnetic field is orientated in the vertical direction, directly opposing gravity. In contrast to magnetoconvection in a laterally periodic/infinite plane layer, where a linear instability of the motionless conducting state classically gives rise to domain-filling arrays of convection rolls (Chandrasekhar Reference Chandrasekhar1961; Busse & Clever Reference Busse and Clever1982), wall-localised onset modes, named wall modes, are observed when sidewalls are introduced to the system with a sufficiently strong magnetic field. Wall modes observed in direct numerical simulations (DNS) (Liu, Krasnov & Schumacher Reference Liu, Krasnov and Schumacher2018; Akhmedagaev et al. Reference Akhmedagaev, Zikanov, Krasnov and Schumacher2020; McCormack et al. Reference McCormack, Teimurazov, Shishkina and Linkmann2023; Xu, Horn & Aurnou Reference Xu, Horn and Aurnou2023; Wu, Chen & Ni Reference Wu, Chen and Ni2025) and experiments (Zürner et al. Reference Zürner, Schindler, Vogt, Eckert and Schumacher2020; Xu et al. Reference Xu, Horn and Aurnou2023) have been thought to originate from a linear instability of the conducting state, originating from the analytic-numerical hybrid approach of Houchens, Witkowski & Walker (Reference Houchens, Witkowski and Walker2002) and the asymptotic (large magnetic field limit) linear stability analysis of Busse (Reference Busse2008). However, recent work suggests discrepancies between linear stability theory, simulations and experiments in the measurement of the onset of convection (Xu et al. Reference Xu, Horn and Aurnou2023), and short-time simulations displaying exponential growth from a perturbed conducting state were observed to have a different modal structure than converged wall-mode equilibria at similar parameter values (McCormack et al. Reference McCormack, Teimurazov, Shishkina and Linkmann2023). In this work, we show for the first time that multiple stable wall-mode solutions exist at identical operating conditions, with fundamental consequences for the transition to turbulence in this flow.

In a general fluid dynamics setting, systems typically either follow a supercritical or subcritical transition to turbulence. A supercritical transition occurs when the laminar flow undergoes successive supercritical bifurcations, transitioning the flow to an increasingly spatio-temporally complex state. Each transition point is clearly identifiable in parameter space, and may be found using a linear stability analysis due to the local nature of the instabilities. Such behaviour has been identified in many flow configurations such as Taylor–Couette flow (TCF) (Taylor Reference Taylor1923; Gollub & Swinney Reference Gollub and Swinney1975; Brandstater & Swinney Reference Brandstater and Swinney1987; Chossat & Iooss Reference Chossat and Iooss1994) and Rayleigh–Bénard convection (RBC) (Rayleigh Reference Rayleigh1916; Malkus & Veronis Reference Malkus and Veronis1958; Chandrasekhar Reference Chandrasekhar1961; Libchaber, Laroche & Fauve Reference Libchaber, Laroche and Fauve1982; Ecke, Mainieri & Sullivan Reference Ecke, Mainieri and Sullivan1991; Bodenschatz, Pesch & Ahlers Reference Bodenschatz, Pesch and Ahlers2000; Borońska & Tuckerman Reference Borońska and Tuckerman2010; Oteski et al. Reference Oteski, Duguet, Pastur and Le Quéré2015). In other cases, however, the transition proceeds subcritically, where finite-amplitude disturbances trigger the transition despite a linearly stable laminar flow. In this case, non-trivial nonlinear solutions must be identified beneath the linear stability threshold to understand the transition. This leads to a different phenomenology, featuring hysteresis and large, unpredictable changes in the system response triggered at a range of parameter values depending on the supplied disturbance. In fluid dynamics and related fields, subcritical transitions and hysteretic effects occur in a wide range of systems. They are prominent concerning the transition to turbulence in wall-bounded shear flows of neutral (Kerswell Reference Kerswell2005; Eckhardt et al. Reference Eckhardt, Schneider, Hof and Westerweel2007; Eckhardt Reference Eckhardt2018; Avila, Barkley & Hof Reference Avila, Barkley and Hof2023; Hof Reference Hof2023) and electrically conducting fluids (for comprehensive review see Zikanov et al. (Reference Zikanov, Krasnov, Boeck, Thess and Rossi2014)), in the transition to elasto-inertial and elastic turbulence for flows with straight streamlines (Morozov & van Saarloos Reference Morozov and van Saarloos2007; Bonn et al. Reference Bonn, Ingremeau, Amarouchene and Kellay2011; Pan et al. Reference Pan, Morozov, Wagner and Arratia2013; Lellep, Linkmann & Morozov Reference Lellep, Linkmann and Morozov2024) and for types of (approximately) homogeneous turbulence (Linkmann & Morozov Reference Linkmann and Morozov2015). Finite-amplitude transitions between different turbulent states, such as those characterised by the presence or absence of a condensate, occur in continuum models describing certain regimes of active turbulence (Linkmann et al. Reference Linkmann, Boffetta, Marchetti and Eckhardt2019, Reference Linkmann, Marchetti, Boffetta and Eckhardt2020b ), two-dimensional turbulence depending on the type of forcing (Linkmann et al. Reference Linkmann, Hohmann and Eckhardt2020a ; Gallet Reference Gallet2024) and in quasi-two-dimensional turbulence occurring in rotating systems or thin fluid layers (Favier, Guervilly & Knobloch Reference Favier, Guervilly and Knobloch2019; de Wit et al. Reference de Wit, Aguirre Guzmán, Clercx and Kunnen2022a , Reference de Wit, van Kan and Alexakisb ; Yokoyama & Takaoka Reference Yokoyama and Takaoka2017). Moreover, there are more complex scenarios, such as the transition to the ultimate regime of thermal convection (Roche Reference Roche2020; Lohse & Shishkina Reference Lohse and Shishkina2023, Reference Lohse and Shishkina2024; Shishkina & Lohse Reference Shishkina and Lohse2024) which occurs through a subcritical transition from laminar to turbulent boundary layers, in the presence of turbulent flow in the bulk.

Liquid-metal MC (i.e. large magnetic diffusivity) transitions supercritically in a laterally periodic layer (Chandrasekhar Reference Chandrasekhar1961; Busse & Clever Reference Busse and Clever1982; Proctor & Weiss Reference Proctor and Weiss1982; Weiss & Proctor Reference Weiss and Proctor2014), and the motionless conducting state undergoes a supercritical bifurcation when laterally confined between sidewalls (Houchens et al. Reference Houchens, Witkowski and Walker2002; Busse Reference Busse2008; Bhattacharya et al. Reference Bhattacharya, Boeck, Krasnov and Schumacher2024). Despite this, we show that non-trivial nonlinear equilibrium solutions exist beneath the wall-mode linear stability threshold, and that the transition proceeds on this subcritical branch of solutions. We further observe that this has a fundamental impact on the organisation of large-scale structures in the turbulent regime. This implies that (a) the transition to turbulence and turbulent solutions in closed-vessel experiments are fundamentally different to the laterally unbounded geoastrophysical flows they often intend to replicate in nature, and (b) we must change the way we think about the transition and the formation of large-scale structures in convective and potentially other nonlinear pattern-forming systems subject to external forces, even when linear stability theory predicts a supercritical transition.

2. Formulation

An incompressible flow of viscous and electrically conducting fluid is driven by an imposed vertical temperature difference, $\delta T$ , and a constant vertical magnetic field $\boldsymbol{B} = B_0\boldsymbol{e}_z$ , where $\boldsymbol{e}_z$ is the vertical unit vector. The equations of motion under the quasistatic approximation (magnetic Reynolds ${Rm} = {U} \ell /\eta \ll \rm 1$ and magnetic Prandtl number ${Pm} = \nu /\eta \ll 1$ , where $U$ and $\ell$ are characteristic velocity and length scales, $\nu$ is the kinematic viscosity and $\eta$ is the magnetic diffusivity) and the Oberbeck–Boussinessq approximation are

(2.1a) \begin{align}& \partial _t\boldsymbol{u} + \boldsymbol{u}\boldsymbol{\cdot }\boldsymbol{\nabla }\boldsymbol{u}+\boldsymbol{\nabla } p = \sqrt {\frac {Pr }{{{\textit{Ra}}}}} \big [\Delta \boldsymbol{u} + {{\textit{Ha}}} ^2(\boldsymbol{j}\times \boldsymbol{e}_z)\big ] + T\boldsymbol{e}_z, \end{align}
(2.1b) \begin{align}&\qquad\qquad\qquad \partial _t T + \boldsymbol{u}\boldsymbol{\cdot }\boldsymbol{\nabla } T = \frac {1}{\sqrt {{{\textit{Ra}}}{\textit{Pr}} }}\Delta T, \end{align}
(2.1c) \begin{align}& \boldsymbol{\nabla } \boldsymbol{\cdot } \boldsymbol{u} = 0, \quad \boldsymbol{j} = -\boldsymbol{\nabla }\phi + (\boldsymbol{u}\times \boldsymbol{e}_z), \quad \Delta \phi = \boldsymbol{\nabla }\boldsymbol{\cdot }(\boldsymbol{u}\times \boldsymbol{e}_z), \end{align}

where $\boldsymbol{u} = [u,v,w]^\textrm T$ is the velocity, $T$ the temperature, $p$ the kinematic pressure, $\boldsymbol{j}$ the electric current density and $\phi$ the electric field potential. Quantities have been non-dimensionalised using the height of the fluid layer $H$ , the temperature difference $\delta T\equiv T_+-T_- \gt 0$ , where $T_+$ and $T_-$ are the temperatures at the bottom and top plates, respectively, the free-fall velocity $u_{\textit{ff}}\equiv (\alpha gH\delta T)^{1/2}$ , the free-fall time $t_{\textit{ff}}\equiv H/u_{\textit{ff}}$ , the pressure $\rho u_{\textit{ff}}^2$ and the external magnetic field strength $B_0$ , where $\alpha$ is the thermal expansion coefficient and $g$ the acceleration due to gravity. The control parameters are the Rayleigh number ${\textit{Ra}}$ , Prandtl number $\textit{Pr}$ and Hartmann number ${\textit{Ha}}$ :

(2.2) \begin{eqnarray} {{\textit{Ra}}}\equiv \,\dfrac {\alpha g \,\delta T\, H^3}{\kappa \nu }, \qquad {\textit{Pr}}\equiv \,\dfrac {\nu }{\kappa }, \qquad {{\textit{Ha}}}\equiv \,B_0 H \sqrt {\dfrac {\sigma }{\rho \nu }} \ , \end{eqnarray}

where $\kappa$ is the thermal diffusivity, $\sigma$ the electrical conductivity and $\rho$ the mass density. We apply no-slip boundary conditions (BCs) for the velocity at all boundaries $\boldsymbol{u}=0$ , and adiabatic BC at the sidewalls, $\partial T/\partial \boldsymbol{n}=0$ , where $\boldsymbol{n}$ is the vector orthogonal to the surface. All solid boundaries are considered electrically insulating $\partial \phi /\partial \boldsymbol{n}=0$ . The spatial domain $\varOmega \subset \mathbb{R}^3$ is a cube with width $W$ and length $L$ , i.e. $H=W=L$ , and is resolved on non-uniform grids using the MC extension of goldfish (Kooij et al. Reference Kooij, Botchev, Frederix, Geurts, Horn, Lohse, van der Poel, Shishkina, Stevens and Verzicco2018; Reiter et al. Reference Reiter, Shishkina, Lohse and Krug2021, Reference Reiter, Zhang and Shishkina2022; McCormack et al. Reference McCormack, Teimurazov, Shishkina and Linkmann2023; Teimurazov et al. Reference Teimurazov, McCormack, Linkmann and Shishkina2024), which uses a third-order Runge–Kutta time integration scheme and a fourth-order finite-volume discretisation. The coordinates of the grid points, discretising the interval $[0,1]$ , are defined as $x_k = [1-(2\tilde {x}_k/(\tilde {x}_1 - \tilde {x}_n))]/2$ for $k = 1, \ldots , n$ , where $\tilde {x}_k$ are generalised Chebyshev nodes $\tilde {x}_k = \cos ((2k + 2k_{\textit{clip}}-1)\pi /(2n + 4k_{\textit{clip}}) )$ with uniformity parameter $k_{\textit{clip}} \in \mathbb{N}$ . In our DNS, we use at least $220^2$ points in the cross-plane direction with $k_{\textit{clip}} = 10$ , and 350 points in the vertical direction with $k_{\textit{clip}} = 1$ in the vertical direction. With this grid, spatial flow fluctuations are resolved to less than two Kolmogorov microscales for all simulations used in the stability analysis, with no less than nine points in the Hartmann boundary layer which forms on the top/bottom boundaries of the domain. Here, the Hartmann layer thickness is defined as $\delta _{\nu }\approx 1/{{\textit{Ha}}}$ which uses the prefactor measured by Teimurazov et al. (Reference Teimurazov, McCormack, Linkmann and Shishkina2024). This layer first forms near onset in the boundary layer underneath the wall modes and thus resolving this layer is critical to correctly analysing the stability of the solutions. Simulations with ${{\textit{Ra}}}\geqslant 10^8$ have been resolved on a finer $250^2\times 400$ grid to explore the long-time behaviour of the system, the results of which have been validated with shorter runs using a $350^2\times 600$ grid. We set $Pr=0.025$ for liquid metals such as gallium-indium-tin, and fix the Hartmann number at ${{\textit{Ha}}}=500$ or ${{\textit{Ha}}}=1000$ , leaving the Rayleigh number ${\textit{Ra}}$ as our bifurcation parameter.

Figure 1. (a) Bifurcation diagram at ${{\textit{Ha}}}=500$ showing the $L^2$ norm of the vertical velocity field $\|w\|_2$ of stable/unstable equilibria (filled/open markers), limit cycles (diamonds), invariant tori (asterisks) on the linear onset branch (LB) stemming from the linear instability at ${{\textit{Ra}}}_{c,L}$ , the mixed symmetry branch (MB) and the subcritical branch (SB). The mean of time-dependent solutions is shown. The grey area represents ${{\textit{Ra}}}\lt {{\textit{Ra}}}_{c,L}$ . (b) Upflow (pink) and downflow (blue) vertical velocity isosurfaces and streamlines (black) of solutions on the various branches shown from the top view at ${{\textit{Ha}}}=500$ at ${{\textit{Ra}}}=8\times 10^5$ and ${{\textit{Ra}}}=2\times 10^6$ with $w=\pm 0.005$ and $\pm 0.01$ , respectively.

3. Multistability of nonlinear wall-mode equilibria

3.1. Linear stability and the linear onset branch

Although a strong magnetic field can completely suppress convection, the motionless conducting state becomes linearly unstable with sufficient thermal driving in an infinitely wide and long domain, with the critical Rayleigh number ${{\textit{Ra}}}_c \sim {{\textit{Ha}}}^2$ (Chandrasekhar Reference Chandrasekhar1961) (shown by the open black star in figure 1 a). However, including insulating sidewalls results in a different linear instability giving rise to wall modes beneath the onset of bulk convection (Houchens et al. Reference Houchens, Witkowski and Walker2002; Busse Reference Busse2008; Bhattacharya et al. Reference Bhattacharya, Boeck, Krasnov and Schumacher2024), even in large-aspect ratio domains (also with finite sidewall conductivity). The critical Rayleigh number for a sidewall in a semi-infinite domain with free-slip top and bottom boundaries is ${{\textit{Ra}}}_c \sim {{\textit{Ha}}}^{3/2}$ for large ${\textit{Ha}}$ (Busse Reference Busse2008) (shown by the filled black star in figure 1 a). Since no stability theory has been performed for the domain and BCs we consider, or any other fully confined domain, we first perform a linear stability analysis for the conducting state. The linear stability threshold is obtained by perturbing the conducting state with random noise of magnitude $O(10^{-10})$ and scanning ${\textit{Ra}}$ to find an approximate threshold. Two equilibria are converged on each side of the threshold with $|\partial _t E_k| = O(10^{-14})$ where $E_k = (1/2)\int _\varOmega (\boldsymbol{u}\boldsymbol{\cdot }\boldsymbol{u}) \,\mathrm{d} \boldsymbol{x}$ , ensuring agreement between the fastest growing/slowest decaying modes. This modal structure (similar to figure 1 b for the linear onset branch (LB) at ${{\textit{Ra}}}=8\times 10^5$ ) is used as the spatial eigenmode to bound the linear stability threshold ${{\textit{Ra}}}_{c,L}$ by bisection. For ${{\textit{Ha}}}=500$ we have $6.89 \times 10^5 \lt {{\textit{Ra}}}_{c,L} \lt 6.90 \times 10^5$ and at ${{\textit{Ha}}}=1000$ we have $ 1.83 \times 10^6 \lt {{\textit{Ra}}}_{c,L} \lt 1.84 \times 10^6$ . Locally, based on these two points, ${{\textit{Ra}}}_{c,L}\sim {{\textit{Ha}}}^{1.412\pm 0.005}$ , where the error corresponds to the uncertainty in ${{\textit{Ra}}}_{c,L}$ at each ${\textit{Ha}}$ . Although this is a local measurement of the scaling at finite ${\textit{Ha}}$ , it is close to the asymptotic result for a single sidewall obtained by Busse (Reference Busse2008) where ${{\textit{Ra}}}_{c}\sim {{\textit{Ha}}}^{3/2}$ at leading order. The stable equilibria born from the linear instability (LB in the bifurcation diagram in figure 1 a) exhibit a discrete fourfold rotational symmetry in the vertical velocity field (figure 1 b (LB)), and is thus invariant under the rotation $r: rw(x,y,z) = \mathcal{R}_{\pi /2} w(x,y,z),$ where $\mathcal{R}_\theta$ represents a rotation $\theta$ about $\boldsymbol{e}_z$ . The solution additionally has reflection symmetries in the $x$ - $y$ plane (about the $x$ and $y$ axes, and about the two diagonals). Thus, these symmetries may be described by a group isomorphic to the dihedral group $D_4$ , generated by $r$ , the identity $\mathrm{id}$ and the reflection $s$ about the $x=y$ diagonal, which we will denote $D_4^+ = \{\mathrm{id}, r, r^2, r^3, s, sr, sr^2, sr^3\}$ . These solutions have thin convective wall modes with a two-layer structure near the sidewalls, with convection almost completely suppressed in the bulk. Each sidewall features two counter-rotating rolls, which lead to the spatially periodic pattern of up/down flows shown in the vertical velocity isosurfaces in figure 1(b) (LB). A secondary solution can be obtained by changing the direction of the rolls and flipping the vertical coordinate $ \textit{z}\kern-5.5pt{\scriptstyle -}\,:\,\textit{z}\kern-5.5pt{\scriptstyle -} w(x,y,z)= -w(x,y,-z)$ , and thus the bifurcation is a supercritical pitchfork, which has broken the $\mathbb{Z}_2$ Boussinesq symmetry of the $D_4 \times \mathbb{Z}_2 = D_4^+ \times \{\mathrm{id},\textit{z}\kern-5.5pt{\scriptstyle -}\}$ conducting state. Increasing ${\textit{Ra}}$ results in protrusions growing from the wall modes into the bulk of the flow, similar to other observed solutions (Liu et al. Reference Liu, Krasnov and Schumacher2018; McCormack et al. Reference McCormack, Teimurazov, Shishkina and Linkmann2023; Xu et al. Reference Xu, Horn and Aurnou2023), and the equilibria eventually lose stability in the range $ 3 \times 10^6 \lt {{\textit{Ra}}} \lt 4\times 10^6$ . Since solutions on this branch are only weakly unstable up to ${{\textit{Ra}}}\approx 10^7$ , we have estimated the continuation of the unstable branch in figure 1(a).

3.2. Subcritical branch

Having determined the threshold for a linear instability, we now assess the nonlinear stability of the conducting state and search for subcritical solutions. We note that the solutions on the LB have a different spatial symmetry than other wall-mode solutions in the same domain (McCormack et al. Reference McCormack, Teimurazov, Shishkina and Linkmann2023). Thus, to try to construct a different branch of solutions, we initialise from a solution obtained far from the threshold of the linear instability ( ${{\textit{Ra}}}=5\times 10^7$ ) at the given ${\textit{Ha}}$ , which displays chaotic dynamics, and then slowly reduce ${\textit{Ra}}$ until we encounter equilibria. The obtained solutions again feature wall-mode structures but indeed have a different symmetry in the vertical velocity field (the same as in McCormack et al. (Reference McCormack, Teimurazov, Shishkina and Linkmann2023)), now invariant under the rotation $\textit{z}\kern-5.5pt{\scriptstyle -} r: (\textit{z}\kern-5.5pt{\scriptstyle -} r) w(x,y,z) = -\mathcal{R}_{\pi /2} w(x,y,-z)$ which features a reflection in sign and vertical coordinate (figure 1 b for the subcritical branch (SB)). This state still exhibits reflection symmetries about the diagonals in the $x{-}y$ plane, but also along the $x$ and $y$ axes, this time accompanied by a reflection in sign and vertical coordinate. Thus, the symmetry group is again isomorphic to $D_4$ but now with the different generators, $r_- = (\textit{z}\kern-5.5pt{\scriptstyle -} r), \mathrm{id}$ , and $s$ , which we will denote $D_4^- = \{\mathrm{id}, r_-, r_-^2, r_-^3, s, sr_-, sr_-^2, sr_-^3\}$ . Importantly, we have found that this solution branch extends beneath the linear stability threshold ${{\textit{Ra}}}_{c,L}$ (figure 1 a) to as low as ${{\textit{Ra}}}=6\times 10^5$ $(\approx 0.87 {{\textit{Ra}}}_c)$ for ${{\textit{Ha}}}=500$ , and ${{\textit{Ra}}}=1.7\times 10^6$ $(\approx 0.93 {{\textit{Ra}}}_c)$ for ${{\textit{Ha}}}=1000$ , and is thus subcritical. These simulations required lengths of approximately 800 or 700 free-fall times (6.5 or 3.4 diffusion times), respectively, to assess stability. Solutions at lower ${\textit{Ra}}$ evolve extremely slowly, and we have not been able to determine whether they are stable. The simplest explanation from the given data is that this SB stems from a saddle-node bifurcation, although we have been unable to determine where the unstable branch from this bifurcation reconnects with the other branches due to computational cost. We note that tests conducted using an under-resolved $60^2\times 80$ grid did not obtain the correct qualitative transition scenario exhibited by the properly resolved simulations we present here, and thus resolving the thin boundary layers on all walls is imperative to assessing the stability of the solutions. Importantly, the SB at both magnetic field strengths extends quite close to the conducting state, suggesting that experimentally small disturbances with the right modal structure could initiate the transition subcritically. Although the magnitude of the subcriticality ( ${{\textit{Ra}}}/{{\textit{Ra}}}_{c,L} \approx 0.9$ in both cases) is quite small, the clear difference in the symmetries of the solutions is promising and should be exploited to measure this phenomenon in experiments. Additionally, the difference between the lowest ${\textit{Ra}}$ subcritical solution found and $Ra_{c,L}$ grows with increased magnetic field strength by a factor of approximately 1.5 from ${{\textit{Ha}}}=500$ to ${{\textit{Ha}}}=1000$ , suggesting that the gap in ${\textit{Ra}}$ between the point of linear instability and the point of the expected saddle-node bifurcation grows with increased magnetic field strength as approximately ${{\textit{Ha}}}^{3/2}$ . Thus, the subcriticality becomes increasingly important at higher ${\textit{Ha}}$ .

3.3. Mixed symmetry branch

A third branch of equilibrium solutions denoted the mixed symmetry branch (MB) has also been identified, first at ${{\textit{Ha}}}=1000$ by trying a number of different initial conditions at parameter values where stable equilibria exist. This branch has then been continued to ${{\textit{Ha}}}=500$ (figure 1 a). The solutions have a symmetry that is mixed between the two other solutions, featuring one/two roll combinations on adjacent walls with matched symmetry on opposing walls (figure 1 b (MB)). The vertical velocity field is invariant under the $180^{\circ }$ rotation $rr_-: (rr_-)w(x,y,z) = -\mathcal{R}_{\pi } w(x,y,-z)$ , a reflection in $y$ , and a reflection in $x$ accompanied by a reflection in sign and vertical coordinate. Thus, the symmetry group is isomorphic to the Klein 4-group $K_4 = \{\mathrm{id}, rr_-, sr^3, sr_-\}$ . These stable solutions are interesting as they suggest that narrow/wide rolls characteristic of the LB/SB, respectively, may coexist. This is relevant to larger-aspect-ratio domains where recent simulations from Wu et al. (Reference Wu, Chen and Ni2025) show near-onset solutions that have various combinations of wide/narrow rolls. This suggests that the results are unlikely to be specific to this geometry or aspect ratio.

3.4. Amplitude equations from symmetry

A model amplitude system which captures the main features of the bifurcation diagram (figure 1 a) may be deduced from symmetries of the system. Associated with each of the three equilibrium solutions found, a secondary solution can be obtained by reversing the direction of the rolls and flipping in the vertical direction $z\mapsto -z$ . As ${\textit{Ra}}$ is increased, the LB first bifurcates from the conducting state through a supercritical pitchfork bifurcation. We then suppose that the MB and SB subsequently bifurcate through subcritical pitchfork bifurcations as ${\textit{Ra}}$ is increased further. Close to onset, when the reduced Rayleigh number $R \sim ({{\textit{Ra}}} - {{\textit{Ra}}}_{c,L})/{{\textit{Ra}}}_{c,L} \ll 1$ , we assume that only these three modes are relevant to the dynamics of the system, which can thus be described by three amplitudes $A_n(t)$ corresponding to the LB, SB and MB ( $n={1,2,3}$ ) equilibria, respectively:

(3.1) \begin{equation} \boldsymbol{\varPsi }(x,y,z,t) = \sum _{n=1}^3 \boldsymbol{\varPsi }_n(x,y,z,t) + \mathrm{s.m.} = \sum _{n=1}^3 A_n(t) \boldsymbol{\psi }_n(x,y,z) + \mathrm{s.m.}, \end{equation}

where $\boldsymbol{\varPsi }$ is the state vector of the system and $\mathrm{s.m.}$ are stable modes which will be neglected. Since for each solution $\boldsymbol{\varPsi}_{\!n}(x,y,z,t)$ there exists a distinct antisymmetric solution $\textit{z}\kern-5.5pt{\scriptstyle -} \boldsymbol{\varPsi }_n(x,y,z,t) = -\boldsymbol{\varPsi }_n(x,y,-z,t)$ , the equation that governs the evolution of the amplitudes must respect this symmetry. Thus, writing a general nonlinear evolution equation for the amplitudes as $\partial _t{\boldsymbol{A}} = \mathcal{N}(\boldsymbol{A})$ , where $\boldsymbol{A} = [A_1, A_2, A_3]^\textrm T$ , we note that $\mathcal{N}$ must be equivariant with respect to the reflection $\varpi \boldsymbol{A} \equiv -\boldsymbol{A}$ , meaning the commutative property $(\varpi \mathcal{N})(\boldsymbol{A}) = (\mathcal{N}\varpi )(\boldsymbol{A)}$ must hold (Crawford & Knobloch Reference Crawford and Knobloch1991). This implies that $\mathcal{N}$ is odd $-\mathcal{N}(\boldsymbol{A}) = \mathcal{N}(-\boldsymbol{A})$ , and thus the evolution equation must have the form $\partial _t{\boldsymbol{A}} = \boldsymbol{\mathcal{F}}(A_k^2,R)\boldsymbol{A}$ , for $k=\{1,2,3\}$ . We note that these equations are also equivariant with respect to the additional symmetries of the different states. The information about these symmetries is contained in the eigenfunctions $\boldsymbol{\psi }_n$ and do not change the amplitudes themselves. We then Taylor expand $\boldsymbol{\mathcal{F}}$ , for $R\ll 1$ , and truncate the system to the lowest order that qualitatively captures the desired dynamics, obtaining

(3.2a) \begin{align} \partial _t{A}_1 &= \big[R - b_1A_2^2 - c_1A_3^2 - d_1A_1^2\big]A_1, &(\textrm{LB}) \end{align}
(3.2b) \begin{align} \partial _t{A}_2 &= \big[(R-a_2) - b_2A_1^2 - c_2A_3^2 + d_2A_2^2 - e_2A_2^4\big]A_2, &(\textrm {SB}) \end{align}
(3.2c) \begin{align} \partial _t{A}_3 &= \big[(R-a_3) - b_3A_1^2 - c_3A_2^2 + d_3A_3^2 - e_3A_3^4\big]A_3, &(\textrm{MB}) \end{align}

where $a_n, b_n, c_n, d_n, e_n\gt 0$ are undetermined real coefficients. The positions of the subcritical pitchfork bifurcations of the SB and MB branches are set by the coefficients $a_2$ and $a_3$ . The saddle-node bifurcations occur at $R_{sn,n} = a_n - d_n^2/4e_n$ at amplitudes $A_n = \pm \sqrt {d_n/2e_n}$ for $n=\{2,3\}$ . The coupling coefficients $b_i, c_i$ determine the bifurcation points, stability and domains of existence of the mixed-mode solutions. For illustration, we have set $a_2=0.5, a_3=0.25, b_1 = 1.5, e_2=b_2=c_n=2, d_2=3, d_3 = 3.5, b_3=6,e_3=8$ . The bifurcation diagram for this choice of parameters is shown in figure 2, exhibiting good qualitative agreement with the bifurcation diagram of the full system in figure 1(a). Thus, although additional complexity may exist in the full system, the behaviour we observe in the DNS can be rationalised by a simplified system, derived from assumptions about the origin of the SB and MB branches, and a consideration of the symmetries of the problem.

Figure 2. Bifurcation diagram of the amplitude (3.2), showing the norm of the amplitudes $\|\boldsymbol{A}\|_2$ as a function of the reduced Rayleigh number $R$ for the various stable/unstable equilibria denoted by solid/dotted lines. Single mode solutions corresponding to the LB, MB and SB states are shown in green, blue and orange, respectively. Mixed mode solutions are shown in grey. Markers show the bifurcation points.

A natural question that arises is whether this behaviour is also expected in a cylindrical geometry, often used in experiments, where the system has a continuous rotational symmetry i.e. $O(2)\times \mathbb{Z}_2$ instead of the discrete $D_4\times \mathbb{Z}_2$ as before. We describe the competition between $N$ unstable modes of azimuthal wavenumber $n$ in cylindrical coordinates by expanding the state vector as

(3.3) \begin{equation} \boldsymbol{\varPsi }(r,\theta ,z,t) = \textrm{Re}\left [\sum _{n=1}^N a_n(t) \textrm{e}^{\textit{in}\theta }\boldsymbol{\psi }_n(r,z)\right ] + \mathrm{s.m.}, \end{equation}

where $\textrm{Re}[\boldsymbol{\cdot }]$ takes the real part of the now complex eigenfunction decomposition. Again, the equation evolving the amplitudes $a_n$ must be equivariant with respect to the symmetries of the system, and thus must commute with rotations $\theta \rightarrow \theta + \theta _0$ , for $\theta_0$ in $\mathbb{R}$ , and the reflection. Thus, $\partial _t{a}_n = \mathcal{F}(|a_k|^2,R)a_n$ , for $k\in \{1,\ldots ,N\}$ . Since the conducting state now bifurcates through a circle pitchfork bifurcation, meaning the solutions are neutrally stable with respect to azimuthal rotations, we write the complex amplitudes in polar form $a_n = A_n \textrm{e}^{\mathrm{i}\varphi _n}$ , noting that due to this neutrality, $\partial _t{\varphi }_n=0$ . Again, Taylor expanding $\mathcal{N}$ , for $R\ll 1$ , we arrive at a system for the real amplitudes $A_n$ which is equivalent to (3.2) for $N$ equations. This means that the continuous symmetry of a cylinder does not preclude the possibility of having the same behaviour we have observed in the cube. Whether the same behaviour occurs as in the cylinder depends on the sign of the coefficients as described before, which must be determined to make qualitative statements about the behaviour of the system. However, we observe that the continuous symmetry of the cylinder does not affect the final structure of the equations we obtain compared with the cube. Furthermore, any finite subgroup is isomorphic to a dihedral $D_n$ or cyclic $\mathbb{Z}_n$ group and thus, the solutions bifurcating from the conducting state through these circle pitchforks will generically have $D_n$ or $\mathbb{Z}_n$ symmetries. This suggests that similar symmetry breaking phenomena observed in the cube, may plausibly occur in the cylinder.

Figure 3. Phase portrait at ${{\textit{Ha}}}=500$ of the (a) invariant 2-torus at ${{\textit{Ra}}}=4\times 10^6$ and (c) chaotic solution at ${{\textit{Ra}}}=10^8$ constructed using time-delay embedding with a time-delay $\tau \approx 4$ . The colour map corresponds to $\|w(t+3\tau )\|_2$ . The unstable equilibrium point (LB) is shown in black. (b,d) Corresponding power spectral density of $\|w(t)\|_2$ , respectively, as a function of the frequency $f$ . Fundamental frequencies are labelled $f_i$ . Zoomed out spectra are shown in the insets.

4. Subcritical transition to turbulence

Although the presence of subcritical solutions is of fundamental interest, the dynamical role that these solutions play in the transition to chaos must be investigated. Thus, we assess the route to chaos for all three branches at ${{\textit{Ha}}}=500$ . As previously mentioned, the equilibria on the LB become unstable in the range $3\times 10^6 \lt {{\textit{Ra}}} \leqslant 4\times 10^6$ , and undergo long transients of approximately 700 free-fall times before ultimately being attracted to solutions on the SB. The MB similarly becomes unstable with increased ${\textit{Ra}}$ . The SB first undergoes a Hopf bifurcation as ${\textit{Ra}}$ is increased in the range $3 \times 10^6 \lt {{\textit{Ra}}} \lt 3.1\times 10^6$ which breaks the $s$ reflection symmetry, producing a limit cycle featuring a periodic flapping of four $r_-$ rotationally symmetric wall-mode extrusions in the horizontal plane. This is equivalent to the secondary bifurcation previously observed at ${{\textit{Ha}}}=1000$ (McCormack et al. Reference McCormack, Teimurazov, Shishkina and Linkmann2023). A secondary Hopf bifurcation (Neimark–Sacker bifurcation on the Poincaré section of the limit cycle) subsequently occurs upon increased ${\textit{Ra}}$ , forming an invariant 2-torus. The invariant 2-torus is characterised by two fundamental frequencies ( $f_1 \approx 0.03542$ and $f_2\approx 0.01028$ at ${{\textit{Ra}}}=4\times 10^6$ ) which are incommensurable (i.e. not rationally related), and thus the solution is not periodic and densely fills $\mathbb{T}^2$ . This property is confirmed by constructing an arbitrary time series using $f_1$ and $f_2$ in a Fourier expansion, the phase portrait of which is seen to become increasingly dense as time is made arbitrarily large. The power density spectrum of the solution at ${{\textit{Ra}}}=4\times 10^6$ is shown in figure 3(b) with additional peaks appearing as harmonics of the fundamental frequencies. The phase portrait of the solution is shown in figure 3(a). The solution continues to feature flapping wall modes, but the additional frequency modulates the protrusions as they flap, moving their tips back and forth from each other, maintaining the $\mathbb{Z}_4=\{\mathrm{id},r_-\}$ symmetry despite the additional wall-mode modulation. Increasing the Rayleigh number to ${{\textit{Ra}}}=5\times 10^6$ leads to a $\mathbb{Z}_4$ to $\mathbb{Z}_2=\{\mathrm{id}, r^2\}$ symmetry break. Here, the solution still features four wall modes flapping back and forth as before, but as the wall-mode dynamics become more vigorous, extending further into the bulk, the solution now has a preferred diagonal where two opposing wall modes dominate over the others. The solution becomes chaotic by ${{\textit{Ra}}}=10^7$ and is accompanied by a large increase in fast time-scale activity, revealing a broadband spectrum. Clear trends are more easily observed at ${{\textit{Ra}}}=10^8$ , whose power density spectrum is shown in figure 3(d) and phase portrait is constructed in figure 3(c). The solution features three incommensurate frequencies and a wide broadband spectrum, which can be compared with the spectrum of the 2-torus at ${{\textit{Ra}}}=4\times 10^6$ in figure 3(b). Thus, the system undergoes a quasiperiodic route to chaos where a finite number of Hopf bifurcations leads to the formation of a chaotic attractor. This is in the spirit of the Ruelle–Takens–Newhouse theorem (Ruelle & Takens Reference Ruelle and Takens1971; Newhouse, Ruelle & Takens Reference Newhouse, Ruelle and Takens1978; Eckmann Reference Eckmann1981), which predicts more generally that three-tori are likely to be structurally unstable. However, more detailed analysis is required to determine precise details of the route in the present system, due to the possibility of phase-locking (tori-periodic orbit transitions) (Arnol’d Reference Arnol’d1961; Reichhardt & Nori Reference Reichhardt and Nori1999), periodic-chaotic-periodic transitions (Knobloch et al. Reference Knobloch, Moore, Toomre and Weiss1986) or other effects. Exploring the behaviour of the system at finer increments of increased nonlinearity ( ${\textit{Ra}}$ ), as has been done in other fluid systems (Knobloch et al. Reference Knobloch, Moore, Toomre and Weiss1986; Oteski et al. Reference Oteski, Duguet, Pastur and Le Quéré2015; Launay et al. Reference Launay, Cambonie, Henry, Pothérat and Botton2019; Wang et al. Reference Wang, Ayats, Deguchi, Meseguer and Mellibovsky2025), may well reveal additional complexity.

At ${{\textit{Ha}}}=1000$ , the wall modes display smaller length scales and are increasingly confined near the sidewall, resulting in a different transition pathway. The SB undergoes a Hopf bifurcation (similar to the ${{\textit{Ha}}}=500$ case), but then displays complex homoclinic behaviour reminiscent of Shilnikov dynamics (McCormack et al. Reference McCormack, Teimurazov, Shishkina and Linkmann2023). Importantly, the transition again occurs on the SB and has no connection to the initial linear instability.

Figure 4. (a) Instantaneous Q-criterion isosurfaces (Q = 3) coloured by the vertical vorticity for the flow at ${{\textit{Ha}}}=500$ , ${{\textit{Ra}}}=10^9$ and (b) the corresponding mean flow ( $w=\pm 0.1$ isosurfaces (pink/blue)). (c) Equilibrium solution on the SB at ${{\textit{Ra}}}=2\times 10^6$ ( $w=\pm 0.01$ isosurfaces).

4.1. Consequences of the subcritical transition

At all values of ${\textit{Ha}}$ studied, the route to chaos originates through a sequence of bifurcations on the SB. Often in supercritical transitions, the equilibrium solution born from the first linear instability in the system imprints the turbulent flow, organising large-scale features (e.g. Taylor vortices in turbulent TCF (Taylor Reference Taylor1923; Grossmann, Lohse & Sun Reference Grossmann, Lohse and Sun2016) or convection rolls in turbulent RBC (Chandrasekhar Reference Chandrasekhar1961; Busse Reference Busse1978; Ahlers, Grossmann & Lohse Reference Ahlers, Grossmann and Lohse2009)). Although the subcriticality is reasonably small in this system at the considered ${\textit{Ha}}$ , it has a significant effect on the flow in the turbulent regime. At ${{\textit{Ra}}}=10^9$ ( ${\approx} 1000\,{{\textit{Ra}}}_{c,L}$ ), the flow field exhibits clear multiscale behaviour with small-scale vortices being observed in an instantaneous snapshot of the flow in figure 4(a). However, unlike the supercritical cases, this turbulent flow has been formed through bifurcations on the SB and thus the turbulent flow retains properties of the subcritical equilibria and not the linear instability. This is clearly observed in the mean flow of the solution which retains the same $D_4^-$ symmetry as the subcritical equilibria, and not the $D_4^+$ symmetry of the solutions born from the linear instability. This means that understanding the subcritical nature of the transition is imperative to understanding properties of turbulent MC, such as the globally organising flow structures.

5. Conclusions

The addition of insulating sidewalls changes the phenomenology of the transition to turbulence in liquid-metal MC, with stable solutions existing beneath the linear stability threshold. This occurs despite the conducting state undergoing a supercritical bifurcation, and the transition proceeding supercritically in a laterally periodic layer. Although the magnitude of the subcriticality is relatively small at the studied magnetic field strengths, our results suggest the transition becomes increasingly subcritical at higher magnetic field strengths, and is thus centrally important in high- ${\textit{Ha}}$ regimes currently inaccessible to simulations and experiments. Multiple stable equilibrium solutions exist near onset, which are distinguished by their spatial symmetries and have different transport properties. The qualitative features of the near onset behaviour can be rationalised by low-dimensional amplitude equations, deduced from symmetries of the system. We have shown that the solutions born from the LB and the MB quickly become unstable, and solutions are attracted to the SB, which is the primary branch involved in the route to chaos. Thus, we show that both the transition and the formation of global flow structures in turbulent regimes can be understood only through the subcriticality, and have no connection to the linear instability, despite linear stability theory predicting a supercritical transition. This must be considered when laboratory experiments with sidewalls are compared with laterally unbounded geophysical or astrophysical flows. Since wall modes are observed in a variety of container geometries and with finite wall conductivity, both numerically and experimentally, we anticipate the subcriticality is insensitive to these properties. Wall modes are also well known to form in rotating convection with sidewalls (Goldstein et al. Reference Goldstein, Knobloch, Mercader and Net1993; Herrmann & Busse Reference Herrmann and Busse1993; Kuo & Cross Reference Kuo and Cross1993; Ning & Ecke Reference Ning and Ecke1993; Goldstein et al. Reference Goldstein, Knobloch, Mercader and Net1994; Liu & Ecke Reference Liu and Ecke1999; Liao, Zhang & Chang Reference Liao, Zhang and Chang2006; Kunnen et al. Reference Kunnen, Stevens, Overkamp, Sun, van Heijst and Clercx2011; Horn & Schmid Reference Horn and Schmid2017; Favier & Knobloch Reference Favier and Knobloch2020; de Wit et al. Reference de Wit, Aguirre Guzmán, Madonia, Cheng, Clercx and Kunnen2020; Zhang, Ecke & Shishkina Reference Zhang, Ecke and Shishkina2021; Wedi et al. Reference Wedi, Moturi, Funfschilling and Weiss2022; de Wit et al. Reference de Wit, Boot, Madonia, Aguirre Guzmán and Kunnen2023; Ravichandran & Wettlaufer Reference Ravichandran and Wettlaufer2024; Zhang et al. Reference Zhang, Reiter, Shishkina and Ecke2024; Vasil et al. Reference Vasil, Burns, Lecoanet, Oishi, Brown and Julien2025), and are robust to varied sidewall geometry (Favier & Knobloch Reference Favier and Knobloch2020). Furthermore, wall-localised modes are known to form in a variety of simple pattern-forming systems on bounded domains, such as the cubic-quintic Swift–Hohenberg equation, which are expected to have wide applicability (Verschueren, Knobloch & Uecker Reference Verschueren, Knobloch and Uecker2021). Thus, one might anticipate that similar results to those obtained here may extend to a wider class of systems, such as rotating convection or rotating MC. Further open questions in this direction relate to obtaining a detailed understanding of these effects for domains with continuous symmetries (e.g. for MC in a cylinder), and for larger aspect ratio domains.

Acknowledgements

We thank A. Morozov and G. Vasil for helpful discussions. M.L. thanks the Isaac Newton Institute for Mathematical Sciences, Cambridge, for support and hospitality during the programme ‘Anti-diffusive dynamics: from sub-cellular to astrophysical scales’ (EPSRC grant EP/R014604/1), where work on this paper was undertaken.

Funding

This work was supported by the Deutsche Forschungsgemeinschaft (SPP1881 ‘Turbulent Superstructures’ and grants Sh405/20, Sh405/22, Li3694/1) and used the ARCHER2 UK National Supercomputing Service (https://www.archer2.ac.uk) with resources provided by the UK Turbulence Consortium (EPSRC grants EP/R029326/1, EP/X035484/1).

Declaration of interests

The authors report no conflict of interest.

References

Ahlers, G., Grossmann, S. & Lohse, D. 2009 Heat transfer and large scale dynamics in turbulent Rayleigh–Bénard convection. Rev. Mod. Phys. 81 (2), 503.10.1103/RevModPhys.81.503CrossRefGoogle Scholar
Akhmedagaev, R., Zikanov, O., Krasnov, D. & Schumacher, J. 2020 Turbulent Rayleigh–Bénard convection in a strong vertical magnetic field. J. Fluid Mech. 895, R4.10.1017/jfm.2020.336CrossRefGoogle Scholar
Arnol’d, V.I. 1961 Small denominators. I. Mapping the circle onto itself. Izv. Akad. Nauk SSSR Ser. Mat. 25, 2186.Google Scholar
Avila, M., Barkley, D. & Hof, B. 2023 Transition to turbulence in pipe flow. Annu. Rev. Fluid Mech. 55, 575602.10.1146/annurev-fluid-120720-025957CrossRefGoogle Scholar
Bhattacharya, S., Boeck, T., Krasnov, D. & Schumacher, J. 2024 Wall-attached convection under strong inclined magnetic fields. J. Fluid Mech. 979, A53.10.1017/jfm.2023.1087CrossRefGoogle Scholar
Bodenschatz, E., Pesch, W. & Ahlers, G. 2000 Recent developments in Rayleigh–Bénard convection. Annu. Rev. Fluid Mech. 32, 709778.10.1146/annurev.fluid.32.1.709CrossRefGoogle Scholar
Bonn, D., Ingremeau, F., Amarouchene, Y. & Kellay, H. 2011 Large velocity fluctuations in small-Reynolds-number pipe flow of polymer solutions. Phys. Rev. E 84 (4), 2.10.1103/PhysRevE.84.045301CrossRefGoogle ScholarPubMed
Brandstater, A. & Swinney, H. 1987 Strange attractors in weakly turbulent Couette–Taylor flow. Phys. Rev. A 35 (5), 2207.10.1103/PhysRevA.35.2207CrossRefGoogle ScholarPubMed
Busse, F.H. 1978 Non-linear properties of thermal convection. Rep. Prog. Phys. 41 (12), 19291967.10.1088/0034-4885/41/12/003CrossRefGoogle Scholar
Busse, F.H. 2008 Asymptotic theory of wall-attached convection in a horizontal fluid layer with a vertical magnetic field. Phys. Fluids 20 (2), 024102.10.1063/1.2837175CrossRefGoogle Scholar
Busse, F.H. & Clever, R.M. 1982 Stability of convection rolls in the presence of a vertical magnetic field. Phys. Fluids 25 (6), 931935.10.1063/1.863845CrossRefGoogle Scholar
Chandrasekhar, S. 1961 Hydrodynamic and Hydromagnetic Stability. Oxford University Press.Google Scholar
Chossat, P. & Iooss, G. 1994 The couette–Taylor problem. In Applied Mathematical Sciences, vol. 102, Springer-Verlag.Google Scholar
Crawford, J.D. & Knobloch, E. 1991 Symmetry and symmetry-breaking bifurcations in fluid dynamics. Annu. Rev. Fluid Mech. 23 (1), 341387.10.1146/annurev.fl.23.010191.002013CrossRefGoogle Scholar
Davidson, P.A. 1999 Magnetohydrodynamics in materials processing. Annu. Rev. Fluid Mech. 31, 273300.10.1146/annurev.fluid.31.1.273CrossRefGoogle Scholar
Ecke, R.E., Mainieri, R. & Sullivan, T.S. 1991 Universality in quasiperiodic Rayleigh–Bénard convection. Phys. Rev. A 44, 81038118.10.1103/PhysRevA.44.8103CrossRefGoogle ScholarPubMed
Eckhardt, B. 2018 Transition to turbulence in shear flows. Phys. A 504, 121129.10.1016/j.physa.2018.01.032CrossRefGoogle Scholar
Eckhardt, B., Schneider, T.M., Hof, B. & Westerweel, J. 2007 Turbulence transition in pipe flow. Annu. Rev. Fluid Mech. 39, 447.10.1146/annurev.fluid.39.050905.110308CrossRefGoogle Scholar
Eckmann, J.-P. 1981 Roads to turbulence in dissipative dynamical systems. Rev. Mod. Phys. 53 (4), 643.10.1103/RevModPhys.53.643CrossRefGoogle Scholar
Favier, B., Guervilly, C. & Knobloch, E. 2019 Subcritical turbulent condensate in rapidly rotating Rayleigh–Bénard convection. J. Fluid Mech. 864, R1.10.1017/jfm.2019.58CrossRefGoogle Scholar
Favier, B. & Knobloch, E. 2020 Robust wall states in rapidly rotating Rayleigh–Bénard convection. J. Fluid Mech. 895, R1.10.1017/jfm.2020.310CrossRefGoogle Scholar
Gallet, B. 2024 Two-dimensional turbulence above topography: condensation transition and selection of minimum enstrophy solutions. J. Fluid Mech. 988, A13.10.1017/jfm.2024.365CrossRefGoogle Scholar
Goldstein, H.F., Knobloch, E., Mercader, I. & Net, M. 1993 Convection in a rotating cylinder. Part 1 Linear theory for moderate Prandtl numbers. J. Fluid Mech. 248, 583604.10.1017/S0022112093000928CrossRefGoogle Scholar
Goldstein, H.F., Knobloch, E., Mercader, I. & Net, M. 1994 Convection in a rotating cylinder. Part 2. Linear theory for low Prandtl numbers. J. Fluid Mech. 262, 293324.10.1017/S0022112094000510CrossRefGoogle Scholar
Gollub, J.P. & Swinney, H.L. 1975 Onset of turbulence in a rotating fluid. Phys. Rev. Lett. 35 (14), 927.10.1103/PhysRevLett.35.927CrossRefGoogle Scholar
Grossmann, S., Lohse, D. & Sun, C. 2016 High–Reynolds number Taylor–Couette turbulence. Annu. Rev. Fluid Mech. 48 (1), 5380.10.1146/annurev-fluid-122414-034353CrossRefGoogle Scholar
Herrmann, J. & Busse, F.H. 1993 Asymptotic theory of wall-attached convection in a rotating fluid layer. J. Fluid Mech. 255, 183194.10.1017/S0022112093002447CrossRefGoogle Scholar
Hof, B. 2023 Directed percolation and the transition to turbulence. Nat. Rev. Phys. 5, 6272.10.1038/s42254-022-00539-yCrossRefGoogle Scholar
Horn, S. & Schmid, P.J. 2017 Prograde, retrograde, and oscillatory modes in rotating rayleigh–Bénard convection. J. Fluid Mech. 831, 182211.10.1017/jfm.2017.631CrossRefGoogle Scholar
Houchens, B.C., Witkowski, L.M. & Walker, J.S. 2002 Rayleigh–Bénard instability in a vertical cylinder with a vertical magnetic field. J. Fluid Mech. 469, 189207.10.1017/S0022112002001623CrossRefGoogle Scholar
Jones, C.A. 2011 Planetary magnetic fields and fluid dynamos. Annu. Rev. Fluid Mech. 43, 583614.10.1146/annurev-fluid-122109-160727CrossRefGoogle Scholar
Kerswell, R.R. 2005 Recent progress in understanding the transition to turbulence. Nonlinearity 18, R17R44.10.1088/0951-7715/18/6/R01CrossRefGoogle Scholar
Knobloch, E., Moore, D.R., Toomre, J. & Weiss, N.O. 1986 Transitions to chaos in two-dimensional double-diffusive convection. J. Fluid Mech. 166, 409448.10.1017/S0022112086000216CrossRefGoogle Scholar
Kooij, G.L., Botchev, M.A., Frederix, E.M.A., Geurts, B.J., Horn, S., Lohse, D., van der Poel, E.P., Shishkina, O., Stevens, R.J.A.M. & Verzicco, R. 2018 Comparison of computational codes for direct numerical simulations of turbulent Rayleigh–Bénard convection. Comput. Fluids 166, 18.10.1016/j.compfluid.2018.01.010CrossRefGoogle Scholar
Kunnen, R., Stevens, R., Overkamp, J., Sun, C., van Heijst, G. & Clercx, H. 2011 The role of Stewartson and Ekman layers in turbulent rotating Rayleigh–Bénard convection. J. Fluid Mech. 688, 422442.10.1017/jfm.2011.383CrossRefGoogle Scholar
Kuo, E.Y. & Cross, M.C. 1993 Traveling-wave wall states in rotating Rayleigh–Bénard convection. Phys. Rev. E 47 (4), R2245.10.1103/PhysRevE.47.R2245CrossRefGoogle ScholarPubMed
Launay, G., Cambonie, T., Henry, D., Pothérat, A. & Botton, V. 2019 Transition to chaos in an acoustically driven cavity flow. Phys. Rev. Fluids 4, 044401.10.1103/PhysRevFluids.4.044401CrossRefGoogle Scholar
Lellep, M., Linkmann, M. & Morozov, A. 2024 Purely elastic turbulence in pressure-driven channel flows. Proc. Natl Acad. Sci. 121, e2318851121.10.1073/pnas.2318851121CrossRefGoogle ScholarPubMed
Liao, X., Zhang, K. & Chang, Y. 2006 On boundary-layer convection in a rotating fluid layer. J. Fluid Mech. 549, 375384.10.1017/S0022112005008189CrossRefGoogle Scholar
Libchaber, A., Laroche, C. & Fauve, S. 1982 Period doubling cascade in mercury, a quantitative measurement. J. Physique Lett. 43 (7), 211216.10.1051/jphyslet:01982004307021100CrossRefGoogle Scholar
Linkmann, M., Boffetta, G., Marchetti, M.C. & Eckhardt, B. 2019 Phase transition to large scale coherent structures in two-dimensional active matter turbulence. Phys. Rev. Lett. 122, 214503.10.1103/PhysRevLett.122.214503CrossRefGoogle ScholarPubMed
Linkmann, M., Hohmann, M. & Eckhardt, B. 2020 a Non-universal transitions to two-dimensional turbulence. J. Fluid Mech. 892, A18.10.1017/jfm.2020.198CrossRefGoogle Scholar
Linkmann, M., Marchetti, M.C., Boffetta, G. & Eckhardt, B. 2020 b Condensate formation and multiscale dynamics in two-dimensional active suspensions. Phys. Rev. E 101, 022609.10.1103/PhysRevE.101.022609CrossRefGoogle ScholarPubMed
Linkmann, M.F. & Morozov, A. 2015 Sudden relaminarization and lifetimes in forced isotropic turbulence. Phys. Rev. Lett. 115, 134502.10.1103/PhysRevLett.115.134502CrossRefGoogle ScholarPubMed
Liu, W., Krasnov, D. & Schumacher, J. 2018 Wall modes in magnetoconvection at high Hartmann numbers. J. Fluid Mech. 849, R21R212.10.1017/jfm.2018.479CrossRefGoogle Scholar
Liu, Y. & Ecke, R.E. 1999 Nonlinear traveling waves in rotating Rayleigh–Bénard convection: stability boundaries and phase diffusion. Phys. Rev. E 59 (4), 4091.10.1103/PhysRevE.59.4091CrossRefGoogle Scholar
Lohse, D. & Shishkina, O. 2023 Ultimate turbulent thermal convection. Phys. Today 76 (11), 2632.10.1063/PT.3.5341CrossRefGoogle Scholar
Lohse, D. & Shishkina, O. 2024 Ultimate Rayleigh–Bénard turbulence. Rev. Mod. Phys. 96 (3), 035001.10.1103/RevModPhys.96.035001CrossRefGoogle Scholar
Malkus, W.V.R. & Veronis, G. 1958 Finite amplitude cellular convection. J. Fluid Mech. 4 (3), 225260.10.1017/S0022112058000410CrossRefGoogle Scholar
McCormack, M., Teimurazov, A., Shishkina, O. & Linkmann, M. 2023 Wall mode dynamics and transition to chaos in magnetoconvection with a vertical magnetic field. J. Fluid Mech. 975, R2.10.1017/jfm.2023.863CrossRefGoogle Scholar
Morozov, A.N. & van Saarloos, W. 2007 An introductory essay on subcritical instabilities and the transition to turbulence in visco-elastic parallel shear flows. Phys. Rep. 447, 112143.10.1016/j.physrep.2007.03.004CrossRefGoogle Scholar
Newhouse, S., Ruelle, D. & Takens, F. 1978 Occurrence of strange axiom a attractors near quasi periodic flows on t m, m $\geqslant$ 3. Comm. Math. Phys. 64, 3540.10.1007/BF01940759CrossRefGoogle Scholar
Borońska, K. & Tuckerman, L.S. 2010 Extreme multiplicity in cylindrical Rayleigh–Bénard convection. II. Bifurcation diagram and symmetry classification. Phys. Rev. E 81, 036321.10.1103/PhysRevE.81.036321CrossRefGoogle ScholarPubMed
Ning, L. & Ecke, R.E. 1993 Rotating Rayleigh–Bénard convection: aspect-ratio dependence of the initial bifurcations. Phys. Rev. E 47 (5), 3326.10.1103/PhysRevE.47.3326CrossRefGoogle ScholarPubMed
Oteski, L., Duguet, Y., Pastur, L. & Le Quéré, P. 2015 Quasiperiodic routes to chaos in confined two-dimensional differential convection. Phys. Rev. E 92 (4), 043020.10.1103/PhysRevE.92.043020CrossRefGoogle ScholarPubMed
Pan, L., Morozov, A., Wagner, C. & Arratia, P.E. 2013 Nonlinear elastic instability in channel flows at low Reynolds numbers. Phys. Rev. Lett. 110, 174502.10.1103/PhysRevLett.110.174502CrossRefGoogle ScholarPubMed
Proctor, M.R.E. & Weiss, N.O. 1982 Magnetoconvection. Rep. Prog. Phys. 45 (11), 1317.10.1088/0034-4885/45/11/003CrossRefGoogle Scholar
Ravichandran, S. & Wettlaufer, J.S. 2024 Prograde and meandering wall modes in rotating Rayleigh–Bénard convection with conducting walls. J. Fluid Mech. 998, A47.10.1017/jfm.2024.901CrossRefGoogle Scholar
Rayleigh, Lord & LIX 1916 On convection currents in a horizontal layer of fluid, when the higher temperature is on the under side. Lond. Edinb. Dublin Philos. Mag. J. Sci. 32 (192), 529546.10.1080/14786441608635602CrossRefGoogle Scholar
Reichhardt, C. & Nori, F. 1999 Phase locking, devil’s staircases, farey trees, and arnold tongues in driven vortex lattices with periodic pinning. Phys. Rev. Lett. 82, 414417.10.1103/PhysRevLett.82.414CrossRefGoogle Scholar
Reiter, P., Shishkina, O., Lohse, D. & Krug, D. 2021 Crossover of the relative heat transport contributions of plume ejecting and impacting zones in turbulent Rayleigh–Bénard convection (a). Europhys. Lett. 134 (3), 34002.10.1209/0295-5075/134/34002CrossRefGoogle Scholar
Reiter, P., Zhang, X. & Shishkina, O. 2022 Flow states and heat transport in Rayleigh–Bénard convection with different sidewall boundary conditions. J. Fluid Mech. 936, A32.10.1017/jfm.2022.56CrossRefGoogle Scholar
Roche, P.E. 2020 The ultimate state of convection: a unifying picture of very high Rayleigh numbers experiments. New J. Phys. 22 (7), 073056.10.1088/1367-2630/ab9449CrossRefGoogle Scholar
Ruelle, D. & Takens, F. 1971 On the nature of turbulence. Commun. Math. Phys. 20, 167192.10.1007/BF01646553CrossRefGoogle Scholar
Shishkina, O. & Lohse, D. 2024 Ultimate regime of Rayleigh–Bénard turbulence: subregimes and their scaling relations for the Nusselt vs Rayleigh and Prandtl numbers. Phys. Rev. Lett. 133, 144001.10.1103/PhysRevLett.133.144001CrossRefGoogle ScholarPubMed
Stefani, F. 2024 Liquid-metal experiments on geophysical and astrophysical phenomena. Nat. Rev. Phys. 6 (7), 409425.10.1038/s42254-024-00724-1CrossRefGoogle Scholar
Taylor, G.I. 1923 VIII. Stability of a viscous liquid contained between two rotating cylinders. Philos. Trans. R. Soc. A 223 (605-615), 289343.Google Scholar
Teimurazov, A., McCormack, M., Linkmann, M. & Shishkina, O. 2024 Unifying heat transport model for the transition between buoyancy-dominated and Lorentz-force-dominated regimes in quasistatic magnetoconvection. J. Fluid Mech. 980, R3.10.1017/jfm.2024.33CrossRefGoogle Scholar
Vasil, G.M., Burns, K.J., Lecoanet, D., Oishi, J.S., Brown, B. & Julien, K. 2025 Rapidly rotating wall-mode convection. J. Fluid Mech. 1017, A37.10.1017/jfm.2025.10449CrossRefGoogle Scholar
Verschueren, N., Knobloch, E. & Uecker, H. 2021 Localized and extended patterns in the cubic-quintic Swift-Hohenberg equation on a disk. Phys. Rev. E 104, 014208.10.1103/PhysRevE.104.014208CrossRefGoogle ScholarPubMed
Wang, B., Ayats, R., Deguchi, K., Meseguer, A. & Mellibovsky, F. 2025 Feigenbaum universality in subcritical Taylor–Couette flow. J. Fluid Mech. 1010, A36.10.1017/jfm.2025.278CrossRefGoogle Scholar
Wedi, M., Moturi, V.M., Funfschilling, D. & Weiss, S. 2022 Experimental evidence for the boundary zonal flow in rotating Rayleigh–Bénard convection. J. Fluid Mech. 939, A14.10.1017/jfm.2022.195CrossRefGoogle Scholar
Weiss, N.O. & Proctor, M.R.E. 2014 Magnetoconvection. Cambridge University Press.10.1017/CBO9780511667459CrossRefGoogle Scholar
de Wit, X.M., Aguirre Guzmán, A.J., Madonia, M., Cheng, J.S., Clercx, H.J.H. & Kunnen, R.P.J. 2020 Turbulent rotating convection confined in a slender cylinder: the sidewall circulation. Phys. Rev. Fluids 5 (2), 023502.10.1103/PhysRevFluids.5.023502CrossRefGoogle Scholar
de Wit, X.M., Aguirre Guzmán, Aés J., Clercx, H.J.H. & Kunnen, R.P.J. 2022 a Discontinuous transitions towards vortex condensates in buoyancy-driven rotating turbulence. J. Fluid Mech. 936, A43.10.1017/jfm.2022.90CrossRefGoogle Scholar
de Wit, X.M., Boot, W.J.M., Madonia, M., Aguirre Guzmán, A.J. & Kunnen, R.P.J. 2023 Robust wall modes and their interplay with bulk turbulence in confined rotating Rayleigh–Bénard convection. Phys. Rev. Fluids 8 (7), 073501.10.1103/PhysRevFluids.8.073501CrossRefGoogle Scholar
de Wit, X.M., van Kan, A. & Alexakis, A. 2022 b Bistability of the large-scale dynamics in quasi-two-dimensional turbulence. J. Fluid Mech. 939, R2.10.1017/jfm.2022.209CrossRefGoogle Scholar
Wu, K., Chen, L. & Ni, M.-J. 2025 Flow and heat transfer mechanism of wall mode in Rayleigh–Bénard convection under strong magnetic fields. Phys. Rev. Fluids 10 (3), 033702.10.1103/PhysRevFluids.10.033702CrossRefGoogle Scholar
Xu, Y., Horn, S. & Aurnou, J.M. 2023 Transition from wall modes to multimodality in liquid gallium magnetoconvection. Phys. Rev. Fluids 8, 103503.10.1103/PhysRevFluids.8.103503CrossRefGoogle Scholar
Yokoyama, N. & Takaoka, M. 2017 Hysteretic transitions between quasi-two-dimensional flow and three-dimensional flow in forced rotating turbulence. Phys. Rev. Fluids 2, 092602.10.1103/PhysRevFluids.2.092602CrossRefGoogle Scholar
Zhang, X., Ecke, R.E. & Shishkina, O. 2021 Boundary zonal flows in rapidly rotating turbulent thermal convection. J. Fluid Mech. 915, A62.10.1017/jfm.2021.74CrossRefGoogle Scholar
Zhang, X., Reiter, P., Shishkina, O. & Ecke, R.E. 2024 Wall modes and the transition to bulk convection in rotating Rayleigh–Bénard convection. Phys. Rev. Fluids 9 (5), 053501.10.1103/PhysRevFluids.9.053501CrossRefGoogle Scholar
Zikanov, O., Krasnov, D., Boeck, T., Thess, A. & Rossi, M. 2014 Laminar-turbulent transition in magnetohydrodynamic duct, pipe, and channel flows. Appl. Mech. Rev. 66, 030802.10.1115/1.4027198CrossRefGoogle Scholar
Zürner, T., Schindler, F., Vogt, T., Eckert, S. & Schumacher, J. 2020 Flow regimes of Rayleigh–Bénard convection in a vertical magnetic field. J. Fluid Mech. 894, A21.10.1017/jfm.2020.264CrossRefGoogle Scholar
Figure 0

Figure 1. (a) Bifurcation diagram at ${{\textit{Ha}}}=500$ showing the $L^2$ norm of the vertical velocity field $\|w\|_2$ of stable/unstable equilibria (filled/open markers), limit cycles (diamonds), invariant tori (asterisks) on the linear onset branch (LB) stemming from the linear instability at ${{\textit{Ra}}}_{c,L}$, the mixed symmetry branch (MB) and the subcritical branch (SB). The mean of time-dependent solutions is shown. The grey area represents ${{\textit{Ra}}}\lt {{\textit{Ra}}}_{c,L}$. (b) Upflow (pink) and downflow (blue) vertical velocity isosurfaces and streamlines (black) of solutions on the various branches shown from the top view at ${{\textit{Ha}}}=500$ at ${{\textit{Ra}}}=8\times 10^5$ and ${{\textit{Ra}}}=2\times 10^6$ with $w=\pm 0.005$ and $\pm 0.01$, respectively.

Figure 1

Figure 2. Bifurcation diagram of the amplitude (3.2), showing the norm of the amplitudes $\|\boldsymbol{A}\|_2$ as a function of the reduced Rayleigh number $R$ for the various stable/unstable equilibria denoted by solid/dotted lines. Single mode solutions corresponding to the LB, MB and SB states are shown in green, blue and orange, respectively. Mixed mode solutions are shown in grey. Markers show the bifurcation points.

Figure 2

Figure 3. Phase portrait at ${{\textit{Ha}}}=500$ of the (a) invariant 2-torus at ${{\textit{Ra}}}=4\times 10^6$ and (c) chaotic solution at ${{\textit{Ra}}}=10^8$ constructed using time-delay embedding with a time-delay $\tau \approx 4$. The colour map corresponds to $\|w(t+3\tau )\|_2$. The unstable equilibrium point (LB) is shown in black. (b,d) Corresponding power spectral density of $\|w(t)\|_2$, respectively, as a function of the frequency $f$. Fundamental frequencies are labelled $f_i$. Zoomed out spectra are shown in the insets.

Figure 3

Figure 4. (a) Instantaneous Q-criterion isosurfaces (Q = 3) coloured by the vertical vorticity for the flow at ${{\textit{Ha}}}=500$, ${{\textit{Ra}}}=10^9$ and (b) the corresponding mean flow ($w=\pm 0.1$ isosurfaces (pink/blue)). (c) Equilibrium solution on the SB at ${{\textit{Ra}}}=2\times 10^6$ ($w=\pm 0.01$ isosurfaces).