1. Introduction
Rayleigh–Bénard convection (RBC) is a buoyancy-driven convection flow in which heat is transferred through a thin, wide fluid layer heated from below and cooled from above. As a classical fundamental fluid dynamics problem, it has many applications in natural and industrial fields, including oceanic and atmospheric convections, cooling and heating devices and room ventilations (Chillà & Schumacher Reference Chillà and Schumacher2012). In this phenomenon, the heat transfer mechanism is significantly affected by the activations of characteristic thermal structures such as plumes, thermals, waves, swirls and large-scale circulation (Zocchi, Moses & Libchaber Reference Zocchi, Moses and Libchaber1990). When the Rayleigh number, which is a parameter that measures the driving buoyancy force relative to the resistive forces characterized by viscosity and thermal diffusivity, is sufficiently large, the flow becomes turbulent. Local small structures, the so-called thermal plume structures, occur randomly in space and time and carry heat and momentum across the channel, significantly affecting turbulent transport or mixing (Zocchi et al. Reference Zocchi, Moses and Libchaber1990; Grossmann & Lohse Reference Grossmann and Lohse2000, Reference Grossmann and Lohse2004; Ching et al. Reference Ching, Guo, Shang, Tong and Xia2004; Shishkina & Wagner Reference Shishkina and Wagner2006, Reference Shishkina and Wagner2007; He, Tong & Xia Reference He, Tong and Xia2007; Zhou, Sun & Xia Reference Zhou, Sun and Xia2007; Shishkina & Wagner Reference Shishkina and Wagner2008; Zhou & Xia Reference Zhou and Xia2010; Huang et al. Reference Huang, Kaczorowski, Ni and Xia2013; Park & Lee Reference Park and Lee2015). Major concerns of these studies have been to determine the scaling relations of heat and turbulent transport with the following flow parameters: Rayleigh number, Prandtl number and geometry. Therefore, the thermal and vortical properties of thermal plume structures are important in this single-phase RBC problem (Shishkina & Wagner Reference Shishkina and Wagner2006, Reference Shishkina and Wagner2007, Reference Shishkina and Wagner2008; Zocchi et al. Reference Zocchi, Moses and Libchaber1990; Grossmann & Lohse Reference Grossmann and Lohse2000, Reference Grossmann and Lohse2004; Ching et al. Reference Ching, Guo, Shang, Tong and Xia2004; He et al. Reference He, Tong and Xia2007; Zhou et al. Reference Zhou, Sun and Xia2007; Zhou & Xia Reference Zhou and Xia2010; Huang et al. Reference Huang, Kaczorowski, Ni and Xia2013; Park & Lee Reference Park and Lee2015).
When small heavy particles or bubbles are laden in turbulent RBC, they can influence turbulence through sedimentation or rising motions (Oresta et al. Reference Oresta, Verzicco, Lohse and Prosperetti2009; Zhong, Funfschilling & Ahlers Reference Zhong, Funfschilling and Ahlers2009; Schmidt et al. Reference Schmidt, Oresta, Toschi, Verzicco, Lohse and Prosperetti2011; Lakkaraju et al. Reference Lakkaraju, Schmidt, Oresta, Toschi, Verzicco, Lohse and Prosperetti2011, Reference Lakkaraju, Stevens, Oresta, Verzicco, Lohse and Prosperetti2013; Oresta & Prosperetti Reference Oresta and Prosperetti2013; Prakhar & Prosperetti Reference Prakhar and Prosperetti2021; Raza, Hirata & Calzavarini Reference Raza, Hirata and Calzavarini2024; Wu et al. Reference Wu, Karzahubayev, Shen and Wang2024). Most previous studies have concentrated on the modulation of heat transfer rather than on revealing the complex interaction mechanisms between different phases. In previous studies (Oresta et al. Reference Oresta, Verzicco, Lohse and Prosperetti2009; Zhong et al. Reference Zhong, Funfschilling and Ahlers2009; Schmidt et al. Reference Schmidt, Oresta, Toschi, Verzicco, Lohse and Prosperetti2011; Lakkaraju et al. Reference Lakkaraju, Schmidt, Oresta, Toschi, Verzicco, Lohse and Prosperetti2011, Reference Lakkaraju, Stevens, Oresta, Verzicco, Lohse and Prosperetti2013), thermal and mechanical effects of bubbles or droplets on heat transfer were investigated in detail. As an extension of the previous results, Oresta & Prosperetti (Reference Oresta and Prosperetti2013) simulated settling particles in a weakly turbulent flow in a slender cylindrical domain. They reported that heat transfer is modified by thermal and mechanical coupling between the particles and turbulence, depending on the particle radius. Although a cylindrical system with a small aspect ratio has some advantages in terms of easily increasing the Rayleigh number, cells with a small aspect ratio have a significant influence on the large-scale flow structure (Wu & Libchaber Reference Wu and Libchaber1992; Funfschilling et al. Reference Funfschilling, Brown, Nikolaenko and Ahlers2005; Hartlep, Tilgner & Busse Reference Hartlep, Tilgner and Busse2005; Sun et al. Reference Sun, Ren, Song and Xia2005; Niemela & Sreenivasan Reference Niemela and Sreenivasan2006; Du Puits, Resagk & Thess Reference Du Puits, Resagk and Thess2007). Recently, Kim et al. (Reference Kim, Nam, Shen, Lee and Chamorro2020, Reference Kim, Kim, Kang, Nam, Lee and Chamorro2021) investigated the dynamics of air bubbles in RBC in a cavity experimentally and numerically. Although they focused on the dispersion of bubbles, the behaviour of the bubbles was highly dependent on the large-scale circulation induced by the geometry of the tank, such as the order-one aspect ratio. MacMillan. et al. (Reference MacMillan., Shaw, Cantrell and Richter2022) reported direct numerical simulation (DNS) results to simulate cloud conditions in the Pi chamber with supersaturation conditions, in which they focused on the behaviour of the aerosol concentration injected into RBC. Recently, the interaction between finite-sized particles and convection in a cavity was numerically investigated using the lattice Boltzmann method (Wu et al. Reference Wu, Karzahubayev, Shen and Wang2024).
Investigations on particle-laden RBC within the large-aspect-ratio domain are rare. Park, O’Keefe & Richter (Reference Park, O’Keefe and Richter2018) reported DNS results of particle-laden RBC in a relatively thin layer with the aspect ratio of
$3:3:1$
. They demonstrated how the thermal and momentum effects of non-isothermal particles affect the flow field through the Nusselt number and turbulent kinetic energy. The major reason for their finding that the thermal interaction affected the flow field more than the momentum interaction was that the range of the settling velocity of the particles considered was much smaller than the buoyancy velocity scale, such that the feedback force to the fluid by the particles was very weak, while the settling time was long enough for the thermal interaction. Recently, linear stability analysis of particle-laden convection in the infinite aspect ratio situation has been attempted (Prakhar & Prosperetti Reference Prakhar and Prosperetti2021; Raza et al. Reference Raza, Hirata and Calzavarini2024), revealing that the mechanical interaction between settling particles and fluid in convection tends to stabilize convection. Therefore, the interaction between laden particles and the background convection flow in a thin layer between two infinite plates has not been fully studied, particularly when momentum coupling due to settling particles is not trivial.
Furthermore, the spontaneously formed cellular patterns in RBC have been the longest-running concern in thermal convective flows since Bénard observed hexagonal cells at a fluid layer of which the top surface was free. However, if the top surface is covered with a solid lid, the boundary condition is symmetric, and the convective patterns for a wide range of parameters represent the collection of rolls. Furthermore, as the Rayleigh number increased, the roll patterns became increasingly chaotic and irregular. In this paper, we report the discovery from DNSs that sedimenting heavy particles create large-scale polygonal cell structures in a turbulent regime. Polygonal cell structures are observable only when thermocapillary effect, temperature-dependent viscosity, large Prandtl number or asymmetric boundary conditions such as no-slip bottom wall and free surface top wall are present in low Rayleigh number RBC (Assenheimer & Steinberg Reference Assenheimer and Steinberg1996; Clever & Busse Reference Clever and Busse1996; Busse & Clever Reference Busse and Clever1998; Getling Reference Getling1998; Hartlep et al. Reference Hartlep, Tilgner and Busse2005). Berdnikov & Kirdiashkin (Reference Berdnikov and Kirdiashkin1979) have reported an observation of hexagonal cells in an experiment with particles suspended for visualization at a very low Rayleigh number. They investigated the spatial structure of the cellular flows of ethyl alcohol between two rigid isothermal boundaries that were rendered visible by aluminium particles. They observed that a high concentration of solid particles temporarily transformed the initial roll pattern of the flow structure into polygonal cells. In connection with this experimental study, Wollkind & Zhang (Reference Wollkind and Zhang1994a
,
Reference Wollkind and Zhangb
) proposed a nonlinear stability analysis of the effects of suspended particles on RBC in the context of a two-fluid formulation. However, these studies focused on stability analysis near the critical regime or at Rayleigh numbers below
$20\,000$
. For single-phase convective flows, Bailon-Cuba, Emran & Schumacher (Reference Bailon-Cuba, Emran and Schumacher2010) and Emran & Schumacher (Reference Emran and Schumacher2015) have observed polygonal cell structures and spiral defects in turbulent RBC in a large-aspect-ratio cylindrical container. However, polygonal cell structures caused purely by the particles in turbulent RBC have not yet been reported.
This study concerns the mechanism of formation of polygonal structures by particle feedback force in soft turbulence on RBC, which has not been investigated yet in detail. Furthermore, the effect of particles on heat transport after the polygonal cell structures reached a steady state was also studied. Section 2 introduces numerical methods used in our study, and the results are presented in § 3 with a possible mechanism for the formation of the polygonal structures and energy budget analysis to shed some idea on the interaction between particles and fluid. Finally, conclusions are presented in § 4.
2. Numerical model
The governing equations for RBC under the standard Boussinesq approximation, modified by the feedback force of the inertial particles, are given by



where
$u_{i}$
is the fluid velocity;
$p$
and
$T$
are the fluid pressure and temperature, respectively; and
$\rho$
,
$\nu$
,
$\beta$
,
$\kappa$
and
$g$
are the fluid density, kinematic viscosity, isobaric thermal expansion coefficient, thermal diffusivity of the fluid and magnitude of gravitational acceleration, respectively. Here
$x$
and
$z$
denote the horizontal and
$y$
vertical directions, respectively. The feedback force
$f_{i}$
, defined as the force exerted on the fluid by the particles, is implemented using a point-force approximation because the particles are assumed to be smaller than the Kolmogorov length scale of the particle-free flow. Thus, the feedback force is represented as the sum of the hydrodynamic drag force acting on the kth particle,
$F_{D}^{k}$
, over the particles around a given grid point,

where
$m_{p}$
is the mass of an individual particle and
$\boldsymbol{x}$
and
$\boldsymbol{x}_{p}^{k}$
are the positions at the Eulerian grid point and Lagrangian particle locations, respectively. In this study, we did not consider the thermal interaction between the particles and fluid to focus on the mechanical interactions. The fluid and particle properties in real units and real dimensions of the considered domain are listed in table 1.
Table 1. Parameters and properties of fluid and particles used in the simulations for
$Ra=10^6$
.

In this study, we considered small, undeformable spherical particles, and the density ratio of a particle to a fluid
$ ( \rho _{p} / \rho _{\!f} )$
was
$710.853$
, which is the value for a water droplet and air at approximately
$1$
atm. Owing to the large density difference between the two phases, additional forces on the particles, including the pressure gradient, added mass and Basset forces, are neglected; thus, the Stokes drag and gravity are the two dominant forces acting on the particles (Armenio & Fiorotto Reference Armenio and Fiorotto2001; Eaton Reference Eaton2009). The equations for the motion of the kth particle are expressed as


where
$d_{p}$
is the particle diameter,
$\tau _{p}=d_{p}^{2}\rho _{p}/(18\rho \nu )$
is the particle response time and
$v_{p,i}^{k}$
and
$u_{i}^{k}$
are the particle velocity and fluid velocity at the particle location, respectively. Alternatively, the fluid velocities are
$u$
,
$v$
and
$w$
and the particle velocities are
$u_{p}$
,
$v_{p}$
and
$w_{p}$
. In the nonlinear correction of the drag force
$1+0.15{\textit{Re}}_{p}^{0.687}$
, the particle Reynolds number is defined as
${\textit{Re}}_{p}=d_{p} | \boldsymbol{v}_{p}^{k} - \boldsymbol{u}^{k} |/\nu$
.
A widely extended horizontal fluid layer between two rigid plates, in which the upper one is maintained at temperature
$T_{0}$
and the lower one is maintained at temperature
$T_{0}+\Delta T$
, is considered. Periodic boundary conditions were imposed in the horizontal direction with an aspect ratio
$\varGamma =L/D$
where
$D$
and
$L$
are the distances between the two rigid plates and the horizontal length, respectively. A pseudospectral method (Moser, Moin & Leonard Reference Moser, Moin and Leonard1983; Kerr Reference Kerr1996; Park & Lee Reference Park and Lee2015; Lee & Lee Reference Lee and Lee2015, Reference Lee and Lee2019) is used to solve the governing equations numerically. The space is discretized with a Chebyshev-tau method in the vertical direction and dealiased Fourier expansion in the horizontal direction. Time advancement was performed using a third-order Runge–Kutta scheme for advection and buoyancy terms, coupled with a Crank–Nicolson scheme for viscous terms (Park & Lee Reference Park and Lee2015; Jang & Lee Reference Jang and Lee2018; Kang & Lee Reference Kang and Lee2022).
The Lagrangian tracking of particles is performed by solving (2.5) and (2.6), in which the fluid velocity at the particle locations,
$u_{i}^{k}$
, is interpolated using a four-point Hermite interpolation scheme in the horizontal direction and a fifth-order Lagrangian polynomial interpolation in the vertical direction (Choi, Yeo & Lee Reference Choi, Yeo and Lee2004; Lee & Lee Reference Lee and Lee2015). Time-marching of the equations was performed using the third-order Runge–Kutta scheme. The initial flow field is a fully developed soft turbulence of convection, and at that time, the particles adapt to the flow sufficiently because of the precalculation of the one-way interaction between the fluid and the particles. Periodic boundary conditions were applied to the particles moving outside the side boundary of the domain. To maintain a constant mass loading of the particles in each simulation case, particles reaching the bottom plate were removed, and the same number of particles were injected from the top plate at a random position with settling velocity. The problem of setting the particle boundary conditions is somewhat ambiguous because of the numerical effect of particles that do not adapt to the surrounding flow after they are injected. One possible method for setting the particle injection velocity is to use the value of the surrounding fluid. However, this was accompanied by a thick, dense layer of particles near the top plate, as shown by the time- and area-averaged number densities of particles in Oresta & Prosperetti (Reference Oresta and Prosperetti2013). This particle-rich layer increased the numerical error by suppressing the horizontal movement of the flow near the top plate. To avoid this, the settling velocity of the particles has been suggested in studies on thermal convection problems (Oresta & Prosperetti Reference Oresta and Prosperetti2013; Gereltbyamba & Lee Reference Gereltbyamba and Lee2018, Reference Gereltbyamba and Lee2019). Additionally, Park et al. (Reference Park, O’Keefe and Richter2018) removed particles reaching the bottom layer and injected them in a randomly chosen location in the lower
$10\,\%$
of the domain with the same velocity as the existing particles; however, this method was not chosen because our study focused on the effect of settling particles on the entire flow domain.
In our simulation, the main dimensionless parameters were the Rayleigh number
$Ra=\beta g D^{3} \Delta T / (\kappa \nu )$
, Prandtl number
$Pr=\nu /\kappa$
and aspect ratio
$\varGamma =L/D$
. They were
$Ra=10^{6}, 10^7$
and
$10^8$
,
$Pr=0.7$
, and
$\varGamma =6$
for
$Ra=10^6$
and
$10^7$
, and
$\varGamma =4$
for
$Ra=10^8$
. The number of grids in the
$x$
,
$y$
and
$z$
directions was
$128\times 129\times 128$
for
$Ra=10^6$
and
$256\times 129\times 256$
for
$Ra=10^7$
and
$384\times 129\times 384$
for
$Ra=10^8$
, which satisfied the global grid resolution requirement for an order-one Prandtl number,
$\kappa _{max}\eta _{K}\leqslant \ 1$
with
$\kappa _{max}$
and
$\eta _K$
denoting the maximum horizontal wavenumber and the Kolmogorov length scale, suggested by Grötzbach (Reference Grötzbach1983), and the boundary layer resolution,
$N_{th}\approx 0.35Ra^{0.15}$
, suggested by Shishkina et al. (Reference Shishkina, Stevens, Grossmann and Lohse2010). As shown in table 2, the resolution test confirmed that the selected resolutions are sufficient to yield converged Nusselt number, the maximum dissipation rate and the minimum Kolmogorov length scale, which are the most sensitive quantities related with the small scale motions. Several previous simulations that followed the resolution criterion provided reliable results for soft-to-hard convective turbulence simulations (Kerr Reference Kerr1996; Bailon-Cuba et al. Reference Bailon-Cuba, Emran and Schumacher2010; Park & Lee Reference Park and Lee2015).
Table 2. Convergence of Nusselt number, maximum dissipation rate and minimum Kolmogorov length scale with the grid resolution for
$Ra=10^6, 10^7$
and
$10^8$
. Dissipation rate is non-dimensionalized by the free-fall velocity and the channel height.

The parameters used in particle-laden simulations for
$Ra=10^6$
are listed in table 3. Three differently sized particles are considered throughout the paper:
$d_{p}/D=0.0015$
,
$0.003$
,
$0.0045$
, which are referred to as small, intermediate and large particles, respectively. The particles were loaded into the carrier phase with particle mass fractions of
$\varPhi _{m}=0.03, 0.06, 0.09, 0.12, 0.18$
. Three Stokes numbers are defined: one based on the fluid free-fall velocity
$U_{\!f}=\sqrt {\beta g \Delta T D}$
; one normalized by the viscosity; the other based on the Kolmogorov time scale of the particle-free flow
$\sqrt {\nu /\epsilon }$
,

where
$\epsilon$
denotes the globally averaged energy dissipation rate. The particle Reynolds number
${\textit{Re}}_{p}$
observed for the largest particle
$d_{p}=0.0045D$
is less than
$50$
; therefore, using the nonlinear drag correction
$1+0.15{\textit{Re}}_{p}^{0.687}$
is reasonable in the particle motion equation (2.5). As a reference value for settling particles, terminal velocity normalized by free-fall velocity
$v_{t}/U_{\!f}$
is given. The terminal velocity is the speed of a particle when its weight is balanced by the resistance of the fluid. For
$Ra=10^7$
and
$10^8$
, the parameters are provided in Appendix A.
Table 3. Particle parameters for
$Ra=10^6$
:
$St_{\!f}$
and
$St_{K}$
are the particle Stokes numbers based on free-fall time and Kolmogorov time scales, respectively;
$v_{t}/U_{\!f}$
is the dimensionless terminal velocity;
$N_{p}$
is the total number of particles;
$\varPhi _{m}$
is the particle mass loading. Here
$\phi _m = \varPhi _m S$
with
$S$
denotes the non-dimensionalized settling velocity,
$v_t D/\nu$
;
$n$
is the number of cells in the domain for the cell structures;
$\lambda$
and
$\kappa$
are the length scale quantifying the size of cells or rolls and the corresponding wavenumber, respectively, as defined by
$\lambda = ( L_x L_z/2n )^{1/2}, k =2\pi /\lambda$
for the cell structures. For the roll structures,
$\lambda$
is the lateral size of a pair of rolls.

3. Results and discussion
3.1. Time-averaged flow structures
A study by Hartlep et al. (Reference Hartlep, Tilgner and Busse2005) reveals the spatial and temporal changes of structures in RBC for Rayleigh numbers up to
$10^{7}$
and Prandtl number in the range of
$0.7$
and
$60$
for large-aspect-ratio cells. In particular, when
$Pr\approx 1$
, straight rolls occupy the domain at the onset of convection. Subsequently, the regularly organized roll-like structures are replaced by a random array of transient plumes as the Rayleigh number increases. When the Rayleigh number increases beyond
$10^{4}$
, the flow becomes chaotic. In our study, at parameters
$Ra=10^{6}$
,
$Pr=0.7$
and
$\varGamma =6$
, the flow regime was soft turbulence, in which the flow was chaotic and the thermal plumes moved randomly. As mentioned by Hartlep, Tilgner & Busse (Reference Hartlep, Tilgner and Busse2003), a long-term average of the fluid velocities converges to zero because the flow runs in different directions at different times. Typical instantaneous temperature distributions caused by thermal plumes are shown in figures 1(a) and 1(b) in terms of isosurfaces and two-dimensional distribution at the central plane, respectively. The red structures drawn on the isosurface at
$T/\Delta T=0.8$
indicate that hot plumes converge horizontally in a sheet-like pattern near the bottom plate and create a fine network across the plate, ejecting thermal plumes vertically, as explained by Zocchi et al. (Reference Zocchi, Moses and Libchaber1990). Then detached thermal plumes accelerated vertically and crossed the bulk region in the shape of a mushroom. Finally, they impinge on the top plate and spread out horizontally. Owing to the symmetric nature of RBC, cold plumes behave similarly in opposite directions. Therefore, the distribution of hot rising and cold falling fluids is symmetric in the vertical direction, as shown in figures 1(a) and 1(b), and this symmetry is referred to as the up–down symmetry.
However, when particles are laden in the flow, we find that a polygonal cellular pattern emerges, as shown in the examples given in figures 1(c) and 1(d). The particle size and mass loading were
$d_{p}=0.003D$
and
$\varPhi _{m}=0.12$
. In this example, the observed pattern consists of square cells in which hot fluid soars up in the centre of each cell and cold fluid sinks in the peripheral region of the cell, breaking the up–down symmetry. This cellular pattern persisted over time. This stationary pattern was caused by the settling of the particles. The occurrence of a hexagonal pattern under symmetric boundary conditions was experimentally observed by Berdnikov & Kirdiashkin (Reference Berdnikov and Kirdiashkin1979). In their experiment, aluminium particles of the size of 10 microns were added for visualization to the 4 mm ethyl alcohol layer with an aspect ratio larger than
$30$
at Rayleigh number of
$10^{4}$
. They observed that the agitated solid particles, while settling, altered the flow pattern to a hexagonal cell structure, with a descending flow in the centre of each cell. Although their Rayleigh number is much lower than ours of
$10^6$
and their Prandtl number is much higher than ours, their hexagonal cells are the only examples of polygonal patterns caused by settling particles reported thus far.

Figure 1. Instantaneous temperature distribution: (a,b) of a particle-free flow; (c,d) of a particle-laden flow at
$Ra=10^6$
when
$\varPhi _{m}=0.12$
,
$d_{p}/D=0.003$
. Isosurface at
$T/\Delta T=0.8$
with cross-sectional distribution at selected vertical planes are shown in (a) and (c). Temperature distribution at the central plane of
$y/D=0.5$
is shown in (b) and (d).

Figure 2. Isosurfaces of time-averaged temperature at
$Ra=10^6$
. Hot thermal plumes are drawn in red colour at
$T/\Delta T = 0.8$
and cold ones are in blue colour at
$T/\Delta T = 0.2$
. The two isosurfaces are more separately drawn than real distance for better visualization.

Figure 3. Contours of time-averaged vertical velocity at an
$xz$
-plane of
$y/D=0.5$
at
$Ra=10^6$
. The red area is where the rising flow occurs, and the blue area is where the falling flow occurs. The range of contours is between
$-0.2U_{\!f}$
(blue) and
$0.2U_{\!f}$
(red).
The polygonal cellular patterns observed in our simulations for the three particle sizes and three mass loadings at
$Ra=10^6$
are illustrated on the isosurfaces of the time-averaged temperature and two-dimensional distribution of the time-averaged vertical velocity at the half-channel height, as shown in figures 2 and 3, respectively. Time averaging for approximately
$1200D/U_{\!f}$
yielded a smooth distribution of isosurfaces at
$T/\Delta T=0.2$
(blue) and 0.8 (red). When a polygonal cell pattern was observed for
$d_p/D = 0.0015$
(small) and 0.003 (intermediate), the hot plumes were always concentrated at the centre of each cell, whereas the cold plumes were distributed on the periphery of each cell, as shown in figure 2. When roll structures were observed for
$d_p/D= 0.0045$
(large), straight rolls were observed for
$\varPhi _m=0.06$
whereas crooked rolls and straight but knotted rolls were obtained for
$\varPhi _m=0.12$
and 0.18, respectively. From the vertical velocity distribution at the channel centre height in figure 3, the shape of the cells could be better identified, and a polygonal structure was observed for cases with small and intermediate particles (
$d_p/D=0.0015$
and 0.003), whereas a roll structure was found for large particles (
$d_p/D=0.0045$
). For all particle sizes, the sizes of the cells or rolls tended to decrease as the mass loading increased. The polygonal cell structures indicated that the up–down symmetry was broken, whereas in the roll structures, the symmetry was maintained. For
$Ra=10^7$
and
$10^8$
, similar polygonal cell structures were observed as illustrated in Appendix A, confirming that this kind of polygonal cell structures is not a phenomenon observable only for low Rayleigh numbers.
To quantify the size of the observed cells or rolls, we introduce a length scale
$\lambda$
and a corresponding wavenumber
$k$
for the cell structures, defined as

where
$n$
is the number of polygonal cells in the entire domain. For roll structures,
$\lambda$
is defined as the horizontal wavelength, which is the lateral size of a pair of rolls. The values of
$n, \lambda$
and
$k$
for all cases considered in this study are listed in table 3. For reference, the corresponding values of
$\lambda$
and
$k$
for the most unstable mode of unladen RBC at the critical Rayleigh number
$Ra_{c} (=1708)$
from linear stability analysis are
$\lambda _c/D=2.016$
and
$k_c D=3.117$
, respectively (Drazin & Reid Reference Drazin and Reid1981). To get some clues on why the settling particles induce such cell or roll structures, we carried out the linear stability analysis of a particle-laden RBC as provided in Appendix B. Because the particle Stokes number
$St (=\tau _p \nu /D^2)$
is much smaller than one as shown in table 3, the so-called flow of particles exists, guaranteeing a smooth distribution of the particle velocity (Fouxon et al. Reference Fouxon, Park, Harduf and Lee2015). This allows the particle feedback force that is the Stokes drag force to be explicitly expressed in terms of the local slip velocity field, as discussed in Appendix B. We found that the main parameter quantifying the effect of particles on fluid motion in the linear stability analysis was
$\phi _m (=\varPhi _m S)$
, where
$S$
denotes the non-dimensionalized particle settling velocity
$\tau ^{\prime}_p g D/\nu$
, which is listed in table 3. The stability analysis indicates that the critical Rayleigh number increases with
$\phi _m$
as shown in Appendix B. Therefore, settling particles tend to stabilize the flow, which is consistent with the previous findings by Prakhar & Prosperetti (Reference Prakhar and Prosperetti2021) and Raza et al. (Reference Raza, Hirata and Calzavarini2024). For the range
$\phi _m = 200 - 2000$
considered in the study,
$Ra_c \simeq 35\,000 - 350\,000$
. The Rayleigh number of our cases, however, is
$Ra=10^6$
, which is not much higher than
$Ra_c$
, suggesting that laminar cells or roll structures are preferably formed because the turbulence is expected to be weak. However, the critical wavenumber predicted by the stability analysis is relatively constant at
$k_c D \simeq 3.7$
or
$\lambda _c/D\simeq 1.7$
for the considered range of
$\phi _m$
as shown in table 7 and figure 18(b), which cannot capture the behaviour of
$k D$
varying with the mass loading listed in table 3. Another limitation of the linear stability analysis is the indeterminacy of the cell shape. Despite these limitations, the linear stability analysis suggests that the settling particles stabilize the flow and thus suppress turbulence, possibly leading to the formation of cell or roll structures. The value of
$\lambda _c/D (=1.7)$
obtained from the linear stability analysis is also compared in the figure 4, indicating that a comparable
$\lambda /D$
is observed only for
$d_p/D=0.0015$
and 0.003 at
$\varPhi _m=0.18$
for which the total number of particles is large. It can be conjectured that the discrete nature of the particle feedback force, which cannot be reflected in the linear stability analysis, is responsible for the larger cell or roll sizes compared with those predicted by the linear stability analysis.

Figure 4. The cell size
$\lambda /D$
shown in table 3 as a function of
$\varPhi _m$
at
$Ra=10^6$
. Fitting lines,
$\lambda /D \sim 0.5 \varPhi _m^{-2/3}$
(solid line) for the cell structure and
$\lambda /D \sim 1.5 \varPhi _m^{-1/3}$
(dashed line) for the roll structure are provided to guide the decreasing trend. The value of
$\lambda /D =1.7$
proposed by the linear stability analysis is shown together in the dash–dot line.

Figure 5. Statistics of time-averaged field when mass loading
$\varPhi _{m}=0.06$
,
$0.12$
and
$0.18$
at three different particle sizes at
$Ra=10^6$
. (a) Volume-averaged temperature in the whole flow domain, (b) horizontally averaged temperature, (c) horizontally averaged horizontal components of velocity, (d) horizontally averaged vertical component of velocity. Red, blue and green coloured symbols or lines denote data for small (
$d_p/D=0.0015$
), intermediate (
$d_p/D=0.003$
) and large (
$d_p/D=0.0045$
) sized particles, respectively. Solid, dashed and dotted lines for each colour indicate mass loadings of 0.06, 0.12 and 0.18, respectively. These legends apply to all figures throughout the paper.
For the quantitative analysis of the shape of cells or rolls, figures 5(a) and 5(b) provide the statistics of the time-averaged temperature
$\overline {T}$
defined by

from which the horizontal and volume averages are defined by

Throughout this paper, the overline and bracket represent the time-and space-averaged operations, respectively. The values of the volume-averaged and horizontally averaged temperatures are shown in figures 5(a) and 5(b) clearly indicate a bias of
$\overline {T}$
towards the hot bottom temperature. The volume-averaged temperature for all cases exhibiting cell or roll structures is greater than 0.5 and the horizontal average is greater than 0.5 in most regions except for a small region near the top wall, breaking the up–down symmetry. The asymmetry is more pronounced for cases showing the polygonal cell structure for
$d_p/D=0.0015$
and 0.003 because hot plumes are concentrated in the centre of each cell and cold plumes are distributed along the edge of each cell, as shown in figure 2. This also confirms that the hot plumes are stronger than the cold plumes, resulting in a volume-averaged temperature greater than 0.5. Furthermore, smaller particles induced stronger hot plumes. As more cells were formed with an increase in mass loading, the asymmetry became stronger. However, when the roll structure was observed for
$d_p/D=0.0045$
, a relatively symmetric temperature distribution was observed.
Figures 5(c) and 5(d) show the distribution of the horizontally averaged statistics of the time-averaged velocity field for the cell and roll structures. Because no mean flow was caused by the particles (
$\langle \overline {u_i}\rangle _{V} = \langle \overline {u_i}\rangle _A=0$
), the standard deviation of the velocity components was investigated. The distributions of the standard deviations of the horizontal and vertical velocity components shown in figures 5(c) and 5(d) clearly indicate active horizontal motion near the walls and mainly vertical motion in the core, which is a typical behaviour of the cell or roll structure. A symmetric distribution was maintained for the roll structure (
$d_p/D=0.0045$
), whereas an asymmetric distribution was observed for the cell structures (
$d_p/D=0.0015$
and 0.003). For
$d_p/D=0.0015$
, the positions exhibiting the maximum vertical velocity and the minimum horizontal velocity were found slightly below the channel centre. Although the cases for
$d_p/D=0.0015$
and 0.003 show similar cell structures at the same mass loading, as shown in figure 3, the statistics of the standard deviation show different behaviours, which might be due to detailed differences in the cell structure, such as the shape of the cross-section of hot plumes. For
$d_p/D=0.0015$
, hot plumes in the centre of a cell have a circular cross-section, but for
$d_p/D=0.003$
, the cross-section of the hot plume has a cross shape, which is similar to the shape of a cold plume, as shown in figure 3.
3.2. Turbulence
In this section, we investigate the turbulence behaviour for cell or roll structures defined by the fluctuations in the time-averaged fields through decomposition,

where
$\overline {u}_i(x,y,z)$
and
$\overline {T}(x,y,z)$
correspond to the fields of the stationary cells or roll structures. In the particle-free case,
$\overline {u}_i=0$
. The horizontal and temporally averaged turbulence intensities for the cell and roll structures are shown in figure 6 with those for the particle-free case. For the particle-free case, the intensities in the horizontal and vertical directions were comparable because the randomly occurring hot or cold plumes and distributions were such that the horizontal components peaked near the top and bottom walls. Conversely, the vertical component exhibited a maximum value in the channel centre. The intensities of the cell or roll structures demonstrate that the horizontal components of the velocity are significantly suppressed by the settling particles, whereas the vertical component of the velocity is moderately suppressed or slightly enhanced, depending on the particle size. This observation indicates that hot or cold plumes randomly occurring in the particle-free case are suppressed by the settling particles. Instead, steady plume activity is caused by particles in the form of a cell or roll structure. The relatively less suppressed vertical component is probably due to the discrete distribution of the particle feedback forces. The nearly uniform distribution of the vertical component across the channel and the monotonic increase in the particle size support this conjecture. Furthermore, it is noticeable that all three intensity components are stronger near the bottom wall than near the top wall for the cell structures for
$d_p/D=0.0015$
and 0.003, whereas they are approximately symmetric for the roll structure for
$d_p/D=0.0045$
. By contrast, the temperature fluctuation is not substantially suppressed by the particles, as shown in figure 6(c). The relatively slight changes in mean temperature shown in figure 5 are responsible for this behaviour. The temperature fluctuation near the bottom wall for
$d_p/D=0.0015$
is suppressed more than that near the top wall, which is the opposite trend to the behaviour of the turbulence intensities shown in figures 6(a) and 6(b). A strong asymmetry is observed in the behaviour of the temperature skewness for
$d_p/D=0.0015$
as shown in figure 6(d). Overall, for the roll structure with
$d_p/D=0.0045$
, most turbulence statistics remain symmetric, whereas for the cell structure with
$d_p/D=0.0015$
and 0.003, the turbulence is not symmetric. The observation of turbulence in the cell structure combined with the time-averaged fields indicated that both the time-averaged flow and turbulence near the bottom wall were more intense than those near the top wall, whereas both the time-averaged temperature and fluctuation near the bottom wall were less intense than those near the top wall.

Figure 6. Standard deviations of fluctuating velocity and temperature when mass loading
$\varPhi _{m}=0.06$
,
$0.12$
and
$0.18$
at different particle sizes at
$Ra=10^6$
: (a,b) for velocity components; (c) for temperature; (d) skewness factor of temperature, respectively.

Figure 7. One-dimensional spectra of fluctuating velocity and temperature at the middle plane
$y/D=0.5$
when mass loading
$\varPhi _{m}=0.06$
,
$0.12$
and
$0.18$
at different particle sizes at
$Ra=10^6$
: (a,b,c) for velocity components; (d) for temperature, respectively.
The scale characteristics of the turbulence were investigated through one-dimensional spectra of the turbulent velocity components and temperature fluctuations on the horizontal plane at
$y/D=0.5$
as shown in figure 7. One-dimensional spectra were defined as twice the one-dimensional Fourier transform of the two-point correlation,

where
$R_{ij}$
and
$R_{T'}$
are two-point correlations,

where
$\kappa _x$
denotes the wavenumber in
$x\hbox{-}$
direction.
$E_{11}=E_{u'}, E_{22}=E_{v'}$
and
$E_{33}=E_{w'}$
in figure 7. From a comparison of the energy spectra with those for the particle-free case, it is easily recognized that the settling particles suppress large-scale motions while enhancing small-scale motions in all cases. In particular, small-scale vertical motion is substantially increased owing to the quick settling of heavy particles. This trend became more pronounced as the particle size and mass loading increased. However, the large-scale horizontal motion tended to be more suppressed with mass loading. We conjecture that the modification of the vertical motion by the particles is rather direct and local because of the small size of the particles, whereas the horizontal motion is affected by the modified pressure to satisfy the divergence-free constraint, which is global. However, the spectrum of the temperature fluctuations was not modified significantly by the settling particles, as shown in figure 7(c). Large-scale fluctuations were slightly suppressed in all the cases.
3.3. Heat flux
In this section, we investigate the behaviour of the heat flux modified by the settling particles through mechanical coupling only, because we did not consider thermal coupling in this study. From (2.3), the heat flux vector
$H_j$
is defined as follows:

In particular, the vertical heat flux is important because it determines the Nusselt number,

where

where the last two equalities hold under stationary conditions. The behaviour of
$Nu$
modified by the settling particles is shown in figure 8 and listed in table 3. The settling particles affected the heat flux and augmented
$Nu$
for
$d_p/D=0.003$
and 0.0045, whereas they suppressed
$Nu$
for
$d_p/D=0.0015$
compared with that for the particle-free case. Given that thermal coupling was not considered, the cell or roll structures formed by the particles modified
$Nu$
significantly through momentum coupling. Roll structures seem to enhance heat transfer more effectively than cell structures. However, similar cell structures affect heat transfer differently depending on the size of the particles: intermediate-sized particles (
$d_p/D=0.003$
) enhance heat transfer with an increase in mass loading, whereas small particles (
$d_p/D=0.0015$
) suppress heat transfer. A similar decrease in
$Nu$
by laden small particles was reported in Oresta & Prosperetti (Reference Oresta and Prosperetti2013) and Park et al. (Reference Park, O’Keefe and Richter2018), but the basic mechanism behind the similar phenomenon is different because no cell or roll structure was observed in their studies.

Figure 8. Nusselt number distribution with mass loading for different particle diameters at
$Ra=10^6$
. The letter near each symbol denotes cell structure: ‘R’ for roll structure, ‘S’ for square cell, ‘P’ for pentagonal cell and ‘H’ for hexagonal cell structures. Nusselt number for the case without particles, 8.14 is shown for comparison.

Figure 9. Time-averaged local heat flux at boundary walls for different particle sizes with fixed mass loading of
$\varPhi _{m}=0.06$
, (a,b) for
$d_{p}/D=0.0015$
, (c,d) for
$d_{p}/D=0.003$
, (e,f) for
$d_{p}/D=0.0045$
at
$Ra=10^6$
; (a,c,e) bottom wall, (b,d,f) top wall.

Figure 10. Horizontally averaged local convective heat flux at
$Ra=10^6$
: (a) total convective heat flux, (b) conductive heat flux, (c) heat flux of the time-averaged flow, (d) turbulent heat flux.
The contributions of the cell or roll structures formed by the settling particles to the vertical heat transfer can be investigated in terms of the time-averaged heat flux distribution at the walls, as shown in figure 9. The edges of the hexagonal or square cells were identified by the region showing a locally high heat flux at the bottom wall (figures 9 a and 9 c) and locally low heat flux at the top wall (figures 9 b and 9 d). At the centre of the cells, a locally low heat flux at the bottom wall and a locally high heat flux at the top wall were observed. This suggests that hot plumes contribute strongly to the heat transfer at the top wall in the centre of each cell, whereas cold plumes enhance the heat transfer at the bottom wall along the edge of each cell. However, when the roll structures are formed, such asymmetry between the hot and cold plumes is not observed in the cell structures, as shown in figures 9(e) and 9( f). Furthermore, it appears that the roll structures are more efficient at transporting heat than the cell structures under the same mass loading of particles.
For a more detailed investigation of the heat transport behaviour, the contribution of the turbulent heat flux was separated from the total heat flux through the following decomposition:

The distributions of the normalized components of heat flux are shown in figure 10. The convective transport of heat in the core region (figure 10
a) determines the heat flux at both boundaries or, equivalently, the total heat flux (figure 10
b). Notably, the convective heat transport increases with mass loading for particles with
$d_p/D=0.003$
and 0.0045, whereas it decreases slightly with mass loading for particles with
$d_p/D=0.0015$
. More detailed behaviour can be investigated in the decomposed elements of the convective heat transport, the heat transport due to the cell or roll structures
$\langle \overline {v} \overline {T} \rangle _A$
and the turbulent heat transport
$\langle \overline {v'T'} \rangle _A$
shown in figures 10(c) and 10(d), respectively. The heat transport due to the cell or roll structures was enhanced with mass loading for all particle sizes, whereas turbulent heat transport was not augmented with mass loading. In the case of particles with
$d_p/D=0.0015$
, the turbulent transport was suppressed even with an increase in mass loading, which was responsible for the suppression of the total heat flux. This is obviously due to the suppression of the turbulence of both the vertical momentum and temperature with an increase in the mass loading for
$d_p/D=0.0015$
shown in figures 6(b) and 6(c), unlike the cases for
$d_p/D=0.003$
and 0.0045. Therefore, the decrease in Nusselt number for
$d_p/D=0.0015$
with the mass loading reported in figure 8 is mainly attributed to the suppression of turbulence by the smallest particles, compensating for the augmented transport by the cell or roll structures.
3.4. Formation of polygonal structures
In this section, we investigate the formation of polygonal cell structures resulting from the particle feedbacks. We focused on the initial temporal behaviour of the emerging flow structures owing to the activation of feedback forces induced by sedimenting particles. Figure 11 provides an example of the sequential representation of the flow structures in the midplane temperature distribution, beginning at
$tD/u_{\!f}=0$
, coinciding with the onset of the feedback force action and extending through
$tD/u_{\!f}=300$
, at which point the manifestation of a square-cell structure becomes apparent. It was noticed that randomly distributed hot and cold plumes in the particle-free flow were disturbed by settling particles in the early period, but they became stronger. Soon after, the hot plumes merged, while the cold plumes aligned and organized into a square cell structure. The formed square cells resembled the initial plume distribution, suggesting that the initial field was responsible for determining the phases of the formed square cells. Another test with a different initial field yielded similar square cell structures with different phases.

Figure 11. Temperature distribution at midplane at different times during formation of cell structures when
$d_{p}/D=0.003$
and
$\varPhi _{m}=0.06$
at
$Ra=10^6$
: (a)
$tD/U_{\!f}=0$
, (b)
$tD/U_{\!f}=4$
, (c)
$tD/U_{\!f}=10$
, (d)
$tD/U_{\!f}=39$
, (e)
$tD/U_{\!f}=100$
, ( f)
$tD/U_{\!f}=300$
.

Figure 12. Transient behaviour of (a) magnitude of the horizontal components of velocity
$(u^2 + w^2)/2$
, (b) magnitude of the wall-normal velocity
$v^2$
, (c) Nusselt number, (d) volume-averaged enstrophy at
$Ra=10^6$
. The constant lines at the later time of each figure are the flow component calculated from the time-averaged field in the steady state.
The transient behaviours of the mean kinetic energy, Nusselt number and enstrophy are demonstrated in figure 12 to provide an idea of how particles interact with the fluid, leading to the formation of cell or roll structures. The volume-averaged horizontal kinetic energy of the particle-free initial flow is quickly suppressed by the settling particles within
$tD/U_f=5$
and then slowly recovers slightly, retaining lower levels than those of the particle-free flow for all cases, as shown in figure 12(a). The flat lines in the lower and right-hand corners indicate the mean kinetic energies of the time-averaged flows. A comparison of the levels of the total kinetic energy and the kinetic energy of the time-averaged flow indicates that the turbulent kinetic energy is relatively small and that the level of turbulent kinetic energy increases with mass loading for all particle sizes. However, the behaviour of the mean vertical kinetic energy was significantly different from that of the horizontal kinetic energy, as shown in figure 12(b). For
$d_p/D=0.0015$
and 0.003, the vertical kinetic energy was initially suppressed substantially and slightly, respectively, whereas for
$d_p/D=0.0045$
, the vertical kinetic energy increased. For all particle sizes, the vertical kinetic energy peaked at approximately
$tD/U_f=10$
and then slowly decreased, converging to a stationary value. From the comparison with the mean vertical kinetic energy of the time-averaged flow, the turbulent kinetic energy of the converged flow is non-negligible owing to the direct interaction between the settling particles and the fluid. Compared with the vertical kinetic energy, however, the horizontal kinetic energy is substantially suppressed by the settling particles, making the flow more laminar. Then, the resulting flow becomes unstable, giving way to cellular convection as suggested by the linear stability analysis. However, the linear stability analysis does not determine the type of cell structure.
The Nusselt number initially dropped quickly to the minimum value at
$tD/U_f=3$
and then suddenly increased to the maximum value at
$tD/U_f=10$
, eventually converging to a stationary value for all particle sizes, as shown in figure 12(c). The peak values of the Nusselt number and vertical kinetic energy are observed at the same instance of
$tD/U_f=10$
, confirming that the vertical motion of the fluid is critical for the vertical transport of heat. The mean enstrophy was significantly increased by particles with sizes of
$d_p/D=0.003$
and 0.0045, whereas it was slightly decreased by particles with
$d_p/D=0.0015$
as shown in figure 12(d).

Figure 13. Thermal plume structures with effective feedback force vectors in the cross-section of plumes at
$Ra=10^6$
: (a) hot plume surface at
$T/\Delta T=0.85$
and a cross-section and (b) cold plume surface at
$T/\Delta T=0.15$
and a cross-section, for
$d_{p}/D=0.003$
and
$\varPhi _{m}=0.06$
(square cell structures); (c) hot plume surface and a cross-section and (d) cold plume surface and a cross-section for
$d_p/D=0.0045$
and
$\varPhi _m=0.06$
(roll structures). A thick solid line in (a) and (c) denotes a contour of
$T/\Delta T=0.85$
and that in (b) and (d) a contour of
$T/\Delta T=0.15$
.
To understand how the cell structure was formed by the settling particles, the effect of the feedback force was investigated. The vertical component of the feedback force plays a critical role in the flow modification. The effective feedback force in the vertical direction defined by
$f^{\prime\prime}_2( \equiv f_2 - \langle f_2 \rangle _A )$
directly affects the flow because the horizontally averaged feedback force
$\langle f_2 \rangle _A$
owing to the fast-settling particles, contributes to the buildup of the vertical mean pressure gradient owing to the confined geometry. An example distribution of the vertical effective feedback force in terms of the negative drag force acting on particles
$-(F^k_{D,2} - \rho \Delta V \langle f_2 \rangle _A /m_p )$
in the cross-sections of hot and cold plumes for the case showing square cell structure (
$d_p/D=0.003, \varPhi _m=0.06$
) is demonstrated in figures 13(a) and 13(b), where
$\Delta V$
is the volume of the numerical cell. The first noticeable difference in the effective feedback force between the hot and cold plumes is in the opposite direction of the forces, which is caused by the fluid motion associated with the hot and cold plumes. When the particles are released at the top wall, where no vertical motion of the fluid is observed, the initial settling velocity of the particles is nearly the same as the terminal velocity in the still fluid. However, particles soon undergo rising and descending fluid motions in the hot and cold plume regions, respectively. Therefore, particles in the hot plume experience more drag because of the increased velocity difference, whereas particles in the cold plume experience less drag because of the decreased velocity difference, resulting in a downward effective feedback force in the hot plume and an upward effective feedback force in the cold plume in most regions, as shown in figures 13(a) and 13(b). However, this situation is reversed when the particles reach the near-bottom region, where the vertical fluid motion is suddenly suppressed by the bottom wall. Therefore, the vertical effective feedback force in the hot plume near the bottom wall is upward, whereas that in the cold plume near the wall is downward, as shown in figures 13(a) and 13(b). Interestingly, the upward force in the hot plume near the bottom wall enhanced the hot plume activity, whereas the downward force in the cold plume near the bottom wall did not affect any activity. This difference might help explain the more concentrated and stronger vertical motion of the hot plume and the widely distributed and weak vertical motion of the cold plume typically found in cell structures. For the roll structures, it is observed that such a difference in the feedback force near the bottom wall is not strong enough to break the symmetry, because the roll structures are found in cases with larger particles with large inertia settling so quickly that they are hardly observed near the bottom wall, as shown in figures 13(c) and 13(d). This suggests that small particles, which have small inertia and thus are more likely affected by the buoyancy force, tend to form polygonal cell structures through the asymmetric feedback forces near the wall, while large particles overcome the buoyancy force due strong inertia and create the symmetric feedback force, yielding symmetric structures such as the roll structure.
3.5. Energy budget analysis
Finally, to further analyse the interaction between the particles and fluid, we investigated the energy exchange between the settling particles and fluid through an energy budget analysis. As we did not consider the thermal interactions between them, we focused on the kinetic energy budget. The kinetic energy of the fluid can be obtained by multiplying
$u_i$
by (2.2) and using (2.1),

which can be averaged over the entire domain and time under statistically stationary conditions, thus yielding

where
$\overline {\epsilon }, \overline {P} (=g\beta \overline {u_2 T})$
and
$\overline { u_i f_i }$
indicate the time-averaged local energy dissipation rate, energy production by the buoyancy force and work done to the fluid by particles, respectively.
On the other hand, the kinetic energy equation of a particle can be obtained by multiplying
$v^k_{p,i}$
using (2.5),



where the summation convention applies only to subscript
$i$
. Equation (3.15) can be averaged over all particles in the domain and time under stationary conditions, yielding

with

where
$\dot {n}_{in}$
is the number of particles inserted into the domain at each time step divided by
$\Delta t$
. Here
$(v^k_{p,2} )_{in}$
and
$(v^k_{p,2} )_{out}$
are the vertical velocities of the particles inserted and eliminated at the top and bottom walls, respectively. Here
$\epsilon ^k_p$
indicates the energy dissipation through friction between
$k$
th particle and the fluid. Here
$-g {v}_{p,2}^k$
is the energy produced by gravity acting on settling particles. The meaning of the tilde above each term (e.g.
$\widetilde {v_{p,2}^{k}}$
) is the average of all the particles in the domain and time. However, in
$\widetilde {(v^k_{p,2} )_{in}}$
and
$\widetilde {(v^k_{p,2} )_{out}}$
, the tilde only refers to the number of particles entering or exiting the channel. Here,
$\tau ^{\prime}_p=\tau _p/(1+0.15 {\textit{Re}}_p^{0.687})$
and the variation in
$\tau ^{\prime}_p$
over the particles is considered in the averaging process. The boundary conditions for the particles leaving the domain at the bottom and particles inserted into the domain at the top suggest that the left-hand side of (3.16) is zero. The balance equation for the particle energy (3.16) then becomes

where
$P_p =-g v_{p,2}^k$
is the production of particle energy owing to gravity.
From (2.4), the global energy transfer between the particles and fluid becomes

where
$\varPhi _m (={N_p m_p}/{\rho V})$
corresponds to the mass loading of the particles, indicating that
$\varPhi _m \widetilde {u^k_i F_{D,i}^k}$
is the energy loss owing to the work done by the particles. This indicates that
$\left \langle u_i f_i \right \rangle _V$
is the rate of energy transfer from the particles to the fluids through a two-way interaction, and its sign is not predetermined. Table 4 lists the values of each term in (3.12) and (3.18) for
$d_p/D=0.0015, 0.003$
and 0.0045, and
$\varPhi _m=0.06, 0.09$
and 0.12, respectively. For the particle energy equation (3.18), each term is multiplied by
$\varPhi _m$
for comparison with the terms in the fluid kinetic energy equation (3.12). The first noticeable difference between the kinetic energy equations for the fluid and particles is that the magnitude of the production or dissipation of particles is at least 300 and up to 3000 times larger than the terms of the fluid for the range of particle parameters in our problem. This is due to the large density ratio of
$\rho _p/\rho =711$
. The balance, which is the sum of all three terms in each equation, ranges from 0.4 % to 8 % of the largest term for the kinetic energy equation of the fluid to 0.01 %–4 % for the kinetic energy equation of the particles. The relatively large balance for large particles was probably caused by a numerical error in the approximation of the delta function distribution of the feedback force in finite grids. The terms responsible for the energy transfer between the particles and the fluid,
$\left \langle \overline {u_i f_i} \right \rangle _V$
in the fluid equation and
$- \varPhi _m \widetilde {u^k_i F_{D,i}^k}$
in the particle equation, which are equivalent to each other, as suggested by (3.19), are comparable to each other, as highlighted in table except for the case with
$d_p/D=0.0015$
. This mismatch is acceptable, given that the magnitude of
$\varPhi _m \widetilde {u^k_i F_{D,i}^k}$
is extremely small and of the order of the balance. Another noticeable observation is that
$\left \langle \overline {u_i f_i} \right \rangle _V$
is negative for
$d_p/D=0.0015$
and positive for
$d_p/D=0.003$
and 0.0045. This implies that the fluid is dragged by the settling large particles, whereas the local fluid motion at the small particle location is opposite to the average particle motion. Therefore, net energy is transferred to the fluid from large particles, whereas energy is transferred to small particles from the fluid on average.
Table 4. The value of each term in kinetic energy equations of fluid (3.12) and of particles (3.18) non-dimensionalized by
$U_f^3/(\sqrt {Ra} D)$
at
$Ra=10^6$
, where
$\sqrt {Ra}(=1000)$
is introduced for better readability and easy comparison. The energy transfer between particles and fluid from both equations is highlighted for comparison. The percentage in the bracket denotes the magnitude of the balance relative to the largest term in the corresponding equation.


Figure 15. Energy feedbacks on the selected horizontal planes at
$Ra=10^6$
: (a,d,g) at
$y/D=0.05$
; (b,e,h) at
$y/D=0.5$
; (c,f,i) at
$y/D=0.95$
.
$d_p/D=0.0015$
and
$\varPhi _m=0.12$
for (a,b,c).
$d_p/D=0.003$
and
$\varPhi _m=0.12$
for (d,e,f).
$d_p/D=0.0045$
and
$\varPhi _m=0.06$
for (g,h,i). The energy feedback is averaged over a short time of
$t U_f/D=0.1$
. Coloured symbols at the side of each panel are matched with the same symbols in figure 14 for convenience of comparison.
The detailed energy transfer between the particles and fluid can be investigated using the local information of
$\overline {u_i f_i}$
. The distributions of the horizontally averaged energy transfer
$\left \langle \overline {u_i f_i} \right \rangle _A$
for the two mass loadings
$\varPhi _m=0.06$
and 0.12 is illustrated in figures 14(a) and 14(b), respectively. The energy transfer from the particle equation
$- \varPhi _m \widetilde {u^k_i F_{D,i}^k}$
is also presented for comparison. Here, the wide tilde denotes the average of the particles at the same height. The oscillatory behaviour near the top wall was caused by the sudden introduction of a feedback force at the moment of release of a particle at the top wall, which was intermittent owing to the finite number of particles. As the particle size decreased, this phenomenon weakened, as the total number of particles increased to maintain the same mass loading. The average of these two energy transfer distributions over the gap height corresponds to both sides of (3.19) or the boldface numbers in table 4. The distributions of
$\left \langle \overline {u_i f_i} \right \rangle _A$
and
$- \varPhi _m \widetilde {u^k_i F_{D,i}^k}$
are similar for large particles with
$d_p/D=0.003$
and 0.0045 and are positive everywhere. However, for small particles with
$d_p/D=0.0015$
, the distributions of the two quantities show some differences near the top and bottom walls, where they are negative. The negative correlation between
$u_i$
and
$f_i (\simeq (u_i-v_{p,i})/\tau ^{\prime}_p)$
implies that
$u_i$
and
$v_{p,i}$
are opposite on average because
$|v_{p,i}| \simeq v_t \gg |u_i|$
; that is, the particle velocity is nearly equal to the settling velocity, which is much larger than the magnitude of the fluid velocity.
The distribution of the energy feedback on the selected horizontal planes, as shown in figure 15 can provide a more detailed energy transfer between the fluid and the particles. For the hexagonal cell structures for
$d_p/D=0.0015$
and
$\varPhi _m=0.12$
(figures 15
a, 15
b and 15
c), the negative
$u_i f_i$
in the hot plume regions at the centre of each hexagonal cell and the positive
$u_i f_i$
in the cold plume regions at the periphery of each cell at all heights are clearly due to the ascending and descending motions of the fluid in the corresponding regions, respectively. In particular, the local energy feedback at the midchannel (figure 15
b) is much stronger than the horizontally averaged value shown in figure 14(b), while the energy feedback near the bottom and top walls (figures 15
a and 15
c) is relatively weak, even resulting in a net negative value. Similarly, for the square cell structures with
$d_p/D=0.003$
and
$\varPhi _m=0.12$
(figures 15
d, 15
e and 15
f), negative
$u_i f_i$
at the centre of each cell and positive
$u_i f_i$
along the periphery are observed. The only difference is that not only in the core region but also in the near-wall regions, the net energy feedback is positive, as shown in figure 14(b). However, for the roll structures for
$d_p/D=0.0045$
and
$\varPhi _m=0.06$
(figures 15
g, 15
h and 15
i), positive
$u_i f_i$
in the descending regions is more dominant than negative
$u_i f_i$
in the ascending region at both the core and near-wall regions, although the vertical fluid motions in the ascending and descending regions are approximately symmetric. In summary, when the particles are large and heavy and hence the settling particles directly drag the fluid, more energy is transferred to the fluid from the particles in the descending region than the energy transferred to particles from the fluid in the ascending region. However, when the particles are small, the energy transfer between the settling particles and the fluid is comparable in both directions, resulting in a net energy transfer from the fluid to the particles.
4. Conclusion
In this study, we numerically investigated particle-laden RBC with a Rayleigh number of
$10^{6}$
and Prandtl number of
$0.7$
in a horizontal channel with an aspect ratio of
$6$
. The particles were assumed to be Lagrangian point particles with diameters smaller than the Kolmogorov length scale of the particle-free flow, and a density much larger than that of the carrier phase. Only the two-way momentum interaction between the particles and fluid was considered, without thermal interaction. The major finding was the formation of various robust polygonal cell structures caused by the settling of the inertial particles. Small- and intermediate-sized particles were observed to modify the flow, resulting in polygonal structures, such as squares, pentagons and hexagons, whereas roll structures were formed by large particles. For both the polygonal and roll structures, the sizes of the cells or rolls decreased with mass loading. Linear stability analysis of particle-laden RBC indicated that the critical Rayleigh number increased with the product of the mass loading and non-dimensional terminal velocity, suggesting that settling small heavy particles tended to stabilize the flow. This explains the suppression of turbulence by the particles and the resulting emergence of cell or roll structures observed in our simulations, although the type of cell structure cannot be determined.
The settling particles also affected the transport of heat, resulting in a modification of the Nusselt number. In particle-free RBC, heat is mainly transported by the activity of hot and cold plumes, which occur randomly in space and time. However, when particles are laden in RBC, the settling motion of particles modifies heat transport in two ways: particles enhance heat transfer through more organized plume activity in the form of robust cells or roll structures, and simultaneously, particles mitigate heat transfer by suppressing turbulence. When particles are of intermediate and large sizes, the enhancement by cell or roll structures is dominant, resulting in a net increase in the Nusselt number with mass loading, whereas small-sized particles decrease the Nusselt number through the suppression of the turbulent flux of heat. Given that thermal interactions were not considered in our study, it is surprising that momentum coupling alone can significantly affect heat transfer through the formation of cell or roll structures.
An investigation of the effective feedback forces that deviate from the horizontally averaged feedback force during the transient period leading to the formation of cells or rolls reveals that the up–down symmetry of hot and cold plume activities is broken by the relatively slow settling of small or intermediate-sized particles near the bottom surface, resulting in polygonal cell structures, whereas the symmetry is roughly retained when large particles settle very quickly, resulting in roll structures. This difference in behaviour near the bottom surface between small and large particles is caused by the different inertia of the particles, resulting in different settling times.
Finally, kinetic energy budget analysis for both the fluid and particles provided insight into the energy exchange between the particles and fluid. In particular, an investigation of the energy feedback between particles and fluid revealed that settling intermediate or large particles drag the fluid so strongly that the net energy is transferred from the particles to the fluid, while settling small particles in the hot plume regions extract more energy than they lose in the cold plume regions, resulting in a net energy transfer from the fluid to the particles, contrary to intuition. The polygonal cell structures formed by settling particles are responsible for this phenomenon.
The major limitation of our investigation is that we did not consider the thermal interaction between the particles and fluid, which might play a role in the dynamics of particle-laden RBC. This finding indicates a possible extension of the current study. Another interesting direction can be taken in line with the consideration of finite-sized particles. More interestingly, an experimental validation of the polygonal cell structures in a particle-laden RBC is encouraged.
Acknowledgements
Authors are grateful to D. Shin for numerically obtaining the critical Rayleigh number and the critical wavenumber in linear stability analysis of particle-laden RBC.
Funding
This work was supported by the National Research Foundation of Korea grant funded by the Korean government (MSIP 2022R1A2C2005538).
Declaration of interests
The authors report no conflict of interest.
Author contributions
S.P. and C.L. derived the theory and S.P. and W.K. performed the simulations and analysis. S.P., W.K. and C.L. contributed equally to reaching the conclusions and writing the paper.
Appendix A. Polygonal cell structures for
$\boldsymbol{Ra=10}^{\boldsymbol{7}}$
and
$\boldsymbol{10}^{\boldsymbol{8}}$
Existence of the polygonal cell structures at higher Rayleigh numbers in the same domain as
$Ra=10^6$
is investigated at
$Ra=10^7$
. For the mass loading of
$\varPhi _m=0.12$
, three sizes of particles,
$d_p/D=0.0015, 0.003$
and 0.0045, are considered. Figure 16 provides contours of time-averaged vertical velocity at
$y/D=0.5$
for the three cases considered, clearly confirming that similar polygonal cell and roll structures are observed although the size of each cell is less uniform compared with
$Ra=10^6$
. The shape of the cells for
$d_p/D=0.0015$
and 0.003 is mostly irregular pentagon and square, respectively, while rolls are observed for
$d_p/D=0.0045$
.
Particle parameters are listed in table 5. Compared with cases for
$Ra=10^6$
, the range of
$St_{\!f}$
and
$St_K$
for the same sizes of particles becomes larger due to an increase in
$U_{\!f}$
and a decrease in the Kolmogorov time scale. The size of cells or rolls represented by
$\lambda$
is comparable with those for
$Ra=10^6$
. Enhancement of Nusselt number is observed for all sizes of particles and the amount of increase is more pronounced than for
$Ra=10^6$
.
Table 5. Particle parameters for
$Ra=10^7$
:
$St_{\!f}$
and
$St_{K}$
are the particle Stokes numbers based on free-fall time and Kolmogorov time scales, respectively;
$v_{t}/U_{\!f}$
is the dimensionless terminal velocity;
$N_{p}$
is the total number of particles;
$\varPhi _{m}$
is the particle mass loading.
$\phi _m = \varPhi _m S$
with
$S$
denoting the non-dimensionalized settling velocity,
$v_t D/\nu$
.
$n$
is the number of cells in the domain for the cell structures.
$\lambda$
and
$\kappa$
are the length scale quantifying the size of cells or rolls and the corresponding wavenumber, respectively, as defined by
$\lambda = ( L_x L_z/2n )^{1/2}, k =2\pi /\lambda$
for the cell structures. For the roll structures,
$\lambda$
is the lateral size of a pair of rolls.


Figure 16. Contours of time-averaged vertical velocity at an
$xz$
-plane of
$y/D=0.5$
at
$Ra=10^7$
. The red area is where the rising flow occurs, and the blue area is where the falling flow occurs. The range of contours is between
$-0.2U_{\!f}$
(blue) and
$0.2U_{\!f}$
(red).
Similar cell structures at
$Ra=10^8$
is examined in the slightly smaller horizontal domain size with
$L_x/D=L_z/D=4$
. For two sizes of particles,
$d_p/D=0.0015$
and 0.003, the two mass loadings of
$\varPhi _m=0.12$
and 0.18 are considered. For the largest size of particle
$d_p/D=0.0045$
, simulations cannot be carried out due to numerical instability, which is probably caused by weak viscous force compared with strong feedback force by the heavy settling particles. Figure 17 shows contours of time-averaged vertical velocity at
$y/D=0.5$
, clearly exhibiting that cell and roll structures are identified although the cell structures for
$d_p/D=0.0015$
are more complicated than the cases for lower Rayleigh numbers and the roll structures for
$d_p/D=0.003$
are rather short and curved.
Particle parameters are listed in table 6. With an increase of Rayleigh number, the range of
$St_{\!f}$
and
$St_K$
becomes larger, indicating that the effect of particles gets stronger. The size of cells or rolls decrease and enhancement of Nusselt number by particles is stronger than the cases with lower Rayleigh numbers.
Table 6. Particle parameters for
$Ra=10^8$
:
$St_{\!f}$
and
$St_{K}$
are the particle Stokes numbers based on free-fall time and Kolmogorov time scales, respectively;
$v_{t}/U_{\!f}$
is the dimensionless terminal velocity;
$N_{p}$
is the total number of particles;
$\varPhi _{m}$
is the particle mass loading.
$\phi _m = \varPhi _m S$
with
$S$
denoting the non-dimensionalized settling velocity,
$v_t D/\nu$
.
$n$
is the number of cells in the domain for the cell structures.
$\lambda$
and
$\kappa$
are the length scale quantifying the size of cells or rolls and the corresponding wavenumber, respectively, as defined by
$\lambda = ( L_x L_z/2n )^{1/2}, k =2\pi /\lambda$
for the cell structures. For the roll structures,
$\lambda$
is the lateral size of a pair of rolls.


Figure 17. Contours of time-averaged vertical velocity at an
$xz$
-plane of
$y/D=0.5$
at
$Ra=10^8$
. The red area is where the rising flow occurs, and the blue area is where the falling flow occurs. The range of contours is between
$-0.2U_{\!f}$
(blue) and
$0.2U_{\!f}$
(red).
Appendix B. Linear stability analysis of particle-laden RBC
In this section, we provide linear stability analysis of RBC modified by settling particles. The base flow considered was
$\boldsymbol{u}=0, T(y)=-\Delta T y/D$
with the particles settling at
$v_t = -\tau ^{\prime}_p g$
. To reflect the particle feedback force in the stability analysis, we note that the particle Stokes number is defined by
$St(=\tau ^{\prime}_p \nu /D^2) \ll 1$
, which allows the flow of particles
$\boldsymbol{v}_p(x,y,z,t)$
. Subsequently, the small disturbances
$\boldsymbol{u}', p', \theta '$
and
$\boldsymbol{v}_p^{\prime}$
satisfy the following equations:




where the last term in (B2) denotes the particle feedback force, which is the Stokes drag force. For small
$\tau ^{\prime}_p$
, from (B4) (Fouxon et al. Reference Fouxon, Park, Harduf and Lee2015),

which can be substituted into (B2), yielding

Non-dimensionalization using
$D, \nu$
and
$\Delta T$
yields for non-dimensionalized disturbances
$\boldsymbol{u}, p$
and
$\theta$
,



where
$S = \tau ^{\prime}_p g D/\nu$
the non-dimensionalized settling velocity. Taking the curl of (B8) twice and using
$\boldsymbol{\nabla }\times (\boldsymbol{\nabla }\times \boldsymbol{u} )= \boldsymbol{\nabla }(\boldsymbol{\nabla }\boldsymbol{\cdot }\boldsymbol{u} ) - {\nabla} ^2 \boldsymbol{u} = - {\nabla} ^2 \boldsymbol{u}$
yields

In particular, the vertical component of (B10) is

where
${\nabla} _1^2 = ({\partial ^2}/{\partial x^2}) + ({\partial ^2}/{\partial z^2})$
denotes the horizontal Laplacian operator. Equation (B11) can be rearranged as follows:

from which eliminating
$\theta$
using (B9) yields

Normal-mode analysis can be performed using


yielding


where
$k^2 = k_x^2+k_z^2$
. Because the presence of the particle feedback force does not affect the principle of stability exchange, the imaginary part of
$\sigma$
remains zero. Therefore, the marginal states are characterized by
$\sigma =0$
, which leads to

where
$\phi _m = \varPhi _m S$
. The boundary conditions for the two free boundaries are
$v=({\partial ^2 v}/{\partial y^2})=\theta =0$
at the walls, implying that

and the boundary conditions for the two rigid boundaries are
$v=({\partial v}/{\partial y})=\theta =0$
at the walls, as follows:

or

The general solution of (B18) can be assumed to take the form

where
$q(k, Ra, \phi _m)$
is the root of the equation

The six roots of (B23) are
$ q_1, q_2, \boldsymbol{\cdots} , q_6$
, which are, in general, real, imaginary or complex numbers, and the general solution is

where
$A_i$
denotes the complex coefficient. For the free–free boundaries, imposing boundary conditions (B19) yields



For the rigid–rigid boundaries, the boundary conditions (B21) lead to
Table 7. Critical Rayleigh number and critical wavenumber for the range of
$\phi _m (=\varPhi _m S)=0 - 2{,}000$
.


Figure 18. Critical Rayleigh number and critical wavenumber for the range of
$\phi _m = 0 - 2000$
.



For non-trivial
$A_i$
, the determinant of the above linear system vanishes, yielding
$Ra(k, \phi _m)$
. Subsequently, the critical Rayleigh number
$Ra_c(\phi _m)$
and corresponding wavenumber
$k_c(\phi _m)$
for a given
$\phi _m$
are obtained. The numerically obtained
$Ra_c$
and
$k_cD$
values for the considered range of
$\phi _m = 0 - 2000$
are listed in table 7 and are graphically illustrated in figure 18. They suggested that the presence of settling particles stabilizes the flow; the critical Rayleigh number linearly increases with
$\phi _m$
for large
$\phi _m$
,
$Ra_c \sim 105 \phi _m$
for the free–free boundaries, and
$Ra_c \sim 174\phi _m$
for the rigid–rigid boundaries. For a large
$\phi _m$
, the critical wavenumber
$k_c D$
converges to 3.0 and 3.74 for the free–free and rigid–rigid boundaries, respectively.