Hostname: page-component-6bb9c88b65-s7dlb Total loading time: 0 Render date: 2025-07-25T09:19:31.561Z Has data issue: false hasContentIssue false

Transient and steady convection in two dimensions

Published online by Cambridge University Press:  21 July 2025

Ambrish Pandey
Affiliation:
Department of Physics, Indian Institute of Technology Roorkee, Roorkee 247667, Uttarakhand, India Center for Astrophysics and Space Science, New York University Abu Dhabi, Abu Dhabi 129188, United Arab Emirates
Katepalli R. Sreenivasan*
Affiliation:
Center for Astrophysics and Space Science, New York University Abu Dhabi, Abu Dhabi 129188, United Arab Emirates Tandon School of Engineering, Department of Physics, and Courant Institute of Mathematical Sciences, New York University, New York, NY 11201, USA
*
Corresponding author: Katepalli R. Sreenivasan, katepalli.sreenivasan@nyu.edu

Abstract

We simulate thermal convection in a two-dimensional square box using the no-slip condition on all boundaries, and isothermal bottom and top walls, and adiabatic sidewalls. We choose 0.1 and 1 for the Prandtl number $Pr$ and vary the Rayleigh number $Ra$ between $10^6$ and $10^{12}$. We particularly study the temporal evolution of integral transport quantities towards their steady states. Perhaps not surprisingly, the velocity field evolves more slowly than the thermal field, and its steady state – which is nominal in the sense that large-amplitude low-frequency oscillations persist around plausible averages – is reached exponentially. We study these oscillation characteristics. The transient time for the velocity field to achieve its nominal steady state increases almost linearly with the Reynolds number. For large $Ra$, the Reynolds number itself scales almost as $Ra^{2/3}\, Pr^{-1}$, and the Nusselt number as $Ra^{2/7}$.

Information

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

1. Introduction

Turbulent flows driven by buoyancy due to inhomogeneity of the temperature are common in nature and applications (Verma Reference Verma2018; Schumacher & Sreenivasan Reference Schumacher and Sreenivasan2020; Lohse & Shishkina Reference Lohse and Shishkina2023). Rayleigh–Bénard convection (RBC) is a paradigm for such flows. The RBC originally referred to shallow horizontally extended layers of fluid, heated from below and cooled from above, and the horizontal walls are smooth unless otherwise specified. In this traditional paradigm, RBC is entirely governed by Prandtl and Rayleigh numbers – where the Prandtl number $Pr$ is the ratio of the kinematic viscosity $\nu$ to the thermal diffusivity $\kappa$ of the fluid, and the Rayleigh number $Ra$ is the ratio of the forcing strength to the dissipative mechanisms. Heat and momentum transport across the convective fluid layer are two global responses to thermal driving in RBC. Heat transport is measured by the Nusselt number $Nu$ , which is the total heat flux relative to that by conduction in the absence of fluid motion, and the momentum transport by an appropriate Reynolds number ${Re}$ , which defines the flow strength.

Because the Rayleigh number is proportional to $H^3$ , where $H$ is the height of the convection apparatus, there has been a tendency in the last 25 or so years to choose as high a value of $H$ as possible while, by necessity, shrinking the horizontal dimension (e.g. Castaing et al. Reference Castaing, Gunaratne, Kadanoff, Libchaber and Heslot1989; Niemela et al. Reference Niemela, Skrbek, Sreenivasan and Donnelly2000). The same is also true of direct numerical simulations (DNS) (e.g. Stevens, Lohse & Verzicco Reference Stevens, Lohse and Verzicco2011; Iyer et al. Reference Iyer, Scheel, Schumacher and Sreenivasan2020). The choice of a low aspect ratio $\Gamma \equiv L/H$ (where $L$ is the horizontal dimension of the apparatus) is common in the quest to achieve very high Rayleigh numbers, but an organised motion that develops in such flows has its own structural morphology (Kadanoff Reference Kadanoff2001; Sreenivasan, Bershadskii & Niemela Reference Sreenivasan, Bershadskii and Niemela2002; Chillà & Schumacher Reference Chillà and Schumacher2012; Foroozani et al. Reference Foroozani, Niemela, Armenio and Sreenivasan2014, Reference Foroozani, Niemela, Armenio and Sreenivasan2017; Verma, Kumar & Pandey Reference Verma, Kumar and Pandey2017) that depends on the aspect ratio and the shape of the apparatus; see also Pandey et al. (Reference Pandey, Krasnov, Schumacher, Samtaney and Sreenivasan2022a)and Stevens et al. (Reference Stevens, Hartmann, Verzicco and Lohse2024). Pure scaling laws in such flows are unlikely for all conditions, yet it is common to fit the Nusselt and Reynolds numbers by power laws with respect to $Ra$ , i.e. $Nu \sim Ra^{\gamma }$ and ${Re} \sim Ra^{\zeta }$ . As pointed out by Doering (Reference Doering2020), fitting such local exponents for small ranges of data is bound to lead to conclusions of uncertain value. Indeed, there are considerable variations of the effective exponents $\gamma$ and $\zeta$ from one study to another, and they have been the subject of extensive reviews – e.g. by Chillà & Schumacher (Reference Chillà and Schumacher2012) and Lohse & Shishkina (Reference Lohse and Shishkina2024). In a series of large simulations, an effort has been made to avoid the constraining effect of the sidewalls by stipulating periodic boundary conditions on them (Samuel et al. Reference Samuel, Bode, Scheel, Sreenivasan and Schumacher2024). These studies mimicking large aspect ratio convection have revealed a very different nature of near-wall velocity from a traditional boundary layer that undergoes laminar–turbulent transition, and have implications for the so-called ultimate state.

There has been the expectation that the riddle of the ultimate state can be solved for the two-dimensional (2-D) case in which the flow is compelled to occur only in a vertical plane (Zhu et al. Reference Zhu, Mathai, Stevens, Verzicco and Lohse2018; Samuel & Verma Reference Samuel and Verma2024; Tiwari, Sharma & Verma Reference Tiwari, Sharma and Verma2025) because of the natural hope that very high Rayleigh numbers can be achieved here for the same computing power. Existing data show a heat transport scaling that is similar to the three-dimensional (3-D) counterpart (Schmalzl, Breuer & Hansen Reference Schmalzl, Breuer and Hansen2004; van der Poel, Stevens & Lohse Reference van der Poel, Stevens and Lohse2013; Pandey et al. Reference Pandey, Kumar, Chatterjee and Verma2016; Zhang, Zhou & Sun Reference Zhang, Zhou and Sun2017; Pandey Reference Pandey2021) but the Nusselt number in two dimensions is smaller when $Pr \geqslant 1$ (van der Poel et al. Reference van der Poel, Stevens and Lohse2013), although essentially the same when $Pr$ is small (Pandey Reference Pandey2021), as for liquid metals. The Reynolds number, on the other hand, shows very different behaviours in two and three dimensions. The magnitudes of momentum transport and the scaling exponent $\zeta$ are consistently higher in two dimensions, with $\zeta \geqslant 0.60$ (Schmalzl et al. Reference Schmalzl, Breuer and Hansen2004; van der Poel et al. Reference van der Poel, Stevens and Lohse2013; Zhang et al. Reference Zhang, Zhou and Sun2017; Wan et al. Reference Wan, Wang, Wang, Xia, Zhou and Sun2020; Pandey Reference Pandey2021). As a summary, the exponent $\gamma$ has been reported to take values in the range $[1/4, 1/3]$ , while $\zeta$ assumes values in the range $[4/9, 2/3]$ (Verma Reference Verma2018). Note that these scaling features are well captured by the model of Reference Grossmann and LohseGrossmann & Lohse (2000); see also Pandey et al. (Reference Pandey, Kumar, Chatterjee and Verma2016).

In this paper, we perform DNS of 2-D convection in a unit box for $Pr = 0.1$ and $1$ , for Rayleigh numbers between $10^6$ and $10^{12}$ , not only for the purposes of exploring flow properties, but also for highlighting the challenges of simulating 2-D convection. We use the no-slip condition on all boundaries, with bottom and top walls isothermal, and sidewalls adiabatic. In particular, we show that the velocity field evolves more slowly than the temperature, and call attention to large fluctuations that occur in what may be regarded as the nominal steady state of the velocity field; it is nominal in the sense that large-amplitude low-frequency oscillations persist around plausible averages. (To avoid excessive repetition, however, we often omit the qualifier ‘nominal’.) We relate these fluctuations to heat transport characteristics, and provide estimates of suitably defined transient times. Heat transport also exhibits persistently wild fluctuations about its mean for strong thermal forcing. The main qualitative conclusion of this study is that such long transients, as well as wildly fluctuating heat transport, at least for aspect ratios of order unity, make the observation of the so-called ultimate state as elusive in two dimensions (Doering, Toppaladoddi & Wettlaufer Reference Doering, Toppaladoddi and Wettlaufer2019) as in three dimensions (Doering Reference Doering2020).

2. Numerical methodology

We perform DNS in a 2-D fluid layer with horizontal dimension $L$ , with an imposed temperature difference $\Delta T$ between the bottom and top plates separated by vertical dimension $H$ . For this work, the aspect ratio is $\Gamma = L/H = 1$ . The following Oberbeck–Boussinesq equations dictating the flow are solved using the Nek5000 solver (which has been used extensively for the simulation of turbulent convection – for details, see Scheel, Emran & Schumacher Reference Scheel, Emran and Schumacher2013):

(2.1) \begin{align}\boldsymbol{\nabla \cdot u} &= 0, \end{align}
(2.2) \begin{align}\frac {\partial \boldsymbol{u}}{\partial t} + \boldsymbol{u \cdot \nabla u} &= - \frac { \boldsymbol{\nabla} p}{\rho _0} + \alpha g (T-T_0) \hat {z} + \nu\nabla^2 \boldsymbol{u}, \end{align}
(2.3) \begin{align}\frac {\partial T}{\partial t} + \boldsymbol{u} \boldsymbol{\cdot} \boldsymbol{\nabla} T &= \kappa\nabla^2 T. \end{align}

Here, $\boldsymbol{u} = (u_x, u_z)$ , $p$ and $T$ are the velocity, pressure and temperature, respectively; $\rho _0$ is the reference density, and $T_0$ is the reference temperature. We non-dimensionalise (2.1)–(2.3) using $H$ , $\Delta T$ , $u_f$ and $t_f$ as the scales for length, temperature, velocity and time, respectively, where $u_f = \sqrt {\alpha g\, \Delta \textit{TH}}$ is the free-fall velocity, and $t_f = H/u_f$ is the free-fall time. The result contains the Prandtl number $Pr$ and the Rayleigh number $Ra = \alpha g\, \Delta \textit{TH}^3/(\nu \kappa )$ , where $\alpha$ is the isobaric thermal expansion coefficient, and $g$ is the acceleration due to gravity. We explore two fluids with $Pr = 0.1$ and $1$ for $Ra$ between $10^6$ and $10^{12}$ . The square domain is decomposed into $N_e$ elements, and each element is further resolved using Lagrangian interpolation polynomials of order $N$ in both horizontal and vertical directions. Thus the entire flow is resolved using $N_e N^2$ mesh cells. The no-slip condition for the velocity field is imposed on all boundaries. Isothermal and adiabatic conditions for the temperature field are imposed on the horizontal plates and sidewalls, respectively. The flow consists of a single large-scale structure covering the entire domain, except for $Pr = 1,\ Ra \leqslant 10^7$ , where two vertically stacked structures occur. The flow morphology is similar to that observed by Labarre, Fauve & Chibbaro (Reference Labarre, Fauve and Chibbaro2023) and Samuel & Verma (Reference Samuel and Verma2024), so we do not show it here.

To resolve the thermal and viscous boundary layers, finer mesh is used near all boundaries. The spatial resolution in the flow is further ensured by computing the Kolmogorov length scale $\eta$ , which is nominally the finest scale in the velocity field, and stipulating that the local vertical grid spacing is $\varDelta_{z}(z)/\eta (z) \lt 1.5$ for all simulations. The kinetic energy and the thermal dissipation rates are defined, respectively, as

(2.4) \begin{eqnarray} \varepsilon _u (\boldsymbol{x}) &= \frac {\nu }{2} \sum _{l,m} \left ( \frac {\partial u_l}{\partial x_m} + \frac {\partial u_m}{\partial x_l} \right )^2 , \end{eqnarray}

(2.5) \begin{eqnarray} \varepsilon _T (\boldsymbol{x}) &=\kappa \left ( \frac {\partial T}{\partial x_l} \right )^2 , \end{eqnarray}

and represent the rates of loss of kinetic energy and thermal energy per unit mass. Here, $l, m = (x,z)$ , and the local Kolmogorov scale is given by $\eta = (\nu ^3/\varepsilon _u)^{1/4}$ . As the intermittent variation of $\varepsilon _u (\boldsymbol{x})$ in the flow leads to variations in the Kolmogorov scale as well, we estimate an average Kolmogorov scale in each horizontal plane using the horizontally and temporally averaged dissipation as

(2.6) \begin{equation} \eta (z) = \frac {\nu ^{3/4}}{\langle \varepsilon _u \rangle _{x,t}^{1/4}(z)}. \end{equation}

The finest length scale in the temperature field, the Batchelor scale $\eta /\sqrt {Pr}$ , is of the same order as $\eta$ , or coarser, in the present work. Thus it is always adequately resolved.

3. Other associated definitions

The Nusselt number in a horizontal plane is computed as (Chillà & Schumacher Reference Chillà and Schumacher2012)

(3.1) \begin{equation} Nu(z) = \frac { \langle u_z T \rangle _{x,t} - \kappa\partial \langle T \rangle _{x,t}/\partial z} { \kappa\kern-0.5pt\Delta T /H}, \end{equation}

where $\langle u_z T \rangle _{x,t}$ is the convective component of the heat flux, and $- \kappa\partial \langle T \rangle _{x,t}/\partial z$ is the diffusive component. As the vertical velocity vanishes at the horizontal plates, this relation yields the definition using the wall temperature gradient as

(3.2) \begin{equation} Nu_W = - \frac {H}{\Delta T} \left . \frac {\partial \langle T \rangle _{x,t}} {\partial z} \right |_{z=0,H} . \end{equation}

Averaging (3.1) along the vertical direction yields the following relation for the global heat transport across the convective layer:

(3.3) \begin{equation} Nu = 1 + \frac {H}{\kappa\kern-0.5pt\Delta T}\, \langle u_z T \rangle _{A,t} . \end{equation}

Here, $\langle \cdot \rangle _{A,t}$ stands for the average over the entire flow domain and simulation time covering the steady state.

In another family of relations, the exact relations in RBC (Howard Reference Howard1972; Shraiman & Siggia Reference Shraiman and Siggia1990) connect the Nusselt number $Nu$ with the globally averaged kinetic energy dissipation rate $\varepsilon _u$ and the thermal dissipation rate $\varepsilon _T$ , as

(3.4) \begin{align}\langle \varepsilon _u \rangle _{A,t} &= \frac {\nu ^3}{H^4}\, \frac {(Nu-1)Ra}{Pr^2} , \end{align}
(3.5) \begin{align}\langle \varepsilon _T \rangle _{A,t} &= \kappa\, \frac {(\Delta T)^2}{H^2}\, Nu. \end{align}

Thus the Nusselt number can also be estimated from mean dissipation rates:

(3.6) \begin{align}Nu_{\varepsilon _u} &= 1 + \frac {H^4}{\nu ^3} \frac {Pr^2}{Ra} \langle \varepsilon _u \rangle _{A,t} , \end{align}
(3.7) \begin{align}Nu_{\varepsilon _T} &= \frac {H^2}{(\Delta T)^2} \frac {\langle \varepsilon _T \rangle _{A,t}}{\kappa }. \end{align}

The agreement among simulated values from different definitions serves as a check on the accuracy and adequacy of spatial and temporal resolutions (Stevens, Verzicco & Lohse Reference Stevens, Verzicco and Lohse2010; Zhang et al. Reference Zhang, Zhou and Sun2017; Pandey et al. Reference Pandey, Krasnov, Schumacher, Samtaney and Sreenivasan2022a ). In table 1, we list the global heat flux estimated from all the methods, and note that the agreement among various estimates of the Nusselt number is indeed very good. Also listed are other crucial parameters of the flow. We discuss below the approach to the final state of the Nusselt number using these definitions and relations, with comments on the scaling of global averages. In (3.1)–(3.3) and (3.6)–(3.7), the Nusselt numbers are globally averaged quantities, though we do not show the averaging symbol explicitly, following standard usage of the past.

Table 1. Important parameters of DNS in a 2-D box with $\Gamma = 1$ . We list the Prandtl number, the Rayleigh number, the total number of mesh cells in the entire flow domain $N_e N^2$ , the Nusselt numbers using (3.3), (3.6), (3.7) and (3.2), respectively, the Reynolds number, and the simulation time after the flow attains a steady state, $t_{sim}$ . Error bars in Nusselt and Reynolds numbers are the corresponding standard deviations.

In the following, we study the evolutions of the instantaneous domain-averaged heat fluxes, which are defined as

(3.8) \begin{align}\overline {Nu} &= 1 + \frac {H}{\kappa\kern-0.5pt\Delta T}\, \langle u_z T \rangle _{A} , \\[-10pt]\nonumber\end{align}
(3.9) \begin{align} \overline {Nu_{\varepsilon _u}} &= 1 + \frac {H^4}{\nu ^3}\, \frac {Pr^2}{Ra}\, \langle \varepsilon _u \rangle _{A} , \end{align}

(3.10) \begin{align} \overline {Nu_{\varepsilon _T}} = \frac {H^2}{(\Delta T)^2}\, \frac {\langle \varepsilon _T \rangle _{A}}{\kappa }. \end{align}

4. Transient characteristics

A common method for initiating high- $Ra$ simulations is to start them from the flow at a lower $Ra$ value. Simulations can also be performed ab initio from the conduction state with random perturbations. In both cases, global heat flux and kinetic energy evolve with time, and one needs to wait some time before a statistically steady state is attained. Figure 1 shows the temporal evolution of integral quantities for the two $Ra$ indicated and $Pr = 0.1$ , with simulations initiated from the conduction state. On the one hand, we observe that $\overline {Nu}$ and $\overline {Nu_{\varepsilon _T}}$ in figure 1(b,d) start oscillating about some mean value shortly after the simulation begins. On the other hand, $\overline {Nu_{\varepsilon _u}}$ in figure 1(c) initially increases and starts to fluctuate about a mean value only after some time has elapsed. The instantaneous domain-averaged non-dimensional kinetic energy $E = \langle (u_x^2 + u_z^2)/2 \rangle _A/u_f^2$ in figure 1(a) offers the best means to determine the time to the steady state. There is no ambiguity about this approach for the lower $Ra$ ; however, aside from the fact that it takes longer at the higher $Ra$ for the energy to achieve its steady state, the latter is somewhat nominal because fluctuations about the average are significant. The situation becomes more so at even higher Rayleigh numbers. Fluctuations in the domain-averaged energy $E$ follow from dynamical considerations; we will show this in § 6.

Figure 1. Evolution of the integral quantities in the transient state for $Pr = 0.1$ , and $Ra = 3 \times 10^8$ (black curves) and $Ra = 3 \times 10^9$ (orange curves). (a) The domain-averaged kinetic energy $E$ increases slowly and takes a few thousand free-fall times to reach the steady state. The transient time is longer for higher $Ra$ . Dashed vertical lines show a quantitative measure of the transient time $t_{\it trns}$ , obtained from (4.3), to be discussed later. (b,c,d) Plots show that the Nusselt number fluctuates rapidly about its mean nearly from the start, but fluctuations have different characters depending on the definition of the Nusselt number. In (b), the fluctuations in $\overline {Nu}$ are strong and of high frequency with no well-defined transient state, and there is an overlap for the two $Ra$ values. In (c), one can approximately identify a transient state in $\overline {Nu_{\varepsilon _u}}$ , which exhibits very strong fluctuations containing both high and low frequencies. Plot (d) shows that $\overline {Nu_{\varepsilon _T}}$ fluctuates similarly to $\overline {Nu}$ in (b), but there is no overlap for the two Rayleigh numbers. The occasional appearance of negative values of $\overline {Nu}$ suggests the likelihood that a small parcel of the coldest fluid from the top wall registers directly at the bottom wall, and vice versa.

We define the average kinetic energy in the steady state, $E_{av}$ , as

(4.1) \begin{equation} E_{av} = \tfrac {1}{2}u_{\it RMS}^2/u_f^2 , \end{equation}

where $u_{\it RMS} = \sqrt { \langle u_x^2 + u_z^2 \rangle _{A,t}}$ is the root mean square (RMS) velocity of the flow in the steady state, and plot it in figure 2 as a function of $Ra$ . Before discussing differences between the two Prandtl numbers, we note the major difference between 2-D and 3-D fluctuations. In the 3-D case, for low and moderate Prandtl numbers, $E_{av}$ is a slowly decreasing function of $Ra$ ; for example, $E_{av}$ from a horizontally periodic cuboid of $\Gamma = 4$ for $Pr = 0.7$ (from Samuel et al. Reference Samuel, Bode, Scheel, Sreenivasan and Schumacher2024), shown as green squares in figure 2, follows a $Ra^{-0.07}$ scaling. The corresponding behaviour for 2-D cases is that the fluctuations increase with $Ra$ . The behaviour is different for different $Pr$ when the Rayleigh numbers are low, but follows an approximately $Ra^{1/3}$ scaling at high Rayleigh numbers, as indicated by the blue and red dashed lines in figure 2. (Our objective is not a detailed study of differences between 2-D and 3-D convection. Some such differences have been pointed out e.g. by van der Poel et al. (Reference van der Poel, Stevens and Lohse2013) and Pandey et al. (Reference Pandey, Kumar, Chatterjee and Verma2016)). The challenges posed by this increasing trend will be discussed below.

Figure 2. Global (area- as well as time-averaged) kinetic energy in the steady state $E_{av}$ as a function of $Ra$ . An increasing trend with $Ra$ is observed in 2-D RBC. In the turbulent regimes ( $Ra \geqslant 10^8$ for $Pr = 0.1$ , and $Ra \gt 10^{9}$ for $Pr = 1$ ), the data approximately follow $E_{av} \sim Ra^{1/3}$ , shown as dashed lines. In contrast, $E_{av}$ in 3-D RBC for $Pr = 0.7$ (taken from Samuel et al. Reference Samuel, Bode, Scheel, Sreenivasan and Schumacher2024) shows a weakly decreasing trend. The transitions observed in this figure are related to transitions in the flow structure; see § 5.3. The statistical error bars in almost all the figures here are comparable to the thickness of the symbols. The exception is figure 6, for which the errors bars are shown explicitly.

It is useful to study the variation of $t_{\it trns}$ (which is the time required for $E$ to attain some stipulated fraction of $E_{av}$ ) with respect to $Ra$ and $Pr$ by some simple scheme. To this end, we plot the evolution of $E$ for $Ra = 3 \times 10^9$ and $10^{10}$ for $Pr = 0.1$ in figure 3. The growth of $E$ for both $Ra$ values is comparable for initial times but the curve for higher $Ra$ continues to grow for a longer period before a nominally steady state is reached, but with conspicuous fluctuations.

Figure 3. Evolution of the domain-averaged kinetic energy $E(t)$ in the transient state for $Pr = 0.1$ and $Ra = 3 \times 10^9$ and $10^{10}$ can be described well by (4.2), and the dashed curves show $E_{\it fit}(t)$ for the growth rate. Vertical lines show that the transient time $t_{\it trns}$ is much longer for $Ra = 10^{10}$ than for $Ra = 3 \times 10^9$ .

Figure 3 suggests that $E(t)$ can be fitted with an equation of the form

(4.2) \begin{equation} [E_{av} - E(t)]/E_{av} = c \exp (-kt), \end{equation}

where $E_{av}$ is the ‘steady-state’ or ‘asymptotic’ mean energy of the flow; the growth rate $k$ depends on the Rayleigh and Prandtl numbers. The factor $c$ captures the finite energy at the initial instant in figure 3: though $E(t)$ grows rapidly only when the convective motion is established, there is a finite $E(t)$ from which it starts.

We fit the suitable segment of the growth curve $E(t)$ with (4.2). The segment at late times is not suitable for fitting the formula because the kinetic energy does not attain a constant value but fluctuates strongly for long periods of time. The scaling factor $c$ is calculated from the data as $c = (E_{av} - E(0))/E_{av}$ , where $E(0)$ is the energy at $t = 0$ . The dashed curves in figure 3 are fits to the data. Having determined the growth rate, we define the transient time $t_{\it trns}$ as the time when the fitted curve $E_{\it fit}(t)$ in figure 3 reaches 96 % of its ‘constant’ value, $E_{av}$ . Thus the transient time is estimated as

(4.3) \begin{equation} t_{\it trns} = \frac {1}{k} \log \frac {c}{0.04} . \end{equation}

The transient time thus estimated is plotted as a function of $Ra$ in figure 4(a). Note that $t_{\it trns}$ increases as a power law; for $Pr = 0.1$ , the best fit yields $t_{\it trns} \sim Ra^{0.59 \,\pm\, 0.03}$ , while we find $t_{\it trns} \sim Ra^{0.71\, \pm\, 0.08}$ for $Pr = 1$ .

We have experimented with different definitions of $t_{\it trns}$ (e.g. by requiring it to reach 90 % of $E_{av}$ ), and find the same trends, though the numbers are different. The fact that the $Ra$ exponent is non-trivial suggests an important role for the boundary layers and their relation to the large structure; those details (as well as the effect of the aspect ratio) are in need of further exploration.

Figure 4. (a) The transient time $t_{\it trns}$ as a function of $Ra$ shows a power law. For a given $Ra$ , $t_{\it trns}$ is longer for lower $Pr$ . (b) The time $t_{\it trns}$ as a function of ${Re}$ is essentially the same for both values of $Pr$ , and exhibits a nearly linear trend. Moreover, $t_{\it trns}$ is nearly the same for both $Pr$ values when the Reynolds numbers are the same. Data correspond only to the turbulent regimes. Sparser data sets for $Pr = 0.021$ are consistent with the Prandtl numbers of these plots.

Figure 5. Inverse growth rate as a function of (a) $Ra$ and (b) ${Re}$ . The trends with ${Re}$ are approximately the same for both Prandtl numbers, and are slightly below linear.

Figure 4(a) further shows that for a given $Ra$ , $t_{\it trns}$ is longer for $Pr = 0.1$ than for $Pr = 1$ . This is expected as $u_{\it RMS}$ – consequently the Reynolds number – is higher for smaller $Pr$ (Schumacher, Götzfried & Scheel Reference Schumacher, Götzfried and Scheel2015; Pandey & Verma Reference Pandey and Verma2016; Pandey et al. Reference Pandey, Krasnov, Schumacher, Samtaney and Sreenivasan2022a ). In figure 4(b), we plot $t_{\it trns}$ as a function of the Reynolds number ${Re}$ , which is defined as

(4.4) \begin{equation} {Re} = u_{\it RMS}H/\nu . \end{equation}

We find that the two $t_{\it trns}$ fall approximately on the same line for both Prandtl numbers, the best fit for which is nearly linear. Thus a 2-D RBC at high $Ra$ has to be simulated for very long times to achieve a statistically steady state. If a steady state is not achieved and $E$ is still growing, then one will find $Nu_{\varepsilon _u} \lt Nu$ (see § 6), even when spatial and temporal resolutions are adequate.

Shown in figure 5 is the variation of the inverse growth rate $k^{-1}$ with respect to $Ra$ (figure 5 a) and ${Re}$ (figure 5 b). It appears that $k^{-1}$ is approximately a linear function of ${Re}$ , and depends only weakly on the Prandtl number.

A longer transient at higher $Ra$ compels researchers to use some workaround to achieve the statistically steady state in a shorter time. For example, one might start with a much coarser resolution and ramp it up to the required level only after the kinetic energy has reached an approximate steady state. It is practically impossible with available computing power to conduct very high $Ra$ simulation with fully-resolved fields in the entire transient state. One can use a coarser mesh during the transient when all degrees of freedom have not yet been excited. However, it is important to ensure that simulations on the coarser mesh lead to the same steady state as that attained in a well-resolved simulation. A concern otherwise would be that the coarse-grid simulation results in kinetic energy different from the properly resolved case, ultimately affecting the scaling of ${Re}$ , particularly because of the presence of the low frequency modes (see § 6). Though we demonstrate in Appendix A that the mean kinetic energy in the steady state and the transient time are essentially the same in coarser simulations, it is not necessarily likely to be a general result. The data in figures 1 and 3 correspond to simulations at the coarser resolution.

5. Scaling exponents for integral transport in the steady state

5.1. Heat transport

We plot $Nu$ , computed from (3.2), as a function of $Ra$ in figure 6(a). The data for high Rayleigh numbers seem to closely follow similar power laws for both $Pr$ values. For $Pr = 1$ , the best fit for $Ra \geqslant 6 \times 10^7$ yields the scaling $Nu = 0.12\, Ra^{0.29\,\pm\,0.003}$ . The exponent is close to $2/7$ proposed for the so-called ‘hard turbulence’ in confined 3-D RBC (Castaing et al. Reference Castaing, Gunaratne, Kadanoff, Libchaber and Heslot1989), also explored in various later studies (Siggia Reference Siggia1994; Chillà & Schumacher Reference Chillà and Schumacher2012; Johnston & Doering Reference Johnston and Doering2009). We plot the normalised Nusselt number $Nu\, Ra^{-2/7}$ versus $Ra$ in figure 6(b), which reveals that the local exponent varies considerably and is only approximately $2/7$ . Furthermore, we note that the Nusselt numbers fluctuate significantly, being significantly larger when $Nu$ is computed using (3.3). To some extent, this result indicates that the local scaling exponents obtained by fitting heat transport data over short ranges of Rayleigh numbers could lead to misleading conclusions. Figure 6(a) also shows that the Nusselt numbers for $Ra \leqslant 10^7$ for $Pr = 1$ do not follow the higher- $Ra$ trend. This is a consequence of differing flow structures observed: the flow for $Ra \leqslant 10^7$ consists of two vertically stacked rolls, whereas for $Ra \gt 10^7$ , a single-roll structure is observed. These findings suggest that the flow is less efficient in transporting heat when the double-roll state occurs, which is in line with previous observations in both two and three dimensions (Xi & Xia Reference Xi and Xia2008; van der Poel, Stevens & Lohse Reference van der Poel, Stevens and Lohse2011; Weiss & Ahlers Reference Weiss and Ahlers2011).

Figure 6. (a) The Nusselt number as a function of $Ra$ for $Pr = 0.1$ (red circles) and $Pr = 1$ (blue triangles). The scaling for high Rayleigh numbers differs only slightly from the low- $Ra$ behaviour. (b) The normalised Nusselt number $Nu\, Ra^{-2/7}$ shows that the $2/7$ scaling is only approximate for moderate Rayleigh numbers and moderate aspect ratios. Wall heat flux from (3.2) is shown here with the error bars representing the standard deviation.

For $Ra \leqslant 3 \times 10^7$ , the Nusselt numbers for $Pr = 0.1$ are higher than those for $Pr = 1$ . Unlike for $Pr = 1$ , the flow for $Pr = 0.1$ shows no double-roll state at lower Rayleigh numbers, which is why a larger difference appears in $Nu$ for the two Prandtl numbers. Figure 6(b) clearly shows that heat transport is smaller for $Pr = 0.1$ than for $Pr = 1$ in the turbulent regime, i.e. for $Ra \gt 3 \times 10^7$ . This feature is similar to that observed in 3-D RBC, where lower- $Pr$ fluids are less efficient at transporting heat when $Pr \lt 1$ (Verzicco & Camussi Reference Verzicco and Camussi1999; van der Poel et al. Reference van der Poel, Stevens and Lohse2013; Pandey & Sreenivasan Reference Pandey and Sreenivasan2021; Pandey et al. Reference Pandey, Krasnov, Sreenivasan and Schumacher2022b ). The Grossmann–Lohse model (Grossmann & Lohse Reference Grossmann and Lohse2000, Reference Grossmann and Lohse2001) also suggests a similar trend.

Figure 7. The Reynolds number based on $u_{\it RMS}$ scales nearly as $Ra^{2/3}$ in the turbulent regime (indicated by the blue solid line), which is distinctively different from scaling $Ra^{1/2}$ reported in 3-D RBC. The green dashed line indicates the ${Re} \sim Ra^{0.46}$ scaling observed for 3-D RBC in a $\Gamma = 4$ box by Samuel et al. (Reference Samuel, Bode, Scheel, Sreenivasan and Schumacher2024).

5.2. Momentum transport

A variety of velocity scales can be defined in turbulent RBC, and can be used to define the Reynolds number ${Re}$ . In cylindrical or cubic domains with $\Gamma \approx 1$ , the most dominant eddy in the flow is in the form of a large-scale circulation, and its velocity is observed to scale with the free-fall velocity (Lam et al. Reference Lam, Shang, Zhou and Xia2002; Xia, Sun & Zhou Reference Xia, Sun and Zhou2003). In DNS, the Reynolds number is often obtained using the RMS velocity and the depth of the convective layer $H$ (see (4.4)) (Scheel & Schumacher Reference Scheel and Schumacher2016; Pandey et al. Reference Pandey, Krasnov, Sreenivasan and Schumacher2022b ). We plot ${Re}\,Pr$ against $Ra$ in figure 7. Also indicated by the dashed line is the power law ${Re}\,Pr = 0.33\, Ra^{0.46}$ , as found by Samuel et al. (Reference Samuel, Bode, Scheel, Sreenivasan and Schumacher2024) for $Pr = 0.7$ . There is some overlap in the magnitude of the Reynolds numbers between 2-D and 3-D RBC, but the scaling exponents differ, especially for high Rayleigh numbers. In the high- $Ra$ regime, we find that the Reynolds number scales as ${Re} = 0.13\, Ra^{0.65\,\pm\,0.01}$ for $Pr = 0.1$ , and ${Re} = 0.02\, Ra^{0.65\,\pm\,0.01}$ for $Pr = 1$ . The data for lower Rayleigh numbers exhibit approximately the ${Re} \sim Ra^{0.55}$ scaling for both $Pr$ values.

As is well known, the Reynolds number decreases with increasing $Pr$ in 3-D RBC (van der Poel et al. Reference van der Poel, Stevens and Lohse2013; Pandey & Sreenivasan Reference Pandey and Sreenivasan2021); similarly, in 2-D RBC, the Reynolds number scales approximately as $Ra^{2/3}\,Pr^{-1}$ , consistent with the result found semi-analytically by Chini & Cox (Reference Chini and Cox2009) and Wen et al. (Reference Wen, Goluskin, LeDuc, Chini and Doering2020). This ${Re}$ scaling further suggests that

(5.1) \begin{equation} \frac {u_{\it RMS}}{u_f} = {Re}\, \sqrt {\frac {Pr}{Ra}} \sim Ra^{1/6}\, Pr^{-1/2} , \end{equation}

and consequently

(5.2) \begin{equation} E_{av} = \tfrac{1}{2} \frac {u^2_{\it RMS}}{u^2_f} \sim Ra^{1/3}\, Pr^{-1}. \end{equation}

The high- $Ra$ data in figure 2 are indeed consistent with this scaling.

5.3. The RMS temperature fluctuation

We compute the RMS temperature fluctuation as

(5.3) \begin{equation} T_{\it RMS} = \sqrt { \langle T^2 \rangle _{A,t} - \langle T \rangle _{A,t}^2}, \end{equation}

and plot $T_{\it RMS}/\Delta T$ as a function of $Ra$ in figure 8(a). It is clear that $T_{\it RMS}/\Delta T$ decreases with increasing $Ra$ , but various regions can be identified in figure 8(a). For $Pr = 0.1$ , $T_{\it RMS}$ for $Ra \leqslant 2 \times 10^7$ exhibits $Ra^{-0.07}$ scaling, but it starts, somewhat abruptly, to follow a steeper $Ra^{-0.12}$ scaling for higher $Ra$ . For $Pr = 1$ , too, we find that $T_{\it RMS}$ for $Ra \lt 10^9$ shows a scaling $Ra^{-0.11}$ , while a scaling $Ra^{-0.14}$ ensues for $Ra \geqslant 10^9$ . Again, the transition between these two regimes is nearly abrupt. The RMS fluctuations for $Ra \leqslant 10^7$ are larger and clearly depart from the trend for higher $Ra$ ; this occurs because of the double-roll state observed for weak thermal forcing at $Pr = 1$ . It is interesting that the scaling $Ra^{-0.14}$ for high $Ra$ agrees well with scalings of fluctuations at the centre of cylindrical RBC cells for $Pr \approx 0.7$ (Castaing et al. Reference Castaing, Gunaratne, Kadanoff, Libchaber and Heslot1989; Niemela et al. Reference Niemela, Skrbek, Sreenivasan and Donnelly2000), as well as with that in the bulk region of a horizontally periodic box for $Pr = 0.7$ , $\Gamma = 4$ (Samuel et al. Reference Samuel, Bode, Scheel, Sreenivasan and Schumacher2024); see also Pandey & Verma (Reference Pandey and Verma2016). We also note that the magnitudes of $T_{\it RMS}$ for moderate Rayleigh numbers ( $10^7 \lt Ra \lt 10^9$ ) are quite similar for the two Prandtl numbers.

Figure 8. (a) The RMS temperature fluctuation $T_{\it RMS}$ decreases with $Ra$ for both $Pr$ values, and has nearly the same magnitude for moderate Rayleigh numbers. Different exponents as well as prefactors are found for moderate and high Rayleigh numbers. (b) Fluctuation $T_{\it RMS}$ as a function of ${Re}$ shows that it scales approximately as ${Re}^{-0.2}$ in the turbulent regime (for ${Re} \gt 10^4$ ) for both Prandtl numbers.

The nearly abrupt transitions in $T_{\it RMS}$ are related to changes in flow morphology and statistics, and in the corresponding heat transport scaling. Data for $Pr = 0.1$ show that the scaling exponent $\gamma$ in the $Nu$ $Ra$ relation changes from ${\approx}0.35$ to ${\approx}0.25$ for $Ra \geqslant 3 \times 10^7$ , with the latter value persisting for nearly two decades of $Ra$ (i.e. up to $Ra \approx 10^9$ ), beyond which it again increases to 0.30. Similarly, for $Pr = 1$ , $\gamma$ changes from ${\approx}0.31$ to ${\approx}2/7$ after the transition at $Ra \approx 10^9$ . Labarre et al. (Reference Labarre, Fauve and Chibbaro2023) recently noted that the ratio between the RMS fluctuations and the mean heat flux increases abruptly in two dimensions once $Ra/Pr$ increases beyond ${\approx}10^9$ . A somewhat similar transition was reported also by Gao et al. (Reference Gao, Tao, Huang, Bao and Xie2024), who found the transition $Ra$ to scale with the Prandtl number as $Pr^{1.41}$ . Our own data show that the relative fluctuations in the heat flux near the wall and in the bulk are enhanced significantly for $Ra \geqslant 3 \times 10^7$ at $Pr = 0.1$ , and for $Ra \geqslant 10^9$ at $Pr = 1$ (see table 1).

The decrease in $T_{\it RMS}/\Delta T$ with $Ra$ is related to the thermal boundary layer, which becomes thinner with $Ra$ (Scheel, Kim & White Reference Scheel, Kim and White2012; Pandey Reference Pandey2021). As $T_{\it RMS}$ represents an average measure of the temperature anomaly in the flow, the dominant contribution to $T_{\it RMS}$ arises from regions occupied by thermal plumes. This is because the temperature within the plumes varies slowly and differs strongly from the ambient temperature, which is approximately the mean temperature $\Delta T/2$ in the flow. As the fraction of the volume occupied by the plumes decreases with $Ra$ , so does their contribution to $T_{\it RMS}/\Delta T$ . A similar magnitude of RMS fluctuations for the two Prandtl numbers (and moderate Rayleigh numbers) is due to the similar nature in the two cases of heat transport (see figure 6), which determines the thickness of the thermal boundary layer.

In figure 8(b), we show $T_{\it RMS}/\Delta T$ as a function of ${Re}$ . Although the data for the two Prandtl numbers are distinct, the scaling regimes reveal themselves clearly. We observe that for large ${Re}$ , the scaling exponent of $T_{\it RMS}/\Delta T$ with respect to ${Re}$ is essentially the same for both $Pr$ values. For ${Re} \gt 10^4$ , temperature RMS exhibits the same scaling, $T_{\it RMS}/\Delta T \sim {Re}^{-0.2}$ , for both $Pr$ values. However, the exponents in moderate Reynolds numbers, to the extent that they can be defined at all, are different, with $T_{\it RMS}/\Delta T$ showing ${Re}^{-0.13}$ and ${Re}^{-0.21}$ for $Pr = 0.1$ and $Pr = 1$ , respectively. They are unlikely to be of fundamental significance.

6. Fluctuation of global quantities

We now discuss the fluctuation of integral quantities in the nominally steady state. Taking the dot product of (2.2) with $\boldsymbol{u}$ , and averaging over the entire domain, we obtain

(6.1) \begin{equation} \frac {\partial }{\partial t} \big\langle u_i^2 / 2 \big\rangle _A = \alpha g\, \langle u_z T \rangle _A - \langle \varepsilon _u \rangle _A . \end{equation}

Recalling that $E = \langle (u_x^2 + u_z^2)/2 \rangle _A/u_f^2$ is the domain-averaged kinetic energy, (6.1) takes the non-dimensional form

(6.2) \begin{equation} \sqrt {Ra\,Pr}\, \frac {\partial E}{\partial t} = \overline {Nu} - \overline {Nu_{\varepsilon _u}} , \end{equation}

where $\overline {Nu}$ and $\overline {Nu_{\varepsilon _u}}$ are the instantaneous domain-averaged heat fluxes defined in (3.8) and (3.9), respectively. Equation (6.2) states that $\overline {Nu}$ and $\overline {Nu_{\varepsilon _u}}$ are not equal to each other whenever $E$ varies with time.

Figure 9. Temporal evolution of the integral quantities in statistically steady state for $Pr = 0.1,\ Ra = 10^{10}$ : (a) domain-averaged scaled kinetic energy $\sqrt {Ra\,Pr} \, E$ is dominated by a slow evolution; (b) $\overline {Nu}$ fluctuates rapidly about its mean; (c) $\overline {Nu_{\varepsilon _u}}$ , in addition to having rapidly fluctuating components, evolves slowly and is related to $E$ (see (6.2)); (d) $\overline {Nu_{\varepsilon _T}}$ fluctuates rapidly, but a weak slowly-varying trend is present. The horizontal dashed line in all the plots indicates the time-averaged quantity. Here, the origin is taken to be $3000 \, t_f$ of figure 3.

In figure 9, we show the temporal evolutions of $\sqrt {Ra\,Pr}\, E,\ \overline {Nu},\ \overline {Nu_{\varepsilon _u}},\ \overline {Nu_{\varepsilon _T}}$ in the nominally steady state for $Pr = 0.1,\ Ra = 10^{10}$ . Each quantity evolves differently from the others. Figure 9(a) shows that domain-averaged kinetic energy $E$ contains sizeable changes with respect to its mean value, indicated by a dashed horizontal line, occurring in 40–60 units of free-fall time. In figure 9(c), $\overline {Nu_{\varepsilon _u}}$ shows a slow variation superimposed by strong rapid fluctuations. In contrast, $\overline {Nu}$ in figure 9(b) fluctuates rapidly around its mean value. In figure 9(d), $\overline {Nu_{\varepsilon _T}}$ also fluctuates rapidly about its mean, with weak (but non-monotonic) trends. Data for $Ra \gt 10^8$ at $Pr = 0.1$ exhibit similar characteristics.

Figure 10. Temporal evolution of the integral quantities in steady state for $Pr = 1,\ Ra = 10^{12}$ . The descriptions are the same as in figure 9.

Coming to the energy balance equation, if we average (6.2) over a finite interval of time, say between an initial (non-dimensional) time $t_{\it init}$ and a final time $t_{\it fin}$ , then we obtain

(6.3) \begin{equation} \sqrt {Ra\,Pr}\, \frac {\Delta E}{\Delta t} = \langle Nu \rangle _{\Delta t} - \langle Nu_{\varepsilon _u} \rangle _{\Delta t}, \end{equation}

where $\Delta E = E(t_{\it fin}) - E(t_{\it init})$ , $\Delta t = t_{\it fin} - t_{\it init}$ , and $\langle \cdot \rangle _{\Delta t}$ denotes an averaging over the time interval $\Delta t$ . As there exist long periods of growth or decay of $E$ (see figure 9 a), we can apply (6.3) to those intervals. For example, focusing on a segment where $E$ grows in figure 9 (the region highlighted by red shading), we find the left-hand side of (6.3) to be $\sqrt {Ra\,Pr}\, \Delta E/\Delta t = 23.8$ . During this period, the average values of the heat fluxes are found to be $\langle Nu \rangle _{\Delta t} = 80.3$ and $\langle Nu_{\varepsilon _u} \rangle _{\Delta t} = 56.4$ , yielding 23.9 for the right-hand side. Thus (6.3) is satisfied perfectly. Similarly, in the blue-shaded region in figure 9 where $E$ decays, we obtain $\sqrt {Ra\,Pr}\, \Delta E/\Delta t = -33,\ \langle Nu \rangle _{\Delta t} = 90.4$ and $\langle Nu_{\varepsilon _u} \rangle _{\Delta t} = 123.5$ ; thus the terms of (6.3) balance perfectly again. We note that the power spectra of the quantities shown in figure 9 show slow oscillations, or low-frequency modes, in $E$ and $Nu_{\varepsilon _u}$ , $Nu$ and $Nu_{\varepsilon _T}$ , although these modes diminish in power compared to the significantly strengthening high frequencies.

On the one hand, due to the rapid fluctuation of $\overline {Nu}$ about its mean, its short-term average does not differ much from the long-term average. For example, $\langle Nu \rangle _{\Delta t} = 80.3$ and $\langle Nu \rangle _{\Delta t} = 90.4$ in the same two intervals are not far from the average 84.7. On the other hand, the presence of a strong low-frequency component in $\overline {Nu_{\varepsilon _u}}$ causes short-term averages to differ significantly from the long-term value. For example, $\langle Nu_{\varepsilon _u} \rangle _{\Delta t} = 56.4$ and $\langle Nu_{\varepsilon _u} \rangle _{\Delta t} = 123.5$ in the growing and decaying periods of $E$ , respectively, differ by up to 50–60 % from $Nu_{\varepsilon _u} = 79.2$ . This applies to all the high- ${Re}$ data in 2-D RBC that we have explored. For example, figure 10 demonstrates the same picture for $Pr = 1,\ Ra = 10^{12}$ , where in the red- and blue-shaded regions, $\langle Nu \rangle _{\Delta t}$ is, respectively, larger and smaller than $\langle Nu_{\varepsilon _u} \rangle _{\Delta t}$ , and (6.3) applies perfectly.

As $Ra$ approaches very high values, the overwhelming resources required to simulate convective flows in two dimensions limit the total simulation time available to gather statistics. As a result, $Nu$ and $Nu_{\varepsilon _u}$ may not converge perfectly even if the sufficiency of spatial and temporal resolutions is ensured. For the simulation at $Pr = 0.1,\ Ra = 10^{10}$ (shown in figure 9), $Nu$ and $Nu_{\varepsilon _u}$ differ by more than 6 % (see table 1). Similarly, for $Pr = 1,\ Ra = 10^{12}$ (shown in figure 10), the difference is also approximately 5 %. For lower $Ra$ , on the other hand, there is better convergence, to within 1–2 %. The convergence of $Nu_{\varepsilon _T}$ and $Nu$ is much better because both $\overline {Nu_{\varepsilon _T}}$ and $\overline {Nu}$ oscillate with comparable rapidity about their long-term averages.

7. Summary and conclusions

Our goal here has been to study the nature of the transient evolution of the DNS of 2-D thermal convection, using the no-slip boundary condition on all the walls, along with isothermal bottom and top walls, and adiabatic sidewalls. We illustrate related features using a square box for Prandtl numbers 0.1 and 1, in the Rayleigh number range between $10^6$ and $10^{12}$ . We particularly study the temporal evolution of integral transport quantities – such as the Nusselt number (defined in three different ways) and the turbulent energy – and discuss their scaling. The nominally steady state is reached exponentially with substantial dependence on Rayleigh and Prandtl numbers. Although there is some degree of common behaviour of transients for all the conditions explored here, there is no strict universality to the details of the exponential approach. We find, perhaps not surprisingly, that the velocity field evolves more slowly than the thermal field. That the velocity field evolves more slowly is not surprising because this evolution involves a few intermediate steps from the onset of heating changes. In the relation ${Re} \sim Ra^{\zeta }$ , if the exponent $\zeta \gt 0.5$ as in 2-D RBC, then the ratio $u_{\it RMS}/u_f$ continues to increase with $Ra$ . It is evident from (5.2) (see also (4.2) and (4.3)) that $t_{\it trns}$ increases with increasing fluctuating energy.

We also call attention to large oscillations of the velocity field in what may be regarded effectively as the steady state. One main conclusion is that these low-frequency oscillations are related to differences between the Nusselt number defined by the correlation of $u_z$ and $T$ , and the Nusselt number based on the energy dissipation (see (6.2)). The time to saturation of the turbulent energy is presumably dependent on the formation of the large structure (Smith & Yakhot Reference Smith and Yakhot1993), which itself would depend on the aspect ratio. The relation between the formation of the large structure and the time to saturation remains unclear at present, but it appears that achieving the so-called ultimate state of convection for smooth boundaries is as elusive in two dimensions as in three. The naive expectation that the reduced dimensionality in two dimensions allows one to settle the question of the ultimate state by attaining high enough $Ra$ is thwarted to some degree by the appearance of strong, low-frequency fluctuations that demand excessively long simulation times for definitive scaling relations.

Acknowledgements

We appreciate long-term collaboration on convection studies with J. Schumacher. To him and to E. Lindborg, D. Lohse, J. Wettlaufer and M. Verma, we are grateful for comments on an earlier draft. Lindborg communicated to us that the scalings observed here are consistent with his theory. This research was carried out on the High Performance Computing resources at New York University Abu Dhabi. The authors also gratefully acknowledge Shaheen II of KAUST, Saudi Arabia (under project nos k1491 and k1624) for providing computational resources.

Funding

This material is based upon work supported by Tamkeen under the NYU Abu Dhabi Research Institute grant G1502, and by the KAUST Office of Sponsored Research under Award URF/1/4342-01. A.P. also acknowledges financial support from SERB, India, under grant SRG/2023/001746, as well as from IIT Roorkee under the FIG scheme. Also, NYU supports the research of K.R.S.

Declaration of interests

The authors report no conflict of interest.

Data availability statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Appendix A. Numerical details and effects of resolution on the transient state

As discussed in § 4, the transient time increases rapidly with $Ra$ in 2-D RBC. Therefore, the steady state for high- $Ra$ flows is challenging to attain because the simulations require hundreds or thousands of free-fall times in the transient state, during which the domain-averaged kinetic energy continues to increase with time. Thus in the transient flow state, conducting simulations with a mesh that resolves all relevant scales in the flow is extremely challenging due to a significant increase in the required computational resources and wait time. Therefore, we start the simulation with conduction temperature profile and random perturbations on a coarse mesh, and continue until the domain-averaged kinetic energy stops growing with time and starts to fluctuate about some mean, whose value depends on $Ra$ and $Pr$ . However, it is important to ensure that the steady state that is attained using a coarser mesh is close to the one that would be attained if a finer mesh is used.

Figure 11. (a) Evolution of the domain-averaged energy $E$ for $Pr = 0.1,\ Ra = 3 \times 10^8$ using three different spatial resolutions. The similarity of the evolutions of $E$ , in the statistical sense, suggests that the transient time and the mean energy in the steady state do not depend on the spatial resolution. (b) Here, $\overline {Nu_{\varepsilon _u}}$ exhibits strong fluctuations, especially at moments when a rapid decay is observed in $E$ .

To verify this, we performed simulations for a few governing parameters with different spatial resolutions, and compared the temporal evolution of the integral quantities. In figure 11(a), we show the evolution of $E$ for $Pr = 0.1,\ Ra = 3 \times 10^8$ in three different simulations, for mesh cells of $300^2$ , $690^2$ and $1150^2$ . We can see that the growth of $E$ in the initial stage (for $t \lt 120 \, t_f$ ) is similar in all simulations, although the evolutions differ slightly for intermediate stages. However, once $E$ stops growing and starts fluctuating about some mean, the average value of energy in the steady state does not depend on the spatial resolution. We find that the mean energy for $t \gt 400 t_f$ in figure 11(a) differs only by at most 2 %. We also show the evolution of $\overline {Nu_{\varepsilon _u}}$ in figure 11(b) for the three simulations, and observe wild fluctuations, especially when $E$ decreases rapidly over a short time interval. Thus the flow properties in the steady state seem to be largely unaffected by the spatial resolution used in the transient state.

Figure 12. Evolution of the kinetic energy in the intermediate stage of the transient state for $Pr = 1,$ $Ra = 10^{12}$ . Computational time is close to one million core-hours for the simulation with $5210^2$ mesh cells – much bigger than one thousand core hours needed for simulation with $690^2$ cells. It is clear that performing high- $Ra$ simulation in the transient state with full resolution is infeasible.

Similarly, figure 12 shows $E$ for intermediate stages of the transient state for $Pr = 1,\ Ra = 10^{12}$ . We start the simulation with a coarse resolution having $690^2$ mesh cells (red curve). While $E$ is still growing, we ramp up the resolution and start another simulation with $5210^2$ mesh cells (green curve), in addition to continuing the original one. Figure 12 shows that the trajectories of $E$ in both simulations are similar, while the computational resources differ substantially. On the one hand, simulation with the coarse mesh consumes only one thousand core hours, and the segment shown in figure 12 was obtained in just eight hours on 128 cores. On the other hand, simulation with the finer mesh takes more than 900 hours on 1024 cores (green curve). As the transient time is nearly $8000 \, t_f$ for these parameters (see figure 4), one would need to run the simulation with $5210^2$ mesh cells for 72 000 h on 1024 cores to attain the steady state, which is clearly impossible at present.

References

Castaing, B., Gunaratne, G., Kadanoff, L., Libchaber, A. & Heslot, F. 1989 Scaling of hard thermal turbulence in Rayleigh–Bénard convection. J. Fluid Mech. 204, 130.10.1017/S0022112089001643CrossRefGoogle Scholar
Chillà, F. & Schumacher, J. 2012 New perspectives in turbulent Rayleigh–Bénard convection. Eur. Phys. J. E 35 (7), 58.10.1140/epje/i2012-12058-1CrossRefGoogle ScholarPubMed
Chini, G.P. & Cox, S.M. 2009 Large Rayleigh number thermal convection: heat flux predictions and strongly nonlinear solutions. Phys. Fluids 21 (8), 083603.10.1063/1.3210777CrossRefGoogle Scholar
Doering, C.R. 2020 Absence of evidence for the ultimate state of turbulent Rayleigh–Bénard convection. Phys. Rev. Lett. 124 (22), 229401.10.1103/PhysRevLett.124.229401CrossRefGoogle ScholarPubMed
Doering, C.R., Toppaladoddi, S. & Wettlaufer, J.S. 2019 Absence of evidence for the ultimate regime in two-dimensional Rayleigh–Bénard convection. Phys. Rev. Lett. 123 (25), 259401.10.1103/PhysRevLett.123.259401CrossRefGoogle ScholarPubMed
Foroozani, N., Niemela, J.J., Armenio, V. & Sreenivasan, K.R. 2014 Influence of container shape on scaling of turbulent fluctuations in convection. Phys. Rev. E 90 (6), 063003.10.1103/PhysRevE.90.063003CrossRefGoogle ScholarPubMed
Foroozani, N., Niemela, J.J., Armenio, V. & Sreenivasan, K.R. 2017 Reorientations of the large-scale flow in turbulent convection in a cube. Phys. Rev. E 95 (3), 033107.10.1103/PhysRevE.95.033107CrossRefGoogle Scholar
Gao, Z.-Y., Tao, X., Huang, S.-D., Bao, Y. & Xie, Y.-C. 2024 Flow state transition induced by emergence of orbiting satellite eddies in two-dimensional turbulent Rayleigh–Bénard convection. J. Fluid Mech. 997, A54.10.1017/jfm.2024.847CrossRefGoogle Scholar
Grossmann, S. & Lohse, D. 2000 Scaling in thermal convection: a unifying theory. J. Fluid Mech. 407, 2756.10.1017/S0022112099007545CrossRefGoogle Scholar
Grossmann, S. & Lohse, D. 2001 Thermal convection for large Prandtl numbers. Phys. Rev. Lett. 86 (15), 33163319.10.1103/PhysRevLett.86.3316CrossRefGoogle ScholarPubMed
Howard, L.N. 1972 Bounds on flow quantities. Annu. Rev. Fluid Mech. 4 (1), 473494.10.1146/annurev.fl.04.010172.002353CrossRefGoogle Scholar
Iyer, K.P., Scheel, J.D., Schumacher, J. & Sreenivasan, K.R. 2020 Classical 1/3 scaling of convection holds up to Ra = $10^{15}$ . Proc. Natl Acad. Sci. USA 117 (14), 75947598.10.1073/pnas.1922794117CrossRefGoogle ScholarPubMed
Johnston, H. & Doering, C.R. 2009 Comparison of turbulent thermal convection between conditions of constant temperature and constant flux. Phys. Rev. Lett. 102 (6), 064501.10.1103/PhysRevLett.102.064501CrossRefGoogle ScholarPubMed
Kadanoff, L.P. 2001 Turbulent heat flow: structures and scaling. Phys. Today 54 (8), 3439.10.1063/1.1404847CrossRefGoogle Scholar
Labarre, V., Fauve, S. & Chibbaro, S. 2023 Heat-flux fluctuations revealing regime transitions in Rayleigh–Bénard convection. Phys. Rev. Fluids 8 (5), 053501.10.1103/PhysRevFluids.8.053501CrossRefGoogle Scholar
Lam, S., Shang, X.-D., Zhou, S.-Q. & Xia, K.-Q. 2002 Prandtl number dependence of the viscous boundary layer and the Reynolds numbers in Rayleigh–Bénard convection. Phys. Rev. E 65 (6), 066306.10.1103/PhysRevE.65.066306CrossRefGoogle ScholarPubMed
Lohse, D. & Shishkina, O. 2023 Ultimate turbulent thermal convection. Phys. Today 76 (11), 2632.10.1063/PT.3.5341CrossRefGoogle Scholar
Lohse, D. & Shishkina, O. 2024 Ultimate Rayleigh–Bénard turbulence. Rev. Mod. Phys. 96 (3), 035001.10.1103/RevModPhys.96.035001CrossRefGoogle Scholar
Niemela, J.J., Skrbek, L., Sreenivasan, K.R. & Donnelly, R.J. 2000 Turbulent convection at very high Rayleigh numbers. Nature 404 (6780), 837840.10.1038/35009036CrossRefGoogle ScholarPubMed
Pandey, A. 2021 Thermal boundary layer structure in low-Prandtl-number turbulent convection. J. Fluid Mech. 910, A13.10.1017/jfm.2020.961CrossRefGoogle Scholar
Pandey, A., Krasnov, D., Schumacher, J., Samtaney, R. & Sreenivasan, K.R. 2022 a Similarities between characteristics of convective turbulence in confined and extended domains. Physica D 442, 133537.10.1016/j.physd.2022.133537CrossRefGoogle Scholar
Pandey, A., Krasnov, D., Sreenivasan, K.R. & Schumacher, J. 2022 b Convective mesoscale turbulence at very low Prandtl numbers. J. Fluid Mech. 948, A23.10.1017/jfm.2022.694CrossRefGoogle Scholar
Pandey, A., Kumar, A., Chatterjee, A.G. & Verma, M.K. 2016 Dynamics of large-scale quantities in Rayleigh–Bénard convection. Phys. Rev. E 94 (5), 053106.10.1103/PhysRevE.94.053106CrossRefGoogle ScholarPubMed
Pandey, A. & Sreenivasan, K.R. 2021 Convective heat transport in slender cells is close to that in wider cells at high Rayleigh and Prandtl numbers. Europhys. Lett. 135 (2), 24001.10.1209/0295-5075/ac1bc9CrossRefGoogle Scholar
Pandey, A. & Verma, M.K. 2016 a Scaling of large-scale quantities in Rayleigh–Bénard convection. Phys. Fluids 28 (9), 095105.10.1063/1.4962307CrossRefGoogle Scholar
Pandey, A., Verma, M.K., Chatterjee, A.G. & Dutta, B. 2016 b Similarities between 2D and 3D convection for large Prandtl number. Pramana – J. Phys. 87 (1), 13.10.1007/s12043-016-1204-zCrossRefGoogle Scholar
van der Poel, E.P., Stevens, R.J.A.M. & Lohse, D. 2011 Connecting flow structures and heat flux in turbulent Rayleigh–Bénard convection. Phys. Rev. E 84 (4), 045303(R).10.1103/PhysRevE.84.045303CrossRefGoogle ScholarPubMed
van der Poel, E.P., Stevens, R.J.A.M. & Lohse, D. 2013 Comparison between two- and three-dimensional Rayleigh–Bénard convection. J. Fluid Mech. 736, 177194.10.1017/jfm.2013.488CrossRefGoogle Scholar
Samuel, R. & Verma, M.K. 2024 Bolgiano–Obukhov scaling in two-dimensional Rayleigh–Bénard convection at extreme Rayleigh numbers. Phys. Rev. Fluids 9 (2), 023502.10.1103/PhysRevFluids.9.023502CrossRefGoogle Scholar
Samuel, R.J., Bode, M., Scheel, J.D., Sreenivasan, K.R. & Schumacher, J. 2024 No sustained mean velocity in the boundary region of plane thermal convection. J. Fluid Mech. 996, A49.10.1017/jfm.2024.853CrossRefGoogle Scholar
Scheel, J.D., Emran, M.S. & Schumacher, J. 2013 Resolving the fine-scale structure in turbulent Rayleigh–Bénard convection. New J. Phys. 15 (11), 113063.10.1088/1367-2630/15/11/113063CrossRefGoogle Scholar
Scheel, J.D., Kim, E. & White, K.R. 2012 Thermal and viscous boundary layers in turbulent Rayleigh–Bénard convection. J. Fluid Mech. 711, 281305.10.1017/jfm.2012.392CrossRefGoogle Scholar
Scheel, J.D. & Schumacher, J. 2016 Global and local statistics in turbulent convection at low Prandtl numbers. J. Fluid Mech. 802, 147173.10.1017/jfm.2016.457CrossRefGoogle Scholar
Schmalzl, J., Breuer, M. & Hansen, U. 2004 On the validity of two-dimensional numerical approaches to time-dependent thermal convection. Europhys. Lett. 67 (3), 390396.10.1209/epl/i2003-10298-4CrossRefGoogle Scholar
Schumacher, J., Götzfried, P. & Scheel, J.D. 2015 Enhanced enstrophy generation for turbulent convection in low-Prandtl-number fluids. Proc. Natl Acad. Sci. USA 112 (31), 95309535.10.1073/pnas.1505111112CrossRefGoogle ScholarPubMed
Schumacher, J. & Sreenivasan, K.R. 2020 Colloquium: unusual dynamics of convection in the Sun. Rev. Mod. Phys. 92 (4), 041001.10.1103/RevModPhys.92.041001CrossRefGoogle Scholar
Shraiman, B.I. & Siggia, E.D. 1990 Heat transport in high-Rayleigh-number convection. Phys. Rev. A 42 (6), 36503653.10.1103/PhysRevA.42.3650CrossRefGoogle ScholarPubMed
Siggia, E.D. 1994 High Rayleigh number convection. Annu. Rev. Fluid Mech. 26 (1), 137168.10.1146/annurev.fl.26.010194.001033CrossRefGoogle Scholar
Smith, L.M. & Yakhot, V. 1993 Bose condensation and small-scale structure generation in a random force driven 2D turbulence. Phys. Rev. Lett. 71 (3), 352355.10.1103/PhysRevLett.71.352CrossRefGoogle Scholar
Sreenivasan, K.R., Bershadskii, A. & Niemela, J.J. 2002 Mean wind and its reversal in thermal convection. Phys. Rev. E 65 (5), 056306.10.1103/PhysRevE.65.056306CrossRefGoogle ScholarPubMed
Stevens, R., Lohse, D. & Verzicco, R. 2011 Prandtl and Rayleigh number dependence of heat transport in high Rayleigh number thermal convection. J. Fluid Mech. 688, 3143.10.1017/jfm.2011.354CrossRefGoogle Scholar
Stevens, R., Verzicco, R. & Lohse, D. 2010 Radial boundary layer structure and Nusselt number in Rayleigh–Bénard convection. J. Fluid Mech. 643, 495507.10.1017/S0022112009992461CrossRefGoogle Scholar
Stevens, R.J., Hartmann, R., Verzicco, R. & Lohse, D. 2024 How wide must Rayleigh–Bénard cells be to prevent finite aspect ratio effects in turbulent flow? J. Fluid Mech. 1000, A58.10.1017/jfm.2024.996CrossRefGoogle Scholar
Tiwari, H., Sharma, L. & Verma, M.K. 2025 Compressible turbulent convection at very high Rayleigh numbers. Intl J. Heat Mass Transfer 242, 126821.10.1016/j.ijheatmasstransfer.2025.126821CrossRefGoogle Scholar
Verma, M.K. 2018 Physics of Buoyant Flows. World Scientific.10.1142/10928CrossRefGoogle Scholar
Verma, M.K., Kumar, A. & Pandey, A. 2017 Phenomenology of buoyancy-driven turbulence: recent results. New J. Phys. 19 (2), 025012.10.1088/1367-2630/aa5d63CrossRefGoogle Scholar
Verzicco, R. & Camussi, R. 1999 Prandtl number effects in convective turbulence. J. Fluid Mech. 383, 5573.10.1017/S0022112098003619CrossRefGoogle Scholar
Wan, Z.-H., Wang, Q., Wang, B., Xia, S.-N., Zhou, Q. & Sun, D.-J. 2020 On non-Oberbeck–Boussinesq effects in Rayleigh–Bénard convection of air for large temperature differences. J. Fluid Mech. 889, A10.10.1017/jfm.2020.66CrossRefGoogle Scholar
Weiss, S. & Ahlers, G. 2011 Turbulent Rayleigh–Bénard convection in a cylindrical container with aspect ratio $\gamma$ = 0.50 and Prandtl number Pr = 4.38. J. Fluid Mech. 676, 540.10.1017/S0022112010005963CrossRefGoogle Scholar
Wen, B., Goluskin, D., LeDuc, M., Chini, G.P. & Doering, C.R. 2020 Steady Rayleigh–Bénard convection between stress-free boundaries. J. Fluid Mech. 905, R4.10.1017/jfm.2020.812CrossRefGoogle Scholar
Xi, H. & Xia, K. 2008 Flow mode transitions in turbulent thermal convection. Phys. Fluids 20 (5), 5104.10.1063/1.2920444CrossRefGoogle Scholar
Xia, K.Q., Sun, C. & Zhou, S.Q. 2003 Particle image velocimetry measurement of the velocity field in turbulent thermal convection. Phys. Rev. E 68 (6), 066303.10.1103/PhysRevE.68.066303CrossRefGoogle ScholarPubMed
Zhang, Y., Zhou, Q. & Sun, C. 2017 Statistics of kinetic and thermal energy dissipation rates in two-dimensional turbulent Rayleigh–Bénard convection. J. Fluid Mech. 814, 165184.10.1017/jfm.2017.19CrossRefGoogle Scholar
Zhu, X., Mathai, V., Stevens, R.J.A.M., Verzicco, R. & Lohse, D. 2018 Transition to the ultimate regime in two-dimensional Rayleigh–Bénard convection. Phys. Rev. Lett. 120 (14), 144502.10.1103/PhysRevLett.120.144502CrossRefGoogle Scholar
Figure 0

Table 1. Important parameters of DNS in a 2-D box with $\Gamma = 1$. We list the Prandtl number, the Rayleigh number, the total number of mesh cells in the entire flow domain $N_e N^2$, the Nusselt numbers using (3.3), (3.6), (3.7) and (3.2), respectively, the Reynolds number, and the simulation time after the flow attains a steady state, $t_{sim}$. Error bars in Nusselt and Reynolds numbers are the corresponding standard deviations.

Figure 1

Figure 1. Evolution of the integral quantities in the transient state for $Pr = 0.1$, and $Ra = 3 \times 10^8$ (black curves) and $Ra = 3 \times 10^9$ (orange curves). (a) The domain-averaged kinetic energy $E$ increases slowly and takes a few thousand free-fall times to reach the steady state. The transient time is longer for higher $Ra$. Dashed vertical lines show a quantitative measure of the transient time $t_{\it trns}$, obtained from (4.3), to be discussed later. (b,c,d) Plots show that the Nusselt number fluctuates rapidly about its mean nearly from the start, but fluctuations have different characters depending on the definition of the Nusselt number. In (b), the fluctuations in $\overline {Nu}$ are strong and of high frequency with no well-defined transient state, and there is an overlap for the two $Ra$ values. In (c), one can approximately identify a transient state in $\overline {Nu_{\varepsilon _u}}$, which exhibits very strong fluctuations containing both high and low frequencies. Plot (d) shows that $\overline {Nu_{\varepsilon _T}}$ fluctuates similarly to $\overline {Nu}$ in (b), but there is no overlap for the two Rayleigh numbers. The occasional appearance of negative values of $\overline {Nu}$ suggests the likelihood that a small parcel of the coldest fluid from the top wall registers directly at the bottom wall, and vice versa.

Figure 2

Figure 2. Global (area- as well as time-averaged) kinetic energy in the steady state $E_{av}$ as a function of $Ra$. An increasing trend with $Ra$ is observed in 2-D RBC. In the turbulent regimes ($Ra \geqslant 10^8$ for $Pr = 0.1$, and $Ra \gt 10^{9}$ for $Pr = 1$), the data approximately follow $E_{av} \sim Ra^{1/3}$, shown as dashed lines. In contrast, $E_{av}$ in 3-D RBC for $Pr = 0.7$ (taken from Samuel et al.2024) shows a weakly decreasing trend. The transitions observed in this figure are related to transitions in the flow structure; see § 5.3. The statistical error bars in almost all the figures here are comparable to the thickness of the symbols. The exception is figure 6, for which the errors bars are shown explicitly.

Figure 3

Figure 3. Evolution of the domain-averaged kinetic energy $E(t)$ in the transient state for $Pr = 0.1$ and $Ra = 3 \times 10^9$ and $10^{10}$ can be described well by (4.2), and the dashed curves show $E_{\it fit}(t)$ for the growth rate. Vertical lines show that the transient time $t_{\it trns}$ is much longer for $Ra = 10^{10}$ than for $Ra = 3 \times 10^9$.

Figure 4

Figure 4. (a) The transient time $t_{\it trns}$ as a function of $Ra$ shows a power law. For a given $Ra$, $t_{\it trns}$ is longer for lower $Pr$. (b) The time $t_{\it trns}$ as a function of ${Re}$ is essentially the same for both values of $Pr$, and exhibits a nearly linear trend. Moreover, $t_{\it trns}$ is nearly the same for both $Pr$ values when the Reynolds numbers are the same. Data correspond only to the turbulent regimes. Sparser data sets for $Pr = 0.021$ are consistent with the Prandtl numbers of these plots.

Figure 5

Figure 5. Inverse growth rate as a function of (a) $Ra$ and (b) ${Re}$. The trends with ${Re}$ are approximately the same for both Prandtl numbers, and are slightly below linear.

Figure 6

Figure 6. (a) The Nusselt number as a function of $Ra$ for $Pr = 0.1$ (red circles) and $Pr = 1$ (blue triangles). The scaling for high Rayleigh numbers differs only slightly from the low-$Ra$ behaviour. (b) The normalised Nusselt number $Nu\, Ra^{-2/7}$ shows that the $2/7$ scaling is only approximate for moderate Rayleigh numbers and moderate aspect ratios. Wall heat flux from (3.2) is shown here with the error bars representing the standard deviation.

Figure 7

Figure 7. The Reynolds number based on $u_{\it RMS}$ scales nearly as $Ra^{2/3}$ in the turbulent regime (indicated by the blue solid line), which is distinctively different from scaling $Ra^{1/2}$ reported in 3-D RBC. The green dashed line indicates the ${Re} \sim Ra^{0.46}$ scaling observed for 3-D RBC in a $\Gamma = 4$ box by Samuel et al. (2024).

Figure 8

Figure 8. (a) The RMS temperature fluctuation $T_{\it RMS}$ decreases with $Ra$ for both $Pr$ values, and has nearly the same magnitude for moderate Rayleigh numbers. Different exponents as well as prefactors are found for moderate and high Rayleigh numbers. (b) Fluctuation $T_{\it RMS}$ as a function of ${Re}$ shows that it scales approximately as ${Re}^{-0.2}$ in the turbulent regime (for ${Re} \gt 10^4$) for both Prandtl numbers.

Figure 9

Figure 9. Temporal evolution of the integral quantities in statistically steady state for $Pr = 0.1,\ Ra = 10^{10}$: (a) domain-averaged scaled kinetic energy $\sqrt {Ra\,Pr} \, E$ is dominated by a slow evolution; (b) $\overline {Nu}$ fluctuates rapidly about its mean; (c) $\overline {Nu_{\varepsilon _u}}$, in addition to having rapidly fluctuating components, evolves slowly and is related to $E$ (see (6.2)); (d) $\overline {Nu_{\varepsilon _T}}$ fluctuates rapidly, but a weak slowly-varying trend is present. The horizontal dashed line in all the plots indicates the time-averaged quantity. Here, the origin is taken to be $3000 \, t_f$ of figure 3.

Figure 10

Figure 10. Temporal evolution of the integral quantities in steady state for $Pr = 1,\ Ra = 10^{12}$. The descriptions are the same as in figure 9.

Figure 11

Figure 11. (a) Evolution of the domain-averaged energy $E$ for $Pr = 0.1,\ Ra = 3 \times 10^8$ using three different spatial resolutions. The similarity of the evolutions of $E$, in the statistical sense, suggests that the transient time and the mean energy in the steady state do not depend on the spatial resolution. (b) Here, $\overline {Nu_{\varepsilon _u}}$ exhibits strong fluctuations, especially at moments when a rapid decay is observed in $E$.

Figure 12

Figure 12. Evolution of the kinetic energy in the intermediate stage of the transient state for $Pr = 1,$$Ra = 10^{12}$. Computational time is close to one million core-hours for the simulation with $5210^2$ mesh cells – much bigger than one thousand core hours needed for simulation with $690^2$ cells. It is clear that performing high-$Ra$ simulation in the transient state with full resolution is infeasible.