Brought to you by:

Prandtl-number Effects in High-Rayleigh-number Spherical Convection

, , , and

Published 2018 March 20 © 2018. The American Astronomical Society. All rights reserved.
, , Citation Ryan J. Orvedahl et al 2018 ApJ 856 13 DOI 10.3847/1538-4357/aaaeb5

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/856/1/13

Abstract

Convection is the predominant mechanism by which energy and angular momentum are transported in the outer portion of the Sun. The resulting overturning motions are also the primary energy source for the solar magnetic field. An accurate solar dynamo model therefore requires a complete description of the convective motions, but these motions remain poorly understood. Studying stellar convection numerically remains challenging; it occurs within a parameter regime that is extreme by computational standards. The fluid properties of the convection zone are characterized in part by the Prandtl number $\Pr $ = ν/κ, where ν is the kinematic viscosity and κ is the thermal diffusion; in stars, $\Pr $ is extremely low, $\Pr $ ≈ 10−7. The influence of $\Pr $ on the convective motions at the heart of the dynamo is not well understood since most numerical studies are limited to using $\Pr $ ≈ 1. We systematically vary $\Pr $ and the degree of thermal forcing, characterized through a Rayleigh number, to explore its influence on the convective dynamics. For sufficiently large thermal driving, the simulations reach a so-called convective free-fall state where diffusion no longer plays an important role in the interior dynamics. Simulations with a lower $\Pr $ generate faster convective flows and broader ranges of scales for equivalent levels of thermal forcing. Characteristics of the spectral distribution of the velocity remain largely insensitive to changes in $\Pr $. Importantly, we find that $\Pr $ plays a key role in determining when the free-fall regime is reached by controlling the thickness of the thermal boundary layer.

Export citation and abstract BibTeX RIS

1. Introduction

Convective motions within the outer one-third of the Sun transport energy from the radiative interior to the photosphere. In the process, these overturning motions, which are thought to drive the solar differential rotation, play a pivotal role in generating the solar magnetic field. Any model of the solar dynamo necessarily requires a description of the Sun's underlying convective motions, and yet those motions remain poorly characterized in spite of the observational coverage enabled by the Sun's proximity to Earth.

On the largest spatial scales, photospheric convection manifests in cellular patterns known as supergranules that were first noted by Hart (1954) and better characterized by Leighton et al. (1962). The horizontal extent of these cells is approximately 35 Mm, and they possess a spectral peak in photospheric Dopplergram power around the spherical harmonic degree  ≈ 120 (e.g., Hathaway et al. 2000, 2015). In addition, smaller-scale motions known as granulation are clearly visible in the photosphere, possessing a characteristic size of about 1 Mm and a clear peak in photospheric Dopplergram power around  ≈ 103 (Bray et al. 1984).

Presently, only the granular component of photospheric convection is reliably captured in direct numerical simulations. Radiative hydrodynamic simulations of solar surface convection that can simulate granulation fail to yield clear evidence for supergranulation (e.g., Stein et al. 2009; Ustyugov 2010). Inconsistencies between numerical models and the Sun have also been observed in the velocity power distribution associated with larger scales of convection (e.g., Miesch et al. 2008). Measurements of deep convective flow speeds, made using time-distance helioseismology, suggest that convection models may overestimate the amplitude of the convection on spatial scales larger than 30 Mm (Hanasoge et al. 2012). The results of Hanasoge et al. (2012) estimate that the convective velocities on spatial scales larger than 70 Mm is at most 5–6 m s−1, about an order of magnitude weaker than that expected from simulations or theoretical arguments (e.g., Miesch et al. 2012). Ring-analysis measurements of the subsurface flows in the near-surface shear layer, however, exhibit good agreement with the models and theory (Greer et al. 2015).

Resolving these discrepancies requires careful comparison of these two different helioseismic techniques, and perhaps improvements to both. A resolution to this problem also requires a better theoretical understanding of convective dynamics under stellar conditions. Exploring stellar convection numerically remains challenging owing to the fact that it occurs within parameter regimes that are considered extreme by modern computational and laboratory standards alike. These regimes can be characterized by several nondimensional parameters. In particular, the Reynolds number measures the relative importance of inertial forces to viscous forces, the Rayleigh number expresses the relative strength of buoyancy driving and diffusive effects, and the Prandtl number specifies the relative importance of viscosity to thermal diffusion. Estimates for the Sun lead to values of the Reynolds, Rayleigh, and Prandtl numbers on the order of 1013, 1020, and 10−7, respectively, indicating that the solar convection zone is highly turbulent (e.g., Ossendrijver 2003).

Such extreme values of nondimensional parameters are largely an expression of the fact that while diffusion in stellar interiors may be active on very small scales, it tends to be negligible at the system scale. Achieving such a situation, wherein diffusion plays no appreciable role in the leading-order force balance, is possible in those parameter regimes already accessible through computational models. Featherstone & Hindman (2016a, 2016b) identified two such regimes by exploring the response of convection to changes in the Rayleigh number and the Ekman number (which expresses the relative importance of the Coriolis force and viscous diffusion), while using a fixed Prandtl number of unity. In so doing, they identified asymptotic scaling laws for spectral properties of the convection that in principle could be extrapolated to the stellar parameter regime. The purpose of this paper is to extend those studies by exploring the response of the convective spectrum to changes in the Prandtl number.

This paper is organized as follows. In Section 1.1, we summarize earlier investigations of the role of the Prandtl number. In Sections 2 and 3, we discuss our numerical model and the parameter space that was explored. Results are discussed in Section 4, followed by a discussion of their implications in Section 5.

1.1. Prandtl Number in Convection

The Prandtl number, $\Pr $ = ν/κ, where ν is the kinematic viscosity and κ is the thermal diffusivity, is known to be small in the dynamo regions of planets and stars (Ossendrijver 2003; Roberts 2007). It is well known from linear theory that the value of the Prandtl number both influences the structure and amplitude of convective motions and controls the critical Rayleigh number required for the onset of rotating convection (e.g., Chandrasekhar 1961). Asymptotic approximations to the governing fluid equations can be carried out based on the size of the Prandtl number. Such approximations yield some insight into the convective dynamics arising under different Prandtl-number regimes. In the large Prandtl-number limit, the influence of inertia is weak and can be neglected; this limit is routinely exploited in studying the convection of planetary mantles where the Prandtl number can be of order 1020 (Schubert et al. 2001). Spiegel (1962) developed an approximate set of equations valid in the limit $\Pr \to 0$ that showed that inertia plays a leading-order role in the convective dynamics (see also Thual 1992). Both of these approximate models have been applied only to nonrotating and incompressible Boussinesq systems to date.

Much less is known about the role of the Prandtl number in compressible convection. Earlier studies of compressible convection have primarily used a Prandtl number of order unity (e.g., Gilman 1977; Gilman & Glatzmaier 1981; Goudard & Dormy 2008; Christensen 2011; Schrinner et al. 2012; Soderlund et al. 2012; Gastine et al. 2015, 2016; Wicht & Meduri 2016). While there are also many studies that make use of nonunity Prandtl numbers (e.g., Brown et al. 2011; Käpylä et al. 2013; Jones 2014; Nelson et al. 2014; Augustson et al. 2015, 2016; Duarte et al. 2016; Brun et al. 2017), no parameter studies that vary the Prandtl number systematically have been carried out. One exception to this trend is the work of O'Mara et al. (2016), who explored the characteristics of high Prandtl number compressible convection. These authors found that high Prandtl number convection tended to possess lower characteristic flow speeds with respect to unity Prandtl-number convection, owing to the enhanced entropy content of its downflow plumes. Finally, we note that recent work using a small Prandtl number with rapid rotation has found that the anelastic approximation can yield spurious behavior (Calkins et al. 2015a) that does not appear in nonrotating anelastic convection (Calkins et al. 2015b). These results raise serious questions regarding the applicability of the anelastic approximation within rotating stellar interiors where it remains to be seen if the convective flows are well-approximated by a Prandtl number of unity.

Systematic parameter space studies of convective dynamics in stellar interiors have, so far, focused largely on the role of buoyancy driving and rotation. Featherstone & Hindman (2016b) investigated the response of the convection to varying Rayleigh numbers and varying degrees of density stratification. Those simulations were nonrotating, hydrodynamic, $\Pr $ = 1, and demonstrated a clear scaling relationship between kinetic energy and Rayleigh number. Those results also suggest that a naive interpretation of model results (by ascribing solar values to all of the problem parameters but the diffusion coefficients) will naturally overestimate the low-wavenumber power in the convective power spectrum. The influence of rotation was investigated using a similar methology by Featherstone & Hindman (2016a) who identified a complementary scaling law relating convective-cell size and rotational influence. When rotation is present, and diffusive effects are negligible, the typical spatial size of convective cells is determined primarily through the Rossby number, which expresses the ratio of the rotation period to a characteristic convective timescale. Their work was also restricted to $\Pr $ = 1.

Through this paper, we extend these studies and examine the effects of Prandtl-number variation on the convective dynamics. We present a series of numerical simulations designed to examine how the structure and amplitude of the convective flow within a stellar interior depends on the Prandtl number and the convective forcing. We vary the Prandtl number and the convective forcing in a systematic way, covering both low and high Prandtl numbers. Effects due to rotation and magnetism are not included. We will show that the convection develops smaller-scale structures as the convective forcing is increased and the Prandtl number is decreased, corresponding to an increase in high-wavenumber power. As the high-wavenumber power increases, the low-wavenumber power decreases and this trend occurs for all of the Prandtl numbers studied. We also show that the Prandtl number has an important influence on the boundary-layer thickness.

2. Numerical Model

This study is based on a series of 3D, nonlinear convection simulations that use the pseudo-spectral convection code Rayleigh (e.g., Featherstone & Hindman 2016b). We employ a spherical geometry and represent the horizontal variation of all variables along spherical surfaces using spherical harmonics Ym(θ, ϕ). Here is the spherical harmonic degree, and m is the azimuthal mode order. In the radial direction, we employ a Chebyshev collocation method, expanding all variables in Chebyshev polynomials Tn(r), where n is the degree of the polynomial.

We are particularly interested in understanding convection in the deep stellar interior, far removed from the photosphere. In this region, plasma motions are subsonic and perturbations to thermodynamic variables are small compared to their mean, horizontally averaged values (represented using overbars). Under these conditions, the anelastic approximation provides a convenient description of the system's thermodynamics (Gough 1969; Gilman & Glatzmaier 1981). The governing evolution equations include the continuity equation

Equation (1)

where $\bar{\rho }$ is the background density and ${\boldsymbol{u}}$ is the fluid velocity. The momentum equation is given by

Equation (2)

where P is the pressure, S is the entropy, cp is specific heat at constant pressure, ${\boldsymbol{g}}$ is the gravitational acceleration, and the viscous stress tensor ${ \mathcal D }$ is given by

Equation (3)

Here, eij is the strain rate tensor and δij is the Kronecker delta. Written in terms of the entropy, the thermal energy equation is given by

Equation (4)

where $\bar{T}$ is the background temperature. Sources of internal heating and cooling are encapsulated in the functional form of Q. A linearized equation of state closes the system and is given by

Equation (5)

assuming the ideal gas law

Equation (6)

where ${ \mathcal R }$ is the gas constant and γ = 5/3 is the adiabatic index.

3. Numerical Experiment

We have constructed a set of 34 model stellar convection zones designed to explore how the convective kinetic energy depends on both the thermal diffusion and the viscous diffusion as characterized by the Prandtl number. The diffusion coefficients are taken to be constant values within each simulation. In particular, they have no variation with radius. Table 1 has a detailed list of all of the model parameters for each run.

Table 1.  List of All of the Simulation Parameters for Each Run

Input Parameters Output Parameters
κ ν ${\mathrm{Ra}}_{F}$ $\Pr $ nmax ${{\ell }}_{\max }$ KE $\widehat{\mathrm{KE}}$ fconv δBL $\mathrm{Re}$ $\mathrm{Re}$peak
(1012 cm2 s−1)         (1038 erg)     (Mm)
10 1 6.88 × 104 0.1 85 1023 34.43 19064.7 0.6620 15.57 248.3 342.1
4 1 4.30 × 105 0.25 85 1023 42.59 23582.9 0.8301 10.36 261.2 376.9
8 2 5.37 × 104 0.25 85 511 46.56 6446.4 0.7026 13.76 132.8 192.8
12 3 1.59 × 104 0.25 85 511 42.08 2588.8 0.5876 16.18 85.1 123.4
16 4 6.72 × 103 0.25 85 263 35.80 1238.4 0.5087 17.68 58.7 83.6
24 6 1.99 × 103 0.25 85 127 21.07 324.8 0.3251 20.78 30.8 44.9
32 8 8.40 × 102 0.25 85 127 9.36 81.6 0.1569 23.26 15.7 23.0
4 2 2.15 × 105 0.5 85 511 42.24 5847.2 0.8403 9.87 129.0 186.0
6 3 6.37 × 104 0.5 85 511 40.46 2489.6 0.7750 11.98 83.7 121.9
8 4 2.68 × 104 0.5 85 263 39.62 1371.2 0.7167 13.71 62.7 91.9
12 6 7.96 × 103 0.5 85 127 34.51 530.8 0.6031 16.18 38.5 56.5
16 8 3.36 × 103 0.5 85 127 24.84 214.8 0.4860 18.18 25.6 38.4
24 12 9.96 × 102 0.5 85 127 79.94 30.4 0.2314 22.21 10.4 16.5
1 1 6.81 × 106 1 85 1023 36.22 20057.7 0.9480 4.96 229.8 320.4
2 2 8.53 × 105 1 85 511 34.73 4807.6 0.9120 6.93 114.1 166.5
4 4 1.06 × 105 1 85 263 33.42 1156.7 0.8470 9.83 56.3 85.6
6 6 3.16 × 104 1 85 127 32.92 506.4 0.7910 12.18 37.5 58.2
8 8 1.33 × 104 1 85 127 31.62 273.5 0.7360 14.09 27.5 42.4
12 12 3.94 × 103 1 85 127 22.08 84.9 0.6230 16.03 15.4 24.0
16 16 1.66 × 103 1 42 127 12.02 26.0 0.4420 19.06 9.0 15.6
24 24 4.93 × 102 1 42 63 2.05 2.0 0.1220 23.77 2.5 4.5
2 4 4.30 × 105 2 85 511 27.22 941.9 0.9145 6.92 53.5 78.0
4 8 5.37 × 104 2 85 263 25.60 221.4 0.8539 10.18 26.6 40.0
6 12 1.59 × 104 2 85 127 24.37 93.7 0.8038 13.43 17.3 26.2
8 16 6.72 × 103 2 85 127 20.11 43.5 0.7398 15.22 12.1 18.7
12 24 1.99 × 103 2 85 127 10.04 9.6 0.5277 18.35 6.1 9.9
16 32 8.40 × 102 2 42 127 4.33 2.3 0.3393 20.65 3.1 5.2
1 4 1.72 × 106 4 85 1023 17.10 591.8 0.9490 5.02 45.3 67.1
2 8 2.15 × 105 4 85 511 20.26 175.2 0.9150 7.11 24.6 37.9
4 16 2.68 × 104 4 85 263 17.12 37.0 0.8641 11.62 11.7 18.5
6 24 7.96 × 103 4 85 127 11.88 11.4 0.7953 14.10 6.9 11.5
8 32 3.36 × 103 4 85 127 8.10 4.3 0.6808 15.86 4.2 7.1
12 48 9.96 × 102 4 85 127 3.66 0.8 0.4438 19.16 1.9 3.2
16 64 4.29 × 102 4 42 127 0.63 0.1 0.0969 24.08 0.6 1.0

Note. Each simulation used a polytropic background state with an adiabatic index of γ = 5/3, three density scale heights across the domain Nρ = 3, and a polytropic index of n = 1.5. The inner and outer radii of each simulation were ri = 5 × 1010 cm and ro = 6.586 × 1010 cm. The variable input parameters are the thermal diffusivity κ, the kinematic viscosity ν, the Rayleigh number, the Prandtl number, the radial resolution, and the azimuthal resolution. The output parameters are the dimensional kinetic energy KE, nondimensional kinetic energy $\widehat{\mathrm{KE}}$, the fractional convective flux fconv, the dimensional thermal boundary-layer thickness ${\delta }_{\mathrm{BL}}$, the Reynolds number, and the peak Reynolds number.

Download table as:  ASCIITypeset image

Each model is constructed using a polytropic background state following Jones et al. (2011). The background states were constructed in a similar fashion to the models presented in Featherstone & Hindman (2016b). We use a polytropic index of n = 1.5, which corresponds to the adiabatic value, and model the innermost three density scale heights of the convection zone. The spherical shell has an aspect ratio of χ = ri/ro = 0.759, corresponding to a dimensional shell depth of 159 Mm, where ri and ro are the inner and outer radii of the domain, respectively.

Our models are fully characterized by two parameters: a Rayleigh number and a Prandtl number. As discussed in Featherstone & Hindman (2016b), a flux Rayleigh number ${\mathrm{Ra}}_{F}$ appropriate for this system may be defined as

Equation (7)

where tildes indicate the volume averages over the full shell, making ${\mathrm{Ra}}_{F}$ a bulk Rayleigh number. In this definition, F is the thermal energy imposed by the radiative heating and H is chosen to be the shell depth, ${r}_{o}-{r}_{i}$. The nondimensionalization was carried out using the shell depth and the viscous diffusion timescale H2/ν.

Heat enters the system through the internal deposition by Q, which drops to zero at the upper boundary. In all of the simulations we adopt a functional form of Q that depends only on the background pressure profile such that

Equation (8)

The normalization is defined so that

Equation (9)

where L is the stellar luminosity. The thermal energy flux $F(r)$ that convection and conduction must transport across a spherical surface at radius r is then given by

Equation (10)

For all of the simulations, we have adopted impenetrable and stress-free boundary conditions on the velocity. The radial entropy gradient is forced to vanish at the lower boundary of the convection zone, and the entropy perturbations are forced to vanish at the upper boundary.

Our numerical experiments span the range of 4 × 102 ≲ ${\mathrm{Ra}}_{F}$ ≲  7 × 106 and 0.1 ≤  $\Pr $ ≤ 4. Each simulation was initialized using a small random thermal perturbation, evolved until the kinetic energy reached a statistically steady state, and further evolved for at least one diffusion time. Since there are two diffusion timescales, the larger of the two was used for this purpose. The larger of the two diffusion times is also the time interval over which a time average is computed when necessary. This averaging interval includes several tens of convective overturning times.

4. Survey of Results

4.1. Kinetic Energy Scaling

We begin our examination of the convective energetics by looking at the integrated kinetic energy KE, defined as

Equation (11)

where the integration is computed over the entire domain and then time averaged.

Figure 1(a) shows the dimensional kinetic energy versus ${\mathrm{Ra}}_{F}$. Different symbols indicate different values of $\Pr $. As the Prandtl number is lowered, the kinetic energy increases for any given ${\mathrm{Ra}}_{F}$. The kinetic energy for those runs with ${\mathrm{Ra}}_{F}$ ≳ 2 × 104 appears to have reached a steady value that is independent of ${\mathrm{Ra}}_{F}$. The level at which the kinetic energy saturates is dependent on the Prandtl number. The saturation of the kinetic energy as the Rayleigh number is changed was also found in Featherstone & Hindman (2016b), although their study was restricted to a Prandtl number of unity.

Figure 1.

Figure 1. (a) Dimensional kinetic energy vs. flux Rayleigh number ${\mathrm{Ra}}_{F}$ for all cases. Colored symbols indicate different Prandtl numbers. Low Prandtl number runs have higher kinetic energies for a given ${\mathrm{Ra}}_{F}$. Beyond the high-${\mathrm{Ra}}_{F}$ cutoff, denoted by the vertical dotted–dashed line, the kinetic energy tends toward an asymptotic value. (b) Reynolds number vs. ${\mathrm{Ra}}_{F}$ $\Pr $−2. The high-${\mathrm{Ra}}_{F}$ region reaches an asymptotic regime with a power law scaling exponent that is very close to one-third. The dashed line is ${({\mathrm{Ra}}_{F}{\Pr }^{-2})}^{0.373}$. Colored symbols are the same as in panel (a).

Standard image High-resolution image

As $\Pr $ is increased, ν becomes larger leading to enhanced viscous dissipation and a smaller Reynolds number. The Reynolds number measures the relative ratio of inertial forces to viscous forces and is given by

Equation (12)

where H and the tilde retain the same meaning as before, representing the shell depth and a volume average, respectively. Larger Prandtl numbers will produce smaller Reynolds numbers for a given ${\mathrm{Ra}}_{F}$.

Figure 1(a) indicates that below some ${\mathrm{Ra}}_{F}$ cutoff, diffusion plays a leading-order role in the force balance. Beyond ${\mathrm{Ra}}_{F}$ ∼ 2 × 104, which we denote the high-${\mathrm{Ra}}_{F}$ regime,6 this is no longer the case; diffusion no longer plays a leading-order role in the global force balance and the kinetic energy remains constant as ${\mathrm{Ra}}_{F}$ is increased. We note that the cutoff for the high-${\mathrm{Ra}}_{F}$ region is based on the $\Pr $ = 1 results of Featherstone & Hindman (2016b). Importantly, the cutoff is a decreasing function of the Prandtl number; lower Prandtl numbers will have a lower high-${\mathrm{Ra}}_{F}$ cutoff. To simplify the analysis, only a single cutoff is used.

The kinetic energy can also be discussed from a nondimensional point of view. To do this, we choose a nondimensional measure of the kinetic energy $\widehat{\mathrm{KE}}$ as

Equation (13)

The nondimensionalization has been carried out using the mass M contained within the spherical shell, the shell depth, and the viscous diffusion timescale. Under such a nondimensionalization, the kinetic energy can be related to the Reynolds number as

Equation (14)

Therefore we do not plot the nondimensional kinetic energy separately, but do list it for each run in Table 1.

If ${\mathrm{Ra}}_{F}$ is large enough to be in the high-${\mathrm{Ra}}_{F}$ regime, then the kinetic energy is largely insensitive to the level of diffusion. This implies that the velocity is not strongly dependent on the level of diffusion and that the Reynolds number should scale with the viscosity as $\mathrm{Re}$ ∼ ν−1. Given our definition of ${\mathrm{Ra}}_{F}$ and $\Pr $, namely that ${\mathrm{Ra}}_{F}$ ∼ ν−1κ−2 and $\Pr $ ∼ νκ−1, the Reynolds number scaling becomes $\mathrm{Re}\,\sim \,{\mathrm{Ra}}_{F}^{1/3}{\Pr }^{-2/3}$. In Figure 1(b), we plot the Reynolds number versus ${\mathrm{Ra}}_{F}$ $\Pr $−2, where the different symbols are the same as in panel (a). Each data point is time averaged. A least-squares fit to the data in the high-${\mathrm{Ra}}_{F}$ region yields a scaling law of

Equation (15)

These results indicate that there are two distinct parameter regimes: one in which diffusion is an important factor in the global force balance (the low-${\mathrm{Ra}}_{F}$ region), and one where diffusion no longer plays an appreciable role in the interior, bulk global force balance (the high-${\mathrm{Ra}}_{F}$ region).

4.2. Spectral Distribution

We find that the kinetic energy may reach a ${\mathrm{Ra}}_{F}$-independent regime, but the flow's morphology is still affected by the level of diffusion. This can be seen in the relative spectral distribution of velocity between the high- and low-${\mathrm{Ra}}_{F}$ systems. Figure 2 shows the velocity power spectra for all of the runs with Prandtl numbers of $\Pr $ = 0.25, $\Pr $ = 1.0, and $\Pr $ = 4.0. Each spectrum has been normalized such that it has unit integrated power. The rows correspond to the different Prandtl numbers. The first column shows the spectra taken at the lower convection zone or r/ro ≈ 0.775. The second column shows the spectra near the upper boundary, or r/ro ≈ 0.985, in the thermal boundary layer. In each panel, all of the spectra with the given Prandtl number are plotted. Each spectrum is colored by ${\mathrm{Ra}}_{F}$ $\Pr $, with high values taking on red tones and low values displaying blue tones. Each spectrum is a time average over several tens of overturnings.

Figure 2.

Figure 2. Time-averaged velocity power spectra and shell slices for cases with $\Pr $ = 0.25, $\Pr $ = 1.0, and $\Pr $ = 4.0. The power spectra have been normalized such that each curve has unit integrated power. Each row represents a different value of $\Pr $. The first two columns correspond to power spectra at a single depth within the convective shell. The dashed black line in the center column has a slope of –5/3 for reference. Within each panel, spectra for all of the cases at that depth and $\Pr $ are displayed. Each curve is colored by ${\mathrm{Ra}}_{F}$ $\Pr $ with low ${\mathrm{Ra}}_{F}$ $\Pr $ in blue tones and high ${\mathrm{Ra}}_{F}$ cases in red tones. As ${\mathrm{Ra}}_{F}$ is increased, the power at low -values increases initially. At high ${\mathrm{Ra}}_{F}$, it decreases as high-wavenumber power is generated at the expense of low-wavenumber power. This trend occurs at all of the Prandtl numbers studied. The last column shows shell slices of the radial velocity taken near the outer boundary at r/ro ≈ 0.985. The larger $\Pr $ run shows wider downflow lanes compared to the small $\Pr $ run.

Standard image High-resolution image

Across all of the cases, as ${\mathrm{Ra}}_{F}$ increases from low values, the point-wise velocity power increases at nearly all spherical harmonic degrees. The -value associated with the peak for each spectra increases with increasing ${\mathrm{Ra}}_{F}$. This indicates that smaller-scale structures become more apparent with increasing ${\mathrm{Ra}}_{F}$. At sufficiently high Rayleigh number, power in the high- portion continues to increase, but the low- portion starts to decrease. This suggests a break down of large-scale coherent structures. The highest Rayleigh number runs have lower power in the large scales compared to the low-${\mathrm{Ra}}_{F}$ counterparts. The low-wavenumber region occurs in the approximate range of  ≲ 10 for most of the simulations in this study. This trend occurs for all values of $\Pr $ that were studied. If viscosity played a significant role in the asymptotic regime, one might expect there to be some variation in the spectra when the Prandtl number is varied. We do not observe large variations between the spectra for the Prandtl numbers within our range indicating that viscosity only plays a minor role in the asymptotic regime.

The third column in Figure 2 plots shell slices of the radial velocity taken near the upper boundary, the same location in radius as the second column. Each slice is a single snapshot in time. The color scale is the same for all three shell slices with red tones indicating positive outward flows and blue tones indicating negative inward flows. The lower Prandtl number run (the top panel) displays larger velocities and more small-scale structures compared to the larger Prandtl cases. The Reynolds numbers of all three slices cover a large range. The top row $\Pr $ = 0.25 run has $\mathrm{Re}$ = 261.2, the middle slice with $\Pr $ = 1.0 has $\mathrm{Re}$ = 56.3, and the bottom slice with $\Pr $ = 4.0 has $\mathrm{Re}$ = 11.7. The large range in Reynolds numbers indicates that the inertial subrange for each simulation is different.

At a sufficiently high Rayleigh number, the integrated KE becomes independent of the level of diffusion (both thermal and momentum diffusion). As ν is decreased and ${\mathrm{Ra}}_{F}$ is increased, the flow becomes more turbulent with smaller-scale structures. To leading-order, once the high-${\mathrm{Ra}}_{F}$ regime is reached, the total integrated dimensional KE is constant as the diffusion is further reduced. This fact is largely independent of $\Pr $ for the Prandtl numbers that were within our range, however, a weak $\Pr $ dependence remains because the inertial subrange is extended as the viscosity is reduced individually. These trends are similar to those found in the simulations of Featherstone & Hindman (2016b), which used a Prandtl number of unity.

4.3. Energy Transport

The energy transport across the layer can be characterized by four radial energy fluxes: the enthalpy flux Fe, the kinetic energy flux FKE, the conductive flux Fc, and the viscous flux Fν, which we define as

Equation (16)

Equation (17)

Equation (18)

Equation (19)

respectively. Note that the conductive flux Fc is associated with the diffusion of entropy perturbations, and it should not be confused with radiative diffusion arising from the reference state temperature gradient; that effect is represented by Q in our models. Averages are taken of these fluxes over several diffusion times, indicated using brackets. We consider the contribution of conduction by looking at the fractional convective flux ${f}_{\mathrm{conv}}$ defined as

Equation (20)

Upon rearranging this quantity, we can write

Equation (21)

Equation (22)

This resembles the traditional Nusselt number, but differs in two important ways. First, the traditional Nusselt number does not include the viscous flux or the kinetic energy flux, both of which we include. Second, the conductive flux that appears in our definition is the established conductive flux, not the conductive flux in the absence of convection. Values of $1/(1-{f}_{\mathrm{conv}})$ that are of order unity translate to a lack of convective heat transport. Large values indicate that convection plays a dominant role over thermal conduction in transporting energy through the shell. Figure 3(a) plots $1/(1-{f}_{\mathrm{conv}})$ as a function of ${\mathrm{Ra}}_{F}$ $\Pr $. Viscosity is not expected to play a large role in the heat transport across the shell, which is why we plot against ${\mathrm{Ra}}_{F}$ $\Pr $ and not simply ${\mathrm{Ra}}_{F}$. The plot shows a clear trend that can be fit using least-squares to obtain

Equation (23)

with a scaling exponent that is approximately 2/7 (∼0.286). Other studies have found that the Nusselt number scales with the Rayleigh number to the 2/9 power (∼0.222; e.g., Gastine et al. 2015), when a flux based Rayleigh number is used. Our results have a steeper exponent because we include the viscous flux and the kinetic energy flux, which act to increase $1/(1-{f}_{\mathrm{conv}})$ at any given Rayleigh number.

Figure 3.

Figure 3. Energy transport and thermal boundary-layer scaling with ${\mathrm{Ra}}_{F}$ $\Pr $. (a) Fractional convective flux $1/(1-{f}_{\mathrm{conv}})$ vs. ${\mathrm{Ra}}_{F}$ $\Pr $. The dashed line is ${({\mathrm{Ra}}_{F}\Pr )}^{0.275}$, very close to a 2/7 scaling. (b) Thermal boundary-layer width plotted vs. ${\mathrm{Ra}}_{F}$ $\Pr $. The dashed line is ${({\mathrm{Ra}}_{F}\Pr )}^{-0.164}$, very close to −1/6.

Standard image High-resolution image

4.4. Boundary-layer Thickness

We can further characterize the role of conduction by determining how ${\mathrm{Ra}}_{F}$ and $\Pr $ control the thermal boundary-layer thickness. We define the thermal boundary-layer thickness in terms of the time-averaged mean entropy, such that

Equation (24)

where the brackets indicate a time-average as before and sup indicates the supremum. Figure 4 shows three examples of the time-averaged mean entropy profiles and the associated boundary-layer location as calculated using Equation (24). We plot the variation of δBL with ${\mathrm{Ra}}_{F}$ $\Pr $ for different Prandtl numbers in Figure 3(b). We plot against ${\mathrm{Ra}}_{F}$ $\Pr $ because we do not expect the thermal boundary to depend on the viscosity and the quantity ${\mathrm{Ra}}_{F}$ $\Pr $ is independent of the viscosity and scales as ${\mathrm{Ra}}_{F}$ $\Pr $ ∼ κ−3. A least-squares fit gives the scaling law

Equation (25)

The scaling exponent is very close to −1/6 (∼ −0.166), indicating that the boundary-layer width scales purely as the thermal diffusivity, i.e., ${\delta }_{\mathrm{BL}}\propto \sqrt{\kappa }$. There is no strong dependence on viscosity within the range of $\Pr $ studied here.

Figure 4.

Figure 4. Time-averaged mean entropy profiles for three cases in the high-${\mathrm{Ra}}_{F}$ regime. The vertical dashed lines indicate the location of the boundary layer as calculated using Equation (24).

Standard image High-resolution image

We note that the highest Rayleigh number run had a boundary-layer width of about 3% of the shell depth (∼5 Mm). Our thermal boundary layer lacks the radiative processes at work in a star like the Sun, but its physical extent is confined to a similarly small region of the convective domain.

Figure 1 showed that a free-fall regime that is independent of both viscosity and thermal diffusion could be obtained for different Prandtl numbers. We find that in order to have a bulk kinetic energy that is independent of both viscosity and thermal diffusion, the thermal boundary-layer thickness must depend on both the Rayleigh number and the Prandtl number with the scaling law given in Equation (25).

For a fixed Prandtl number, this scaling relation suggests that the boundary-layer thickness will continually decrease as the Rayleigh number is increased. In global simulations that employ explicit diffusivities, the Prandtl number is therefore critical in maintaining a boundary layer that is confined to a small region of the convective domain without becoming vanishingly small as the Rayleigh number is increased.

Certainly, simulations with a conductive boundary layer cannot match every feature of the Sun's thermal boundary layer, for the simple reason that the Sun's boundary layer is regulated by radiative transfer instead of thermal conduction. In a global simulation without radiative transfer, one hopes that the microphysics of the cooling layer can be ignored and only the gross properties of the boundary layer (thickness and entropy contrast) are important. When the boundary layer is conductive, the entropy contrast and thickness are inherently linked. Thus, the best one could hope to do is achieve a physically realistic thickness; this requires that the product $H{({\mathrm{Ra}}_{F}\Pr )}^{-1/6}$ take on the desired thickness, i.e., as the Rayleigh number is increased, the Prandtl number must be decreased. This realistic thickness will probably not, however, coincide with a convective power spectrum that possesses a realistic inertial range.

5. Perspectives and Conclusions

The results presented here have interesting consequences for several aspects of stellar/solar convection zone dynamics. Many of these results will depend on rotation and magnetism, both of which were omitted in this study.

We find that simulations with a lower $\Pr $ number have faster flows and a broader range of scales compared to those of high $\Pr $ models with the same ${\mathrm{Ra}}_{F}$. The higher $\Pr $ models have more viscous dissipation, resulting in slower flows (equivalent to a lower Reynolds number). This is consistent with the results found in O'Mara et al. (2016).

The higher ${\mathrm{Ra}}_{F}$ simulations obtain a free-fall state where diffusion no longer plays an important role in the interior bulk global force balance. In this state the kinetic energy becomes independent of both viscosity and thermal diffusion. Similar results were found in Featherstone & Hindman (2016b), although their study was restricted to $\Pr $ = 1.

The boundary-layer thickness scaling suggests that most simulations may be achieving the correct driving scale with a modest Rayleigh number of about 106 and a Prandtl number of unity. The obtained scaling also suggests that in order to maintain a boundary layer whose physical extent is confined to a small region of the convective domain in a global simulation without radiative transfer, the Prandtl number should be decreased as the Rayleigh number is increased.

We did not find that the Prandtl number substantially alters the earlier observed behavior in the spectral distribution of the velocity for $\Pr $ unity simulations (e.g., Featherstone & Hindman 2016b). As the ${\mathrm{Ra}}_{F}$ is increased, the convection develops smaller-scale structures and a corresponding increase in the high-wavenumber power. The high-wavenumber power increases, but the low-wavenumber power decreases indicating the break down of coherent large-scale structures. The spectral range of  ≤ 10 appears to be the most sensitive region of the power spectrum. This occurs for all of the Prandtl numbers studied here.

Our results indicate that care needs to be taken when interpreting convection simulations and comparing the results to observations of real solar/stellar systems. Simulations that do not access a high enough ${\mathrm{Ra}}_{F}$ may overestimate the low-wavenumber power that is accessible to helioseismology. Stellar convection simulations must run with parameters that place it in the high-Rayleigh-number regime in order to correctly capture the integrated kinetic energy and the large-scale motions of the flow. We used rather modest levels of diffusion (κ ≤ 4 × 1012  cm2 s−1) to put our simulations in the high-Rayleigh-number regimes.

Most importantly, our simulations did not include rotation or magnetism. Featherstone & Hindman (2016a) looked at the effects of rotation, but restricted their study to hydrodynamical, $\Pr $ = 1 simulations. It will be important to examine how our Prandtl number findings are modified by rotation and magnetism before we can fully trust comparisons of power spectra between observations and simulations.

This work was supported by NASA grants NNX09AB04G, NNX14AC05G, NNX11AJ36G, NNX17AM01G, and NNX14AG05G. M.A.C. was supported by National Science Foundation award number EAR-1620649. N.A.F. and the development of Rayleigh were further supported by the Computational Infrastructure for Geodynamics (CIG), which is supported by the National Science Foundation award numbers NSF-0949446 and NSF-1550901. Resources supporting this work were provided by the NASA High-End Computing (HEC) Program through the NASA Advanced Supercomputing (NAS) Division at the Ames Research Center.

Footnotes

  • This is the same cutoff used in Featherstone & Hindman (2016b). Their reported ${\mathrm{Ra}}_{F}$ are too high by a factor of π; we correct for that here.

Please wait… references are loading.
10.3847/1538-4357/aaaeb5