1. Introduction
Particle-laden flows are ubiquitous in nature and industries, such as in the dispersion of water droplets in atmospheric clouds (Shaw Reference Shaw2003; Abade, Grabowski & Pawlowska Reference Abade, Grabowski and Pawlowska2018), aerosols and dust particles in the atmosphere (Chandrakar et al. Reference Chandrakar, Saito, Yang, Cantrell, Gotoh and Shaw2020), crystals settling in the Earth’s magma chambers (Martin & Nokes Reference Martin and Nokes1988; Koyaguchi et al. Reference Koyaguchi, Hallworth, Huppert and Stephen J. Sparks1990; Molina, Burgisser & Oppenheimer Reference Molina, Burgisser and Oppenheimer2015), mixing fuel droplets in combustion chambers (Huang et al. Reference Huang, Zhao, Xu, Li and Zhang2021; Cao et al. Reference Cao, Liu, Wang and Weng2024), spray coating (Pendar & Páscoa Reference Pendar and Páscoa2021), ejection of solid particles and ash due to volcanic eruption (Schwaiger, Denlinger & Mastin Reference Schwaiger, Denlinger and Mastin2012; Yarushina, Bercovici & Michaut Reference Yarushina, Bercovici and Michaut2015), dispersion of pollen grains (Helbig et al. Reference Helbig, Vogel, Vogel and Fiedler2004; Robichaud & Comtois Reference Robichaud and Comtois2021), and dynamics of phytoplankton in ocean waters (Squires & Yamazaki Reference Squires and Yamazaki1995; Ruiz, Macías & Peters Reference Ruiz, Macías and Peters2004). Due to this broad spectrum of applications, particle-laden flows are of great interest to fluid dynamicists and geophysicists. Hence to get a comprehensive understanding of these flows, various experimental and numerical techniques have been proposed in the literature (Kiger & Lasheras Reference Kiger and Lasheras1997; Hwang & Eaton Reference Hwang and Eaton2006; Zhong, Funfschilling & Ahlers Reference Zhong, Funfschilling and Ahlers2009; Brandt & Coletti Reference Brandt and Coletti2022; Srinivas & Tomar Reference Srinivas and Tomar2025).
Microphysical mechanisms of rain initiation involve three stages, namely: activation of droplets on cloud condensation nuclei (size of the order of
$0.05\,\unicode{x03BC}\mbox{m}$
), increase in the droplet size by condensation (reaches a size of the order of
$10\,\unicode{x03BC}\mbox{m}$
) and by coalescence (Beard & Ochs Reference Beard and Ochs1993). Generally, droplets of radius larger than
$50\,\unicode{x03BC}\mbox{m}$
settle under gravity and continue to grow via collisions with droplets in their path to reach a raindrop size of the order of
$80{-}100\,\unicode{x03BC}\mbox{m}$
. It has been shown that theoretically, droplets should take times of the order of hours to reach
$50\,\unicode{x03BC}\mbox{m}$
in size, despite the actual observations showing merely 15–20 min. Using tangling clustering instability proposed by Elperin et al. (Reference Elperin, Kleeorin, Liberman and Rogachevskii2013) for thermally stratified flow, Elperin et al. (Reference Elperin, Kleeorin, Krasovitov, Kulmala, Liberman, Rogachevskii and Zilitinkevich2015) provided the explanation for rapid droplet growth from
$1\,\unicode{x03BC}\mbox{m}$
to
$40{-}50\,\unicode{x03BC}\mbox{m}$
within 15–20 min. Recently, Poydenot & Andreotti (Reference Poydenot and Andreotti2024) showed the formation of raindrops through the attractive van der Waals and electrostatic interactions. However, the formation of local particle clusters from their initially uniform distribution in a thermally stratified flow is not well understood. In the present study, we use Rayleigh–Bénard convection as the model for the thermally stratified flow. It should be mentioned that in this work, we only take into account the momentum and thermal interaction among the fluid and the particulate phases, ignoring phase change phenomena such as condensation and evaporation. Therefore, only qualitative inferences for droplet dynamics in clouds can be made from the present study. Nevertheless, the analysis presented in this work is useful for understanding atmospheric phenomena like dust storms and aerosol dispersion.
The effect of settling particles on a weakly turbulent Rayleigh–Bénard convection has been studied by Oresta & Prosperetti (Reference Oresta and Prosperetti2013). They showed that the mechanical coupling between the particles and fluid increases the Nusselt number with increasing particle size. However, thermal coupling between the fluid and particles tends to make the fluid temperature uniform and thus reduces the strength of the convection of the underlying fluid flow. They also reported an unusual kind of reverse one-way coupling in the sense that the underlying flow was affected more significantly than the settling particles. Later, Oresta, Fornarelli & Prosperetti (Reference Oresta, Fornarelli and Prosperetti2014) reviewed the mathematical formulations and the numerical methods for the particle and bubble-laden Rayleigh–Bénard convection. To understand the settling of dense crystals in magma chambers and planetary scale magma oceans, Patočka et al. (Reference Patočka, Calzavarini and Tosi2020) conducted a numerical study on the rate of settling of particles in a rectangular two-dimensional Rayleigh–Bénard convection with Rayleigh number up to
$10^{12}$
and Prandtl number ranging from
$10$
to
$50$
. Four distinct settling regimes, namely, stone-like, bi-linear, transitional and dust-like regimes, have been observed based on the ratio of particle terminal speed and flow root mean square velocity. Similarly, Srinivas & Tomar (Reference Srinivas and Tomar2024) analysed the effect of particle size on the particle cloud patterns in Rayleigh–Bénard convection. However, in these studies, the coupling between the fluid and the particles is one-way, where the fluid flow field affects the particle trajectories while the effect of particles on the underlying flow is ignored. Recently, Denzel, Bragg & Richter (Reference Denzel, Bragg and Richter2023) developed a stochastic model to predict the residence times of the particles in a turbulent Rayleigh–Bénard convection using Euler–Lagrange formulation.
A linear stability analysis of a fluid confined between more realistic rigid surfaces and heated from below was performed by Sir Harold Jeffreys (Jeffreys Reference Jeffreys1928). The critical Rayleigh number
${\textit{Ra}}_c$
and critical wavenumber
$k_c$
were obtained as 1708 and 3.117, respectively. However, we expect, based on various numerical and experimental studies (Hetsroni & Sokolov Reference Hetsroni and Sokolov1971; Sun & Faeth Reference Sun and Faeth1986; Gore & Crowe Reference Gore and Crowe1989; Lance, Marie & Bataille Reference Lance, Marie and Bataille1991), that the collective dynamics of small particles or bubbles can affect the underlying flow significantly. However, most of the studies have been performed for isothermal systems (Balachandar & Eaton Reference Balachandar and Eaton2010; Kuerten Reference Kuerten2016; Maxey Reference Maxey2017). The effect of the suspended particles on the convective heat transfer, particularly Rayleigh–Bénard convection, is gaining importance in recent times and is the focus of the present work. Prakhar & Prosperetti (Reference Prakhar and Prosperetti2021) formulated an Euler–Euler or two-fluid model to study the effect of highly dense point particles on the stability of Rayleigh–Bénard convection in a horizontally unbounded cell. They reported that the addition of particles into the flow increases the underlying dimensionless parameter space and stabilises the flow significantly. Recently, Raza, Hirata & Calzavarini (Reference Raza, Hirata and Calzavarini2024) extended this stability analysis to the suspension of bubbles in Rayleigh–Bénard convection modelled as particles by including additional forces such as added mass in the particle momentum balance equations.
In the linear analysis, we initially neglect nonlinear perturbation terms that are very small in magnitude. However, as soon as time
$t=\mathcal{O} (1/c_i )$
, they become
$\mathcal{O} (1 )$
in magnitude and cannot be neglected. Here,
$c_i$
is the growth rate. In other words, as the Rayleigh number is
${\textit{Ra}} = {g \beta\, \Delta {\textit{TH}}^3}/({\nu _{\!f} \alpha _{\!f}}) \gt {\textit{Ra}}_c$
, the initial infinitesimally small perturbations grow exponentially and reach a magnitude that can affect the mean flow, which makes the linear stability predictions unreliable. Here,
$g$
is the gravitational acceleration,
$\beta$
is the thermal expansion coefficient,
$H$
is the gap between the plates,
$\Delta T$
is the difference in the temperature of the bottom hot plate and the top cold plate,
$\nu _{\!f}$
is the kinematic viscosity of the fluid, and
$\alpha _{\!f}$
is the thermal diffusivity of the fluid. The nonlinear terms quench the exponential linear growth and lead to a steady or oscillating solution for
${\textit{Ra}}$
slightly above
${\textit{Ra}}_c$
(Cross & Hohenberg Reference Cross and Hohenberg1993; Cross & Greenside Reference Cross and Greenside2009). More generally, as Hof et al. (Reference Hof, Westerweel, Schneider and Eckhardt2006) mentioned, with an increase in the flow velocity, the transition from the smooth laminar to the highly disordered turbulent flow can occur through a series of instabilities during which the system can progressively lead to complicated states (Niemela et al. Reference Niemela, Skrbek, Sreenivasan and Donnelly2000), or it may occur abruptly (Grossmann Reference Grossmann2000; Hof et al. Reference Hof, Van, Westerweel, Nieuwstadt, Faisst, Eckhardt, Wedin, Kerswell and Waleffe2004). Out of these two routes, Busse (Reference Busse2003) described a sequence of bifurcations to complex fluid flow that occurs from the simple laminar flow. Beyond the onset of convection, linear stability, in general, cannot predict the nature of instabilities and secondary flow patterns that occur in the flow field. The temporal evolution of the perturbation amplitude can be studied using nonlinear analysis. There are three approaches available for nonlinear analysis: (i) weakly nonlinear stability analysis, (ii) direct numerical simulations (DNS), and (iii) the deflation technique. Near the onset of convection, weakly nonlinear stability analysis gives valuable insights into the nature of instabilities and the secondary flow patterns with minimum computational cost compared to DNS. The deflation technique is a recent method devised to obtain the non-trivial distinct solutions of nonlinear partial differential equations (Farrell, Birkisson & Funke Reference Farrell, Birkisson and Funke2015, Reference Farrell, Beentjes and Birkisson2016). Using this technique, Boullé et al. (Reference Boullé, Dallas and Farrell2022) performed the bifurcation analysis of steady states of two-dimensional Rayleigh–Bénard convection with no-slip boundary conditions.
In the current work, we use the Euler–Euler formulation given by Prakhar & Prosperetti (Reference Prakhar and Prosperetti2021), and perform a weakly nonlinear stability analysis to study the effect of particles on the underlying flow beyond the onset of convection near the critical point. As the perturbation amplitude grows, the nonlinear terms become significant and might lead to the phenomenon of preferential concentration of particles, which is completely absent in the linear stability analysis. Hence we study the effect of nonlinear terms on the secondary flow patterns beyond the onset of convection.
The rest of the paper is organised as follows. The problem statement and the mathematical model are given in § 2. In § 3, the linear stability analysis is provided, together with the governing equations of the basic flow and linear disturbances, and a brief review of the linear stability findings. The full formulation of the amplitude equation using weakly nonlinear stability analysis and the analysis of mutual energy exchange between fluid and particles is provided in § 4. (The numerical procedure for the present work is shown in § A.5.) The results and discussion are presented in § 5. Finally, a brief summary of the current study is given in § 6.
2. Problem statement and mathematical model
The physical model and domain with schematic representation, as shown in figure 1, consists of a continuous fluid domain
$\varOmega =\left \{(x^*,z^*)\in \mathbb{R}\times (-H/2, H/2)\right \}$
, along with a dispersed particulate phase. Here, the bottom and the top surfaces are treated as isothermal walls maintained at temperatures
$T_h$
and
$T_c$
, respectively. The present study considers the monodispersion of very tiny spherical particles at relatively small volume fractions (
$\leqslant 10^{-3}$
). There are two alternative methods, called Euler–Euler and Euler–Lagrange formulations, available for the particles in the millimetre range for which the homogeneous equilibrium model fails. However, the choice of a particular method depends on various length scales, such as particle size
$d_{\!p}$
, the smallest flow length scale
$\varDelta$
, inter-particle separation
$\lambda$
, and particle number density
$n$
. When
$d_{\!p}\ll \varDelta \sim \lambda$
and
$n\varDelta ^3\gg 1$
, the Euler–Lagrange formulation is preferred, whereas if
$d_{\!p}\ll \lambda \ll \varDelta$
and
$n\Delta ^3\gg 1$
, then the Euler–Euler, also called the two-fluid formulation, is preferred. The inter-particle separation
$\lambda$
is related to the non-dimensional particle size
$\delta =(d_{\!p}/H)$
and volume fraction as
$(\lambda /H)\sim (\delta /\varPhi _0^{1/3})$
, and, the condition becomes
$\delta \ll (\delta /\varPhi _0^{1/3})\ll (\varDelta /H)$
. Here, in the present work, we consider the particle volume fraction
$\sim 10^{-3}$
, thus the first inequality,
$\delta \ll (\delta /\varPhi _0^{1/3})$
, is satisfied. However, the second inequality is satisfied only when the particle size is
$\delta \ll (\varDelta /H)/\varPhi _0^{1/3}$
. Therefore, when
$(\varDelta /H)\sim 10^{-1}$
and
$\varPhi _0\sim 10^{-3}$
, the particle size is
$\delta \ll 1$
. Hence in the present work, we choose
$\delta =0.01,0.05$
. We adopt the two-fluid model given by Prakhar & Prosperetti (Reference Prakhar and Prosperetti2021), in which the particles are introduced steadily and uniformly at the top surface at their terminal velocity with a fixed temperature. The momentum equation for the particulate phase is given by the volume-averaged Maxey–Riley–Gatignol equation (Gatignol Reference Gatignol1983; Maxey & Riley Reference Maxey and Riley1983), and the thermal energy equation ensures the balance between the rate of change of the sensible heat of the particles and the rate of convective heat transfer between particulate and fluid. The dimensional independent variables
$ (x^*,z^*,t^* )$
and dependent variables
$ (u_x^*,u_z^*,p^*,T^*,v_x^*,v_z^*,T_{\!p}^* )$
are non-dimensionalised as follows:
\begin{equation} \left .\begin{array}{l} \left (x,z,t\right )=\left (x^*/H,z^*/H,t^*U/H\right )\!,\\[8pt]\left (u_x,u_z,p,\theta \right )=\left (u_x^*/U,u_z^*/U,p^*/\left(\rho _{\!f}U^2\right)\!,(T-T_c)/\Delta T\right )\!,\\[8pt](v_x,v_z,\theta _{\!p})=\left (v_x^*/U,v_z^*/U,(T_{\!p}-T_c)/\Delta T\right )\!, \end{array}\!\!\right \} \end{equation}
where
$x$
,
$z$
and
$t$
are the non-dimensional horizontal coordinate, vertical coordinate and time, respectively. Furthermore,
$u_x$
,
$u_z$
,
$p$
and
$\theta$
are the fluid horizontal velocity component, vertical velocity component, pressure and temperature, respectively. Similarly,
$v_x$
,
$v_z$
and
$\theta _{\!p}$
are dimensionless particulate phase horizontal velocity component, vertical velocity component and temperature, respectively. Here, we use the distance between the two horizontal surfaces
$H$
as a length scale, fluid free-fall velocity
$U=\sqrt {g\beta\, \Delta \textit{TH}}$
as a velocity scale, and the temperature difference
$\Delta T=T_h-T_c$
as a temperature scale, where
$\rho _{\!f}$
,
$g$
and
$\beta$
are fluid density, the acceleration due to gravity and the volume expansion coefficient of fluid, respectively. The resulting governing equations in non-dimensional form are given as

Figure 1. Schematic of the particle-laden Rayleigh–Bénard convection.
for all
$ (x,z,t )\in \left \{\mathbb{R}\times (-1/2, 1/2 )\times (0, \infty )\right \}$
. The boundary conditions are
for all
${ (x,t )}\in \left \{\mathbb{R}\times { (0,\infty )}\right \}$
, where
$\phi$
is the particle volume fraction field, and the non-dimensional parameters present in the problem, Rayleigh number
${\textit{Ra}}$
, Prandtl number
$ \textit{Pr} $
, density ratio
$R$
, specific heat capacity ratio
$E$
, mechanical Stokes number
$St_m$
, thermal Stokes number
$St_{\textit{th}}$
, initial particle temperature
$\varTheta _{\textit{pt}}$
, and non-dimensional particle terminal speed
$v_0$
, are defined as
\begin{align} &\textit{Ra}=g\beta\, \Delta {\textit{TH}}^3/(\alpha _{\!f}\nu _{\!f}), \quad \textit{Pr}=\nu _{\!f}/\alpha _{\!f}, \quad R=\rho _{\!p}/\rho _{\!f}, \nonumber \\&{\textit{St}}_m=\frac {\tau _{\!p}}{\tau _{\!f}}=\frac {{R}\delta ^2}{18{C}_d} \sqrt {\frac {\textit{Ra}}{\textit{Pr}}},\quad {\textit{St}}_{\textit{th}}=\frac {\tau _{\textit{th}}}{\tau _{\!f}}=\frac {{E}\delta ^2}{6{\textit{Nu}}_{\!p}}\sqrt {{{\textit{Ra}}{\textit{Pr}}}}, \nonumber \\&E=R(C_{\textit{pp}}/C_{\textit{pf}}), \quad v_0=\frac {{\textit{Re}}_{\!p}}{\delta }\sqrt {\frac {\textit{Pr}}{\textit{Ra}}}, \quad \varTheta _{\textit{pt}}=\frac {T_{\textit{pt}}-T_c}{T_h-T_c}, \end{align}
where
$\alpha _{\!f}=\kappa _{\!f}/\rho _{\!f}C_{\textit{pf}}$
is the fluid thermal diffusion coefficient,
$\nu _{\!f}=\mu _{\!f}/\rho _{\!f}$
is the fluid kinematic viscosity,
$\kappa _{\!f}$
is the fluid thermal conductivity,
$C_{\textit{pf}}$
is the fluid specific heat constant,
$\mu _{\!f}$
is fluid dynamic viscosity,
$\rho _{\!p}$
is the particle density, and
$C_{\textit{pp}}$
is the particle specific heat. Particle diameter
$d_{\!p}$
in the non-dimensional form is represented as
$\delta =d_{\!p}/H$
. Introduction of particles into the flow leads to three distinct time scales:
$\tau _{\!f}=H/U$
is the flow time scale,
$\tau _{\!p}=\rho _{\!p}d_{\!p}^2/(18\mu _{\!f}C_d({\textit{Re}}_{\!p}))$
is the particle mechanical relaxation time scale, and
$\tau _{\textit{th}}=\rho _{\!p}C_{\textit{pp}}d_{\!p}^2/(6\kappa _{\!f}\,{\textit{Nu}}_{\!p}({\textit{Re}}_{\!p},\textit{Pr}))$
is the particle thermal relaxation time scale. Here,
${C}_d=1+0.15\,{\textit{Re}}_{\!p}^{0.687}$
is the Schiller–Naumann correction factor to the Stokes drag when the particle Reynolds number
${\textit{Re}}_{\!p}=\rho _{\!f}v_\tau d_{\!p}/\mu _{\!f}$
is greater than unity and less than 800 (Clift, Grace & Weber Reference Clift, Grace and Weber2005),
$v_\tau ={ (1-\rho _{\!f}/\rho _{\!p} )}\tau _{\!p}g$
is the settling speed of a particle in the quiescent fluid, and
${\textit{Nu}}_{\!p}=2.0+0.6\,{\textit{Re}}_{\!p}^{1/2}\,\textit{Pr}^{1/3}$
is the particle Nusselt number. The last terms on the right-hand sides of (2.3) and (2.4) represent the mechanical two-way coupling between the fluid and particles, whereas the thermal two-way coupling is captured by the right-hand-side last term of (2.5).
3. Linear stability analysis
We assume that the fluid–particle system is at a steady state, with particles settling uniformly with their terminal velocity in a quiescent fluid. Under these conditions, the above governing equations (2.2)–(2.9) reduced to a system of ordinary differential equations, which are represented in operator form as
for
$z\in { (-1/2,1/2 )}$
with boundary conditions
where
$P_0$
,
$\varTheta _0$
and
$\varTheta _{p0}$
are the basic state fluid pressure, basic state fluid temperature and basic state particle temperature field, respectively, and
$\varPhi _0$
is the initial uniform particle volume fraction. Using the above boundary conditions, the solution to (3.1)–(3.3) yields the base state given by
where
$\mathscr{A}$
and
$\mathscr{B}$
are integration constants given by
\begin{align} \mathscr{A}&=-\frac {(1-\varTheta _{\textit{pt}}){\textit{lm}}_2+\varTheta _{\textit{pt}}\left [1-{\left (1-{\textit{lm}}_2\right )}\,e^{-m_2}\right ]}{{\textit{lm}}_1\left [1-{\left (1-{\textit{lm}}_2\right )}\,e^{-m_2}\right ]-{\textit{lm}}_2\left [1-{\left (1-{\textit{lm}}_1\right )}\,e^{-m_1}\right ]}, \end{align}
\begin{align} \mathscr{B}&=+\frac {(1-\varTheta _{\textit{pt}}){\textit{lm}}_1+\varTheta _{\textit{pt}}\left [1-{\left (1-{\textit{lm}}_1\right )}\,e^{-m_1}\right ]}{{\textit{lm}}_1\left [1-{\left (1-{\textit{lm}}_2\right )}\,e^{-m_2}\right ]-{\textit{lm}}_2\left [1-{\left (1-{\textit{lm}}_1\right )}\,e^{-m_1}\right ]}. \end{align}
Here,
$m_1$
and
$m_2$
are given by
\begin{align} m_1=\frac {1}{2l}\left [1+\sqrt {1+4{\left (\frac {l}{{L}}\right )}^2}\right ]\!,\quad m_2=\frac {1}{2l}\left [1-\sqrt {1+4{\left (\frac {l}{{L}}\right )}^2}\right ]\!, \end{align}
where
$L$
is the spatial non-dimensional length scale over which the effect of a particle on the surrounding fluid temperature is significant (Prakhar & Prosperetti Reference Prakhar and Prosperetti2021), given by
\begin{equation} {L}={\left (\frac {{\textit{St}}_{\textit{th}}}{{E}\varPhi _0\sqrt {{{\textit{Ra}}{\textit{Pr}}}}}\right )}^{1/2}=\frac {\delta }{\sqrt {6\varPhi _0\,{\textit{Nu}}_{\!p}}}. \end{equation}
Here,
$l=v_0{\textit{St}}_{\textit{th}}$
is the non-dimensional length scale that the particle must traverse for its temperature to be locally equal to that of the fluid (Prakhar & Prosperetti Reference Prakhar and Prosperetti2021), related to other parameters as
The classical normal mode analysis (Drazin & Reid Reference Drazin and Reid2004) is performed to examine the stability of the basic flow mentioned above. In general, the linear stability analysis is carried out by decomposing all the dependent variables into a steady basic state and the respective infinitesimal perturbations (represented by a superscript prime):
\begin{align} {\big (u_x,u_z,\theta ,v_x,v_z,\theta _{\!p},p,\phi \big )}={}&{\big (0,0,\varTheta _0(z),0,-v_0,\varTheta _{p0}(z),P_{0}(z),\varPhi _0\big )}\nonumber\\&{}+{\big (u_x^{\prime},u_z^{\prime},\theta ',v_x^{\prime},v_z^{\prime},\theta _{\!p}^{\prime},p',\phi '\big )}. \end{align}
After neglecting the higher-order terms and retaining only the first-order terms in
$ \{u_x^{\prime},u_z^{\prime},\theta ',v_x^{\prime},v_z^{\prime},\theta _{\!p}^{\prime},p',\phi ' \}$
, we obtain
The above system of equations is linear homogeneous and has the coefficient functions of spatial variables only, not dependent on time. Hence they satisfy the solution in the normal mode form given by
\begin{align} &{\big (u_x^{\prime},u_z^{\prime},\theta ',v_x^{\prime},v_z^{\prime},\theta _{\!p}^{\prime},p',\phi '\big )}^{\textrm T}\nonumber \\&\quad=e^{{\textrm i}k{ (x-ct )}}{\big (\hat {u}_x(z),\hat {u}_z(z),\hat {\theta }(z),\hat {v}_x(z),\hat {v}_z(z),\hat {\theta }_{\!p}(z),\hat {p}(z),\hat {\phi }(z)\big )}^{\textrm T}, \end{align}
where
$k$
is the wavenumber, and
$c=c_r+\textrm{i}c_r$
is the complex wave speed corresponding to
$k$
. The sign of
$c_i$
determines the growth or decay of the disturbance. That is, as
$c_i\lt 0$
or
$c_i=0$
or
$c_i\gt 0$
, the flow is stable or neutrally stable or unstable, respectively. Substitution of (3.22) in (3.14)–(3.21) leads to linearised disturbance equations in operator form given in the § A.1. The boundary conditions for corresponding equations are given by
The system of linear equations (A1)–(A8) together with the boundary conditions gives rise to a generalised eigenvalue problem given by
where
$\boldsymbol{q}={ (\hat {u}_x,\hat {u}_z,\hat {\theta },\hat {v}_x,\hat {v}_z,\hat {\theta }_{\!p},\hat {p},\hat {\phi } )}$
is the eigenfunction corresponding to the eigenvalue
$c$
, and
$ \mathbb{A},\mathbb{B} $
are the square complex matrices. The fundamental disturbance is given by
$\boldsymbol{q}\,e^{{\textrm i}k{ (x-c_rt )}}$
, where
$\boldsymbol{q}$
is the eigenfunction related to the least stable eigenvalue, and
$c_r$
is the corresponding wave speed. (Several parts of the weakly nonlinear analysis will employ the frequent use of the fundamental disturbance.)
The bifurcation point, also known as the critical point
$(k, Ra)$
, and the shape of the emerging disturbances are both determined using the linear stability theory. It offers no information regarding the actual magnitude of the disturbances (amplitude) away from the critical value. A weakly nonlinear stability study is necessary to examine the amplitude of such disturbances. The outcomes of the linear stability analysis are also necessary for a nonlinear analysis.
There have been two studies (Prakhar & Prosperetti Reference Prakhar and Prosperetti2021; Raza et al. Reference Raza, Hirata and Calzavarini2024) on the linear stability analysis of the present problem in the literature. Using a two-fluid model, Prakhar & Prosperetti (Reference Prakhar and Prosperetti2021) studied the stability threshold of particle-laden Rayleigh–Bénard convection when the particles are much denser than the underlying fluid. Raza et al. (Reference Raza, Hirata and Calzavarini2024) extended the two-fluid model by Prakhar & Prosperetti (Reference Prakhar and Prosperetti2021) to lighter particles such as bubbles by adding an added-mass term. Using dimensional analysis, we can show that eight independent dimensionless parameters exist in this problem, such as Rayleigh number
${\textit{Ra}}$
, Prandtl number
$ \textit{Pr} $
, density ratio
$R$
, heat capacity ratio
$E$
, particle Reynolds number
${\textit{Re}}_{\!p}$
, particle diameter
$\delta$
, particle injection temperature
$\varTheta _{\textit{pt}}$
, and initial particle volume fraction
$\varPhi _0$
. We fix the dimensionless parameters such as Prandtl number
$ \textit{Pr}=0.71$
, density ratio
$R=800$
, and specific heat capacity ratio
$E=3385$
, such that the fluid–particle system represents the water droplets suspended in dry air.
We present the linear stability results for the air–water droplet system. The effect of initial particle volume fraction on the critical Rayleigh number
${\textit{Ra}}_c$
and the critical wavenumber
$k_c$
for different particle Reynolds numbers
${\textit{Re}}_{\!p}$
and the particle sizes
$\delta$
is shown in figure 2(a,b). The critical Rayleigh number increases with
$\varPhi _0$
for all
${\textit{Re}}_{\!p}$
and
$\delta$
. This stabilisation of the flow under the presence of particles is due to the dissipative nature of the mechanical and thermal two-way coupling source terms in (2.3)–(2.4) and (2.5), respectively. The strength of the mechanical source terms is proportional to
$({R}\varPhi _0/{\textit{St}}_m)\sim \varPhi _0{C}_d/\delta ^2$
, independent of the density ratio
$R$
. Similarly, the strength of the thermal source term is proportional to
${E}\varPhi _0/{\textit{St}}_{\textit{th}}\sim \varPhi _0\,{\textit{Nu}}_{\!p}/\delta ^2$
, which is independent of the heat capacity ratio
$E$
. Hence with an increase in the
$\varPhi _0$
value (decrease in
$\delta$
value) in either source term, the dissipation increases and thereby increases the stability (increase in
${\textit{Ra}}_c$
value). However, the increase in
${\textit{Ra}}_c$
with
$\varPhi _0$
is more significant for smaller particles (
$\delta =0.01$
), due to the strong
$1/\delta ^2$
dependence of the two-way coupling source terms, than for the large particles (
$\delta =0.05$
). Moreover, the weak dependency of the mechanical and thermal source terms on particle Reynolds number
${\textit{Re}}_{\!p}$
through drag force coefficient
${C}_d$
and the Nusselt number
${\textit{Nu}}_{\!p}$
explains the small increase in
${\textit{Ra}}_c$
with the increase in
${\textit{Re}}_{\!p}$
. Similarly, from figure 2(b), it is clear that the critical wavenumber
$k_c$
increases with an increase in
$\varPhi _0$
for all
${\textit{Re}}_{\!p}$
and
$\delta$
. The plausible explanation for this can be obtained using the inter-particle distance
$\lambda /H\sim (1-\varPhi _0)^{1/3}\delta /\varPhi _0^{1/3}\approx \delta /\varPhi _0^{1/3}$
for the dilute suspensions
$(\varPhi _0\ll 1)$
(Prakhar & Prosperetti Reference Prakhar and Prosperetti2021). Hence the critical wavenumber can be expected to vary as
$k_c\sim {\varPhi _0^{1/3}}/{\delta }$
, which explains the increase of
$k_c$
with an increase in
$\varPhi _0$
for all
${\textit{Re}}_{\!p}$
and
$\delta$
values. We note that the increase in
$k_c$
is significant for the small particles (
$\delta =0.01$
) with small Reynolds number
${\textit{Re}}_{\!p}=1$
rather than for the large particles (
$\delta =0.05$
) with large Reynolds number
${\textit{Re}}_{\!p}=10$
as shown in figure 2(b).

Figure 2. Effect of initial particle volume fraction on critical Rayleigh number
${\textit{Ra}}_c$
and critical wavenumber
$k_c$
: (a) variation in critical Rayleigh number
${\textit{Ra}}_c$
, and (b) variation in critical wavenumber
$k_c$
with initial undisturbed particle volume fraction
$\varPhi _0$
for two different particle Reynolds numbers
${\textit{Re}}_{\!p}$
and particle sizes
$\delta$
, and other parameters kept at
$\varTheta _{\textit{pt}}=0$
,
${R}=800$
,
${E}=3385$
and
${Pr}=0.71$
for both graphs. Here, the circles represent data from Prakhar & Prosperetti (Reference Prakhar and Prosperetti2021).

Figure 3. Effect of particle injection temperature
$\varTheta _{\textit{pt}}$
on critical parameters, base-state fluid temperature and its stratification. (a) Variation in critical Rayleigh number
${\textit{Ra}}_c$
, where the dots represent the data from Prakhar & Prosperetti (Reference Prakhar and Prosperetti2021). (b) Variation in critical wavenumber
$k_c$
. (c) Variation in base-state fluid temperature. (d) Variation in unstably stratified layer thickness
$\delta _{st}$
with particle injection temperature. Here,
${\textit{Re}}_{\!p}=1$
,
$\delta =0.01$
,
$\varPhi _0={10^{-4}}$
,
${R}=800$
,
${E}=3385$
and
$ \textit{Pr}=0.71$
for all graphs.
The effect of particle injection temperature
$\varTheta _{\textit{pt}}$
on the critical parameters and the base-state fluid temperature is shown in figure 3(a–d). In figure 3(a), the present data are compared with the existing work by Prakhar & Prosperetti (Reference Prakhar and Prosperetti2021). It is observed that with an increase in
$\varTheta _{\textit{pt}}$
, the critical Rayleigh number
${\textit{Ra}}_c$
increases in the initial part of figure 3(a). The explanation for this is that the particles act as the sources of heat, and by increasing their temperature, they tend to reduce the thermal stratification inside the domain, leading to an increase in the stability of the flow. However, from figure 3(c), it is clear that as the particle temperature increases beyond
$\varTheta _{\textit{pt}}\gt 1$
, the strong negative base-state temperature gradients start to appear in the upper part of the domain, and favour the instability. The effect of
$\varTheta _{\textit{pt}}$
on the critical wavenumber is shown in figure 3(b). As
$\varTheta _{\textit{pt}}$
is increased from
$-1$
to 2,
$k_c$
decreases until
$\varTheta _{\textit{pt}}\approx 0.5$
, then increases monotonically. This can be explained by looking at the variation in unstably stratified layer thickness
$\delta _{st}$
in the base-state temperature profile shown in figure 3(c) and 3(d). For
$\varTheta _{\textit{pt}}=-1$
in figure 3(c), from
$z\approx -0.26$
to
$-0.5$
, the fluid is unstably stratified over thickness
$\delta _{st}$
, and in the remaining domain,
$\varTheta _{\textit{pt}}$
has a symmetric distribution. A similar distribution exists for
$\varTheta _{\textit{pt}}=2$
, but the unstably stratified layer exists near the cold top surface. The thickness of this unstably stratified layer increases for
$\varTheta _{\textit{pt}}=-1$
to
$\varTheta _{\textit{pt}}\approx 0$
, maintains a constant total height
$\delta _{st}=1$
from
$\varTheta _{\textit{pt}}\approx 0$
to
$\varTheta _{\textit{pt}}\approx 1.25$
, and subsequently increases monotonically for
$\varTheta _{\textit{pt}}\approx 1.25$
to
$\varTheta _{\textit{pt}}=2$
. From the definition of
$\delta _{st}$
, we note that the non-dimensional temperature difference across all these unstably stratified layers is 1. Hence an increase in
$\delta _{st}$
value with the same temperature difference results in a lower temperature gradient, favouring stability. This explains the initial increase in
${\textit{Ra}}_c$
value in figure 3(a). From
$\varTheta _{\textit{pt}}\approx 1.25$
to
$\varTheta _{\textit{pt}}=2$
, the unstable stratified layer thickness reduces, maintaining the same temperature difference across its length, which increases the negative temperature gradient and favours the instability. Hence
${\textit{Ra}}_c$
reduces from
$\varTheta _{\textit{pt}}\approx 1.25$
onwards. It should be noted that
$k_c$
represents the length scale for the onset of convection, and it changes with particle volume fraction
$\varPhi _0$
and size
$\delta$
. However, in this case, we keep
$\varPhi _0$
and
$\delta$
constant, and vary only
$\varTheta _{\textit{pt}}$
. Hence the explanation for the non-monotonic variation of
$k_c$
with
$\varTheta _{\textit{pt}}$
can be deduced from the variation of
$\delta _{st}$
with
$\varTheta _{\textit{pt}}$
shown in figure 3(d). As
$\delta _{st}$
increases, the length scale at the onset of convection increases, leading to a decrease in
$k_c$
, and vice versa.
4. Formulation of finite amplitude equations
In the present work, the finite amplitude expansions are given based on the analysis of Stuart (Reference Stuart1960), Yao & Rogers (Reference Yao and Rogers1992), Khandelwal & Bera (Reference Khandelwal and Bera2015), Sharma, Khandelwal & Bera (Reference Sharma, Khandelwal and Bera2018) and Aleria, Khan & Bera (Reference Aleria, Khan and Bera2024). The Fourier expansion of fluid temperature
$\theta$
in separable form is
\begin{align} \theta (x,z,t)&=\varTheta (z,\tau )\,\mathbb{E}^0+\hat {\theta }_1\left (z,\tau \right )\,\mathbb{E}^1+\hat {\theta }_2\left (z,\tau \right )\,\mathbb{E}^2+\cdots +\text{c.c.} \nonumber \\&=\mathbb{E}^0\big [\varTheta _0(z)+c_i\,|B(\tau )|^2\,\varTheta _1(z)+\mathcal{O}\big (c_i^2\big )\big ] \nonumber\\&\quad {}+\mathbb{E}^1\left [c_i^{1/2}\,B(\tau )\,\theta _{10}(z)+c_i^{3/2}\,B(\tau )\,|B(\tau )|^2\,\theta _{11}(z)+\mathcal{O}\big (c_i^{5/2}\big )\right ] \nonumber\\&\quad {}+\mathbb{E}^2\big [c_i\,B(\tau )^2\,\theta _{20}(z)+\mathcal{O}\big (c_i^2\big )\big ]+\cdots +\text{c.c.,} \end{align}
where
$\mathbb{E}^j=\exp (j [\mbox{i}k_c (x-c_rt ) ])$
for
$j=0,1,2$
,
$k_c$
is the wavenumber corresponding to the critical Rayleigh number
${\textit{Ra}}_c$
, and
$c_r$
is the real part of
$c$
corresponding to the most unstable disturbance. The amplitude function
$B$
will be obtained from the Landau equation, and c.c. stands for the complex conjugate of the complex-valued functions (the second and third terms).
The evolution equation for the slowly varying amplitude
$B$
is derived using the method of multiple time scales. Hence two distinct time scales (a fast time scale
$t$
and a slow time scale
$\tau $
) are used in the nonlinear stability analysis of the flow. As in linear stability analysis, the exponential variation of disturbance amplitude is associated with the fast time scale. As the disturbance amplitude grows exponentially and attains a finite amplitude, the nonlinear terms become important. This leads to the deviation of the temporal exponential growth of the disturbances. Hence another time scale (the slow time scale) is needed to capture the effect of nonlinearities of different orders. The slow time scale is defined as
$\tau =c_it$
, with
$c_i$
being the growth rate of the most unstable disturbance obtained from the linear stability analysis. Hence the derivative with respect to
$t$
of a function
$f(t)=F(t,\tau )$
is given by
Thus for a function of the form
$F(t, \tau )$
, both
$t$
and
$\tau$
are treated as if they are independent variables. The justification of (4.1) is as follows. From the linear theory, it is known that the growth rate
$c_i$
is linearly proportional to
$ ({Ra}-{\textit{Ra}}_c )/{\textit{Ra}}_c$
, the Rayleigh number difference from the neutral curve. Using a finite amplitude stability analysis for the Rayleigh–Bénard convection (Schlüter et al. Reference Schlüter, Lortz and Busse1965) and for the non-isothermal flow in a vertical annulus, Yao & Rogers (Reference Yao and Rogers1992) showed that the square of equilibrium amplitude
$A_e^2$
is proportional to
$ ({Ra}-{\textit{Ra}}_c )/{\textit{Ra}}_c$
, and consequently the equilibrium amplitude is proportional to
$c_i^{1/2}$
, with a proportionality constant that is a weak function of
${\textit{Ra}}$
near
${\textit{Ra}}_c$
. Hence the amplitude of the disturbance fluid temperature
${\theta }_{10}$
with wavenumber
$k_c$
is of the order of
$c_i^{1/2}$
. The amplitude of the disturbances of higher harmonics is chosen based on the number of interactions between lower-order harmonics. For example, the harmonic with wavenumber
$2k_c$
needs two disturbances with wavenumber
$k_c$
, which gives its amplitude of order
$c_i$
. Similarly, the higher-order correction to the base state results from the Reynolds stress due to the interaction of
$\mathbb{E}^1$
and its complex conjugate
$\mathbb{E}^{-1}$
with an amplitude of order
$c_i$
(Stewartson & Stuart Reference Stewartson and Stuart1971).
4.1. Derivation of the cubic Landau equation
Substitute the perturbation series similar to (4.1) for all the dependent variables in (2.2)–(2.9), and separate the harmonic components. The resulting equations are shown in § A.2. The first Landau coefficient does not depend on the higher-order harmonics (
$\mathbb{E}^3,\mathbb{E}^4$
etc.). Therefore, those terms are ignored in the series (4.1). The systems of equations at different harmonics (A9)–(A13), (A15)–(A22) and (A23)–(A30) are solved at increase orders of
$c_i$
. At
$\mathcal{O} (c_i^0 )$
,
$\mathbb{E}^0$
harmonic equations are identical to the basic-flow equations (3.1)–(3.3). At
$\mathcal{O} (c_i )^{1/2}$
,
$\mathbb{E}^1$
equations are similar to that of linear stability (A1)–(A8), and the
$\mathbb{E}^0$
and
$\mathbb{E}^2$
do not contribute at this order. The functions
$u_{x10}$
,
$u_{z10}$
,
$\theta _{10}$
,
$v_{x10}$
,
$v_{z10}$
,
$\theta _{p10}$
and
$\phi _{10}$
are the eigenfunctions of linear stability equations at a particular wavenumber and Rayleigh number
${\textit{Ra}}$
. At
$\mathcal{O} (c_i )$
,
$\mathbb{E}^0$
and
$\mathbb{E}^2$
generate the system of non-homogeneous equations for basic-flow distortion functions
$\varTheta _1$
,
$\varTheta _{p1}$
,
$V_1$
and
$\varPhi _1$
, and the functions
$u_{x20}$
,
$u_{z20}$
,
$\theta _{20}$
,
$v_{x20}$
,
$v_{z20}$
,
$\theta _{p20}$
and
$\phi _{20}$
, respectively. These equations contain the non-homogeneous part made up of known functions
$u_{x10}$
,
$u_{z10}$
,
$\theta _{10}$
,
$v_{x10}$
,
$v_{z10}$
,
$\theta _{p10}$
,
$\phi _{10}$
and their derivatives, which are obtained at lower-order computations. At
$\mathcal{O} (c_i^{3/2} )$
, the
$\mathbb{E}^1$
harmonic results in a non-homogeneous system of equations with linear stability operators acting on the functions
$u_{x11}$
,
$u_{z11}$
,
$\theta _{11}$
,
$v_{x11}$
,
$v_{z11}$
,
$\theta _{p11}$
and
$\phi _{11}$
. However, the right-hand-side non-homogeneous part contains the terms proportional to the unknown terms
${\textrm d}B/{\textrm d}\tau$
,
$B$
and
$B\,|B|^2$
, with coefficients depending on functions known from the lower-order analysis. The Fredholm alternative solvability condition is imposed on the non-homogeneous right-hand side of the
$\mathbb{E}^1$
harmonic at
$\mathcal{O} (c_i^{3/2} )$
to obtain a non-trivial solution. To impose the solvability condition, the solution to the homogeneous adjoint system (see § A.4) corresponding to the linear stability operator is required. Accordingly, the non-homogeneous right-hand side must be orthogonal to the adjoint functions
$\hat {u}_x^\dagger$
,
$\hat {u}_z^\dagger$
,
$\hat {\theta }^\dagger$
,
$\hat {v}_x^\dagger$
,
$\hat {v}_z^\dagger$
,
$\hat {\theta }_{p}^\dagger$
and
$\hat {\phi }^\dagger$
. The orthogonality is imposed by multiplying the right-hand side of the system of linear equations of
$\mathbb{E}^1$
harmonic functions at
$\mathcal{O} (c_i^{3/2} )$
with
$\displaystyle { (\hat {p}^\dagger , \hat {u}_{x}^\dagger ,\hat {u}_z^\dagger ,\hat {\theta }^\dagger , \,\hat {\phi }^\dagger ,\hat {v}_x^\dagger ,\hat {v}_z^\dagger ,\hat {\theta }_{\!p}^\dagger )}$
, and integrating with respect to
$z$
from
$-1/2$
to
$1/2$
, which gives the cubic Landau equation
where
$a_1$
is the Landau constant defined as
which represents the first correction to the growth rate given by linear stability analysis.
By changing the variables as
$A=c_i^{1/2}B$
and
$\tau =c_it$
in (4.3),
and its corresponding complex conjugate equation is
where
$k_cc_i$
is the growth rate from the linear stability analysis, and
$A$
is the physical amplitude of the wave. Multiply (4.5) with
$\tilde {A}$
and (4.6) with
$A$
and add them:
where
${\textrm{Re}}\{a_1\}$
is the real part of
$a_1$
. Equation (4.7) has an equilibrium amplitude
$A_e$
as a solution if
Consequently, three possible equilibrium amplitudes exist:
Here,
$A_e=0$
represents the steady base flow, which is stable for
${\textit{Ra}}\lt {\textit{Ra}}_c$
and unstable for
${\textit{Ra}}\gt {\textit{Ra}}_c$
, and
${\textit{Ra}}_c$
is the critical Rayleigh number obtained from linear stability analysis. From (4.9), the existence of a finite amplitude solution is guaranteed if
$k_cc_i$
and
${\textrm{Re}}\{a_1\}$
have opposite signs (Drazin & Reid Reference Drazin and Reid2004; Shukla & Alam Reference Shukla and Alam2011). Therefore, there are two possibilities: first, the growth rate is positive (for
${\textit{Ra}}\gt {\textit{Ra}}_c$
), and the real part of the Landau constant is negative; second, the growth rate is negative (for
${\textit{Ra}}\lt {\textit{Ra}}_c$
), and the real part of the Landau constant is positive. The first possibility leads to a supercritical pitchfork bifurcation, whereas the second possibility leads to a subcritical pitchfork bifurcation.
Using equilibrium amplitude definition (4.9b ), (4.7) can be rewritten as
where
$A_e^2=-{ (k_cc_i )}/{\textrm{Re}}\{a_1\}$
. The solution for the above equation is
\begin{equation} |A|^2=\frac {A_e^2}{1+\displaystyle {\left (\frac {A_e^2}{|A_0|^2}-1\right )}e^{-2k_cc_it}}, \end{equation}
where
$|A_0|$
is the value of
$|A|$
at
$t=0$
. From (4.11), it is clear that when
${\textit{Ra}}\gt {\textit{Ra}}_c$
, as
$t\to \infty$
,
$|A|\to A_e$
. Hence, irrespective of the initial amplitude
$|A_0|$
, the amplitude eventually tends to the equilibrium amplitude
$A_e$
. Multiply (4.5) with
$1/\tilde {A}$
and (4.6) with
$-A/\tilde {A}^2$
, and add them:
where
$\textrm{Im} \{a_1\}$
is the imaginary part of
$a_1$
. The above equation can be rewritten as
\begin{equation} \frac {\textrm{d}\displaystyle {\left (\frac {A}{\tilde {A}}\right )}}{\displaystyle {\left (\frac {A}{\tilde {A}}\right )}}=2{\textrm i}\,\textrm{Im} \{a_1\}\,|A|^2\,{\textrm d}t. \end{equation}
Hence the solution for
$A$
is given by
where
$|A|{ (t )}$
is given by the square root of (4.11). The closed-form solution for amplitude function
$A{ (t )}$
is obtained by integrating the above equation:
where
$|A|(t)$
is obtained from (4.11).
4.2. Heat flux balance
The solution to the Landau equation given by (4.11) and (4.15) shows the existence of a steady-state solution as
$t\to \infty$
. Hence in this subsection, we derive an equation for the average heat flux balance at steady state. At steady state, integrating the fluid energy (2.5) in the domain
${ (x,z )}\in \left \{{ (-0.5,0.5 )}\times { (-\pi /k_c,\pi /k_c )}\right \}$
yields
\begin{align} &\int _{z=-0.5}^{0.5}\int _{x=-\pi /k_c}^{\pi /k_c}{\left (u_x\frac {\partial \theta }{\partial x}+u_z\frac {\partial \theta }{\partial z}\right )}\,{\textrm d}x\,{\textrm d}z \nonumber\\ &\quad =\int _{z=-0.5}^{0.5}\int _{x=-\pi /k_c}^{\pi /k_c}{\left (\frac {1}{\sqrt {{{\textit{Ra}}{\textit{Pr}}}}}{\left (\frac {\partial ^2 {\theta }}{\partial {x}^2}+\frac {\partial ^2 {\theta }}{\partial {z}^2}\right )}-\frac {{E}\phi }{{\textit{St}}_{\textit{th}}}{\left (\theta _{\!p}-\theta _{\!p}\right )}\right )}\,{\textrm d}x\,{\textrm d}z. \end{align}
Since
$u_x$
,
$u_z$
,
$\theta$
and
$\theta _{\!p}$
are periodic functions in
$x$
with period
$2\pi /k$
, the non-zero contribution comes only from the
$\mathbb{E}^0$
harmonic. Hence (4.16) is essentially equivalent to the integration of
$\mathbb{E}^0$
harmonic (A10) in
$z\in { (-0.5,0.5 )}$
. Therefore, after substituting
$A_e^2=c_i\,|B|^2$
,
$\varTheta =\varTheta _0+A_e^2\varTheta _1$
and
$\varTheta _{\!p}=\varTheta _{p0}+A_e^2\varTheta _{p1}$
in (A10), we have
\begin{align} &\int _{z=-0.5}^{0.5}\frac {\textrm{d}^2 \varTheta }{\textrm{d}z^2}\,{\textrm d}z\\ \nonumber &={E}\sqrt {{{\textit{Ra}}{\textit{Pr}}}}\int _{z=-0.5}^{0.5}\left \{\varPhi _0\frac {\varTheta -\varTheta _{\!p}}{{\textit{St}}_{\textit{th}}} + A_e^2\varPhi _1\frac {\varTheta _0-\varTheta _{p0}}{{\textit{St}}_{\textit{th}}}+A_e^2\frac {\textrm{d} }{\textrm{d} z} {\big (\tilde {\theta }_{10}u_{z10}+\theta _{10}\tilde {u}_{z10}\big )}\right \}{\textrm d}z, \end{align}
where the integral of the last term on the right-hand side goes to zero because
$u_{z10}=\theta _{10}=0$
at
$z=-0.5$
and 0.5. From (A13), we have the identities
Hence (4.17) simplifies to
where the bracket is defined for the function
$f(z)$
as
$[\![ f ]\!]=f(0.5)-f(-0.5)$
, and
${\textit{Nu}}_h={\textit{Nu}}(-0.5)=- {\textrm{d} \varTheta }/{\textrm{d} z}\big |_{z=-0.5}$
and
${\textit{Nu}}_c={\textit{Nu}}(0.5)=- {\textrm{d} \varTheta }/{\textrm{d} z}\big |_{z=0.5}$
are the Nusselt numbers (or non-dimensional heat fluxes) at the hot and cold surfaces, respectively. It can be shown that the solution to the linearised volume fraction equation is always zero (see § A.6). Hence from (A11),
for
$z\in { (-0.5,0.5 )}$
. Therefore, from (4.20) and (4.21), the net flux balance is given by
where the proportionality constant
$v_0{E}\varPhi _0\sqrt {{{\textit{Ra}}{\textit{Pr}}}}$
is equal to
${L}^2/l$
. Here,
$L$
and
$l$
are the non-dimensional length scales given by (3.11) and (3.12), respectively. Thus the alternate form of the above equation is given by
where
$Q_{ph}^{\prime \prime }=Q_{\!p}^{\prime \prime }(-0.5)=(l/{L}^2)\,\varTheta _{\!p}(-0.5)$
and
$Q_{pc}^{\prime \prime }=Q_{\!p}^{\prime \prime }(0.5)=(l/{L}^2)\,\varTheta _{\!p}(0.5)$
are the non-dimensional particle sensible heat fluxes at hot and cold surfaces, respectively. The physical significance of (4.23) is that at steady state, the net heat flux released from the hot and cold surfaces is equal to the net sensible heat exchanged by the particles. Therefore, for the single-phase and particle-laden flows without thermal energy two-way coupling, the heat flux at the hot surface is exactly the heat flux at the cold surface. However, for the problems with thermal energy two-way coupling, the two phases exchange the heat by following (4.23), and heat fluxes
$Nu$
at hot and cold need not be equal.
5. Results and discussion
We carry out the analysis of the effect of the nonlinear interactions of different harmonics on the equilibrium amplitude, rate of heat transfer and the pattern of secondary flow. All the analysis is done near the bifurcation point such that the perturbation series (4.1) is valid. We define a new parameter called the reduced Rayleigh number,
$\delta _{Ra}={ (Ra-{\textit{Ra}}_c )}/{\textit{Ra}}_c$
, which quantifies the deviation from the bifurcation point
${\textit{Ra}}_c$
.
5.1. Effect of particle volume fraction
The grid independence test on the growth rate
$c_i$
, the real part of the Landau constant
${\textrm{Re}}\{a_1\}$
, and the equilibrium amplitude
$A_e$
for
$\delta _{Ra}\in [0, 0.75]$
at three different particle volume fractions with corresponding critical wavenumbers is shown in table 1. With an increase in the degree of the Chebyshev polynomial
$N$
, the values of
$c_i$
,
${\textrm{Re}}\{a_1\}$
and
$A_e$
are not changed consistently for each
$\varPhi _0$
value. Hence we fix
$N=50$
for all the computations in the present work. Moreover, it can be seen that the growth rate
$c_i$
is positive and decreases with an increase in
$\varPhi _0$
,
${\textrm{Re}}\{a_1\}\lt 0$
, and
$A_e$
also decreases with an increase in
$\varPhi _0$
. These observations are more apparent in figure 4(a–c), in which
$c_i$
,
${\textrm{Re}}\{a_1\}$
,
$A_e$
and
$A_e/\sqrt {c_i}$
are plotted against
$\delta _{Ra}$
at four different
$\varPhi _0$
values. Figure 4(a) shows the linear variation of
$c_i$
with
$\delta _{Ra}$
near the critical point. The decrease in
$c_i$
value with an increase in
$\varPhi _0$
reveals that the stability of the flow increases with an increase in particle volume fraction. The real part of the Landau constant satisfies
${\textrm{Re}}\{a_1\}\lt 0$
for
$\delta _{Ra}\in [0,0.75]$
at all
$\varPhi _0$
values (see figure 4
b). Hence
${\textrm{Re}}\{a_1\}\lt 0$
together with
$c_i\gt 0$
for
$\delta _{Ra}\gt 0$
shows the supercritical pitchfork bifurcation at the critical point. Therefore, the stable, positive and finite equilibrium amplitude
$A_e$
is guaranteed (Drazin & Reid Reference Drazin and Reid2004; Shukla & Alam Reference Shukla and Alam2011). In the present work on particle-laden Rayleigh–Bénard convection, the real part of the wave speed
$c_r$
of the most dominant disturbance, and the imaginary part of the Landau constant
$\textrm{Im} \{a_1\}$
are always zero. Hence from (4.1) and (4.15), we expect the steady-state equilibrium solutions near the critical point
$ \delta _{Ra}\in (0,0.75) $
as
$t\to \infty$
. Here,
$c_r=\textrm{Im} \{a_1\}=0$
can be argued physically from the symmetries of the present problem. The real part of the complex wave speed
$c_r$
represents the speed and the direction of propagation of the disturbance in the flow. However, there is no preferential direction for the flow due to the absence of a non-zero base state fluid velocity. In other words, the clockwise and anticlockwise steady rolls are the solutions to the problem, which is possible only if
$c_r=0$
for the most unstable disturbance. Moreover, the present problem is invariant under translation along the horizontal direction due to its infinite extent
$x\in (-\infty ,\infty )$
, and
$\textrm{Im} \{a_1\}$
represents the phase angle of amplitude (see (4.14)), which captures a finite spatial shift along the horizontal direction that leads to
$\textrm{Im} \{a_1\}=0$
.

Figure 4. Effect of particle volume fraction near the bifurcation point: (a) variation of growth rate, (b) real part of Landau constant, (c) equilibrium amplitude, and (d) ratio of equilibrium amplitude and the square root of growth rate with reduced Rayleigh number
$\delta _{Ra}$
, for other parameters kept at
$\delta =0.01$
,
${\textit{Re}}_{\!p}=1$
,
${R}=800$
,
${E}=3385$
and
$ \textit{Pr}=0.71$
.
Table 1. Grid independence test on growth rate
$c_i$
, real part of Landau constant
${\textrm{Re}}\{a_1\}$
and equilibrium amplitude
$A_e$
for
${\textit{Re}}_{\!p}=1$
,
$\delta =0.01$
,
${R}=800$
,
${E}=3385$
,
$ \textit{Pr}=0.71$
,
$\varTheta _{\textit{pt}}=0$
and
$\varPhi _0=\{10^{-5}, 10^{-4}, 10^{-3}\}$
at
$\delta _{Ra}=0.1$
.

The supercritical bifurcation at the critical point leads to an equilibrium amplitude
$A_e$
, which varies with control parameter
$\delta _{Ra}$
without any hysteresis as shown in figure 4(c). Since
$c_i\propto \delta _{Ra}$
, and from (4.9b
) it is clear that
$A_e\propto \sqrt {c_i}$
,
$A_e$
shows a square root dependency on
$\delta _{Ra}$
near the critical point, as shown in figure 4(c). Moreover, with increase in
$\varPhi _0$
, for a given
$\delta_{Ra}$
, the growth rate
$c_i$
decreases with a nearly constant
${\textrm{Re}}\{a_1\}$
for
$\varPhi _0=10^{-6}, 10^{-5}$
, and
$10^{-4}$
. However, for
$\varPhi _0=10^{-3}$
, a significant decrease in
${\textrm{Re}}\{a_1\}$
is observed. Therefore, from (4.9b
) for the equilibrium amplitude,
$A_e$
decreases with an increase in
$\varPhi _0$
. In the present study,
$A_e\sim \sqrt {c_i}$
is the primary assumption in writing perturbation series (4.1) for all the dependent variables, which is shown in figure 4(d). The nature of nonlinearity with supercritical bifurcation is to quench or saturate the exponential growth predicted by the linear analysis (Hoyle Reference Hoyle2006; Cross & Greenside Reference Cross and Greenside2009), and leads to a finite amplitude equilibrium solution. In the present work, the magnitude of nonlinear effects is proportional to
${\textrm{Re}}\{a_1\}$
in the cubic Landau equation (4.7). The growth rate
$c_i$
decreases with an increase in particle volume fraction, as shown in figure 4(a), beyond the critical point. Hence the stability of the system is increasing with an increase in
$\varPhi _0$
. The same physical phenomenon of increasing stabilisation is reflected in the weakly nonlinear regime through an increase in the magnitude of
${\textrm{Re}}\{a_1\}$
, as shown in figure 4(b). Here, the more negative
${\textrm{Re}}\{a_1\}$
, the more rapid quenching/saturation of exponential growth beyond the critical point.
The effect of particle volume fraction
$\varPhi _0$
on the heat transfer measured as the mean value of
$[\![{\textit{Nu}}]\!]$
for
$\delta _{Ra}\in { (0.0,0.1 )}$
is shown in table 2. The Nusselt numbers at the hot
${\textit{Nu}}_h$
and cold
${\textit{Nu}}_c$
surfaces are unequal due to the thermal energy coupling between the particles and fluid. As shown in (4.23), the difference between
${\textit{Nu}}_h$
and
${\textit{Nu}}_c$
is balanced by the convective sensible heat flux by the particles at hot (
$Q^{\prime \prime }_{ph}$
) and cold (
$Q^{\prime \prime }_{pc}$
) surfaces. Here, the particles are introduced into the domain at the cold surface temperature (
$\varTheta _{\textit{pt}}=0$
). Hence sensible heat flux by particles at the cold surface is
$Q^{\prime \prime }_{pc}=0$
. As the particle concentration increases, the difference between the heat fluxes at the hot and cold surfaces for fluid and particles increases. Thus the Nusselt number at the hot surface increases, while at the cold surface, it tends to zero when the particle concentration increases. Physically, this means that the particles absorb all the heat flux emerging from the hot surface while settling down under gravity.
Table 2. Effect of particle volume fraction on heat transfer near the bifurcation point. Here,
$\overline {[\![{\textit{Nu}}]\!]}$
represents the average value of
$[\![{\textit{Nu}}]\!]$
at a given particle volume fraction
$\varPhi _0$
and
$\delta _{Ra}\in {(0,0.1)}$
. Other parameters are kept constant at
$\delta =0.01$
,
${\textit{Re}}_{\!p}=1$
,
${R}=800$
,
${E}=3385$
,
$\varTheta _{\textit{pt}}=0$
and
$ \textit{Pr}=0.71$
.

The variation in the average Nusselt number
$({\textit{Nu}}_h+{\textit{Nu}}_c)/2$
with
$\delta _{Ra}$
near the bifurcation point for different particle volume fractions is shown in figure 5. As expected,
$({\textit{Nu}}_h+{\textit{Nu}}_c)/2$
increases for all
$\varPhi _0$
with an increase in
$\delta _{Ra}$
.

Figure 5. Variation in heat transfer near the bifurcation point at different particle volume fractions and with other parameters kept constant at
$\delta =0.01$
,
${\textit{Re}}_{\!p}=1$
,
${R}=800$
,
${E}=3385$
,
$\varTheta _{\textit{pt}}=0$
and
$ \textit{Pr}=0.71$
. Here, the heat transfer is measured as the average value of
${\textit{Nu}}_h$
and
${\textit{Nu}}_c$
.
5.2. Effect of particle injection temperature
Figure 6 depicts the effect of particle injection temperature
$\varTheta _{\textit{pt}}$
on the growth rate
$c_i$
, the real part of the Landau constant
${\textrm{Re}}\{a_1\}$
, equilibrium amplitude
$A_e$
, and the ratio of equilibrium amplitude and the square root of growth rate at a fixed particle volume fraction
$ \varPhi _0=10^{-4} $
, and other parameters kept at
$\delta =0.01$
,
${\textrm{Re}}_{\!p}=1$
,
${R}=800$
,
${E}=3385$
and
$ \textit{Pr}=0.71$
. Similar to the non-monotonic effect of
$\varTheta _{\textit{pt}}$
on the onset of instability (see figure 3),
$\varTheta _{\textit{pt}}$
shows the non-monotonic effect on the growth rate, the real part of the Landau constant, and the equilibrium amplitude. As described in figure 3, an increase in
$\varTheta _{\textit{pt}}$
favours stability in the initial part, and the further increase in
$\varTheta _{\textit{pt}}$
causes flow to favour instability. Hence this explains the decrease in
$c_i$
in the initial part, and the increase in
$c_i$
in the remaining part of the curves in figure 6(a). Figure 6(b) shows that for all
$\varTheta _{\textit{pt}}\in (-1,2)$
,
${\textrm{Re}}\{a_1\}\lt 0$
, and together with
$c_i\gt 0$
, the flow exhibits a supercritical pitchfork bifurcation. The equilibrium amplitude
$A_e$
has a non-monotonic variation with respect to
$\varTheta _{\textit{pt}}$
: as
$\varTheta _{\textit{pt}}$
increases from
$-1$
,
$A_e$
decreases and reaches a minimum at approximately 1, and again increases. Finally, figure 6(d) shows
$A_e\sim \sqrt {c_i}$
, which is the key underlying assumption while writing the perturbation series of the form given by (4.1).

Figure 6. Effect of particle initial temperature
$\varTheta _{\textit{pt}}$
at
$\delta _{Ra}=0.1$
: (a) variation of growth rate
$c_i$
, (b) variation of the real part of the Landau constant
${\textrm{Re}}\{a_1\}$
, (c) variation of equilibrium amplitude
$A_e$
, and (d) variation of the ratio of equilibrium amplitude and the square root of the growth rate. The other parameters are kept at
$\delta =0.01$
,
${\textit{Re}}_{\!p}=1$
,
${R}=800$
,
${E}=3385$
,
$ \textit{Pr}=0.71$
and
$\varPhi _0=10^{-4}$
.

Figure 7. Time history of amplitude function
$|A|$
at
$\delta _{Ra}=0.1$
with different initial amplitudes
$|A_0|$
at
$R=800$
,
$E=3385$
,
$\varPhi _0=10^{-4}$
,
$\varTheta _{\textit{pt}}=0$
,
${\textit{Re}}_{\!p}=1$
and
$ \textit{Pr}=0.701$
. Here, the amplitude
$|A|$
is normalised by the equilibrium amplitude
$A_e$
.

Figure 8. Secondary flow pattern obtained for particle-laden Rayleigh–Bénard convection: (a,c,e,g,i,k,m) the linear analysis, (b,d,f,h,j,l,n) the nonlinear analysis, at reduced Rayleigh number
$\delta _{Ra}=0.1$
, with critical wavenumber
$k_c\approx 3.77$
, and other parameters
${R}=800$
,
${E}=3385$
,
$\varPhi _0=10^{-4}$
,
$\varTheta _{\textit{pt}}=0$
,
${\textit{Re}}_{\!p}=1$
and
${Pr}=0.701$
.

Figure 9. Secondary flow pattern obtained for particle-laden Rayleigh–Bénard convection at (a,d,g,j,m,p,s)
$t=20$
, (b,e,h,k,n,q,t)
$t=60$
and (c,f,i,l,o,r,u)
$t=320$
, for reduced Rayleigh number
$\delta _{Ra}=0.1$
, with critical wavenumber
$k_c\approx 3.77$
, and other parameters fixed at
${R}=800$
,
${E}=3385$
,
$\varPhi _0=10^{-4}$
,
$\varTheta _{\textit{pt}}=0$
,
${\textit{Re}}_{\!p}=1$
,
${Pr}=0.701$
, and the initial amplitude
$A_0$
is taken as
$0.1A_e$
.
5.3. Secondary flow pattern
Figure 7 shows the variation of perturbation amplitude normalised by the equilibrium amplitude with time for different initial conditions given by (4.11). Due to the supercritical pitchfork bifurcation, for the given control parameter, all the perturbation amplitudes with different initial conditions tend to approach the same equilibrium amplitude as
$t\to \infty$
. The perturbation series given by (4.1) can be rewritten for the secondary flow
$\theta ^\prime (x,z,t)=\theta (x,z,t)-\varTheta _0(z)$
as
where
$A(t)$
is given by (4.15). Using similar expressions for the remaining variables, we obtain the contour plots of secondary flow patterns at
$\delta _{Ra}=0.1$
as shown in figure 8. The secondary flow patterns obtained from the linear stability are shown in the left-hand column, and those obtained from the nonlinear stability analysis are shown in the right-hand column. Due to the presence of particles, the top–bottom symmetry is absent even in the contours obtained from the linear stability analysis. The velocity and temperature profiles for the fluid and particles are qualitatively different due to the momentum and thermal inertia of the particles. It is expected that the distortion of the flow patterns increases with an increase in
$\delta _{Ra}$
due to a corresponding increase in the equilibrium amplitude. The width of the contour corresponding to the vertical component of velocity
$u_z$
is larger for negative values than for positive values. Moreover, at a steady state, the net mass flux across any horizontal plane must be zero to satisfy the continuity equation. Hence the magnitude of the vertical component of velocity in upward flow regions should be higher than that in downward flow.
From the linear stability analysis, we showed that the perturbations in particle volume fractions and divergence of particle velocity field are always zero in the domain (see § A.6). However, when the perturbations are finite, the contribution from the nonlinear terms gives a non-zero distortion function
$\varPhi _1$
, and the second harmonic function
$\mathbb{E}^2$
, which results in the tendency of particles to become spatially non-uniform as shown in figure 8(n). The time evolution of the secondary flow pattern obtained from the present weakly nonlinear analysis before reaching a steady state is shown in figure 9. As shown in figure 7, for initial amplitudes close to zero, the growth amplitude
$A$
in time is approximately exponential with negligible effect of nonlinear terms. However, as time progresses (
$t\gtrsim 80$
), the nonlinear terms become significant and saturate the exponential growth, leading to a steady-state solution. Here, there is no oscillating solution for this particle-laden Rayleigh–Bénard system, because we observed that
$c_r=0$
always for the most unstable mode, which is also reported in the previous work by Prakhar & Prosperetti (Reference Prakhar and Prosperetti2021). Hence the right-hand panels in figure 9 show the steady-state secondary flow patterns, which are exactly the same as the right-hand panels in figure 8.
6. Conclusions
In this work, we have performed a weakly nonlinear stability analysis of particle-laden Rayleigh–Bénard convection. We assume the particles to be initially uniformly distributed in space and moving with the settling velocity. At the top cold surface, new particles are injected at a uniform particle volume fraction with their terminal velocity, and settled particles at the bottom hot surface are collected. First, using the linear stability, we show that for a given particle size
$\delta$
, initial temperature
$\varTheta _{\textit{pt}}$
and particle Reynolds number
${\textit{Re}}_{\!p}$
, the critical Rayleigh number
${\textit{Ra}}_c$
significantly increases with an increase in particle volume fraction
$\varPhi _0$
, whereas the initial particle temperature
$\varTheta _{\textit{pt}}$
significantly affects the base-state temperature profile and has a non-monotonic effect on
${\textit{Ra}}_c$
with the maximum at
$\varTheta _{\textit{pt}} \sim 1$
.
We derived a Landau equation to study the evolution of the unstable modes in the weakly nonlinear regime in the vicinity of the critical point. In particular, we show that the growth rate reduces with an increase in particle volume fraction. We note that the Landau constant satisfies
${\textrm{Re}}\{a_1\}\lt 0$
for all
$\varPhi _0$
, and together with
$c_i\gt 0$
, it leads to a supercritical pitchfork bifurcation with an equilibrium amplitude
$A_e=\sqrt {-k_cc_i/{\textrm{Re}}\{a_1\}}$
. The amplitude
$A_e$
reduces with an increase in
$\varPhi _0$
.
Further, at steady state, the Nusselt numbers at the cold and hot surfaces are not equal, like in a single-phase Rayleigh–Bénard convection. This difference in Nusselt numbers at the cold and hot surfaces is exactly balanced by the net sensible heat flux convected by the particles; thus with an increase in
$\varPhi _0$
, the difference in the Nusselt numbers at cold and hot surfaces increases. However,
$c_i$
,
${\textrm{Re}}\{a_1\}$
and
$A_e$
have a non-monotonic dependence on
$\varTheta _{\textit{pt}}$
. Moreover, for the entire range of values for
$\varTheta _{\textit{pt}}$
considered in the present work, the analysis showed the supercritical bifurcation for the flow.
Finally, we have analysed the variation in the secondary flow pattern due to the nonlinear interactions of different harmonics. Unlike in single-phase convection, particle-laden Rayleigh–Bénard convection does not show top–bottom symmetry due to the directional settling of the particles under gravity, even in the linear regime. The nonlinear interaction of the fundamental modes generates Reynolds stress or distortion of the base-state temperature and particle volume fraction fields. Interestingly, we observe local clustering of particles due to these nonlinear interactions of fundamental modes. We note that the linear stability analysis does not show particle clustering (see Prakhar & Prosperetti Reference Prakhar and Prosperetti2021).
Declaration of interests
The authors report no conflict of interest.
Appendix A
A.1. Linear disturbance equations
Substitution of the normal modes for all the dependent variables into the linearised governing equations (3.14)–(3.21) leads to a system of linear ordinary differential equations. The resulting equations are represented in the linear operator form as
\begin{align} &\quad \mathscr{L}_x{\left (k,c,\hat {u}_x,\hat {p},\hat {v}_x,\varPhi _0,{\textit{Ra}},{\textit{Pr}},{R},{\textit{St}}_m\right )}\nonumber \\ &\qquad =\textrm{i}ck\hat {u}_x+\sqrt {\frac {\textit{Pr}}{\textit{Ra}}}{\left (\frac {\textrm{d}^2}{\textrm{d}z^2}-k^2\right )}\hat {u}_x-\textrm{i}k\hat {p}-{\left (\frac {\varPhi _0{R}}{{\textit{St}}_m}\right )}{\left (\hat {u}_x-\hat {v}_x\right )}=0, \end{align}
\begin{align} &\quad \mathscr{L}_z{\big (k,c,\hat {u}_z,\hat {p},\hat {\theta },\hat {v}_z,\hat {\phi },v_0,\varPhi _0,{\textit{Ra}},{\textit{Pr}},{R},{\textit{St}}_m\big )}\nonumber \\&\qquad =\textrm{i}ck\hat {u}_z+\sqrt {\frac {\textit{Pr}}{\textit{Ra}}}{\left (\frac {\textrm{d}^2}{\textrm{d}z^2}-k^2\right )}\hat {u}_z-\frac {\textrm{d} \hat {p}}{\textrm{d} z}+\hat {\theta }-{\left (\frac {\varPhi _0{R}}{{\textit{St}}_m}\right )}{\left (\hat {u}_z-\hat {v}_z\right )}-{\left (\frac {{R}v_0}{{\textit{St}}_m}\right )}\hat {\phi }=0, \end{align}
\begin{align} & \quad \mathscr{L}_\theta {\big (k,c,\hat {u}_z,\hat {\theta },\hat {\theta }_{\!p},\hat {\phi },\varTheta _0,\varTheta _{p0},\varPhi _0,{\textit{Ra}},{\textit{Pr}},{E},{\textit{St}}_{\textit{th}}\big )}\nonumber \\&\qquad =\textrm{i}ck\hat {\theta }+\frac {1}{\sqrt {{{\textit{Ra}}{\textit{Pr}}}}}{\left (\frac {\textrm{d}^2}{\textrm{d}z^2}-k^2\right )}\hat {\theta }-\frac {\textrm{d} \varTheta _0}{\textrm{d} z}\hat {u}_z-\frac {{E}\varPhi _0}{{\textit{St}}_{\textit{th}}}{\big (\hat {\theta }-\hat {\theta _{\!p}}\big )}-\frac {{E}{\big (\varTheta _0-\varTheta _{p0}\big )}}{{\textit{St}}_{\textit{th}}}\hat {\phi }=0, \end{align}
for
$z\in { (-1/2,1/2 )}$
.
A.2. Equations at different hormonic components
The equations for the harmonic
$\mathbb{E}^0$
are given as
\begin{align} &\mathcal{L}_z{\big (P_0,\varTheta _0,v_0,\varPhi _0,{R},{\textit{St}}_m\big )} +c_i\,|B|^2\left \{\mathcal{L}_z{\left (P_1,\varTheta _1,-V_1,\varPhi _0,{R},{\textit{St}}_m\right )}+\frac {{R}v_0}{{\textit{St}}_m}\varPhi _1\right \} \nonumber\\&\quad {}+c_i\,|B|^2\left \{2\frac {\textrm{d}}{\textrm{d} z} \left (\tilde {u}_{z10}u_{z10}\right )+\frac {{R}}{{\textit{St}}_m}\big \{\phi _{10}{\left (\tilde {u}_{z10}-\tilde {v}_{z10}\right )}+\tilde {\phi }_{10}{\left (u_{z10}-v_{z10}\right )}\big \}\right \}=\mathcal{O}\big (c_i^2\big ), \end{align}
\begin{align} &\mathcal{L}_\theta {\left (\varTheta _0,\varTheta _{p0},\varPhi _0,{\textit{Ra}},{\textit{Pr}},{E},{\textit{St}}_{\textit{th}}\right )} \nonumber \\&\quad {}+c_i\,|B|^2\left \{\mathcal{L}_\theta {\left (\varTheta _1,\varTheta _{p1},\varPhi _0,{\textit{Ra}},{\textit{Pr}},{E},{\textit{St}}_{\textit{th}}\right )}-\frac {{E}{\left (\varTheta _0-\varTheta _{p0}\right )}}{{\textit{St}}_{\textit{th}}}\varPhi _1\right \} \nonumber \\& \quad {}-c_i\,|B|^2\left \{\frac {\textrm{d} }{\textrm{d} z} \left (\tilde {\theta }_{10}u_{z10}+\theta _{10}\tilde {u}_{z10}\right )+\frac {{E}}{{\textit{St}}_{\textit{th}}}{\left (\phi _{10}{\big (\tilde {\theta }_{10}-\tilde {\theta }_{p10}\big )}+\tilde {\phi }_{10}{\left (\theta _{10}-\theta _{p10}\right )}\right )}\right \}\nonumber\\&\quad=\mathcal{O}\big (c_i^2\big ), \end{align}
\begin{align} &\mathcal{L}_{p\theta }{\left (\varTheta _0,\varTheta _{p0},v_0,{\textit{St}}_{\textit{th}}\right )}+c_i\,|B|^2\left \{\mathcal{L}_{p\theta }{\left (\varTheta _1,\varTheta _{p1},v_0,{\textit{St}}_{\textit{th}}\right )}-V_1\frac {\textrm{d} \varTheta _{p0}}{\textrm{d} z}\right \} \nonumber \\ &\quad {}-c_i|B|^2\left \{v_{z10}\frac {\textrm{d} \tilde {\theta }_{p10}}{\textrm{d} z}+\tilde {v}_{z10}\frac {\textrm{d} \theta _{p10}}{\textrm{d} z}+\textrm{i}k_c{\big (\tilde {v}_{x10}\theta _{p10}-v_{x10}\tilde {\theta }_{p10}\big )}\right \}= \mathcal{O}\big (c_i^2\big ) \end{align}
for
$z\in { (-1/2,1/2 )}$
, where the operators
$\mathcal{L}_z$
,
$\mathcal{L}_\theta$
and
$\mathcal{L}_{p\theta }$
are given by (3.1), (3.2) and (3.3), respectively. The boundary conditions for
$\varTheta _1$
,
$V_1$
,
$\varTheta _{p1}$
and
$\varPhi _1$
are given by
Similarly, the equations for the harmonic
$\mathbb{E}^1$
are given by
\begin{align} &{\left (c_i\right )}^{1/2}B\,\mathscr{L}_x{\left (k_c,c,u_{x10},p_{10},v_{x10},\varPhi _0,{\textit{Ra}},{\textit{Pr}},{R},{\textit{St}}_m\right )}\nonumber \\ &\quad {}+ {\left (c_i\right )}^{3/2}B\,|B|^2\,\mathscr{L}_x{\left (k_c,c,u_{x11},p_{11},v_{x11},\varPhi _0,{\textit{Ra}},{\textit{Pr}},{R},{\textit{St}}_m\right )}\nonumber \\ &\quad {}-{\left (c_i\right )}^{3/2}{\left (B\,|B|^2\mathcal{G}_x+\left \{\frac {\textrm{d} B}{\textrm{d} \tau }-k_cB\right \}u_{x10}\right )}=\mathcal{O}\big (c_i^{5/2}\big ), \end{align}
\begin{align} &{\left (c_i\right )}^{1/2}B\,\mathscr{L}_z{\left (k_c,c,u_{z10},p_{10},\theta _{10},v_{z10},\phi _{10},v_0,\varPhi _0,{\textit{Ra}},{\textit{Pr}},{R},{\textit{St}}_m\right )}\nonumber \\ &\quad {}+ {\left (c_i\right )}^{3/2}B\,|B|^2\,\mathscr{L}_z{\left (k_c,c,u_{z11},p_{11},\theta _{11},v_{z11},\phi _{11},v_0,\varPhi _0,{\textit{Ra}},{\textit{Pr}},{R},{\textit{St}}_m\right )}\nonumber \\ &\quad {}-{\left (c_i\right )}^{3/2}{\left (B\,|B|^2\mathcal{G}_z+\left \{\frac {\textrm{d} B}{\textrm{d} \tau }-k_cB\right \}u_{z10}\right )}=\mathcal{O}\big (c_i^{5/2}\big ), \end{align}
\begin{align} &{\left (c_i\right )}^{1/2}B\,\mathscr{L}_\theta {\left (k_c,c,u_{z10},\theta _{10},\theta _{p10},\phi _{10},\varTheta _{0},\varTheta _{p0},\varPhi _0,{\textit{Ra}},{\textit{Pr}},{E},{\textit{St}}_{\textit{th}}\right )}\nonumber \\ &\quad {}+{\left (c_i\right )}^{3/2}B\,|B|^2\,\mathscr{L}_\theta {\left (k_c,c,u_{z11},\theta _{11},\theta _{p11},\phi _{11},\varTheta _0,\varTheta _{p0},\varPhi _0,{\textit{Ra}},{\textit{Pr}},{E},{\textit{St}}_{\textit{th}}\right )}\nonumber \\ &\quad {}-{\left (c_i\right )}^{3/2}{\left (B\,|B|^2\mathcal{G}_\theta +\left \{\frac {\textrm{d} B}{\textrm{d} \tau }-k_cB\right \}\theta _{10}\right )}=\mathcal{O}\big (c_i^{5/2}\big ), \end{align}
\begin{align} &{\left (c_i\right )}^{1/2}B\mathscr{L}_{\phi }{\left (k_c,c,v_{x10},v_{z10},\phi _{10},v_0,\varPhi _0\right )}\nonumber \\ &\quad {}+{\left (c_i\right )}^{3/2}B\,|B|^2\,\mathscr{L}_{\phi }{\left (k_c,c,v_{x11},v_{z11},\phi _{11},v_0,\varPhi _0\right )}\nonumber \\ &\quad {}-{\left (c_i\right )}^{3/2}{\left (B\,|B|^2\mathcal{G}_\phi +\left \{\frac {\textrm{d} B}{\textrm{d} \tau }-k_cB\right \}\phi _{10}\right )}=\mathcal{O}\big (c_i^{5/2}\big ), \end{align}
\begin{align} &{\left (c_i\right )}^{1/2}\mathscr{L}_{px}{\left (k_c,c,u_{x10},v_{x10},v_0,{\textit{St}}_m\right )}+{(c_i )}^{3/2}B|B|^2\,\mathscr{L}_{px}{\left (k_c,c,u_{x11},v_{x11},v_0,{\textit{St}}_m\right )}\nonumber \\ &\quad {}-{\left (c_i\right )}^{3/2}{\left (B\,|B|^2\mathcal{G}_{px}+\left \{\frac {\textrm{d} B}{\textrm{d} \tau }-k_cB\right \}v_{x10}\right )}=\mathcal{O}\big (c_i^{5/2}\big ), \end{align}
\begin{align} &{\left (c_i\right )}^{1/2}B\mathscr{L}_{pz}{\left (k_c,c,u_{z10},v_{z10},v_0,{\textit{St}}_m\right )}+{(c_i)}^{3/2}B|B|^2\,\mathscr{L}_{pz}{\left (k_c,c,u_{z11},v_{z11},v_0,{\textit{St}}_m\right )}\nonumber \\ &\quad {}-{\left (c_i\right )}^{3/2}{\left (B\,|B|^2\mathcal{G}_{pz}+\left \{\frac {\textrm{d} B}{\textrm{d} \tau }-k_cB\right \}v_{z10}\right )}=\mathcal{O}\big (c_i^{5/2}\big ), \end{align}
\begin{align} & \kern-70pt {\left (c_i\right )}^{1/2}B\mathscr{L}_{p\theta }{\left (k_c,c,\theta _{10},v_{z10},\theta _{p10},v_0,\varTheta _{p0},{\textit{St}}_{\textit{th}}\right )}\nonumber \\& \kern-70pt \quad {}+{\left (c_i\right )}^{3/2}B\,|B|^2\,\mathscr{L}_{p\theta }{\left (k_c,c,\theta _{11},v_{z11},\theta _{p11},v_0,\varTheta _{p0},{\textit{St}}_{\textit{th}}\right )}\nonumber \\&\kern-70pt \quad {}-{\left (c_i\right )}^{3/2}{\left (B\,|B|^2\mathcal{G}_{p\theta }+\left \{\frac {\textrm{d} B}{\textrm{d} \tau }-k_cB\right \}\theta _{p10}\right )}=\mathcal{O}\big (c_i^{5/2}\big ) \end{align}
for
$z\in { (-1/2,1/2 )}$
, where the operators
$\mathscr{L}_0$
,
$\mathscr{L}_x$
,
$\mathscr{L}_z$
,
$\mathscr{L}_\theta$
,
$\mathscr{L}_{\phi }$
,
$\mathscr{L}_{px}$
,
$\mathscr{L}_{pz}$
and
$\mathscr{L}_{p\theta }$
are the linear stability operators given by (A1)–(A8) with boundary conditions given by (3.23) for both
$ (u_{x10},u_{z10},\theta _{10},v_{x10},v_{z10},\theta _{p10},\phi _{10} )$
and
$ (u_{x11},u_{z11},\theta _{11},v_{x11},v_{z11},\theta _{p11},\phi _{11} )$
. Here, the scalar functions
$\mathcal{G}_x$
,
$\mathcal{G}_z$
,
$\mathcal{G}_\theta$
,
$\mathcal{G}_{px}$
,
$\mathcal{G}_{pz}$
,
$\mathcal{G}_\phi$
and
$\mathcal{G}_{p\theta }$
are given by (A31)–(A37) defined in § A.3.
The equations for the harmonic
$\mathbb{E}^2$
are given by
\begin{align} & c_iB^2\,\mathscr{L}_x{\left (2k_c,u_{x20},p_{20},v_{x20},\varPhi _0,{\textit{Ra}},{\textit{Pr}},{R},{\textit{St}}_m\right )}\nonumber \\ &\quad {}-c_iB^2\left \{u_{z10}\frac {\textrm{d} u_{x10}}{\textrm{d} z}-u_{x10}\frac {\textrm{d} u_{z10}}{\textrm{d} z}+\frac {{R}\phi _{10}}{{\textit{St}}_m}{\left (u_{x10}-v_{x10}\right )}\right \}=\mathcal{O}\big (c_i^2\big ), \end{align}
\begin{align} & c_iB^2\left \{\mathscr{L}_z{\left (2k_c,c,u_{z20},p_{20},\theta _{20},v_{z20},\varPhi _0,{\textit{Ra}},{\textit{Pr}},{R},{\textit{St}}_m\right )}-\frac {{R}\phi _{10}}{{\textit{St}}_m}{\left (u_{z10}-u_{x10}\right )}\right \}\nonumber \\ &\quad {}=\mathcal{O}\big (c_i^2\big ), \end{align}
\begin{align} &c_iB^2\,\mathscr{L}_{\theta }{\left (2k_c,c,u_{z20},\theta _{20},\theta _{p20},\phi _{20},\varTheta _0,\varTheta _{p0},\varPhi _0,{\textit{Ra}},{\textit{Pr}},{E},{\textit{St}}_{\textit{th}}\right )}\nonumber \\ &\quad {}-c_iB^2\left \{u_{z10}\frac {\textrm{d} \theta _{10}}{\textrm{d} z}-\theta _{10}\frac {\textrm{d} u_{z10}}{\textrm{d} z}+\frac {{E}\phi _{10}}{{\textit{St}}_{\textit{th}}}{\left (\theta _{10}-\theta _{p10}\right )}\right \}=\mathcal{O}\big(c_i^2\big), \end{align}
\begin{align} & c_iB^2\,\mathscr{L}_{\phi }{\left (2k_c,c,v_{x20},v_{z20},\phi _{20},v_0,\varPhi _0\right )} -c_iB^2\left \{2\textrm{i}k_c\phi _{10}v_{x10}+\frac {\textrm{d} }{\textrm{d} z} \left (\phi _{10}v_{z10}\right )\right \}\nonumber \\ &\quad {}=\mathcal{O}\big(c_i^2\big), \end{align}
\begin{align} & c_iB^2\left \{\mathscr{L}_{p\theta }{\left (2k_c,c,\theta _{20},v_{z20},\theta _{p20},v_0,\varTheta _{p0},{\textit{St}}_{\textit{th}}\right )}-{\left (\mbox{i}k_cv_{x10}\theta _{p10}+v_{z10}\frac {\textrm{d} \theta _{p10}}{\textrm{d} z}\right )}\right \} \nonumber \\ &\quad =\mathcal{O}\big (c_i^2\big ) \end{align}
for
$z\in (-1/2,1/2)$
, and the boundary conditions are similar to (3.23). Here,
${}\sim{}$
represents the complex conjugate.
A.3. Scalar functions
In the derivation of the Landau equation, the following scalar functions emerge due to the nonlinear interaction of different modes:
\begin{align} \mathcal{G}_x&=\tilde {u}_{z10}\frac {\textrm{d} u_{x20}}{\textrm{d} z}+u_{z20}\frac {\textrm{d} \tilde {u}_{x10}}{\textrm{d} z}-\textrm{i}k_cu_{x20}\tilde {u}_{x10}+2u_{x20}\frac {\textrm{d} \tilde {u}_{z10}}{\textrm{d} z} \nonumber \\&\quad {}+\frac {{R}}{{\textit{St}}_m}{\big (\varPhi _1\left \{u_{x10}-v_{x10}\right \}+\phi _{20}\left \{\tilde {u}_{x10}-\tilde {v}_{x10}\right \}+\tilde {\phi }_{10}\left \{u_{x20}-v_{x20}\right \}\big )}, \end{align}
\begin{align} \mathcal{G}_z&=\tilde {u}_{z10}\frac {\textrm{d} u_{z20}}{\textrm{d} z}+u_{z20}\frac {\textrm{d} \tilde {u}_{z10}}{\textrm{d} z}-\textrm{i}k_cu_{x20}\tilde {u}_{z10}+2u_{z20}\frac {\textrm{d} \tilde {u}_{z10}}{\textrm{d} z}\nonumber \\&\quad {}+\frac {{R}}{{\textit{St}}_m}{\big (\varPhi _1\left \{u_{z10}-v_{z10}\right \}+\phi _{20}\left \{\tilde {u}_{z10}-\tilde {v}_{z10}\right \}+\tilde {\phi }_{10}\left \{u_{z20}-v_{z20}\right \}-\phi _{10}V_1\big )}, \end{align}
\begin{align} \mathcal{G}_\theta &=\tilde {u}_{z10}\frac {\textrm{d} \theta _{20}}{\textrm{d} z}+u_{z20}\frac {\textrm{d} \tilde {\theta }_{10}}{\textrm{d} z}-\textrm{i}k_cu_{x20}\tilde {\theta }_{10}+2\theta _{20}\frac {\textrm{d} \tilde {u}_{z10}}{\textrm{d} z}+u_{z10}\frac {\textrm{d} \varTheta _1}{\textrm{d} z}\nonumber \\&\quad{}+\frac {{E}}{{\textit{St}}_{\textit{th}}}{\big (\phi _{10}{\left (\varTheta _1-\varTheta _{p1}\right )}+\tilde {\phi }_{10}{\left (\theta _{20}-\theta _{p20}\right )}+\varPhi _1{\left (\theta _{10}-\theta _{p10}\right )}+\phi _{20}{\big (\tilde {\theta }_{10}-\tilde {\theta }_{p10}\big )}\big )}, \end{align}
\begin{align} \mathcal{G}_\phi &=\frac {\textrm{d} }{\textrm{d} z}\big (\tilde {v}_{z10}\phi _{20}\big )+\frac {\textrm{d} }{\textrm{d} z} \big (v_{z20}\tilde {\phi }_{10}\big )+\textrm{i}k_c{\big ( v_{x20}\tilde {\phi }_{10}+\tilde {v}_{x10}\phi _{20}+\varPhi _1v_{x10}\big )}\nonumber \\&\quad {}+\frac {\textrm{d} }{\textrm{d} z} \left (\varPhi _1v_{z10}\right )+\frac {\textrm{d} }{\textrm{d} z} \left (\phi _{10}V_1\right )\!, \end{align}
A.4. Linear adjoint equations
The adjoint of the linear operator
$\mathscr{L}$
is defined as
where
$\boldsymbol{X}^\dagger$
is the eigenvector corresponding to the adjoint operator
$\mathscr{L}^\dagger$
. The definition for the inner product used in (A38) for the two real-valued vector functions
$\boldsymbol{X}(z)=[x_1\quad x_2\quad x_3\quad x_4\quad x_5\quad x_6\quad x_7\quad x_8]^{\textrm T}$
and
$\boldsymbol{Y}(z)=[y_1\quad y_2\quad y_3\quad y_4\quad y_5\quad y_6 y_7\quad y_8]^{\textrm T}$
is given by
\begin{equation} \left \langle \boldsymbol{X},\boldsymbol{Y}\right \rangle =\displaystyle \int _{z=-1/2}^{1/2} \left [\boldsymbol{X}\right ]^{\textrm T}\boldsymbol{Y} \,{\textrm d}z=\int _{z=-1/2}^{1/2} \sum _{j=1}^8 {x}_j(z)\,y_j(z)\,{\textrm d}z. \end{equation}
Using the definition for the adjoint in (A38), the adjoint linear system corresponding to the linear stability (A1)–(A8) is given by
\begin{align} &\mathscr{L}_x^\dagger {\big (k_c,c,u_x^\dagger ,\hat {p}^\dagger ,\hat {v}_x^\dagger ,\varPhi _0,{\textit{Ra}},{\textit{Pr}},{R},{\textit{St}}_m\big )}\nonumber\\&\quad =\textrm{i}ck_c\hat {u}_x^\dagger +\sqrt {\frac {\textit{Pr}}{\textit{Ra}}}{\left (\frac {\textrm{d}^2}{\textrm{d}z^2}-k_c^2\right )}\hat {u}_x^\dagger +\textrm{i}k_c\hat {p}^\dagger -{\left (\frac {\varPhi _0{R}}{{\textit{St}}_m}\right )}\hat {u}_x^\dagger +\frac {\hat {v}_x^\dagger }{{\textit{St}}_m}=0, \end{align}
\begin{align} & \mathscr{L}_z^\dagger {\big (k_c,c,u_z^\dagger ,\hat {p}^\dagger ,\hat {\theta }^\dagger ,\hat {v}_z^\dagger ,\varTheta _0,\varPhi _0,{\textit{Ra}},{\textit{Pr}},{R},{\textit{St}}_m\big )} \nonumber \\&\quad=\textrm{i}ck_c\hat {u}_z^\dagger +\sqrt {\frac {\textit{Pr}}{\textit{Ra}}}{\left (\!\frac {\textrm{d}^2}{\textrm{d}z^2}-k_c^2\!\right )}\hat {u}_z^\dagger -\frac {\textrm{d} \hat {p}}{\textrm{d} z}^\dagger -\frac {\textrm{d} \varTheta _0}{\textrm{d} z}\hat {\theta }^\dagger -{\left (\frac {\varPhi _0{R}}{{\textit{St}}_m}\right )}\hat {u}_z^\dagger +\frac {\hat {v}_z^\dagger }{{\textit{St}}_m}=0, \end{align}
\begin{align} & \mathscr{L}_\theta ^\dagger {\big (k_c,c,\hat {u}_z^\dagger ,\hat {\theta }^\dagger ,\hat {\theta }_{\!p}^\dagger ,\varPhi _0,{\textit{Ra}},{\textit{Pr}},{\textit{St}}_{\textit{th}},{E}\big )} \nonumber \\&\quad=\textrm{i}ck_c\hat {\theta }^\dagger +\frac {1}{\sqrt {{{\textit{Ra}}{\textit{Pr}}}}}{\left (\!\frac {\textrm{d}^2}{\textrm{d}z^2}-k_c^2\!\right )}\hat {\theta }^\dagger +\hat {u}_z^\dagger -\frac {{E}\varPhi _0}{{\textit{St}}_{\textit{th}}}\hat {\theta }^\dagger +\frac {\hat {\theta }_{\!p}^\dagger }{{\textit{St}}_{\textit{th}}}=0, \end{align}
\begin{align} & \mathscr{L}_{\phi }^\dagger {\big (k_c,c,\hat {u}_z^\dagger ,\hat {\theta }^\dagger ,\hat {\phi }^\dagger ,\varTheta _0,v_0,\varTheta _{p0},{\textit{St}}_m,{R},{\textit{St}}_{\textit{th}},{E}\big )} \nonumber \\ &\quad=\textrm{i}ck_c\hat {\phi }^\dagger -v_0\frac {\textrm{d} \hat {\phi }^\dagger }{\textrm{d} z}-\frac {{R}v_0}{{\textit{St}}_m}\hat {u}_z^\dagger -\frac {{E}{\left (\varTheta _0-\varTheta _{p0}\right )}}{{\textit{St}}_{\textit{th}}}\hat {\theta }^\dagger =0, \end{align}
\begin{align} & \mathscr{L}_{px}^\dagger {\big (k_c,c,\hat {u}_x^\dagger ,\hat {v}_x^\dagger ,\hat {\phi }^\dagger ,v_0,\varPhi _0,{R},{\textit{St}}_m\big )} \nonumber \\ &\quad=\textrm{i}ck_c\hat {v}_x^\dagger -v_0\frac {\textrm{d} \hat {v}_x^\dagger }{\textrm{d} z}-\frac {\hat {v}_x^\dagger }{{\textit{St}}_m}-\textrm{i}k_c\varPhi _0\hat {\phi }^\dagger +\frac {\varPhi _0{R}}{{\textit{St}}_m}\hat {u}_x^\dagger =0, \end{align}
\begin{align} & \mathscr{L}_{pz}^\dagger {\big (k_c,c,\hat {u}_z^\dagger ,\hat {v}_z^\dagger ,\hat {\theta }_{\!p}^\dagger ,\hat {\phi }^\dagger ,\varTheta _0,v_0,\varPhi _0,{R},{\textit{St}}_m\big )}\nonumber\\&\quad =\textrm{i}ck_c\hat {v}_z^\dagger -v_0\frac {\textrm{d} \hat {v}_z^\dagger }{\textrm{d} z}-\frac {\hat {v}_z^\dagger }{{\textit{St}}_m}+\varPhi _0\frac {\textrm{d} \hat {\phi }^\dagger }{\textrm{d} z} +\frac {\varPhi _0{R}}{{\textit{St}}_m}\hat {u}_z^\dagger -\frac {\textrm{d} \varTheta _{p0}}{\textrm{d} z}\hat {\theta }_{\!p}^\dagger =0, \end{align}
for
$z\in { (-1/2,1/2 )}$
with boundary conditions given by
The adjoint eigenfunctions are normalised such that
A.5. Numerical procedure
The basic state equations are solved analytically, and the linear and nonlinear stability equations in the present work are solved using the spectral Chebyshev collocation method (Boyd Reference Boyd2001). The underlying system of linear and nonlinear equations, along with the boundary conditions, is transformed into the Chebyshev polynomial domain, i.e.
$[-1,1]$
, using the transformation
$\xi =2z$
. Here, the continuous variable
$\xi$
is discretised onto collocated Gauss–Labatto points
$\xi _j$
given by
where
${N}-1$
is the degree of the Chebyshev polynomial. All derivatives are obtained using the MATLAB differential matrix suite by Weideman & Reddy (Reference Weideman and Reddy2000). The generalised eigenvalue problem of the form
$\mathbb{A}_d\boldsymbol{q}_d=c\mathbb{B}_d\boldsymbol{q}_d$
is solved using the QZ algorithm (Moler & Stewart Reference Moler and Stewart1973) in MATLAB using the eig command for the linear stability analysis. Here,
$\boldsymbol{q}_d$
is a discretised eigenvector, and
$\mathbb{A}_d$
and
$\mathbb{B}_d$
are the discretised complex square matrices. The system of adjoint equations of the linear stability problem is solved using a similar spectral method. In nonlinear stability analysis, we need to solve a system of non-homogeneous equations of the form
$\mathcal{A}{X}=b$
, here
$\mathcal{A}=\mathbb{A}_d-c\mathbb{B}_d$
. For the harmonic
$\mathbb{E}^2$
, the wavenumber
$k_c$
is replaced by
$2k_c$
, and the complex wave speed
$c$
of the fundamental harmonic
$\mathbb{E}^1$
is not an eigenvalue of this system. Hence the matrix
$\mathcal{A}=\mathbb{A}_d-c\mathbb{B}_d$
is non-singular and can be inverted, and gives a unique solution for the harmonic
$\mathbb{E}^2$
. Similarly, the distortion functions also result in a system of non-singular, non-homogeneous equations, which are solved in a similar manner. For the harmonic
$\mathbb{E}^1$
at
$\mathcal{O} (c_i^{3/2} )$
, the system
$\mathcal{A}X_{11}=b$
turns out to be singular, and the existence of the solution is guaranteed by applying the Fredholm alternative. However, the solution
$X_{11}$
is non-unique; hence one of the solutions is obtained using the singular value decomposition (SVD) built into the MATLAB software. All the integrals required for obtaining the Landau constant
$a_1$
and normalisation are evaluated using the Gauss–Chebyshev quadrature formula.
The numerical code developed for solving the linear stability equations is validated by comparing the results with earlier work by Prakhar & Prosperetti (Reference Prakhar and Prosperetti2021). The results generated by the code are in excellent agreement with the published results. For the nonlinear stability calculations, the results remain consistent when the order of the polynomial
$N$
is 50 or more (shown in table 1). Therefore, for all the computations, the polynomial order is taken to be 50.
A.6. Solution to linearised particle concentration
From the above, the linearised equations for the particle volume fraction and the divergence of the particle velocity field for the particle-laden Rayleigh–Bénard convection are given by
for
${ (x,z,t )}\in \left \{\mathbb{R}\times (-1/2,1/2)\times { (0,\infty )}\right \}$
. Since the above equations are linear, and the coefficients are only functions of spatial variables
$(x,z)$
, the general solution can be expressed as
where
$\alpha \in \mathbb{C}$
in general. By substituting (A53)–(A54) in (A51)–(A52), we obtain
where
$f_t(x)=f(x,1/2)$
and
$g_t(x)=g(x,1/2)$
are the boundary conditions at the top boundary
$z=1/2$
. If we specify the boundary conditions
then
$f_t=g_t=0$
, which results in
$\phi '={\partial v_x^{\prime}}/{\partial x}+{\partial v_z^{\prime}}/{\partial z}=0$
for
${ (x,z,t )}\in \{\mathbb{R}\times (-1/2, 1/2)\times { (0,\infty )} \}$
. Hence the particles initialised uniformly with concentration
$\varPhi _0$
remain uniform in the domain forever in the presence of infinitesimal perturbations (Prakhar & Prosperetti Reference Prakhar and Prosperetti2021).








































































































