OCTOBER 7, 2017, REVISION: 1.258 Preprint typeset using LATEX style emulateapj v. 08/13/06

STELLAR MIXING LENGTH THEORY WITH ENTROPY RAIN AXEL BRANDENBURG Laboratory for Atmospheric and Space Physics, University of Colorado, Boulder, CO 80303, USA JILA and Department of Astrophysical and Planetary Sciences, University of Colorado, Boulder, CO 80303, USA Nordita, KTH Royal Institute of Technology and Stockholm University, Roslagstullsbacken 23, SE-10691 Stockholm, Sweden Department of Astronomy, AlbaNova University Center, Stockholm University, SE-10691 Stockholm, Sweden October 7, 2017, Revision: 1.258

ABSTRACT The effects of a non-gradient flux term originating from the motion of convective elements with entropy perturbations of either sign are investigated and incorporated into a modified version of stellar mixing length theory (MLT). Such a term, first studied by Deardorff in the meteorological context, might represent the effects of cold intense downdrafts caused by the rapid cooling in the granulation layer at the top of the convection zone of late-type stars. These intense downdrafts were first seen in the strongly stratified simulations of Stein & Nordlund in the late 1980s. These downdrafts transport heat nonlocally, a phenomenon referred to as entropy rain. Moreover, the Deardorff term can cause upward enthalpy transport even in a weakly Schwarzschild-stably stratified layer. In that case, no giant cell convection would be excited. This is interesting in view of recent observations, which could be explained if the dominant flow structures were of small scale even at larger depths. To study this possibility, three distinct flow structures are examined: one in which convective structures have similar size and mutual separation at all depths, one in which the separation increases with depth, but their size is still unchanged, and one in which both size and separation increase with depth, which is the standard flow structure. It is concluded that the third possibility with fewer and thicker downdrafts in deeper layers remains the most plausible, but it may be unable to explain the suspected absence of large-scale flows with speeds and scales expected from MLT. Subject headings: convection — turbulence — Sun: granulation

1. INTRODUCTION length scales corresponding to spherical harmonic degrees of Late-type stars such as our Sun have outer convection 30 or below (corresponding to scales of about 140 Mm and zones. The observed solar granulation is a surface manifes- larger), the rms velocity actually increases with depth. This tation of their existence. Solar granulation and the first few in itself is remarkable, because velocity perturbations from megameters (Mm) of the convection zone have been modeled deeper layers were expected to be transmitted to the surface successfully using mixing length theory (MLT) and numerical almost unimpededly (Stix 1981), unlike perturbations in the simulations with realistic physics included (Stein & Nordlund heat flux, which are efficiently being screened by the convec- 1989, 1998; Vogler¨ et al. 2005; Gudiksen et al. 2011; Freytag tion (Spruit 1977). Thus, an increase in horizontal velocity et al. 2012). As a function of depth, the simulations repro- power with depth is unexpected. However, van Ballegooijen duce some essential features predicted by MLT, in particular (1986) pointed out that for the scale and depth of giant cells, the depth dependence of the turbulent root mean square (rms) the screening might be large enough to allow giant cell con- vection of 100 m s−1 to result in only 10ms−1 at the surface, velocity, u (F /ρ)1/3, where F is the enthalpy rms enth enth which would be compatible with observations (Hathaway et flux and ρ is the≈ density. Simulations also seem to confirm an al. 2013). The presence of the near-surface shear layer of the important hypothesis of MLT regarding the gradual increase Sun (Schou et al. 1998), might enhance the screening further. of the typical convective time and length scales with depth. Given that the deeply focused kernels used by Greer et al. Given such agreements, there was never any reason to ques- (2015) can still have 5–10% sensitivity to Doppler shifts aris- tion our basic understanding of convection. ing from flows in the upper layers, such near-surface flows In recent years, local helioseismology has allowed us to de- could still leave an imprint on the signal. This effect would termine subsurface flow velocities (Duvall et al. 1997; Gizon be exaggerated further if the near-surface flows were stronger & Birch 2005). Helioseismic observations by Hanasoge et compared to those in deeper layers, as was theoretically ex- al. (2010, 2012) using the deep-focusing time-distance tech- pected. On the other hand, if convection in the surface layers nique have not, however, detected large-scale convection ve- is of smaller scale, the signal from those layers will to a large locities at the expected levels (Miesch et al. 2012); see also extent be averaged out despite its larger amplitude. It would Gizon & Birch (2012) for a comparison with both global sim- obviously be important to examine this more thoroughly by ulations and radiation hydrodynamics simulations with real- applying the kernels of Greer et al. (2015) to realistic simu- istic radiation and ionization physics included. Greer et al. lations. Unfortunately, this has not yet been attempted. Con- (2015) have suggested that the approach of Hanasoge et al. versely, the deep-focusing technique of Hanasoge et al. (2012) (2012) may remove too much signal over the time span of the could be extended to allow for imaging of the deeper flow measurements. Using instead ring diagram analysis (Gizon structures and thereby a direct comparison of individual tur- & Birch 2005) with appropriately assembled averaging ker- bulent eddies with those detected by Greer et al. (2015). nels to focus on the deeper layers, Greer et al. (2015) instead It should be mentioned that small flow speeds of giant cell find values of the turbulent rms velocities that are consistent convection have been found by correlation tracking of super- with conventional wisdom. Moreover, they find that at large granule proper motions (Hathaway et al. 2013). The typical 2 velocities are of the order of 20ms−1 at spherical degrees term, which thus significantly reduces the rms velocity of the of around 10. Those speeds are still an order of magnitude resolved flow field. They speculate that the flow–temperature above the helioseismic upper limits of Hanasoge et al. (2012), correlation entering the enthalpy flux may be larger in the Sun, but Hanasoge & Sreenivasan (2014) argue that these surface possibly being caused by a magnetic field that may maintain measurements translate to lower flow speeds in the deeper and flow correlations and boost the transport of convective flux by denser layers when considering mass conservation. This ar- smaller scales. This has been partially confirmed by Hotta et gument may be too naive and would obviously be in conflict al. (2015), who have discussed the possible role of small-scale with the results of Greer et al. (2015), which show an increase magnetic fields in suppressing the formation of large-scale with depth at low spherical harmonic degrees. On the other flows. Furthermore, Featherstone & Hindman (2016) have hand, Bogart et al. (2015) have argued that the flows reported found that with increasing Rayleigh numbers, there is more by Hathaway et al. (2013) may be non-convective in nature kinetic energy at small scales and less at large scales such that and, in fact, magnetically driven and perhaps related to the the total kinetic energy is unchanged by this rearrangement of torsional oscillations. energy. Quantitatively, the horizontal flow speed at different scales Given the importance of this subject, it is worthwhile re- is characterized by the kinetic energy spectrum, E(k), where viewing possible shortcomings in our theoretical understand- k is the wavenumber or inverse length scale. Stein et al. ing and numerical modeling of stellar convection. Both global (2007) have shown that the spectra of surface Dopplergrams and local surface simulations of solar-type convection would and correlation tracking collapse with those of Stein & Nord- become numerically unstable with just the physical values of lund onto a single graph such that the spectral velocity and radiative diffusivity. Therefore, the radiative [kE(k)]1/2 is proportional to k, i.e., E(k) k; see also flux (in the optically thick layers) is modified in one of two Nordlund et al. (2009). This is a remarkable∝ agreement be- possible ways. (i) The contribution from temperature fluctua- tween simulations and observations. While the origin of such tions is greatly enhanced, i.e., a spectrum is theoretically not understood, it should be em- F = K∇T K∇T K ∇(T T ), (1) phasized that these statements concern the horizontal veloc- rad − → − − SGS − ity at the surface and do not address the controversy regard- ing the spectrum at larger depths, where the flow speeds at where T and T are the actual and horizontally averaged tem- large length scales may still be either larger (e.g. Greer et al. peratures, respectively, K is the radiative conductivity, and 2015) or smaller (e.g. Featherstone & Hindman 2016) than KSGS is a subgrid scale (SGS) conductivity. The latter is those at the surface. If most of the kinetic energy were to enhanced by many orders of magnitude relative to K. Fur- reside on small scales throughout the convection zone even thermore, numerical diffusion operators often do not translate at a depth of several tens of megameters and beneath, E(k) in any obvious way to the physical operators. (ii) Alterna- is expected to decrease toward smaller wavenumbers k either tively, in direct numerical simulations (DNS), one uses physi- like white noise k2 or maybe even with a Batchelor spec- cal viscosity and diffusivity operators, i.e., the replacement in trum k4 (Davidson∝ 2004). However, even in simulations Equation (1) is not invoked, but the coefficients are enhanced ∝ (K Kenh) and exceed the physical ones by many orders of of forced isothermally stratified turbulence, in which there is → no larger-scale driving from thermal buoyancy, there is a shal- magnitude. Both approaches are problematic. In many global DNS with enhanced coefficients (e.g., lower spectrum, E(k) k3/2 (Losada et al. 2013). By con- ∝ Kapyl¨ a¨ et al. 2013), the lower boundary is closed, so at the trast, if one imagines the flow to be anelastic so that the mass bottom of the domain all the energy is carried by radiation u ∇ ψ flux ρ = can be written as a vectorial stream func- alone. By choosing an enhanced radiative diffusivity, the tion ψ, and if ×ψ is given by white noise, one should expect a 4 radiative flux is increased by a corresponding amount, and k Batchelor spectrum. This is qualitatively similar to the re- therefore also the total flux. The luminosity in those simula- sults of granule tracking, which reveal an intermediate scaling tions can exceed the solar value by several orders of magni- proportional to k3 (Rieutord et al. 2008; Roudier et al. 2012). 4 tude. This has a series of consequences. Most notably, the While the steeper k spectrum for k

2008), or could the stratification still be stable to the excitation given by Equation (2). However, in limited density and tem- of large-scale flows? perature ranges, certain values of a and b can tentatively be We postpone addressing the two aforementioned issues specified, e.g., a = 0.5 and b = 7...18. Clearly, for all these connected with the qualitative reasoning of Spruit (1997) to values the constraint (4) is violated, so the hydrostatic strati- the end of the paper (Section 6), and begin by highlighting the fication is no longer polytropic, but solutions can still be con- less commonly known fact that in the non-convecting refer- structed numerically (Barekat & Brandenburg 2014) and they ence model, only the layer in the top 1 Mm is convectively un- demonstrate, not surprisingly, that the stratification is highly stable (Section 2). We then explain the nature of the Deardorff unstable. What is more surprising is the fact that even with a flux (Section 3), present a correspondingly modified mixing combined opacity law of the form length model (Section 4) and give some illustrative numerical −1 −1 −1 solutions (Section 5). We discuss alternative explanations for κ = κKr + κH− , (6) the lack of giant cell convection in Section 6 and present our where κKr and κH− are given by Equation (2) with suitable conclusions in Section 7. exponents a and b, the solutions of the non-convective ref- 2. A HIGHLY UNSTABLE SURFACE LAYER erence state is unstable only over a depth of approximately 1 Mm. We return to this at the end of Section 5, where we The main argument for the existence of a highly unstable present numerical solutions. surface layer comes from the consideration of the associated Of course, as stated earlier, the non-convective reference non-convective reference solution, where the flux is forced state is only of academic interest. Already with standard to be transported by radiation only, i.e. F = Frad. This is MLT (Biermann 1932; Vitense 1953), which allows for a non- something that is not normally considered in stellar physics, vanishing enthalpy flux, one finds a vastly extended convec- because we know that such a solution would be unstable to tion zone with a depth of the order of 100 Mm (Biermann convection and would therefore never be realized. However, 1938). However, the question now is how this can be affected as mentioned in the introduction, this solution is of certain by the presence of the Deardorff flux. This will be the subject academic interest. We postpone the discussion of an explicit of the rest of this paper. If the Deardorff flux were to become numerical solution to Section 5 and present in this section the dominant and the stratification subadiabatic, it would become basic argument only at a qualitative level, making reference locally stable to the onset of convection. One might further to earlier numerical calculations by Barekat & Brandenburg speculate that the typical scale would no longer be controlled (2014) for a simple model. They considered an opacity law of by the local pressure scale height, but it might be imprinted the form from the downdraft pattern just beneath the surface and be a b κ = κ0(ρ/ρ0) (T/T0) , (2) therefore comparable to the granulation scale or at least the supergranulation scale (Cossette & Rast 2016). This can have where a and b are adjustable parameters, ρ0 and T0 are ref- erence values for density and temperature, respectively, and other important consequences that will also be addressed in this paper. It should be pointed out, however, that the sur- κ0 gives the overall magnitude of the opacity. The essential point to note here is that the exponents a and b determine the face simulations have so far not produced evidence for sub- gradient of specific entropy in a purely non-convecting refer- adiabatic stratification. On the other hand, the models pre- ence model. In thermodynamic equilibrium, the radiative flux sented below predict subadiabatic stratification only a certain must be constant, i.e., distance below the surface, depending on ill-known input pa- rameters. Frad = K dT/dz = const, (3) − THE DEARDORFF FLUX 3 3. where K = 16σSBT /(3κρ) is the radiative conductivity 3.1. Derivation with σSB being the Stefan–Boltzmann constant, and z is the vertical coordinate in a Cartesian coordinate system. For the In the meteorological context, counter-gradient heat flux simple opacity law (2), but with terms have been noticed for a long time (Ertel 1942; Priestley & Swinbank 1947; Deardorff 1966). They appear naturally b< 4+ a, (4) when calculating the enthalpy flux Fenth using the τ approxi- the optically thick regime is characterized by a constant tem- mation in its minimalistic form (e.g. Blackman & Field 2003). perature gradient and therefore K = const (see Appendix A). In this approach, one computes the time derivative of Fenth. n We have then a polytropic stratification with ρ T , where In the absence of ionization effects, Fenth can be written as ∝ n = (3 b)/(1 + a) (5) Fenth = ρuzcP T, (7) − is the polytropic index. It is larger than 1 when Equation (4) where the overbar denotes horizontal averaging. In standard is obeyed, i.e., when the pressure decreases− with height, as MLT, the enthalpy flux is usually referred to as the convec- is required for a physically meaningful solution. For a ra- tive flux and the kinetic energy flux vanishes because of the tio of specific heats of γ = 5/3, the value of n for marginal assumed perfect symmetry between up- and downflows. In Schwarzschild stability is ncrit = 1/(γ 1) = 3/2. In most of deeper layers especially, however, this is not justified (Stein the solar convection zone the dominant− opacity is the bound- et al. 2009), so this restriction will later be relaxed. free absorption owing to the absorption of light during ion- In the case of strongly stratified layers, it is convenient to ization of a bound electron, which is well described by the use pressure P and specific entropy S as thermodynamic vari- Kramers-type opacity law with a = 1 and b = 7/2, so ables, because we later want to ignore pressure fluctuations n = 3.25, which corresponds to a Schwarzschild-stable− solu- on the grounds that pressure disturbances are quickly equi- tion. Only near the surface, at temperatures typically below librated by sound waves. Furthermore, we will restrict our- 15, 000 K, the dominant opacity is the H− opacity. It can no selves to second order correlations in Equation (7) and thus longer be approximated by a simple power law of the type to fluctuations only in the correlation of specific entropy and 5 velocity. Up to some reference value, we have derivative of Fenth can be neglected. (Note, however, that this does not imply that the cooling term will be neglected.) S/c = ln T ln P, (8) P − ∇ad Since the relaxation term in Equation (15) is similar to the where = 1 1/γ, and γ = c /c is the ratio of specific cooling term in Equation (14), we can combine the two by ∇ad − P V heats at constant pressure and constant volume, respectively, introducing the reduced relaxation time τred defined through all constants for a perfect gas with a fixed degree of ionization. −1 −1 −1 (16) Ignoring pressure variations and linearizing δ ln T = δT/T , τred = τcool + τ . we can replace cP δT by TδS. In the following, we denote We can then solve for Fenth, which appears on the right-hand fluctuations by lower case characters, i.e., S = S + s, where sides (rhs) of Equations (14) and (15), and find F enth = F G+ s δS are used interchangeably. As argued above, other fluc- F D, where ≡ U 0 tuations are omitted. We also ignore mean flows ( = ), so F = 1 τ u2 ρ T ∇S, (17) there are only velocity fluctuations (U = u). Focusing thus G − 3 red rms on the dominant contribution proportional to u s, we have F = τ s2 g ρ T /c (18) z D − red P Fenth = ρ T uzs. (9) are the ordinary gradient and the new Deardorff fluxes, re- spectively, and anisotropies have been ignored for the benefit Next, we write the time derivative of Fenth as of simpler notation. Thus, we write u u 1 δ u2 , where i j ≈ 3 ij rms ∂Fenth/∂t = ρ T uzs˙ + u˙ zs , (10) urms is the rms velocity of the turbulence. Equations (17) and (18) are written in vectorial forms to highlight the direc- where dots denote partial time derivatives and changes of the tions of the fluxes: F G is counter-gradient and F D is coun- background state have been neglected. Using the governing 1 2 tergravity. In Equation (17), the term 3 τredurms χt is the equations for ui and s, we have (see Appendix B) turbulent thermal diffusivity. ≡ s˙ = u S s/τ ..., (11) In view of astrophysical applications, we replace the spe- − j∇j − cool cific entropy gradient by the commonly defined superadia- u˙ = g s/c + ..., (12) i − i P batic gradient, i.e., where is the th component of the gravitational acceleration gi i d(S/c )/dz =( )/H , (19) in Cartesian coordinates, namely g = (0, 0, g), the ellipses − P ∇ − ∇ad P − refer to terms that are nonlinear in the fluctuations, where = dln T /dln P is the double-logarithmic tempera- ∇ −1 2 2 ture gradient and H = (dln P /dz)−1 is the pressure scale τ = ιcγ kf with ι = ℓkf /(3 + ℓ k ) (13) P cool f height. Thus, we arrive at− is the inverse heating and cooling time owing to radiation 3 1 2 Fenth = ρcP T (τredurms/HP )( ad + D) , (20) (Unno & Spiegel 1966; Edwards 1990), cγ = 16σSBT /ρcP 3 ∇ − ∇ ∇ is the photon diffusion speed (Barekat & Brandenburg 2014), where is a new contribution to standard MLT, which re- ∇D ℓ = 1/κρ is the photon mean-free path, and kf is the typi- sults from FD. Using the terms on the rhs of Equation (18), cal wavenumber of the fluctuations, which may be associated we can write it explicitly as with the inverse mixing length used routinely in MLT. We 2 2 −2 mention at this point that the nonlocal nature of the entropy D = (3/γ)(s /cP ) Ma , (21) ∇ rain must lead to yet another contribution in Equation (11). where Ma = urms/cs is the Mach number of the turbulence We expect this to be the result of the nonlinear term (indi- 2 and cs is the sound speed with cs = γgHP . Note that this cated by the ellipses), of which a part later gives rise to the Mach number dependence arises in order to cancel the cor- D term. This will be motivated further at the end of this sec- 2 ∇ responding urms factor in the definition (20). The Deardorff tion and in Section 3.3, where it will be included in the final term is a contribution to the flux that is always directed out- expression for . ∇D ward and results from the transport of fluid elements with en- Using Equations (11) and (12) in Equation (10), we have tropy fluctuations of either sign, as is illustrated in Figure 1. enth enth ∂Fi 2 Fi =ρ T uiuj j S gis /cP + i, (14) 3.2. Physical interpretation ∂t − ∇ − − τcool T   We recall that in a stably (unstably) stratified layer, shown 2 where the s term is the Deardorff flux and the i refer to in Figure 1(a) and (b), and under the assumption of pressure triple correlations that will be approximated by theT quadratic equilibrium, the entropy perturbation of a blob with respect correlation Fenth as to its current surroundings decreases (increases) after a small ascent. By contrast, in the marginally or nearly marginally = F enth/τ, (15) Ti − i stratified cases, the perturbation remains constant if one ig- where τ is a relaxation time due to turbulence (Blackman & nores mixing with the surroundings; see panels (c) and (d). Field 2003), which will later be identified with the turnover Therefore, a positive entropy perturbation (s > 0) always time. This procedure, in which correlations with pressure corresponds to a positive temperature perturbation and a neg- fluctuations are also neglected, is called the minimal τ ap- ative density perturbation. Thus, the fluid parcel is buoyant proximation. Important aspects of the closure assumption and moves upward (uz > 0), so uzs > 0, giving a positive (15) have been verified numerically for passive scalar trans- contribution to Fenth. Likewise, for s< 0, the parcel is cooler port (Brandenburg et al. 2004). Among other things, they and heavier and sinks (uz < 0), and again, uzs> 0, giving a found that the time derivative on the left-hand side of Equa- positive contribution to Fenth. tion (14) restores causality by turning the otherwise parabolic In laboratory convection, strong positive entropy perturba- heat equation into a hyperbolic wave equation. Here, how- tions can be driven at the lower boundary, and strong nega- ever, we are interested in slow variations such that the time tive ones at the top boundary. In the Sun, however, only the 6

called coherent structures (De Roode et al. 2004) and those may not simply be the vortex-like structures envisaged above. It is clear, however, that the structures seen in the Earth’s at- mosphere are driven by the on the ground, and sometimes also by the upper inversion layer (De Roode et al. 2004). Pleim (2007) discusses the Deardorff flux in con- nection with other parameterizations of nonlocal fluxes such as the transilient matrix approach (Stull 1984, 1993). The close connection between counter-gradient fluxes and nonlo- cal transport has been elaborated upon by van Dop & Verver (2001) and Buske et al. (2007). Thus, while the approach pre- sented in Section 3.1 may capture the physical phenomenon of the Deardorff flux qualitatively correctly, it is quite possi- ble that the nature of the underlying coherent structures may require additional refinements.

3.3. Depth dependence of the Deardorff term To estimate the resulting depth dependence of the Deardorff term in Equation (21), we must know how the s2 associated with the Deardorff term varies with depth. If the entropy rain was purely of the form of Hill vortex-like structures, as dis- cussed above, their filling factor fs would decrease with in- creasing depth and density like −ζ fs ρ , (22) FIG. 1.— Sketch illustrating overshoot in a stably stratified layer (a), growth ∝ of perturbations in an unstably stratified layer (b), buoyant rise of a blob with where ζ = 0.8 has been found for spherical vortex struc- positive entropy perturbation (c), and descent of a blob with negative entropy tures descending in an isothermally stratified layer; see Ap- perturbation and hence negative buoyancy (d). In cases (b)–(d), the turbulent flux is upward (increasing z). Case (d) is relevant to entropy rain. pendix C. However, Equation (23) is not contingent on Hill vortices, but is a consequence of downdrafts along a density strong radiative cooling in the photosphere provides a signifi- gradient. For purely spherical compression one would expect cant source of entropy perturbations. This became clear with ζ = 2/3, while for horizontal compression one has ζ = 1. the emergence of the realistic solar convection simulations of If we neglect non-ideal (radiative or viscous) effects, as Stein & Nordlund (1989). These simulations showed for the well as entrainment between up- and downflows, the differ- ence ∆S = S↑ S↓ between the entropy of the slowly ris- first time the strong vertical asymmetry of solar convection − resulting from the physics of ionization and strong cooling ing surroundings, S↑, and that in the downward propagating via the H− opacity. This led to the notion of entropy rain vortex, S↓, would remain constant and equal to the entropy and Spruit’s description of solar convection as a nonlocal phe- deficit ∆S0 suffered by overturning motions at the surface, nomenon that is driven solely by surface cooling. Our consid- which is the only location where radiative losses are signif- erations suggest that the resulting (one-dimensional) mean- icant. On sufficiently short length scales, however, radiative field energy flux is parameterized not just in terms of the local heating from the surroundings would erode this entropy dif- superadiabatic gradient. ference and lead to a decrease of the effective ∆S. We model In analogy with laboratory convection (Heslot et al. 1987; this by considering ζ an adjustable parameter that is increased Castaing et al. 1989), Spruit refers to the downdrafts as relative to the value 0.8 that we expect in the absence of ra- threads. Interestingly, the laboratory experiments show that diation, i.e., fs decreases more strongly. The resulting frac- these threads persist even at large Rayleigh numbers. For such tional entropy difference, fs ∆S0, therefore decreases faster threads to persist, one could imagine them to be stabilized with depth than for ζ = 0.8. by their intrinsic vorticity, akin to a Hill vortex (Hill 1894). As will become clear from the considerations below, if there These are vortex rings that can stay concentrated over large is no entrainment from the upflows into the downdrafts and all distances. Vortex rings have also been reported by Stein et of the downflows were confined to the small surface area of al. (2009), but many of those are associated with mushroom the vortex structures, the resulting downward-directed kinetic shapes, indicating that they widen and soon break up and mix energy flux would increase with depth and eventually exceed with their surroundings. Numerical studies (see Appendix C) the enthalpy flux. This would be unphysical, because con- suggest that Hill vortices also survive in a highly stratified vection should lead to outward energy transport. However, if isothermal layer, and that their persistence increases signifi- there is entrainment with associated mixing, the actual filling cantly with resolution and decreasing viscosity. On the other factor (fractional area) of downflows, which we denote by f, hand, the presence of background turbulence provides an ef- would be larger and the mean entropy difference ∆S would fective turbulent viscosity, which contributes to a premature be diluted correspondingly such that break up of small-scale vortices. However, those studies were − f ∆S = f ∆S ρ ζ ∆S . (23) done with an isothermal equation of state and thus ignore the s 0 ∝ 0 effects of continued driving by a persistent negative entropy We recall that non-ideal effects can be captured by choosing perturbation between the vortex and its surroundings. ζ > 0.8 in the prescription (23). The physics of entrainment Studies of the Deardorff flux in the meteorological context has been modeled in the stellar context by Rieutord & Zahn suggest that the flux-carrying plumes are associated with so- (1995) and the extent of entrainment has been quantified in the 7 realistic surface simulations by Trampedach & Stein (2011), TABLE 1 who determined the entrainment length scale as the typical φkin, −U ↓/urms, AND U ↑/urms FOR SELECTED VALUES OF f . scale height of the mass flux in the up- and downflows sepa- rately. They call this scale the mass mixing length and find its f 1/2 1/3 0.14 0.015 0.0006 value to be comparable to H . φkin 0 0.35 1 4 20 P − To estimate the depth dependence of various horizontal av- U ↓/urms 1 1.4 2.5 8 40 U ↑/urms 1 0.7 0.4 0.12 0.025 erages, and to compute enthalpy and kinetic energy fluxes, with φenth = kf HP /(aMLT ad). This yields φenth 20, we now compute various averages in the two-stream approx- which is rather large. By contrast,∇ Brandenburg et al. (2005≈ ) imation, in which all horizontal averages are the sum of the −3/2 determined a quantity k such that φ = ku 4. fraction 1 f of the value in upflows and the fraction f of u enth ≈ − We see that the presence of a kinetic energy flux just mod- the value in downflows. Thus, for the mean specific entropy ifies the usual expression for the convective flux, which then stratification we have becomes the sum of enthalpy and kinetic energy fluxes; see S = (1 f)S↑ + fS↓ = S↑ f ∆S, (24) Section 3.1. This is compatible with recent simulations of − − R. F. Stein (2016, private communication), in which the frac- where S↑ and S↓ are the mean specific entropies in up- and tional kinetic energy flux increaes toward the deeper parts > downflows, and ∆S = S↑ S↓ is their difference. We expect ( 40 Mm) of the domain. − ∼When f becomes small (< 0.14), φkin exceeds unity and that S↑ S↑ will be constant if there is no heating in the ≈ for f < 0.015, φkin exceeds the estimate φenth 4 found by upwellings, while S↓ (1 fs/f)S↑ +(fs/f)S↓ will be ≈ ≈ − Brandenburg et al. (2005), so the sum of enthalpy and kinetic dominated by contributions from S↑ due to entrainment. energy fluxes may become negative, which appears unphysi- To calculate s2, we must first compute the fluctuating quan- cal. In that case, the idea of reconciling the results of Hana- tity s = S S and then average the squared values in up- and soge et al. (2012) with such convection transporting the solar downflows,− which, using Equation (24), gives luminosity could be problematic, unless both up- and down- flows were to occur on sufficiently small scales. In the case 2 2 2 2 s = (1 f)(S↑ S) + f (S↓ S) = fˆ(∆S) , (25) − − − f = 0.14, the resulting U ↓ would still not be particularly fast and only 2.5 times larger than u ; for f = 0.015 we have where fˆ = (1 f)f has been introduced as a shorthand. rms Analogously to the− specific entropy, we find for the velocity U ↓ 8urms; see Table 1. Conversely, if the cold entropy blobs− ≈ were much smaller and were still to contribute signifi- U z = (1 f)U ↑ + fU ↓ = U ↑ f ∆U, (26) cantly to the total energy flux, i.e., if they were much faster, − − this might not be compatible with an upward directed total en- where U ↑ > 0 and U ↓ < 0 are the mean up- and downflow ergy transport. A possible alternative might be suggested by velocities with ∆U = U ↑ U ↓. Since the densities in up- and the compressible simulations of Cattaneo et al. (1991), where downflows are nearly the− same, especially in deeper layers, the downward-directed kinetic energy flux in the downdrafts we have U z 0 from mass conservation. With this, we find was found to balance the upward directed enthalpy flux in the ≈ downdrafts. This would imply that convection would trans- 2 2 2 2 ˆ 2 urms uz = (1 f) U ↑ + f U ↓ = f (∆U) , (27) port energy only in the upwellings. This result has however ≡ − not been confirmed in realistic surface simulations (Stein et 1/2 and thus U ↑/urms =[f/(1 f)] . Similarly, we calculate al. 1992). − Equation (23) shows how decreases with depth, but its 3 3 fs u3 = (1 f) U + f U = fˆ(1 2f) (∆U)3, (28) actual value and that of remains undetermined. In the z − ↑ ↓ − − ∇D following, we propose a quantitative prescription for D by which is negative for f < 1/2, and finally estimating its value within the top few hundred kilometers.∇ In those top layers, we expect D to reach its maximum value, uzs = (1 f) U ↑S↑ + f U ↓S↓ = fˆ∆U ∆S. (29) max ∇ − D , and it should be a certain fraction of ad. In Appendix∇ E we show that max ∇ − ∇. Thus, the kinetic energy flux, ρu2u /2, which reduces to D /( ad)max 3 z In deeper layers, where∇ the local∇ − value ∇ of ≈ has 3 ad ρ uz/2 in this two-stream approximation, is given by 2 ∇ − ∇2 become small, D should scale with s = fˆ(∆S) , and 3 ∇ Fkin = φkin ρu , (30) therefore, using Equation (23), it should be proportional to − rms 2 −2ζ −2 fs (z) ρ , in addition to an Ma factor; see Equa- 1/2 ∝ where φkin = (1/2 f)/fˆ is a positive prefactor (corre- tion (21). Thus, we have sponding to downward− kinetic energy flux) if f < 1/2. Stein = f 2 max, (32) et al. (2009) find f 1/3, nearly independently of depth, ∇D s ∇D ≈ where we have used for f the expression which yields φkin √2/4 0.35; see Table 1, where we s ≈ ≈ 1/2 ˜ list φ and U ↓/u = [(1 f)/f] for selected val- −ζ kin − rms − fs = fs0(ρ/ρ∗) (for ρ > ρ∗). (33) ues of f. The enthalpy flux is proportional to uzs and, using Equations (9) and (29) together with Equations (25) and (27), Here, fs0 is a prefactor determining the strength of the result- ing Deardorff flux and ρ∗ is the density at the point at which we find Fenth = ρ Turmssrms. In Appendix D we show that 2 ad is equal to its maximum value (which is just below Tsrms = urmskf HP /(aMLT ad), where aMLT = 1/8 is a ∇ − ∇ ˜ geometric factor in standard MLT.∇ 2 This leads to the photosphere). The new exponent ζ = ζ ∆ζ takes the scaling of Mach number with density into account,− i.e., 3 Fenth = φenth ρurms (31) Ma ρ −∆ζ , (34) ∝ 2 Not to be confused with the parameter αmix of Section 4.2 below. where ∆ζ (> 0) will be computed in Section 4.3. 8

3.4. Kinetic energy flux where Fconv = Fenth + Fkin is the flux that arises from con- The importance of the kinetic energy flux has been stressed vection. In MLT, it is customary to express these fluxes in for some time as a property of compressible stratified convec- terms of nablas. For a given double-logarithmic temperature gradient , the radiative flux is evidently tion (Hurlburt et al. 1984; Cattaneo et al. 1991). This flux ∇ is related to the asymmetry of up- and downflows (Stein & Kg F = , (40) Nordlund 1989) and is non-vanishing when f = 1/2; see rad c ∇ Section 3.3. It is neglected in standard MLT, as has6 been dis- P ∇ad where characterizes the actual temperature gradient. We cussed by Arnett et al. (2015). Just like the Deardorff flux, ∇ the kinetic energy flux is also a non-gradient flux, but it is al- can also define a hypothetical radiative temperature gradient that would result if all the energy were carried by radia- ways directed downward for f < 1/2 and must therefore be ∇rad overcome by the enthalpy flux so that energy can still be trans- tion, so we can write ported outwards. However, since the lowest order correlations Kg in are triple correlations, there is no approximation Ftot = rad, (41) Fkin τ c ∇ treatment analogous to that of the Deardorff flux. However, P ∇ad 3 which follows from Equation (38). We note in passing that the by comparing Fkin = φkin ρurms with the enthalpy flux in Equation (20), it is possible− to define a corresponding nabla ratio between Ftot and the radiative flux carried by the adia- 1 2 batic temperature gradient, Kg/c , is known in laboratory term via Fkin = ρcP T (τredu /HP ) kin. This yields P − 3 rms ∇ and theoretical studies of convection as the Nusselt number = 3 φ u H /τ c T, (35) (Hurlburt et al. 1984), whose local value is thus equal to ∇kin kin rms P red P which is obtained analogously to in Equation (21). The Nu = rad/ ad. (42) ∇D ∇ ∇ prefactor is here defined with a positive sign, so the total non- Finally, as explained in Section 3, we have from Equation (20) radiative flux caused by the turbulence is proportional to + . ∇− Fenth = F0 ( ad + D) , (43) ∇ad ∇D − ∇kin ∇ − ∇ ∇ Obviously, if kin is strictly proportional to ad, the 1 2 ∇ ∇ − ∇ with F0 = 3 ρcP Tτredurms/HP , and Fkin being essentially addition of the kin term does not modify standard MLT, pro- ∇ proportional to Fenth; cf. Equations (30) and (31). Flux equi- vided that kin depends just on the local value of urms, which librium then implies in turn is, again,∇ assumed to depend just on the local value of ad. This is different for D, which depends on the en- rad = + ǫ ( ad + D) , (44) tropy∇ − ∇ deficit produced by cooling∇ near the surface and in this ∇ ∇ ∇ − ∇ ∇ where ǫ = F c (1 φ /φ )/Kg. However, the way on the value of at the position where ρ = ρ . 0 P ad kin enth ad ∗ expression for ∇involves− the still unknown values of The term is therefore∇ − truly ∇ nonlocal. In the following, we F0 urms D and , which will be discussed in Section 4.2. combine∇ kinetic and enthalpy fluxes into a total contribution τred Fconv = Fenth + Fkin which arises from convection. 4.2. Relaxation time and mixing length 4. MODIFIED MIXING LENGTH MODEL In convection the relevant time scale is the turnover time, which we write as . We argue that should be 4.1. Stratification and flux balance τ = 1/urmskf kf estimated via the separation between the entropy rain struc- To construct an equilibrium model, we begin by consider- tures and not their thickness. This becomes important in the ing first the case without convection, so the flux F is carried picture in which the downdraft threads of the entropy rain by radiation alone. Hydrostatic and thermal equilibrium then merge with neighboring ones to form a tree-like structure, imply dP /dz = ρg and KdT /dz = F , or, alternatively as seen in the surface simulations (Stein & Nordlund 1989; for the logarithmic− gradients, − Spruit 1997), leading therefore to different scalings of the two length scales (separation and thickness of structures) with dln P /dz = ρg/P, (36) − depth. It is then not obvious which of these scales are more dln T /dz = F/(KT ). (37) relevant for determining the kf that is relevant for mixing. In − view of the uncertainty regarding the choice of kf , as well as The double-logarithmic temperature gradient is obtained by for comparison with standard MLT, we consider models with dividing the two equations through each other, i.e., a fixed value of kf as well as the more conventional case in which kf HP const. To capture the various cases in one dln T F P FcP ad ≈ = = = ∇ , (38) expression, we assume in the following ∇ dln P KT ρg Kg β −1 1−β kf = α H (kf0HP ) , (45) where we have used the perfect gas equation of state in the mix P where β = 0 corresponds to kf = kf0 with a fixed value kf0 of form P /T ρ = cP cV = cP (1 1/γ) = cP ad. If the energy is no longer− carried by radiation− alone, ∇cannot be the wavenumber of the entropy rain and β = 1 corresponds to computed from Equation (38), but we have to invoke∇ a suit- kf HP = αmix, where we have allowed for the possibility of able theory of convection. In standard MLT, one obtains as a free mixing length parameter, αmix, as is commonly done a solution of a cubic equation (Vitense 1953; Kippenhahn∇ & in standard MLT. It is not to be confused with the parameter Weigert 1990). In the following, we consider a modification aMLT that was introduced in Section 3.3 and Appendix D. that accounts for the possibility of a Deardorff flux. Negative values of β would correspond to shrinking scales, Flux balance implies that the sum of the radiative, enthalpy, but will not be considered here. Likewise, non-integer values and kinetic energy fluxes equals the total flux, i.e., of β are conceivable, but will also not be considered here. Returning now to the discussion of the relaxation time τ Ftot = Frad + Fconv. (39) in the beginning of this subsection, instead of associating 9

FIG. 2.— Illustrative flow structures (upper row) and corresponding horizontal power spectra (lower row) associated with the three combinations of β and β˜ considered in this paper. Light shades correspond to large logarithmic power, which is seen to extend over large values of k for β = 0 (Cases I and II) and is confined to progressively smaller k at larger depths when β = 1 (Case III). −1 it with the turnover time (urmskf ) , we will allow for the tion (43) in the form possibility of an additional dilution factor φ(z) and write 2 −β˜ −(1−β˜) −1 F =(σ/3γ ) ρu c α (k H ) , (46) τ = (urmskf φ) . Here, φ(z) increases with depth in a 0 ∇ad rms s mix f0 P similar fashion as the scale height, so we assume, in anal- where ogy with our treatment of k in Equation (45), the expres- f σ τ u k φ = u /(u + ιc /φ) (47) β−β˜ red rms f rms rms γ sion φ = (kf0HP /αmix) . This dilution factor only en- ≡ ters in expressions involving the turbulent diffusivity, such as quantifies the radiative heat exchange between convective el- in Equation (17), and hence terms involving τ or τ . For ements and the surroundings. The ι term defined in Equa- red √ β = 1, the case β˜ = 1 corresponds to the usual MLT concept, tion (13), has a maximum at ℓkf = 3, which typically oc- ˜ curs near the surface (e.g. Barekat & Brandenburg 2014). while β = 0 corresponds to a value of φ that increases with In Equation (46), the local value of the turbulent rms veloc- depth. ity always depends on the actual flux transported, and there- In Figure 2 we present illustrative flow structures as well fore it must also depend on ad + D. On dimensional as their depth-dependent horizontal power spectra associated ∇ − ∇ ∇3 grounds, since Fenth is proportional to u , we have with the three combinations of β and β˜ considered here. At rms the top of the domain, the size and separation of flow struc- u = c ( + )1/2 , (48) rms 0 ∇ − ∇ad ∇D tures are the same in all three cases. They remain constant 3/2 ˜ so Fenth and Fkin are proportional to ( ad + D) . with depth in the case β = β = 0 (Case I), while for β = 0 Standard mixing length arguments can∇ be − used ∇ to show∇ that ˜ and β = 1 (Case II), the separation increases with depth, the prefactor in Equation (48) is a fraction of cs. As shown in but the thickness is still constant. Finally, for β = β˜ = 1 Appendix F, we have ′ ′ (Case III), both thickness and separation increase with depth. −β −(1−β ) The filling factor decreases with depth in the case β = 0 c0/cs = σaMLT/3γαmix (kf0HP ) , (49) ˜ ′ p and β = 1, which could potentially make the kinetic energy where β = (β + β˜)/2. It is then clear that Fkin scales with ′ ˜ −3(1−β ) flux divergent with depth, while for both β = β = 0 and HP like (kf0HP ) . This is not the case for Fenth, β = β˜ = 1, the filling factor is independent of height. 1−β˜ however, because of the factor (kf0HP ) which, together ′ With these preparations in place, we can write F0 in Equa- 1−β with the (kf0HP ) factor, gives the scaling proportional to ′′ 2(1−β ) ′′ ′ (kf0HP ) , where β =(β + β˜)/2. 10

4.3. Equation for the superadiabatic gradient

Now that we know F0, we can solve the equation for the su- peradiabatic gradient. This leads to an equation that is similar to the cubic equation for , which is familiar from standard MLT (Kippenhahn & Weigert∇ 1990), ξ = + ǫ∗ ( + ) , (50) ∇rad ∇ ∇ − ∇ad ∇D where ξ = 3/2, and ǫ∗ = (1 φkin/φenth)(ǫenth + ǫkin) has contributions from the enthalpy− and kinetic energy fluxes. These expressions are similar to ǫ in Equation (44), except that ǫenth is evaluated with c0 in place of urms, i.e., ′′ ′′ 3/2 1/2 3 −2β −2(1−β ) ǫenth =(σ/3γ) aMLT (cs /χg) αmix (kf0HP ) , (51) where χ = K/ρcP is the radiative diffusivity. This shows that ǫ∗ is essentially a Peclet´ number based on cs. The contribution ∇ 5 from the kinetic energy flux is given by FIG. 3.— Solution of Equation (50) for ξ = 3/2 and rad = 10 (solid black), compared with the approximation (56) for q = ξ−1 (dashed blue) −β −(1−β) and (dash–dotted red), and (thin orange). The dotted line gives ǫkin/ǫenth = φkinaMLT ad α (kf0HP ) . (52) q = 1 ξ = 1 − ∇ mix an additional unphysical solution of Equation (50) for ξ = 3/2. The limiting ∇ ∇ ∇ We note in passing that ǫ∗ is also related to the Rayleigh cases = ad and rad are shown as thin horizontal green lines. number, which is commonly defined in laboratory and nu- It shows that rad for ǫ∗ 0 (stable surface layers) and merical studies of convection; see Appendix G. Furthermore, ∇ → ∇ → ad for ǫ∗ 1 (deeper layers). Next, to discuss the because of convection and the resulting bulk mixing, S is ∇ → ∇ ≫ general case ξ = 1, we define ∆ = and ∆ = now approximately constant, and therefore, unlike in the non- e 6 ∇ ∇− ∇ad ∇rad . For ∆ < 0 we have ∆ = ∆ , while convecting reference solution with K = const (Section 2), ∇rad − ∇ad ∇rad ∇e ∇rad can no longer be constant, but it reaches a minimum at the for ∆ rad > 0 and ∆ 1, a useful approximation is K ∇ e ∇ ≪ point where κ is maximum, which turns out to be at a depth of 1/ξ 1/ξ−1 1/ξ about 1 Mm in the convection models presented below. Since ∆ ∆ rad q∆ rad + ǫ∗ , (56) ∇≈ ∇ . ∇  ǫ∗ is inversely proportional to K, it reaches a maximum at −1 that depth and falls off both toward the top and the bottom of where q = ξ . It agrees with Equation (55) in the special case ξ = 1, where ∆ = ∆ rad/(1 + ǫ∗). In Figure 3 we the convection zone. ∇ ∇ 5 plot versus ǫ∗ for rad = 10 . The approximation yields An essential difference between Equation (50) and the usual ∇ ∇ −3 one in MLT is the presence of arising from the Dear- > rad for ǫ∗ < 10 , which is unphysical. This can be ∇D mitigated∇ ∇ by choosing q = 1; see Figure 3. dorff flux. Within the usual MLT, where D = 0, one finds that is slightly above , but now it∇ might instead be For the relevant case of large values of ǫ∗, we have ∇ ∇ad slightly above ad D. There are indeed two possibili- 1/ξ ∇ − ∇ ∆ ( rad/ǫ∗) . (57) ties for convecting solutions (Fconv > 0), one corresponding ∇≈ ∇ to a Schwarzschild-stable solution, This relation is useful because, even though both rad and ǫ∗ depend on K, their ratio does not and is given by∇ < < (stable), (53) ad D ad ′′ ∇ − ∇ ∇ ∇ 2−2β ′′ and one that is Schwarzschild unstable, rad 3 ad Ftot/ρ (kf0HP ) −(1+2β ) ∇ = ∇ 3 T , ǫ∗ σ c0g/kf ∝ ρcs ∝ ad D < ad < (unstable). (54) (58) ′′ ∇ − ∇ ∇ ∇ −(1+2β )/ξ Which of the two possibilities is attained depends on the value where β′′ =(β+3β˜)/4. Thus, ∆ T gives the ∇∝ of D and also on details of the solution. As will be discussed scaling of ∆ for the bulk of the convection zone. Therefore, ∇ ∇ −m in Section 6 below, entropy rain convection may actually still looking at Equation (48), we find Ma T with be Schwarzschild unstable without exciting giant cell con- ∝ vection if small-scale turbulent viscosity and diffusivity are m = 1 β′ + (1/2+ β′′)/ξ. (59) strong enough so that the local turbulent Rayleigh number for − For ξ = 3/2, using the relation 3β′ 2β′′ = β, we find the deeper layers is subcritical. − that m = (4 β)/3 is independent of the value of β˜. Thus, In this connection, we note that in standard MLT, one in- − cludes the effects of radiative cooling of the convective ele- we have m = 4/3 with β = 0 and m = 1 with β = 1. For isentropic stratification, this implies for the Mach number, ments in a different manner than here. Instead of ad, the effective buoyancy force is written as ′,∇ where − ∇ ′ given by Equation (34), the following scaling: ∆ζ = 8/9 with ∇ − ∇ ∇ β = 0 and ∆ζ = 2/3 with β = 1. always lies between and ad (Vitense 1953). Thus, one has < ′ < ,∇ which∇ resembles Equation (54) with a ∇ad ∇ ∇ 4.4. Relative importance of the Deardorff term negative value of D. To understand the∇ nature of the solutions of Equation (50), In Section 3.3 we have considered the depth dependence of it is instructive to treat ξ as an adjustable parameter. For given the Deardorff term via Equations (32) and (33). However, for values of rad and ad ad D, the case ξ = 1 yields D to be important at increasing depths, it must exceed the ∇ ∇ ≡ ∇ − ∇ sub∇ adiabatic gradient , because otherwise it would ∇ad − ∇ e rad + ǫ∗ ad not be possible for the Deardorff term to make ad + (ǫ∗)= ∇ ∇ . (55) ∇ − ∇ D positive. From Equations (57) and (58) we see that ∆ ∇ 1+ ǫ∗ e ∇ ∇ 11

TABLE 2 Fenth in Table 3, we see that only β enters, so the scaling of COMPARISON OF ζ∆∇, 2ζ˜ = 2ζ − 2∆ζ, AND THEIR DIFFERENCE FOR urms is fully characterized by that of small blobs with negative VARIOUS COMBINATIONS OF β AND β˜ USING ζ = 1. buoyancy.

′ ′′ β β˜ β β ζ∆∇ 2ζ˜ ζ∆∇ − 2ζ˜ 5. NUMERICAL SOLUTIONS CaseI 0 0 0 0 4/9 2/9 2/9 In this section we present numerical solutions to demon- CaseII 0 1 1/2 3/4 10/9 2/9 8/9 strate the effect of the term on the resulting stratifica- 1 0 1/2 1/4 2/3 2/3 0 ∇D CaseIII 1 1 1 1 4/3 2/3 2/3 tion. At the end of this section, we also compare with the non-convecting reference solution mentioned in the introduc- tion. We should emphasize that, although we use solar param- eters, our models cannot represent the Sun, because ioniza- TABLE 3 tion effects have been ignored (we take µ = 0.6 for the mean ILLUSTRATION OF SCALING RELATIONSHIPS WITH kf0HP . molecular weight in c c = /µ, where is the uni- p − v R R ∝ 3 1−β versal gas constant). A rather simple opacity law of the form Fenth urms (kf0HP ) buoyancy force 4 2 −1 −5 −3 1−β˜ of Equation (6) with κ0 = 10 cm g , ρ0 = 10 g cm , Fenth ∝ urms∆∇ / (kf0HP ) mean-field expression ′ 1/2 1−β ′ ˜ T0 = 13, 000 K, and a = 0.5, b = 18, for the exponents in the urms ∝ (∆∇) /(kf0HP ) β =(β+β)/2 ′′ 3/2 2(1−β ) ′′ ′ ˜ ˜ power-law expression for κH− has been used. We also neglect Fenth ∝ (∆∇) /(kf0HP ) β =(β +β)/2=(β+3β)/4 ′ F ∝ (∆∇)3/2/(k H )3(1−β ) kinetic energy flux the departure from plane-parallel geometry, so our model can kin f0 P only give qualitative indications. The system of two differential equations (36) and (37) de- depends on T in a power-law fashion. We are particularly couples by using ln P as the independent variable. We thus interested in the conditions under which ∆ falls off faster integrate ∇ than D, because that would ensure that the D term remains dln T /dln P = , (61) important∇ even at larger depths. ∇ ∇ For ξ = 3/2, and since the convection zone is nearly isen- using Equations (32), (33), and (50) to compute . As 2/3 1/4 ∇ tropically stratified (T ρ ), we have an initial condition we use T top = Teff /2 at a suffi- ∝ 5 −2 ciently low pressure (here P top = 10 dyncm ), so as −ζ∆∇ 4 ′′ ∆ ρ with ζ∆∇ = (1+2β ). (60) to capture the initially isothermal part of the atmosphere; ∇∝ 9 see, e.g., Bohm-Vitense¨ (1958) or Mihalas (1978). Here, 1/4 In Table 2 we compare for various combinations of β and β˜ Teff = (Ftot/σSB) is the effective temperature. We ap- the exponents for ∆ and D, i.e., 2ζ˜ = 2ζ and 2∆ζ, where proximate the urms term in Equation (47) by using the value ∆ζ = 2m/3. In all∇ cases∇ we have chosen ζ = 1, i.e., we from the previous step. The Deardorff term is characterized allow for moderately non-ideal (radiative) effects relative to by the assumed value of f and the value of ζ˜ = ζ ∆ζ s0 − the ideal case with ζ = 0.8. We see that the difference is (Section 3.3), where ζ = 1 and ∆ζ is a function of β and β˜ positive and non-vanishing in all cases, except for β = 1 and (Section 4.3). Since we integrate from the top downward, no β˜ = 0. In the following, we present solutions for all of the prior knowledge of ρ∗ is needed, because the Deardorff term remaining three cases. Before doing this, let us recapitulate is invoked only after has reached its peak value. ∇ − ∇ad what led to the threefold dominance of β˜ over β. For better Geometrical and optical depths are obtained respectively as illustration, we summarize in Table 3 the various relationships that led to the scaling of ∆ with k H . z = (P /ρg)dln P and τ = (κP /g)dln P, (62) ∇ f0 P − Z Z In the first expression for Fenth, β enters because it char- acterizes the relation between the buoyancy force propor- where the integration of τ starts at ln P top and that of z at 2 tional to δT/T and advection proportional to kf urms; see Ap- the position where τ = 1, which is referred to as the surface.− pendix D. For thin threads, we expect the relevant k to be f The factor κP /g, which is the same as HP , is retained be- large, i.e., β = 0 (Cases I and II in Figure 2). Next, in the cause of the similarity with that in the expression for τ. The second expression for Fenth, we have used a mean-field ex- z coordinate is used in some of our plots. pression to relate Fenth to ∆ via a turbulent diffusivity pro- In the following, we present solutions for the three combi- portional to τu2 u /k∇φ, where, as argued above, the rms rms f nations of β and β˜ sketched in Figure 2. We recall that only dilution factor φ has≈ entered. It is this expression that is clos- Cases I and II (β = 0 with β˜ = 0 or 1) correspond to small est to conventional MLT, because here we expect ˜ . β = 1 length scales in the deeper layers (see also the power spectra The remaining two relationships in Table 3 explain why the in Figure 2) and are therefore of interest when trying to rec- ˜ β term appears three times more dominantly than the β term. oncile the non-detection of convective motions by Hanasoge By equating the first two expressions for Fenth in Table 3, we et al. (2012, 2016) at the theoretically expected levels. We did find first of all the relation between u and ∆ , where β rms already emphasize that Case II with β = 0 and β˜ = 1 is likely ˜ ′ ∇ ˜ and β contribute with equal shares through β = (β + β)/2. to lead to large kinetic energy fluxes. However, based on the However, to see the scaling of ∆ , we need to go back to the results presented below, it turns out that in the case β = β˜ = 0 second expression for , because∇ it changes only weakly Fenth (Case I), which was favored by these two requirements (small ′ ˜ with depth. Now, β and β contribute with equal shares, and length scale and non-divergent kinetic energy flux), the Dear- this means that β˜ has now become three times more dominant dorff term is unlikely to have a significant effect, because for ′′ than β through the expression β = (β + 3β˜)/4. Thus, for β˜ = 0, the gradient term in the enthalpy flux becomes rather β = 0 and β˜ = 1, ∆ shows nearly the standard scaling with inefficient and must therefore be compensated for by a cor- ∇ HP . Furthermore, looking again at the first expression for respondingly larger superadiabatic gradient, and thus, only a 12

FIG. 4.— Profiles of S/cP , ∇−∇ad, and urms for fs0 = 0 (∇D = 0), FIG. 5.— Same as Figure 4, but for β = 0, β˜ = 1 (Case II) with ζ˜ = 1/9, as well as fs0 = 0.2, 0.3, and 0.5 for β = β˜ = 0 (Case I) with ζ˜ = 1/9. and fs0 = 0, 0.1, and 0.2. The location of the surface (τ = 1) is indicated by vertical dash-dotted lines and geometric depths below the surface are indicated in the middle panel, starting with tick marks at 100, 200, and 500 km, and continuing with 1, riod of buoyancy oscillations would be of the order of days. ∇−∇ 2, and 5Mm, etc. The inset in the middle panel shows ad over a This is comparable to or less than the turnover time τ. Indeed, narrower range as a function of −z. one finds that τN /∆ 1 exceeds unity in the BV ≈ ∇D ∇− deeper parts, which is ap consequence of D falling off with rather large Deardorff flux (fs0 0.5) can make the resulting ∇ stratification sufficiently subadiabatic.≥ This is the case shown ρ with a smaller power than ∆ , as discussed in Section 4.4. Nevertheless, the resulting decrease∇ in S/c with depth re- in Figure 4, where we present profiles of S/cP , ad, and P ∇−∇ mains always small (10−3 or less) compared with the value urms for fs0 = 0 (no Deardorff term), as well as fs0 = 0.2, of ∆S /c produced at the surface ( u2 /a c2 0.3, and 0.5. The urms profiles are basically the same for 0 P ≈ rms MLT∇ad s ≈ −1/3 0.01). Thus, based on this, the descending low-entropy blobs all values of fs0 and fall off like P . This is expected, would reach the bottom of the convection zone before their 1/2−m because urms T and m = 4/3; see Equation (59), negative buoyancy is neutralized by the decreasing average ∝ 1/2 entropy. In other words, they will never perform any actual os- where we have used cs T . Thus, for isentropic stratifi- ∝(1−2m)/5 −1/3 cillations before reaching the bottom of the convection zone, cation we have urms P = P . i.e. where . ∝ Fconv = 0 Next, we show in Figures 5 and 6 Cases II and III with β˜ = It may be interesting to note that the depth of the con- 1 and fs0 = 0, 0.1, and 0.2. For both β = 0 and β = 1, the vection zone increases slightly with increasing values of fs0. Deardorff flux now has a stronger effect and is able to make Of course, our model is idealized and represents the Sun at the deeper parts of the domain Schwarzschild stable even for best only approximately. Furthermore, as emphasized in Sec- fs0 = 0.1. Those layers would then be Schwarzschild stable tion 1, the depth of the convection zone is well determined and no longer a source of giant cells. Again, the profiles of seismically, and this should be reproduced by a solar model urms are similar regardless of the value of fs0, but fall off with realistic atomic physics and appropriately chosen ad- −1/5 justable parameters. However, it is known from the work of more slowly for β = 1 (urms P ), compared to β = 0 −1/3 ∝ Serenelli et al. (2009) that a slight expansion of the solar con- (urms P ). This agrees with our theory, because for vection zone would actually be required to compensate for the ∝ (1−2m)/5 −1/5 m = 1 we have urms P = P . shrinking that follows from the downward revision of solar Given that the stratification∝ is stable in the deeper parts, we abundances (Asplund et al. 2004) which is based on three- can calculate the Brunt-Vais¨ al¨ a¨ frequency of buoyancy oscil- dimensional convective atmosphere simulations, compared to 2 lations, NBV, which is given by NBV = ( ad)g/HP . previous analysis based on one-dimensional semi-empirical At intermediate and larger depths of the convection− ∇ − ∇ zone, we models (Grevesse & Sauval 1998). −2 −2 −4 have g/HP < 10 s and ad < 10 , so the pe- Finally, we compare in Figure 7 the standard convective so- ∼ ∇ − ∇ ∼ 13

with the non-convective radiative reference solution. Not sur- prisingly, owing to the absence of convection, the same flux can now only be transported with a greatly enhanced nega- tive (unstable) entropy gradient near the surface. However, this layer is now extremely thin (1.15 Mm) and the peak in rad/ ad is about 100 km below the τ = 1 surface. Note also∇ that∇ its peak value ( 105) is below that in the presence of convection, where it reaches≈ a maximum of 4 106 at a depth of 1 Mm. As shown in Equation (42),≈ this× value can be interpreted≈ as the local Nusselt number. Note also that the result for β = β˜ = 0 (when is weak) is rather similar ∇D to that for β = β˜ = 1; see the dashed and dotted lines in Figure 7. Our calculations have demonstrated that for β˜ = 1, regard- less of the value of β, bulk mixing changes the non-convecting reference state to a nearly isentropic one. However, whether the mean entropy gradient is slightly stably or slightly unsta- bly stratified depends on the presence of the Deardorff flux. In the model with β˜ = 0, however, bulk mixing is rather inef- ficient and the stratification would be Schwarzschild unstable unless an unrealistically large Deardorff flux is invoked. At the end of the introduction, we discussed that the entropy rain itself might create an unstable stratification. Let us now return to this question with more detailed estimates. This will be done in the following section.

6. ALTERNATIVE CONSIDERATIONS Assuming that the upflows are perfectly isentropic, Spruit (1997) argues that the low-entropy material from the top with its decreasing entropy filling factor fs in deeper layers neces- FIG. 6.— Same as Figure 4, but for β = β˜ = 1 (Case III) ζ˜ = 1/3, and f = 0, 0.1, and 0.2. sarily leads to a negative mean entropy gradient. Specifically, s0 using Equation (24), one obtains

dS/cP (∆S)0 dfs 2 (∆S)0 HP = = fs > 0, (63) − dz cP dln P 5 cP

where we have used dln fs/dln ρ = 2/3 and dln ρ/dln P = 3/5 for an isentropic layer with γ = 5/3. Thus, the stratifica- tion would be Schwarzschild unstable. This is also borne out by the solar simulations with realistic physics, although the computational domains are sufficiently shallow so that the ra- diative flux is still small in the deeper parts and usually even neglected altogether. Toward the bottom of the convection zone, however, radiation becomes progressively more impor- tant and the mean entropy gradient in the upflows may no longer be vanishing. To estimate the mean entropy gradient in the upflows, we may balance the steady state entropy advection with the neg- ative radiative flux divergence, i.e.

ρ T U ↑ dS↑/dz (dF /dz)↑. (64) ≈− rad The sign of dFrad/dz is negative and thus compatible with a positive dS↑/dz in the upflows, but it would only be large ∇ ∇ FIG. 7.— Comparison of S/cP (top) and rad/ ad (bottom) between the enough near the bottom of the convection zone. Higher up, non-convective radiative reference solution (∇ = ∇rad) and standard con- dFrad/dz becomes smaller and eventually unimportant. On vective solutions (∇D = 0) with β = β˜ = 0 (red, dashed) and β = β˜ = 1 (blue, dotted). In both panels, the location of the τ = 1 surface is indicated the other hand, U ↑/urms can be rather small (see Table 1). by vertical dash-dotted lines and geometric depths below the surface are in- Furthermore, our considerations neglect the fact that the gas in dicated for the non-convective solution, starting with tick marks at 50, 100, the upflows expands, so only a fraction of the gas can ascend and 200 km, etc. The location of the a priori unstable layer (∇rad > ∇ad) is marked in the lower panel by a small gray strip. For the other solutions, before it begins to occupy the available surface area. There- the depths are different; see those in Figure 6 for the solution with fs0 = 0, fore, the rest of the gas would have to remain stagnant and which is the same as here for β = β˜ = 1. continue to heat up. In reality, of course, there would be con- tinuous entrainment, resulting in a finite, but still low, effec- lution (β = β˜ = 1; same as the case with fs0 = 0 in Figure 6) tive upward velocity. 14

ciently strong turbulent viscosity and turbulent thermal diffu- sivity coefficients. These turbulent coefficients define a turbu- lent Rayleigh number for a layer of thickness d, gd4 ∆ Rat = ∇ . (66) νtχt  HP z∗ This number would then have to be still below the critical value for convection. Equation (66) differs from the usual one defined through Equation (G1) in Appendix G in that, first, ν νt and χ χt have been substituted, and sec- ond, the→ superadiabatic→ gradient of the non-convecting refer- ence solution is now replaced by the actual one. This might be a plausible alternative to explaining the absence of giant cell convection if the idea of turning the stratification from Schwarzschild unstable to Schwarzschild stable through the term were to turn out untenable. ↑ − D FIG. 8.— Dependence of U eff /urms on depths z for Case I with fs0 = ∇We recapitulate that in this alternate explanation, the strat- 0.5 (solid red line), as well as Cases II and III with fs0 = 0.1 (dashed blue and dotted green lines, respectively). ification is Schwarzschild unstable, i.e. Rat > 0, correspond- crit ing to Equation (54), but Rat < Rat is still below a certain We now use selected solutions obtained in Section 5 to es- critical value, so it would be stable by a turbulent version of timate the effective fractional upward velocity for radiative the Rayleigh-Benard´ criterion. Tuominen et al. (1994) found heating/cooling to dominate over advection, defined as crit Rat 300 for a vanishing rotation rate. However, they also ↑ ≈ crit U /u =( dF /dz) / (u ρ T dS/dz). (65) estimated that Rat > Rat for plausible solar parameters, eff rms − rad rms so the direct adoption of this idea would be problematic, too. Here we have used T dS/dz = g ∆ / ad. Figure 8 shows However, one might speculate that a more accurate treatment this quantity for Cases I–III, which are∇ shown∇ in Figures 4–6. could lead to stability when allowing, for example, for spa- It turns out that for Case I with fs0 = 0.5, the effective frac- tial nonlocality of the turbulent transport (Brandenburg et al. tional upward velocity is around 0.01, while for Case III with 2008; Rheinhardt & Brandenburg 2012). fs0 = 0.1 it can be even larger. The value 0.01 is compatible 7. CONCLUSIONS ↑ with the U /urms values in Table 1 if we assume f = 0.01 In the present work we have suggested that the enthalpy flux ↑ ↑ ↑ in stellar mixing length models should contain an extra non- (so U /urms = 0.1) and U eff /U = 0.1. For Case II with fs0 = 0.1, the effective fractional upward velocity is around local contribution so that the enthalpy flux is no longer pro- −3, which may be unrealistically small. portional to the local superadiabatic gradient, ad, but 10 ∇ − ∇ Based on the considerations above, we may conclude that to ad + D, where D is a new nonlocal contribution ∇ − ∇ ∇ ∇ the case for a Schwarzschild-stable entropy gradient depends that was first identified by Deardorff (1966) in the meteoro- on model assumptions for the implementation of the Dear- logical context. The significance of this term lies in the fact dorff flux that are potentially in conflict with the mean en- that it provides an alternative to the usual local entropy gradi- tropy gradient expected from the two-stream model. Whether ent term and can transport enthalpy flux outwards—even in a or not there is really a potential conflict depends not only on slightly stably stratified layer. the model parameters, but also on Equation (64) itself. This We have presented a modified formulation of stellar MLT that includes the term, in addition to a term result- is because the T factor in the entropy equation is outside ∇D ∇kin the derivative, making it impossible to derive the total flux ing from the kinetic energy flux. The formalism and the fi- balance assumed in Equation (39); see also Rogachevskii & nal results are similar to those of conventional MLT in that Kleeorin (2015) for related details. one also arrives here at a cubic equation for , but the term is now replaced by + ∇ . This In view of these complications, it is worthwhile to discuss ∇ − ∇ad ∇ − ∇ad ∇D − ∇kin alternative ways of avoiding giant cell convection. For this new formulation implies that convection can carry a finite flux while is still negative and therefore the stratification purpose, we have to address the global problem of using some ∇ − ∇ad coarse-grained form of effective mean-field equations. When is Schwarzschild stable, i.e., Equation (53) is obeyed. Conse- considering the equations of mean-field hydrodynamics, in quently, if confirmed, no large length scales are being excited. which the small-scale enthalpy and fluxes are pa- The present formulation allows for different treatments of rameterized in terms of negative mean entropy and mean ve- the length scales governing buoyant elements on the one hand locity gradients, respectively, one finds solar differential rota- and the time and length scales associated with mixing on the tion as a result of non-diffusive contributions to the Reynolds other. When both are independent of depth (β = β˜ = 0), stress, but in certain parameter regimes a new instability was mixing becomes inefficient at larger depths. Thus, to carry a found to develop (Tuominen & Rudiger¨ 1989; Rudiger¨ 1989; certain fraction of the enthalpy flux, the superadiabatic gradi- Rudiger¨ & Spahn 1992). This instability was later identified ent needs to be larger than otherwise, making it harder for the as one that is analogous to Rayleigh-Benard´ convection, but Deardorff term to revert the sign of ad. On the other ∇ − ∇ now for an already convecting mean state (Tuominen et al. hand, for a tree-like hierarchy of many downdrafts merging 1994). This instability would lead to giant cell convection. into fewer thin ones at greater depths (β = 0, β˜ = 1), The existence of giant cell convection is under debate, but the increasing length scale associated with increasing separa- assuming that it does not exist in the Sun, one might either tion enhances vertical mixing, making the stratification nearly hypothesize that the mean-field equations are too simplified isentropic without the Deardorff term, and slightly subadia- or that mean-field convection could be suppressed by suffi- batic with a weak Deardorff term. In that case, however, the 15

filling factor of the downflows decreases with depth as ρ−ζ , slightly stable. This would be particularly useful in global which may imply an unrealistically large downward kinetic simulations that would otherwise have no entropy rain. energy flux. This leaves us with the standard flow topology As stimulating as the results of Hanasoge et al. (2012) are, (β = β˜ = 1), where both size and separation of structures they do require further scrutiny and call for the resolution of increase with depth. The Deardorff flux can still cause the the existing conflicts with other helioseismic studies such as stratification to have a subadiabatic gradient, so no giant cell those of Greer et al. (2015). Alternatively, global helioseismic convection would be excited locally, but the flow structures techniques for detecting giant cell convection (Lavely & Ritz- would be large and should be helioseismically detectable, as woller 1993; Chatterjee & Antia 2009; Woodard 2014) can has been found by Greer et al. (2015) using ring-diagram local provide another independent way of detecting deep larger- helioseismology. scale flows (Woodard 2016). Realistic simulations should It would be useful to explore the thermodynamic aspects of eventually agree with helioseismic results of flows in deeper the present model more thoroughly and to connect with re- layers of the Sun, but at the moment it is still unclear whether lated approaches. An example is the work by Rempel (2004), a subgrid scale treatment as in Equation (1) adequately cap- who studied a semianalytic overshoot model that was driven tures the small-scale flows that can be responsible for the nonlocally by downdraft plumes, similar to what was sug- Deardorff flux and whether they would in principle be able gested by Spruit (1997). Rempel also finds an extended suba- to predict the subtle departures from superadiabatic stratifica- diabatic layer in large parts of the model. Furthermore, there tion on subthermal time scales. are similarities to the nonlocal mixing length model by Xiong & Deng (2001), who, again, find an extended deeper layer that is subadiabatically stratified in the deeper parts of their I thank the referee for his/her criticism that has led to model. In this connection we emphasize the main difference many improvements, and Evan Anders, Jean-Francois Cos- between nonlocal turbulence owing to the Deardorff flux and sette, Ben Greer, Ake˚ Nordlund, Mark Rast, Matthias Rem- usual overshoot: in the latter case the enthalpy flux would pel, Matthias Rheinhardt, Bob Stein, Peter Sullivan, Regner go inward as a consequence of the reversed entropy gradient, Trampedach, and Jorn¨ Warnecke for interesting discussions while in the present model the Deardorff flux goes outward. and comments. I also wish to acknowledge Juri Toomre for Future work could proceed along two separate paths. On mentioning to me the need for modeling entropy rain during the one hand, one must establish the detailed physics leading my time in Boulder some 24 years ago. This work was sup- to the D term using models with reduced opacity, in which ported in part by the Swedish Research Council grant No. reliable∇ DNS are still possible, i.e., no SGS terms are added 2012-5797, and the Research Council of Norway under the and the primitive equations are solved as stated, without in- FRINATEK grant 231444. This work utilized the Janus super- voking Equation (1). On the other hand, one could study suit- computer, which is supported by the National Science Foun- ably parameterized large eddy simulations that either include dation (award number CNS-0821794), the University of Col- a nonlocal Deardorff term of the form given by Equation (32), orado Boulder, the University of Colorado Denver, and the as discussed in Section 3, or that explicitly release entropy National Center for Atmospheric Research. The Janus super- rain at the surface such that the resulting stratification is still computer is operated by the University of Colorado Boulder.

APPENDIX POLYTROPIC STRATIFICATION FROM KRAMERS OPACITY We show here that, for the non-convecting reference solution, using the Kramers-like opacity law of Equation (2), but not the combined opacity law of Equation (6), we have K const in the deeper, optically thick layers. Dividing Equation (3) by the equation for hydrostatic equilibrium, dP/dz = ρg→, we have − dT F F (ρ/ρ )a F (P/P )a rad rad 0 rad 0 (A1) = = 3−b = 3+a−b , dP Kρg K0ρ0g (T/T0) K0ρ0g (T/T0) 3 where K0 = 16σSBT0 /(3κ0ρ0) is a constant and P/P0 = (ρ/ρ0)(T/T0) is the ideal gas equation with a suitably defined constant P0 = (cp cv)ρ0T0. Here, ρ0 and T0 are reference values that were defined in Equation (2). Equation (A1) can be integrated to give − (T/T )4+a−b =(n + 1) (0) (P/P )1+a +(T /T )4+a−b, (A2) 0 ∇rad 0 top 0 where (0) = F P /(K T ρ g), which is defined analogously to the without superscript (0) in Equations (38) and (41), ∇rad rad 0 0 0 0 ∇rad and Ttop is an integration constant that is specified such that T Ttop as P 0. Note also that 4+ a b = (n + 1)(1 + a), where n was defined in Equation (5) as the polytropic index, so the→ ratio of 4+→a b to 1+ a is just n +− 1, which enters in front − of the (0) term in Equation (A2). Since K T 3−b/ρ1+a T 4+a−b/P 1+a, we have K const = K for T T . ∇rad ∝ ∝ → 0 ≫ top DERIVATION OF EQUATIONS (10) AND (11) To obtain Equations (11) and (12), which are used to derive the Deardorff flux term in the τ approximation, we start with the equations for specific entropy and velocity in the form (see, e.g., Barekat & Brandenburg 2014) DS ρT = ∇ F , (B1) Dt − · rad DU ρ = ∇P + ρg, (B2) Dt − 16

1/2 FIG. C1.— Velocity vectors superimposed on a color scale representation of the vorticity ω at times 0, 10, and 20 in units of (HP /g) , as well as the resulting filling factor versus density. The negative slope is ζ = 0.8. Note that the frame of view changes as the vortex descends and shrinks. where D/Dt = ∂/∂t + U ∇ is the advective derivative, F rad is the radiative flux, and viscosity has been omitted. Subtracting those equations from their averaged· ones, we obtain the following set of equations ∂s ∂s ∂S 1 + U j + uj = ∇ δF rad + s, (B3) ∂t ∂xj ∂xj −ρ T · N

∂ui ∂ui ∂U i 1 + U j + uj = ∇p + gi(p/γP s/cP )+ ui, (B4) ∂t ∂xj ∂xj −ρ − N where we have used δρ/ρ = p/γP s/c , which can be obtained from Equation (8) and the perfect gas law by linearization. − P Pressure fluctuations will again be neglected and (∇ δF )/ρ T will be replaced by s/τ , as explained in Section 3. · rad − cool Assuming U = 0 and omitting the nonlinear terms and we arrive at Equations (11) and (12). Ns Nui FILLING FACTOR FOR A DESCENDING HILL VORTEX

The solution for a Hill vortex with radius aH and propagation velocity uH is given by a stream function Ψ in spherical coordi- nates (r,θ,φ) as (e.g. Moffatt & Moore 1978) 2 2 1 3uH(1 r /aH) ̟ (for raH). − H We apply it in Cartesian coordinates as the initial condition for the mass flux as ρu = ∇ (Ψφˆ), where φˆ =( y/̟,x/̟, 0) is the unit vector in the toroidal direction of the vortex, using ̟2 = x2 + y2 and r2 = ×̟2 + z2 for cylindrical− radius ̟ and spherical radius r. We adopt here an isothermal equation of state, i.e., there is no buoyancy force in this problem. For non- isothermal calculations, but in two dimensions, we refer to the work of Rast (1998). Density and pressure fall off exponentially with height and both the scale height and the sound speed are independent of height. No analytic solution exists in that case, so the Hill vortex solution is at best approximate. We consider a domain of size L L 4L with L/2 < x,y/HP < L/2 and × × − −5 7L/2

MLT RELATION BETWEEN urms AND srms

To find the relation between urms and the rms values of temperature or entropy fluctuations, it is customary in standard MLT ∇ 2 to approximate the steady state momentum equation, u u g δT/T , by urmskf = aMLT gδT/T , where aMLT 1/8 is a commonly adopted geometric factor (e.g., Spruit 1974).· This leads≈− to ≈ 2 2 srms/cP = δT/T = urmskf /aMLTg = γMa kf HP /aMLT, (D1) where we have used c2 = γgH , and thus (s /c ) Ma−2 = γk H /a . Furthermore, using c T = gH , we have s P rms p f P MLT ∇ad p P Ts = u2 k H /(a ), which is used to derive Equation (31). rms rms f P MLT∇ad ESTIMATE FOR THE SURFACE VALUE OF ∇D The purpose of this section is to show that is a certain fraction of in the top layers, as stated in Equation (32). To ∇D ∇ − ∇ad have an estimate for s2, we multiply Equation (11) by s and average, so we get 1 ∂s2 = u s S s2/τ + . (E1) 2 ∂t − j ∇j − cool Ts 3 https://github.com/pencil-code 17

As for Fenth in Equations (14) and (15), we have a triple correlation term, which is here s = su ∇s. Again, we adopt the τ 2 T − · approximation and replace s by a damping term of the form s = s /τ which, together with the τcool term, combines to give T T − 2 2 τred as the relevant time scale. Assuming a statistically steady state, ∂s /∂t = 0, we derive the following expression for s : s2 = τ u s S (in the surface layers). (E2) − red j ∇j This shows that fluctuations of specific entropy are produced when there is an outward flux (uzs > 0) and a locally negative (unstable) mean entropy gradient; see also Garaud et al. (2010) for a similar derivation. Inserting this into Equation (18), and using Equation (9), we obtain F = τ 2 g (F ∇S/c ) (in the surface layers). (E3) D red enth · P 2 Here, both g and ∇S point downward, so F D points upward and we can write FD = λFenth, where λ =(τred/τff ) ( ad) 1/2 ∇ − ∇ is a coefficient that itself is proportional to the superadiabatic gradient, and τff = (HP /g) is the free-fall time (after which a fluid parcel at rest has reached a depth of H /2). Since F = F + F , this implies F = (1 λ)F and therefore λ< 1 P enth G D G − enth in the highly unstable layer at the top, where both FG and Fenth are positive. We expect FD to be largest just a few hundred kilometers below the photosphere, so λ should be maximum in the upper parts. To calculate the fraction /( ), we use ∇D ∇ − ∇ad the fact that Fenth = FG + FD, together with the part of Equation (20) that relates to FD, to write F = 1 ρc T (τ u2 /H ) =(τ /τ )2( )F . (E4) D 3 P red rms P ∇D red ff ∇ − ∇ad enth Starting with Equation (31) and expressing τred in terms of σ using Equation (47), we find

˜ ˜ max/( ) = 3σφ α−β (k H )−(1−β) 3, (E5) ∇D ∇ − ∇ad max enth∇ad mix f0 P ≈ where we have used β˜ = 1 (appropriate to the near-surface layers), ignored radiative cooling (σ = 1), and assumed αmix = 1.6 (Stix 2002), as well as φenth 4 that was found in simulations of Brandenburg et al. (2005), as discussed in Section 3.3. We emphasize that these equations≈ only characterize the initiation of entropy rain. They cannot be used to compute the Deardorff flux in the deeper layers where we have instead invoked Spruit’s concept of nonlocal convection in the form of threads. The s2 associated with those threads is likely to result from a part of the triple correlation term = su ∇s, which gives Ts − · rise to a negative divergence of the flux of s2 of the form H us2 on the rhs of Equation (E1). This flux (which should not 2 ≡ be confused with an energy flux) should point downward (s is largest when uz < 0) and be strongest in the upper layers, so ∇ H < 0, leading to a positive contribution from ∇ H on the rhs of Equation (E1). This may explain the s2 associated with · − · the D term. To∇ estimate the ratio F /F in the deeper layers, we use Equation (E1) in the steady state with 1/τ 0 and = D G cool → Ts ∇ H (2ζ ∆ζ)s2u /γH and obtain − · ≈ − rms P 0= u s S + (2ζ ∆ζ)s2u /γH (in deeper layers). (E6) − j ∇j − rms P 1 2 2 Multiplying by 3 τredurms (ρ T ) , using Equations (9) and (17), and expanding the fraction by g/cp, we have

1 3 2 0= Fenth FG + 3 ρurms(2ζ ∆ζ) τreds g ρ T cpT /(γgHP ). (E7) −   This implies that FG < 0 in the deeper layers. The second term in the parentheses is FD (with a plus sign, because g > 0 is a scalar here); see Equation (18). Furthermore, we use c T /(γgH ) = 1/(γ 1) and ρu3 = F /φ ; see Equation (31). p P − rms enth enth The Fenth terms on both sides cancel, so we have F / F = 3(γ 1)/ [φ (2ζ ∆ζ)] 0.3 ... 0.5, (E8) D | G| − enth − ≈ where 2ζ ∆ζ = 2ζ˜ + ∆ζ = 10/9 for Cases I and II, and 14/9 for Case III (Section 4.3), and φ = 4 has been assumed. − enth

DERIVATION OF EXPRESSION FOR c0

To find the coefficient c0 given by Equation (49), we use Equation (D1) in the form c δT = u2 k H / a . (F1) P rms f P ∇ad MLT Inserting this into Equation (7), and using Equation (45), yields

F = ρu3 αβ (k H )1−β/ a . (F2) enth rms mix f0 P ∇ad MLT Equating this expression with Equation (43), using Equation (46), we can derive the desired expression for urms in the form 2 2 σ cs aMLT ′ (F3) urms = 2−2β ( ad + D) , 3 γ (kf0HP ) ∇ − ∇ ∇ ′ ˜ 2 where β =(β + β)/2 and cs = γgHP have been used. We thus find the coefficient c0 as stated in Equation (49). 18

RAYLEIGH NUMBER

The purpose of this section is to show that ǫ∗, as defined in Equation (51), is related to the Rayleigh number Ra. In laboratory and numerical studies of convection, it is customary to define Ra as (e.g. Kapyl¨ a¨ et al. 2009) non−conv gd4 dS/c gd4 non−conv gd4 Ra = P ∇rad − ∇ad ∇rad − ∇ad , (G1) νχ − dz z∗ ≈ νχ  HP max ≈ νχ  HP max which is usually evaluated in the middle of the layer at z = z∗. In the Sun, however, the maximum value is more relevant. The superscript “non-conv” indicates that the entropy gradient is taken for the non-convecting reference state, d is the thickness of the layer and ν is the viscosity. Introducing the Prandtl number Pr = ν/χ and using the definition of Nu in Equation (42), we have Pr Ra gd4 = ∇ad . (G2) Nu 1 χ2H − P On the other hand, using Equation (51), we find σ3a (c3/χg)2 σ3a H3g σ3a (H /d)4 gd4 2 MLT s MLT P MLT P ad (G3) ǫ∗ = 3 4(1−β′′) = 4(1−β′′) 2 = 4(1−β′′) 2∇ , 27γ (k H ) 27(k H ) χ 27 (k H ) χ HP f0 P f0 P ∇ad f0 P and therefore 3 2 σ aMLT/27 ad Pr Ra ǫ = ∇ ′′ , (G4) ∗ (k d)4(k H )−4β Nu 1 f0 f0 P − 2 which shows that Ra is proportional to ǫ∗.

REFERENCES Arnett, W. D., Meakin, C., Viallet, M., Campbell, S. W., Lattanzio, J. C., & Greer, B. J., Hindman, B. W., Featherstone, N. A., & Toomre, J. 2015, ApJ, Mocak,´ M. 2015, ApJ, 809, 30 803, L17 Asplund, M., Grevesse, N., Sauval, A. J., Allende Prieto, C., & Kiselman, D. Grevesse, N., & Sauval, A. J. 1998, Spa. Sci. Rev., 85, 161 2004, A&A, 417, 751 Gudiksen, B. V., Carlsson, M., Hansteen, V. H., Hayek, W., Leenaarts, J., Barekat, A., & Brandenburg, A. 2014, A&A, 571, A68 Martınez-Sykora, J. 2011, A&A, 531, A154 Basu, S. 1997, MNRAS, 288, 572 Hanasoge, S. M., & Sreenivasan, K. R. 2014, Sol. Phys., 289, 3403 Biermann, L. 1932, Z. Astrophys, 5, 117 Hanasoge, S. M., Duvall, T. L., Jr., & DeRosa, M. L. 2010, ApJ, 712, L98 Biermann, L. 1938, Astron. Nachr., 264, 395 Hanasoge, S. M., Duvall, T. L., & Sreenivasan, K. R. 2012, Proc. Nat. Acad. Blackman, E. G., & Field, G. B. 2003, Phys. Fluids, 15, L73 Sci., 109, 11928 Bogart, R. S., Baldner, C. S., & Basu, S. 2015, ApJ, 807, 125 Hanasoge, S., Gizon, L., & Sreenivasan, K. R. 2016, Ann. Rev. Fluid Mech., Bohm-Vitense,¨ E. 1958, Z. Astrophys., 46, 108 48, 191 Brandenburg, A., Chan, K. L., Nordlund, A.,˚ & Stein, R. F. 2005, Astron. Hathaway, D. H., Upton, L., & Colegrove, O. 2013, Science, 342, 1217 Nachr., 326, 681 Heslot, F., Castaing, B., Libchaber, A. 1987, Phys. Rev. A, 36, 5870 Brandenburg, A., Jennings, R. L., Nordlund, A.,˚ Rieutord, M., Stein, R. F., & Hill, M. J. M. 1894, Phil. Trans. R. Soc. Lond. A, 185, 213 Tuominen, I. 1996, J. Fluid Mech., 306, 325 Hotta, H., Rempel, M., & Yokoyama, T. 2015, ApJ, 803, 42 Brandenburg, A., Kapyl¨ a,¨ P. J., & Mohammed, A. 2004, Phys. Fluids, 16, Hurlburt, N.E., Toomre, J., Massaguer, J.M. 1984, ApJ, 282, 557 1020 Hurlburt, N. E., Toomre, J., Massaguer, J. M. 1986, ApJ, 311, 563 Brandenburg, A., Radler,¨ K.-H., & Schrinner, M. 2008, A&A, 482, 739 Hurlburt, N. E., Toomre, J., Massaguer, J. M., & Zahn, J. P. 1994, ApJ, 241, Buske, D., Vilhena, M. T., Moreira, D. M., & Tirabassi, T. 2007, Environ. 245 Fluid Mech., 7, 43 Kapyl¨ a,¨ P. J., Korpi, M. J., & Brandenburg, A. 2009, ApJ, 697, 1153 Castaing, B., Gunaratne, G., Heslot, F., Kadanoff, L., Libchaber, A., Thomae, Kapyl¨ a,¨ P. J., Mantere, M. J., Cole, E., Warnecke, J., & Brandenburg, A. 2013, S., Wu, X.-Z., Zaleski, S., & Zanetti, G. 1989, J. Fluid Mech., 204, 1 ApJ, 778, 41 Cattaneo, F., Brummell, N. H., Toomre, J., Malagoli, A., & Hurlburt, N. E. Kippenhahn, R., & Weigert, A. 1990, Stellar structure and evolution 1991, ApJ, 370, 282 (Springer: Berlin) Chatterjee, P., & Antia, H. M. 2009, ApJ, 707, 208 Lavely, E. M., & Ritzwoller, M. H. 1993, Sol. Phys., 403, 810 Christensen-Dalsgaard, J., Gough, D. O., & Thompson, M. J. 1991, Sol. Lord, J. W. 2014, Deep convection, magnetism and solar supergranulation Phys., 378, 413 (PhD thesis, University of Colorado at Boulder) Cossette, J.-F., & Rast, M. P. 2016, ApJ, 829, L17 Lord, J. W., Cameron, R. H., Rast, M. P., Rempel, M., & Roudier, T. 2014, Davidson, P. A. 2004, Turbulence: an introduction for scientists and engineers ApJ, 793, 24 (Oxford: Oxford University Press) Losada, I. R., Brandenburg, A., Kleeorin, N., & Rogachevskii, I. 2013, A&A, Deardorff, J. W. 1966, J. Atmos. Sci., 23, 503 556, A83 Deardorff, J. W. 1972, J. Geophys. Res., 77, 5900 Miesch, M. S., Brun, A. S., De Rosa, M. L., & Toomre, J. 2008, ApJ, 673, De Roode, S. R., Jonker, H. J. J., Duynkerke, P. G., & Stevens, B. 2004, 557 Boundary-Layer Meteorology, 112, 179 Miesch, M. S., Featherstone, N. A., Rempel, M., & Trampedach, R. 2012, Duvall, T. L., Jr., Kosovichev, A. G., Scherrer, P. H., Bogart, R. S., Bush, R. ApJ, 757, 128 I., de Forest, C., Hoeksema, J. T., Schou, J., Saba, J. L. R., Tarbell, T. D., Mihalas, D. 1978, Stellar Atmospheres (W. H. Freeman: San Francisco) Title, A. M., Wolfson, C. J., & Milford, P. N. 1997, Sol. Phys., 170, 63 Moffatt, H. K., & Moore, D. W. 1978, J. Fluid Mech., 87, 749 Edwards, J. M. 1990, MNRAS, 242, 224 Nordlund, A,˚ Stein, R. F., & Asplund, M. 2009, Liv. Rev. Sol. Phys., 6, 2 Ertel, H. 1942, Meteorol. Zeitschr., 59, 250 Pleim, J. E. 2007, J. Appl. Met. Climate, 46, 1383 Featherstone, N. A., & Hindman, B. W. 2016, ApJ, 818, 32 Priestley, C. H. B., & Swinbank, W. C. 1947, Proc. Roy. Soc. Lond. A, 189, Freytag, B., Steffen, M., Ludwig, H.-G., Wedemeyer-Bohm,¨ S., 543 Schaffenberger, W., & Steiner, O. 2012, J. Comput. Phys., 231, 919 Rast, M. P. 1998, J. Fluid Mech., 369, 125 Garaud, P., Ogilvie, G. I., Miller, N., & Stellmach, S. 2010, MNRAS, 407, Rempel, M. 2004, ApJ, 607, 1046 2451 Rheinhardt, M., & Brandenburg, A. 2012, Astron. Nachr., 333, 71 Gizon, L., & Birch, A. C. 2005, Liv. Rev. Sol. Phys., 2, 6 Rieutord, M., & Zahn, J.-P. 1995, A&A, 296, 127 Gizon, L., & Birch, A. C. 2012, Proc. Natl. Acad. Sci., 109, 11896 Rieutord, M., Meunier, N., Roudier, T., Rondi, S., Beigbeder, F., & Pares,` L. 2008, A&A, 479, L17 19

Rogachevskii, I., & Kleeorin, N. 2015, J. Plasma Phys., 81, 395810504 Stix, M. 1981, A&A, 93, 339 Roudier, T., Rieutord, M., Malherbe, J. M., Renon, N., Berger, T., Frank, Z., Stix, M. 2002, The Sun: An introduction (Springer-Verlag, Berlin) Prat, V., Gizon, L., & Svanda,ˇ M. 2012, A&A, 540, A88 Stull, R. B. 1984, J. Atmos. Sci., 41, 3351 Rudiger,¨ G. 1989, Differential rotation and stellar convection: Sun and solar- Stull, R. B. 1993, Boundary Layer Meteorology, 62, 21 type stars (Gordon & Breach, New York) Trampedach, R., & Stein, R. F. 2011, ApJ, 731, 78 Rudiger,¨ G., & Spahn, F. 1992, Sol. Phys., 138, 1 Tuominen, I., Brandenburg, A., Moss, D., & Rieutord, M. 1994, A&A, 284, Schou, J., Antia, H. M., Basu, S., Bogart, R. S., et al. 1998, ApJ, 505, 390 259 Serenelli, A. M., Basu, S., Ferguson, J. W., & Asplund, M. 2009, ApJ, 705, Tuominen, I., & Rudiger,¨ G. 1989, A&A, 217, 217 L123 Unno, W., & Spiegel, E. A. 1966, Publ. Astron. Soc. Japan, 18, 85 Spruit, H. C. 1974, Sol. Phys., 34, 277 van Ballegooijen, A. A. 1986, ApJ, 304, 828 Spruit, H. C. 1977, A&A, 55, 151 van Dop, H., & Verver, G. 2001, J. Atmos. Sci., 58, 2240 Spruit, H. 1997, Mem. Soc. Astron. Ital., 68, 397 Vitense, E. 1953, Z. Astrophys., 32, 135 Stein, R. F., Brandenburg, A., Nordlund, A.˚ 1992, in Cool Stars, Stellar Vogler,¨ A., Shelyag, S., Schussler,¨ M., Cattaneo, F., Emonet, T., & Linde, T. Systems, and the Sun, ed. M. S. Giampapa & J. A. Bookbinder (ASP Conf. 2005, A&A, 429, 335 Series, Vol. 26), 148 Woodard, M. 2014, Sol. Phys., 289, 1085 Stein, R. F., Benson, D., & Nordlund, A.˚ 2007, in New Solar Physics with Woodard, M. 2016, MNRAS, 460, 3292 Solar-B Mission, ASP Conf. Ser., Vol. 369, ed. K. Shibata, S. Nagata, & T. Xiong, D. R., & Deng, L. 2001, MNRAS, 327, 1137 Sakurai (San Francisco: Astron. Soc. Pac.), 87 Stein, R. F., & Nordlund, A.˚ 1989, ApJ, 342, L95 Stein, R. F., & Nordlund, A.˚ 1998, ApJ, 499, 914 Stein, R. F., Nordlund, A.,˚ Georgoviani, D., Benson, D., & Schaffenberger, W. 2009, in Solar-Stellar Dynamos as Revealed by Helio- and Astero- seismology: GONG 2008/SOHO 21, ASP Conf. Ser., Vol. 416, ed. M. Dikpati, et al. (San Francisco: Astron. Soc. Pac.), 421

$Header: /var/cvs/brandenb/tex/hydro/EntropyRain/paper.tex,v 1.258 2016/10/21 23:57:32 brandenb Exp $