<<

Convective Differential in and II: Observational and Numerical Tests Adam Jermyn, Shashikumar Chitre, Pierre Lesaffre, Christopher Tout

To cite this version:

Adam Jermyn, Shashikumar Chitre, Pierre Lesaffre, Christopher Tout. Convective Differential Ro- tation in Stars and Planets II: Observational and Numerical Tests. Monthly Notices of the Royal Astronomical Society, Oxford University Press (OUP): Policy P - Oxford Open Option A, 2020, ￿10.1093/mnras/staa2576￿. ￿hal-03021934￿

HAL Id: hal-03021934 https://hal.archives-ouvertes.fr/hal-03021934 Submitted on 24 Nov 2020

HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est archive for the deposit and dissemination of sci- destinée au dépôt et à la de documents entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non, lished or not. The documents may come from émanant des établissements d’enseignement et de teaching and research institutions in France or recherche français ou étrangers, des laboratoires abroad, or from public or private research centers. publics ou privés. MNRAS 000,1–27 (2020) Preprint 24 August 2020 Compiled using MNRAS LATEX style file v3.0

Convective Differential Rotation in Stars and Planets II: Observational and Numerical Tests

Adam S. Jermyn1,2? Shashikumar M. Chitre,2,3 Pierre Lesaffre4 and Christopher A. Tout2 1Center for Computational , Flatiron Institute, New York, New York, 10010, USA 2Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 0HA, UK 3Centre for Basic Sciences, University of Mumbai, India 4Ecole´ Normale Sup´erieure 24 rue Lhomond, 75231 Paris, France

Accepted XXX. Received YYY; in original form ZZZ

ABSTRACT Differential rotation is central to a great many mysteries in stars and planets. In Part I we predicted the order of magnitude and scaling of the differential rotation in both hydrodynamic and magnetohydrodynamic convection zones. Our results apply to both slowly- and rapidly-rotating systems, and provide a general picture of differential rotation in stars and fluid planets. We further calculated the scalings of the meridional circulation, entropy gradient and baroclinicity. In this companion paper we compare these predictions with a variety of observations and numerical simulations. With a few exceptions we find that these are consistent in both the slowly-rotating and rapidly- rotating limits. Our results help to localize core-envelope shear in stars, suggest a rotation-dependent frequency shift in the internal waves of massive stars and potentially explain observed deviations from von Zeipel’s gravity darkening in late-type stars. Key words: convection - : rotation - stars: rotation - stars: evolution - stars: interiors

1 INTRODUCTION rotation profile (Thompson 1991; Schou et al. 1998). In addi- tion, some compact objects now provide strong constraints Driven largely by high-cadence precision photometry from on the rotation profiles of their progenitors. For instance, the CoRoT (Roxburgh & COROT Team 1999), Ke- gravitational waves allow the spins of merging bi- pler (Gilliland et al. 2010) and TESS (Ricker et al. 2015) naries to be measured (Kushnir et al. 2016; The LIGO Scien- missions, the ability of observations to reveal tific Collaboration et al. 2018; Zackay et al. 2019). Assuming has increased dramatically in recent years. Asteroseismology no significant spin changes owing to accretion or supernovae, now enables strong constraints to be placed on radial rota- such measurements then indicate the angular momenta of tion profiles (Aerts et al. 2019), revealing large differential the cores of the progenitor stars. Similarly, photometry of rotation in both red giant (Beck et al. 2012) and Sun-like white dwarfs provides rotation rates and hence constrains stars (Ouazzani et al. 2019). Similarly, spot timing has the spins of cores of lower-mass stars (Hermes et al. 2017). provided measurements of latitudinal shear at the surfaces of stars (Donahue et al. 1996; Reinhold & Gizon 2015; Lurie The growing diversity and depth of observations makes arXiv:2008.09126v1 [astro-ph.SR] 20 Aug 2020 et al. 2017). the distribution of a key theoretical Spectral deconvolution (Donati & Collier Cameron question which has driven the development of mean field 1997) and spectropolarimetry (Reiners & Schmitt 2003) also turbulence theories (Kitchatinov & Ruediger 1995), provide a handle on latitudinal differential rotation and, be- balance arguments (Brun et al. 2010; Balbus et al. cause these do not require the presence of spots, they can 2012) and numerical simulations (K¨apyl¨a 2011; Miesch & be applied to a wider range of stars. Importantly, the use Toomre 2009). Importantly, rotation and differential rota- of spectra rather than spots also removes uncertainty in the tion play an active role in the structure and evolution of spot . stars. For instance by inducing mixing (Eddington 1929; The breadth of these observations of other stars comple- Maeder 1998; Chaboyer & Zahn 1992) and generating mag- ments the depth of those of the Sun, which reveal its detailed netic fields (Spruit 2002) and activity (Wright & Drake 2016). Previously (Jermyn et al. 2020, hereinafter Paper I) we ? E-mail: [email protected] provided predictions for the magnitude of differential rota-

© 2020 The Authors 2 Adam S. Jermyn

Meridional Ω Convection Circulation

R uθ vc, N r ur ∂RΩ θ z ξ p vA ∂zΩ s

Baroclinicity Magnetism

Figure 1. The rotation, coordinate system and differential ro- Figure 2. Key concepts in our theory are shown schematically: tation are shown schematically: (top) the mean angular veloc- (upper-right) the meridional circulation components ur and uθ ; ity Ω; (upper-right) the cylindrical radius R, vertical direction (upper-left) turbulent eddies move at the convection speed 3c along the rotation axis z, spherical radius r and polar angle θ; and with the Brunt-V¨ ¨ais¨al¨a frequency N; (lower-left) surfaces of (upper-left) an example of cylindrical radial differential rotation constant and entropy meet at an angle of approximately the baroclinicity ξ, resulting in surface variations; (∂R Ω); (lower) an example of cylindrical vertical differential ro- (lower-right) a magnetic field with Alfv´envelocity 3 . tation (∂z Ω). A

2 OVERVIEW Rotation breaks spherical symmetry. This means that rotat- ing stars may be very different from non-rotating stars. In Paper I we studied the response of convection zones to both tion in the convection zones of stars and gaseous planets. slow and rapid rotation. We focused on systems with effi- These predictions are order-of-magnitude scaling relations cient convection zones, meaning those in which most of the based on considerations of the asymptotic scalings of dif- is transported by and very little by diffusion. ferent physical processes. Along the way we also predicted In such systems we expect the microscopic thermal diffusiv- the scaling of other quantities such as the baroclinicity and ity, composition diffusivity and viscosity to all be irrelevant, meridional circulation. In this companion paper we show which allows us to specify the problem fully by giving just that our predictions are generally in good agreemnt with a the geometry of the , the total angular mo- variety of different observations and numerical simulations. mentum and the profile of the Brunt-V¨ ¨ais¨al¨a frequency The greatest disagreements are with the simulations, which γ − 1 arise primarily when these are highly diffusive, highlighting N2 = − g · ∇s, (1) the importance of developed turbulence in angular momen- γ tum transport. where g is the acceleration due to gravity, which need not In the next section we define our notation and review be radial, s is the dimensionless entropy1 the key assumptions and results of Paper I. We then proceed 1 to compare our predictions to observations of both radial s = (ln P − γ ln ρ), (2) and latitudinal shear in low mass, solar-type and red giant γ − 1 stars, as well as measurements of latitudinal shear in where P is the pressure, ρ is the , and γ is the adia- (section3). We also compare our predictions of baro- batic index. clinicity with measurements of the solar latitudinal temper- Unfortunately, turbulence is difficult to analyze. Be- ature gradient. Following this we turn to differential rotation cause of this we just attempted to understand the orders in hydrodynamic and MHD simulations in section4, where of magnitude and scaling relations involved in rotating con- we also examine related quantities such as the baroclinicity, vection. With this, the problem simplifies to one specified convection speed and, for MHD simulations, the magnetic just by the geometry, the ratio of the angular velocity Ω to energy density. In section5 we describe further tests which the magnitude of the Brunt-V¨ ¨ais¨al¨a frequency |N|, and the could be done given more observations and numerical sim- ulations. We conclude with a discussion of the results and their astrophysical implications in section7. 1 See Appendix B of Paper I.

MNRAS 000,1–27 (2020) Convective Differential Rotation II 3 ratio of the pressure scale height as slowly as possible given these constraints. The resulting 2 scaling laws indicate that the Taylor-Proudman term ∂Ω h ≡ |∇ ln P|−1 (3) ∂z balances azimuthal turbulent stresses and thereby sets the to the spherical radius r. In the same spirit we further as- differential rotation, and that advection of angular momen- sume that the precise geometry does not matter, and treat tum balances meridional turbulent stresses to set the merid- spherical shells and fully spherical convection zones alike. ional circulation. Our analysis began in the non-rotating limit, which on To summarize, our predictions were derived under a few average must be spherically symmetric. In that limit the con- assumptions. vection zone is parameterized by the scale height h and the (i) Dimensionless factors arising from geometry are of or- Brunt-V¨ ¨ais¨al¨a frequency |N|. From these we then expect to der unity unless symmetries require them to be otherwise. determine the convection speed 3 , which is the root-mean- c (ii) All external perturbing forces, such as tides or exter- square of the difference between the velocity of a fluid ele- nal heating, are negligible in the regions of interest. ment and its mean velocity, as well as the Alfv´en velocity (iii) The material is non-degenerate, compressible and not B -dominated. 3A ≡ √ , (4) 4πρ (iv) All microscopic (i.e. non-turbulent) diffusivities are negligible, such that where B is the root-mean-squared magnetic field strength. These quantities are shown in the upper-left and lower-right (a) convection is efficient, so the gas is nearly isen- panels of Fig.2. In ordinary mixing length theory for efficient tropic, convection 3c ≈ h|N|, and 3A is usually expected to be of the (b) the Reynolds and Rayleigh numbers are much same order (Cantiello & Braithwaite 2019), so we assume larger than critical, and both of these hold in the non-rotating limit. (c) magnetohydrodynamical processes are ideal. We next turned to the limit of slow rotation (Ω  |N|). In this limit rotation breaks spherical symmetry. This allows (v) The system is axisymmetric in a time-averaged sense. there to be steady meridional circulation currents, differen- (vi) Convection is subsonic. tial rotation, and misaligned entropy and pressure gradients, (vii) The system is chemically homogeneous. shown in Fig.1 and the upper-right and lower-left panels of (viii) The pressure scale height h is less than or of the Fig.2. For convenience in working with these quantities we same order as the radius r. defined the baroclinicity Most of these predictions hold in all of the systems with which we test our predictions. The two exceptions are the eφ · (∇ ln P × ∇s) ξ ≡ , (5) assumption that convection is subsonic, which is marginal |∇ ln P||∇s| in systems with shallow surface convection zones, and that which is directly proportional to the thermal wind term in hlar, which fails to hold in near the centres of core- the equation (Part 1 equation 3) and, when small, convecting stars. However, when these fail to hold they usu- is approximately the angle between the pressure and entropy ally do so over a small volume, so we do not expect either gradients. of them to prevent us from comparing our predictions with To determine how each of these quantities scales with Ω observations or numerical simulations. we used symmetry arguments to constrain the possible scal- ings of the different components of the turbulent Reynolds . So, for instance, because mapping Ω → −Ω is equiva- 3 OBSERVATIONAL TESTS lent to mapping the azimuthal coordinate φ → −φ, we know that the Reynolds stress component Trθ must be an even In Paper I we studied differential rotation in both slowly and function of Ω because it is even under φ → −φ. If the stress is rapidly rotating convection zones in both the hydrodynamic analytic in a region around Ω = 0 this means that Trθ scales and magnetohydrodynamic (MHD) limits. Our results are at least as fast as Ω2. We then assumed that each quan- given in Table1. In the limit of slow rotation the scaling is tity scales at the lowest allowed order and obtained scalings the same for both hydrodynamic and MHD convection so for each of the quantities shown in Figs.1 and2, including those cases are grouped together. the shear and baroclinicity. These scalings suggest, that in It is important to emphasize that we have only pre- steady state, the thermal wind term, mechanical forcing by dicted the scalings of these various quantities but not their turbulent stress, and losses owing to turbulent viscosity are actual magnitudes. We expect that in each case there are all of the same order and serve to set the shear. likely factors of order unity in these relations, such that for α In the limit of rapid rotation we then assumed that there a quantity Q scaling as (Ω/|N|0) we have are no remaining symmetries, so that all components of the  Ω β stress and turbulent diffusivity and so on are of the same or- Q = λ Q , (6) Q 0 | | der. We supplemented this with the scaling laws of Stevenson N 0 (1979) for the convection speed and Christensen & Aubert for some λQ of order unity and dimensional Q0 which gives (2006) for the Alfv´envelocity. Importantly these scalings the characteristic scale in terms of h and |N|0. For clarity imply that the Brunt-V¨ ¨ais¨al¨a frequency depends on the ro- we have non-dimensionalized all quantities appearing in Ta- tation rate in this limit, so we denote the non-rotating fre- ble1, so that in each case Q0 = 1. quency as |N|0. To compute Ω/|N|0 from observations we used given We were then able to determine the lowest possible or- values where this quantity was provided directly, and in der of each quantity in Ω−1, and assumed that each declines all other cases we first computed stellar models to match

MNRAS 000,1–27 (2020) 4 Adam S. Jermyn

Table 1. The scalings of the differential rotation, meridional circulation, baroclinicity, Brunt-V¨ ¨ais¨al¨a frequency, convective velocity, and the ratio of magnetic to kinetic energy are given for the three regimes of interest in terms of the non-rotating Brunt-V¨ ¨ais¨al¨a frequency |N |0.

|R∇Ω| |R∂R Ω| |R∂z Ω| |r ∂r Ω| |∂θ Ω| Case Ω Ω Ω Ω Ω

Slow (Ω  |N |0) 1 1 1 1 1 −3/5 −3/5 −6/5 −3/5 −3/5  Ω   Ω   Ω   Ω   Ω  Fast Hydro.(Ω  |N |0) | N |0 | N |0 | N |0 | N |0 | N |0 −3/4 −3/4 −3/2 −3/4 −3/4  Ω   Ω   Ω   Ω   Ω  Fast MHD (Ω  |N |0) | N |0 | N |0 | N |0 | N |0 | N |0

32 ur uθ | N | 3c A Case h| N | h| N | ξ | N | h| N | 2 0 0 0 0 3c 2 2 2 h  Ω   Ω   Ω  Slow (Ω  |N |0) 1 1 1 r | N |0 | N |0 | N | −7/5 −7/5 2/5 −1/5 h  Ω   Ω   Ω   Ω  Fast Hydro.(Ω  |N |0) 1 N/A r | N |0 | N |0 | N |0 | N |0 −1/2 −1/2 1/4 −1/2 3/2 h  Ω   Ω   Ω   Ω   Ω  Fast MHD (Ω  |N |0) 1 r | N |0 | N |0 | N |0 | N |0 | N |0 the observed characteristics of each star using the Mod- eral these measurements should be interpreted as indicating ules for Experiments in Stellar Astrophysics (MESA Paxton the average shear over the surface of a star, though different et al. 2011, 2013, 2015, 2018; Paxton et al. 2019) software techniques are sensitive to different . For instance instrument. We then constructed a volume-weighted average measurements are most sensitive to shear over typi- Brunt-V¨ ¨ais¨al¨a frequency given as cal active latitudes. While we have made an attempt to stan- dardize these data, we urge the reader to see AppendixA ∫ ln |N | dr max(h,r) for details on how data from each source were interpreted. |N| ≈ |N|avg ≡ exp © ª . (7) 0 ­ ∫ dr ® With the exception of contraints from the Juno mission, max(h,r) « ¬ the top panel mostly probes the limits of slow and moder- A geometric average was chosen because the rotation profile ate rotation. While there are offsets between measurements we predict varies as a power-law in |N|. The weighting of made with different techniques, there is general agreement dr/max(h, r) was chosen because the shear we calculate is per that the shear is of order unity for slow rotators and starts to unit ln r, but near the cores of stars r < h and the convective decline roughly as predicted as the rotation rate increases. cells become large relative to r. In that limit h becomes the The data from Benomar et al.(2018) are notable outliers relevant length-scale, so we weight by h instead of r. which lie well above measurements of objects at comparable As described in AppendixA, equation (7) is the form we rotation rates but otherwise the data seem generally consis- have used whenever a volume average of |N| was needed. In tent both internally and with our predictions. AppendixC we study the sensitivity of our analysis to this The Juno data provide our strongest check on the averaging and find that, while there is a potentially large rapidly-rotating limit (Kaspi et al. 2018). The surface data systematic offset in |N|0 depending on the averaging used, are more consistent with the hydrodynamic regime, while 2 up to a factor of 10 , the shape of the data are robust to the upper limits from deeper regions are more consistent different choices of averaging. An additional factor of order with the MHD regime. This is in agreement with previous 2 uncertainty arises from uncertainties in matching stellar (Kirk & Stevenson 1987; Liu et al. 2008) and the find- models to the observations, and particularly in uncertainties ing that the transition between shallow and deep differential in the mixing length parameter. These effects are discussed rotation in Jupiter is associated with a large change in elec- in more detail in AppendixD. trical conductivity (Guillot et al. 2018). Nonetheless, even In what follows we endeavour to test the relations in Ta- for these rapid rotation rates the difference between these ble1 through many different means. We use data from dif- scalings is only a factor of a few and this makes definitive ferent sources and so must standardize them. Details of how statements about different regimes difficult. the data from various sources were obtained, standardized The lower panel of Fig.3 shows starspot measurements and processed are provided in AppendixA, and all scripts of latitudinal shear. The data appear consistent with our ex- and files needed to reproduce this analysis are available on pected scaling relations. In particular the slopes of the data https://zenodo.org/record/3992228. both of Reinhold & Gizon(2015) and Lurie et al.(2017) roughly match our predictions, though with considerable scatter which makes it difficult to say for certain, and the 3.1 Latitudinal Shear three points from other sources (Frasca et al. 2011; Bonanno We begin by considering observations of latitudinal shear et al. 2014; Davenport et al. 2015) scale similarly. There ap- |∂θ ln Ω|, shown in Fig.3 alongside our predictions for the pears to be an overall offset of a factor of 0.1 to 0.3 in the MHD and hydrodynamic limits where relevant. The top magnitude of the shear between the data and our predic- panel shows observations from a variety of sources includ- tions, so we suggest using our scaling relations with a pref- ing asteroseismology but excluding starspot measurements, actor of β|R∇Ω| ≈ 0.2. This is consistent with the offset we while the bottom panel shows only starspot data. In gen- see with the Sun, though the data of Benomar et al.(2018)

MNRAS 000,1–27 (2020) Convective Differential Rotation II 5

Latitudinal Observations predictions do not depend on the depth of the convection 101 zone that does not mean that there is no such dependence. A further test is provided by the empirical scaling rela- tions of Balona & Abedigamba(2016) who find that 100 −n ∂θ ln Ω ∝ Ω , (8)

10 1 where n = 1.1 for K, 0.8 for G and 0.6 for F stars. Their

| sample is comprised mostly of stars with rotation periods n l 10 2 of order 10 d. Except perhaps very near the surface, this | is generally more rapid than the convective turnover time Solar, Antia & Basu (2008) Jupiter (Surface, Juno) in these stars, so they are primarily in the rapid regime in 10 3 Prediction (MHD) which, for ionized systems, we predict n = −3/4 = −0.75, Prediction (Hydrodynamic) which falls in the middle of the observed range. Jupiter (Deep, Juno) Benomar et al. (2018) The trend of increasing n, and hence increasing relative 10 4 Bazot et al. (2019) shear, as stellar temperature rises could be due to a combi- Donati et al. (2008) Ammler-von Eiff & Reiners (2012) nation of the shallowing convection zone and an increasing Brunt-V¨ ¨ais¨al¨a frequency. As this happens the mean convec- 2 0 2 4 6 10 10 10 10 10 tive become faster and the stars approach the slow /|N|0 regime, where we predict n = 0. So it seems likely that the Latitudinal Spot Observations trend in n is a result of the transition from one regime to 100 Prediction (MHD) the other. Bonanno et al. (2014) Frasca et al. (2011) Reinhold & Gizon (2015) Lurie et al. (2017) 1 10 Davenport et al. (2015) 3.2 Radial Shear

2 We now turn to the (spherical) radial shear |r∂r ln Ω|. In | 10

n this case the observations come from a combination of he- l

| lioseismology, asteroseismology and gravity multipole mea- surements of Jupiter (Fig.4). The asteroseismology mea- 3 10 surements are sensitive to the the total shear across the convection zone and so we have converted this into an av- erage shear and plotted them at the average |N|0 computed 10 4 by equation (7). The helioseismic and Juno results are plot- ted with profiles of |N|0 from models of the Sun and Jupiter described in Appendices A5 and A6. 10 2 100 102 104 106 As before, the fastest rotator for which we have data /|N|0 is Jupiter (Kaspi et al. 2018) and they show a preference in the outer layers for the hydrodynamic regime and in the Figure 3. The relative latitudinal shear |∂θ ln Ω| is shown as inner layers for the MHD regime, though in both cases the a function of Ω/|N |0 for observed convecting stars and Jupiter preference is weak. In the slowly rotating regime the solar alongside our prediction, which is constant for Ω < |N |0 and data provide our main test and are broadly consistent with −3/4 −3/5 scales as (Ω/|N |0) (MHD) and (Ω/|N |0) (hydrodynamic) our prediction of no significant scaling (Antia & Basu 2001). | | for Ω > N 0. Shapes indicate the origin of the data: squares are The available asteroseismic measurements are mostly for asteroseismology, triangles are for spectroscopy, stars are for for slowly-rotating or moderately-rotating systems which spectropolarimetry and circles are for star spot measurements. For clarity the star spot measurements are shown separately in lie close enough to the predicted break in the power-law the lower panel. Helioseismic data and constraints from Juno are (Ω ≈ |N|) that it is not clear which regime they probe. shown as lines. Juno measurements of the shear at a depth Nonetheless they are broadly consistent with our expecta- of less than about 3000km are shown separately from Juno upper tions and show slight evidence of a trend of decreasing rel- limits on that at greater depths. ative shear with increasing rotation rate. Note that some caution is warranted in the interpretation of these data be- cause the asteroseismic measurements of differential rotation probe a part of the star which often includes a stably strati- and Ammler-von Eiff & Reiners(2012) suggest no such off- fied region. There is thus some uncertainty in attributing the set. This could be due to differences in which regions of the shear purely to the convection zone, though that is what we star are probed by different techniques or it could represent have done here. systematic differences of which we are unaware in either the The solar data as well as the bounds set by Nielsen et al. different measurement techniques or the objects observed. (2017) and Deheuvels et al.(2015) suggest that the overall For instance different samples may contain different frac- scale of our predictions is too large by a factor of 3 or so. This tions of fully convective stars, stars with deep convection agrees with the offset we saw in the starspot measurements zones, and others with shallow convection zones. While our so we think it likely that the scale β|R∇Ω| ≈ 0.3.

MNRAS 000,1–27 (2020) 6 Adam S. Jermyn

Radial Shear Baroclinic Angle

1 10 101

100 100 10 1

2 10 1 10

| 10 3 n l r 2 r 10 | 10 4

10 5 10 3

10 6

10 2 100 102 104 106 10 4 /|N|0

Prediction Solar, Altrock & Canfield 1972

10 2 100 102 104 106

/|N|0

Solar, Antia & Basu (2008) Jupiter (Deep, Juno) Jupiter (Surface, Juno) Deheuvels et al. (2015) Prediction (MHD) Klion & Quataert (2017) Prediction (Hydrodynamic) Nielsen et al. (2017) Figure 5. The angle ξ between the and the en- tropy gradient is shown from our predictions. It was also inferred for the Sun from the equator-pole temperature difference and the resulting upper bound is shown with a circle for the surface and a triangle for the inferred average (Altrock & Canfield 1972). Our 2 prediction scales as (Ω/|N |0) for slow rotation (Ω < |N |0) and is Figure 4. The relative radial shear |r∂r ln Ω| ≈ |R∇ ln Ω| is shown constant for rapid rotation (Ω > |N |0). as a function of Ω/|N |0 for observed convecting stars and Jupiter alongside our prediction, which is constant for Ω < |N |0 and (Ω/| | )−3/4 (Ω/| | )−3/5 scales as N 0 (MHD) and N 0 (hydrodynamic) field(1972), who found a temperature difference of 1.5 ± for Ω > |N | . Squares denote asteroseismic results. Helioseismic 0 0.6 K. We have translated done this with two different meth- data and constraints from Juno are shown as solid lines. Juno measurements of the shear at a depth of less than about 3000km ods described in Appendix A5. In the first approach we as- are shown separately from Juno upper limits on that at greater sumed that ξ is determined by the local properties of the depths. convection zone, so we have plotted the resulting data at the local Ω/|N|0 near the surface of the solar convection zone. In the second approach we assumed that the relative change in temperature between the pole and the equator persists 3.3 Baroclinicity throughout the convection zone and computed a volume- In order to compute the differential rotation in convection weighted average ξ and an average Ω/|N|0, with the average zones we also needed to compute the angle ξ between the Brunt-V¨ ¨ais¨al¨a frequency given by equation (7). pressure and entropy gradients. This baroclinic angle is im- Bearing in mind the large uncertainties on this measure- 2 portant because it determines the importance of the thermal ment, we find good agreement with our predicted Ω scaling wind term ∇p×∇ρ in the . It is also closely for slow rotators. The solar data do not probe the rapidly related to the temperature difference between the pole and rotating limit so for that we rely on simulations in Sec- the equator of a star because it is proportional to the tem- tion 4.3. The scaling we find suggests that the baroclinic an- perature gradient along isobars. Because of this, we may gle in convection zones is indeed driven by rotation-induced compare the observed temperature difference on the Sun to anisotropy in the convective heat flux (Jermyn et al. 2018b). our predictions for ξ. Unfortunately the pole-equator temperature difference 3.4 Magnetic Activity on the Sun has proven difficult to measure due to signifi- cant observational uncertainties and the small expected sig- Our prediction is that magnetic field strength is indepen- nal (Takeda & UeNo 2017). Rast et al.(2008) summarizes dent of rotation rate for Ω < |N| and increases with increas- a variety of efforts, most of which either conclude that the ing rotation rate for Ω > |N|. This is somewhat at odds difference is zero to within uncertainties or provided a figure with observations of magnetic activity, which show increas- of order 1.5 K. ing X-ray luminosity with increasing rotation rate for slowly Bearing in mind these uncertainties, in Fig.5 we have rotating stars (Ω < |N|) and a plateau for rapidly rotating inferred ξ from the measurements made by Altrock & Can- stars (Ω > |N|, Wright et al. 2011). This poses two chal-

MNRAS 000,1–27 (2020) Convective Differential Rotation II 7 lenges. First, if our prediction is correct for slowly-rotating Setting F to be constant then yields |N| ∝ Ω2/5. The analysis stars, how can tracers of magnetic activity such as X-ray is similar in magnetized convection zones, but begins with luminosity increase without the magnetic field strength in- F ≈ ρ3 32 . (11) creasing? Secondly, if our prediction is correct for rapidly- c A rotating stars, how can tracers of activity plateau while the Inserting our scaling for 3c and the scalings 3A ≈ h|N| yields magnetic field strength increases? 3 4 −1 We think the first challenge is most likely resolved by F ≈ ρh |N| Ω (12) changes in the geometry of the magnetic field as a function and so |N| ∝ Ω1/4, as given in Table1. of rotation rate. In numerical simulations of convective dy- Our predictions in terms of |N| are given in Table2. namos the magnetic field becomes increasingly ordered on Once more we emphasize that we have only predicted the large scales as the rotation rate increases (see e.g. Chris- scalings of these various quantities but not their actual mag- tensen & Aubert 2006). If activity relies on the magnetic nitudes. We expect that in each case there are likely factors field having dipolar or low-order multipolar structure then of order unity in these relations, such that for a quantity Q the activity can increase with rotation rate even if the mag- scaling as (Ω/|N|)α we have netic field strength is constant, as we predict for slowly- rotating stars. This is supported by observations which sug-  Ω β Q = λ Q (13) gest that the X-ray luminosity increases with increasing Q 0 |N| large-scale magnetic field (Vidotto et al. 2014, figure 6). The second challenge could be explained in a few dif- for some λQ of order unity and dimensional Q0 which gives ferent ways. First, surface convection zones are often ineffi- the characteristic scale in terms of h and |N|. For clarity cient, meaning that most of the heat flux is carried radia- we have non-dimensionalized all quantities appearing in Ta- tively rather than by advection. As a result increasing the ble1, so that in each case Q0 = 1. rotation rate would not increase |N| as we have predicted We now test the relations in Table2. Details of how in the case of efficient convection zones. Because we pre- the data from various sources were obtained, standardized and processed are provided in AppendixA, and all scripts dict that 3A ≈ h|N| this suggests that, even when Ω  |N|, the surface magnetic field strength does not increase with and files needed to reproduce this analysis are available on increasing rotation rate. The entire effect of increasing ac- Zenodo.org. tivity with increasing rotation rate would then be due to the field geometry becoming dipolar. This eventually saturates around Ω ≈ |N|, producing a plateau. Note that this expla- 4.1 Hydrodynamic Simulations nation requires that the surface magnetic field be generated Because simulations tend to probe different regions of pa- near the surface, such that it is sensitive to the properties rameter space they provide a complementary set of tests. In of near-surface convection. this case simulations tend to fill in the slowly-rotating and A further possibility is that the magnetic activity de- moderately-rotating limits more than the observations and pends primarily on a spot coverage fraction which saturates this allows us to test whether the scaling indeed changes and so no further increase in field strength can produce more between these limits. activity (Vilhu 1984). Testing such a proposition is difficult, The results of a variety of different hydrodynamic con- but if it holds it would resolve the apparent conflict between vection simulations performed for different geometries, with our predictions and the observations. different software instruments, in both the anelastic2 and fully compressible limits, and with different choices of di- mensionless parameters are shown in Fig.6. Squares indicate 4 NUMERICAL SIMULATIONS radial shear, circles indicate latitudinal shear and triangles indicate the root-mean square of the shear integrated over We now turn to numerical simulations. Because these typi- the domain. Filled shapes indicate simulations performed cally report the , we can directly compute on deep spherical shell domains3, which are meant to mimic |N| and so we rephrase our scaling relations in terms of the full spheres, while open shapes indicate those performed in Brunt-V¨ ¨ais¨al¨a frequency which is actually realized in the thinner shellular domains. system, rather than the non-rotating one we used in sec- There is significant scatter between simulations per- tion3. The reason that |N| is not the same as |N| is that 0 formed by different groups. Some of this may result from rotation generally makes convection less efficient at trans- the use of different software instruments, though several of porting heat. This means that the entropy gradient has to the instruments used do yield identical results on identical be steeper for convection to carry the same heat flux in a rotating system than in a non-rotating one. The relation between |N| and |N|0 is derived in detail 2 The anelastic approximation is the limit as the in Paper I. Briefly, in non-magnetized convection zones, we and both go to zero, meaning that characteristic compute the heat flux as velocities are much less than both the sound speed and the free- ρcpT fall speed across the domain (Masmoudi 2007). This allows the F ≈ D|∇s|, (9) pressure and gravity terms in the Navier-Stokes equation to be µ written in terms of an entropy perturbation away from a back- 2 where D ≈ 3c /|N| is the turbulent diffusivity. Inserting our ground state, and imposes the constraint that ∇ · (ρ¯v) = 0, where 3/2 −1/2 predicted scaling for 3c ∝ |N| Ω then yields ρ¯ is the background density profile. 3 3 5 −2 These are domains which just exclude a small region in the F ≈ ρh |N| Ω . (10) centre of the sphere for numerical or algorithmic reasons.

MNRAS 000,1–27 (2020) 8 Adam S. Jermyn

Table 2. The scalings of the differential rotation, meridional circulation, baroclinicity, convective velocity, and the ratio of magnetic to kinetic energy are given for the three regimes of interest. Note that the latitudinal and spherical radial differential rotation are each formed of a of the cylindrical vertical and radial differential rotation. Because the cylindrical radial shear is larger than the vertical shear, both spherical components of the differential rotation share the scaling of the former.

|R∇Ω| |R∂R Ω| |R∂z Ω| |r ∂r Ω| |∂θ Ω| Case Ω Ω Ω Ω Ω

Slow (Ω  |N |0) 1 1 1 1 1  −1  −1  −2  −1  −1 Ω  | | Ω Ω Ω Ω Ω Fast Hydro.( N ) | N | | N | | N | | N | | N |  −1  −1  −2  −1  −1 Ω  | | Ω Ω Ω Ω Ω Fast MHD ( N ) | N | | N | | N | | N | | N |

32 ur uθ 3c A Case h| N | h| N | ξ h| N | 2 3c  2  2  2 Ω  | | h Ω Ω Ω Slow ( N ) r | N | | N | | N | 1 1  −3  −3 Ω  | | h Ω Ω | N | Fast Hydro.( N ) r | N | | N | 1 Ω N/A  −1  −1  −1 2 Fast MHD (Ω  |N |) h Ω Ω ≈ Ω 1 | N | Ω r | N | | N | | N | Ω | N |2

HD Simulations HD Simulations 105 101 101 100 104 100 10 1 3 | 10 10 1 10 2 n e l P 3 R

| 10 2 2 10 | 10 4 n 10 l 3 1 R

| 10 10 10 5

6 4 10 10 100 10 2 100 102 104 106 10 5 /|N|

2 10 6 Pe > 10 105 101 10 2 100 102 104 106 100 104 /|N| 10 1 3 Prediction Gastine et al. (2012) | 10 10 2 n e

Augustson et al. (2012) Gastine et al. (2013) l P 3 Augustson et al. (2013) Gilman (1979-79) R | 10 102 Augustson et al. (2016) Guerrero et al. (2013) 4 Aurnou et al. (2007) Käpylä et al. (2011) 10 101 Brown et al. (2008) Kaspi et al. (2009) 10 5 Browning et al. (2004) Mabuchi et al. (2015) 6 Brun et al. (2009) Matt et al. (2011) 10 100 Brun et al. (2017) Rogers (2015) 10 2 100 102 104 106 Featherstone et al. (2015) /|N|

Figure 6. The relative differential rotation |R∇ ln Ω| is shown Figure 7. The relative differential rotation |R∇ ln Ω| is shown alongside our prediction as a function of Ω/|N | for a variety alongside our prediction as a function of Ω/|N | for a variety of of hydrodynamic convection simulations. Squares indicate radial hydrodynamic convection simulations, coloured by P´eclet number shear, circles indicate latitudinal shear and triangles indicate the Pe. The upper panel shows data from all simulations for which we root-mean square of the shear integrated over the domain. Filled could compute Pe while the lower shows just those simulations shapes indicate simulations performed on deep spherical shell do- with Pe > 102. The slopes of the power-law predictions are indi- mains while open shapes indicate those performed in shellular cated in the relevant regimes. domains. Our prediction is constant for slow rotation (Ω < |N |) and scales as (Ω/|N |)−1 for rapid rotation (Ω > |N |).

MNRAS 000,1–27 (2020) Convective Differential Rotation II 9 problems (Jones et al. 2011) It could also be due to imperfec- shear to plateau towards the slow rotation limit and to fall tions in our standardization of the data. In particular, differ- in the rapidly rotating limit. The existence of the plateau ent symbols in Fig.6 correspond to different ways in which suggests that the scaling of the off-diagonal components of the differential rotation was reported, and so reflect differ- the Reynolds stress is indeed as we have argued in Part I, ent kinds of spatial averaging as well as different weightings with the rθ component scaling quadratically and the rφ and of the radial and latitudinal components of ∇Ω. We do not θφ components scaling linearly. see any systematic trends across different reported measures, In the rapidly rotating limit our scaling appears to ap- though, so this also seems unlikely to be the cause of a ma- proximately bound the results of simulations. That is, some jority of the scatter. simulations report shears significantly below our predictions There is also some scatter within the data from indi- while very few report anything significantly above them and vidual sources. Different choices of boundary conditions or the largest shears found in simulations appear to track our geometry can contribute to this scatter (see e.g. Yadav et al. scaling reasonably well. 2016), but it seems likely that a large fraction of it results A second test of our predictions comes from the con- from the use of different Prandtl, Rayleigh or Reynolds num- vective velocities. Fig.8 shows our predictions for the ratio bers, and we expect such differences to become smaller as 3c/h|N| alongside the results of the subset of these simula- the latter two numbers become larger. tions which reported enough information for us to calculate To probe these effects across simulations we use the P´e- this ratio. Note that we have used |N| rather than |N|0 in clet number the denominator to emphasize the suppression relative to 3d the predictions of non-rotating mixing length theory. Pe ≡ , (14) α We see that the data all exhibit the same trend of a plateau towards slow rotation and a decline towards rapid where α is the thermal diffusivity, d is a characteristic length rotation. There is again considerable scatter between simula- scale, 3 is the root-mean square of the velocity in the rotating tions performed by different groups. This is again primarily frame with zero net angular momentum. When the P´eclet due to many of these simulations having a low P´eclet num- number is sufficiently large convection is efficient because the ber. We can see this in Fig.9, which shows the full suite of heat transport comes to be dominated by advection rather simulations as well as cuts with Pe > 102 and Pe > 103. As than diffusion. We choose this number despite it being an we restrict the P´eclet number to be ever-larger the scatter output from simulations rather than an input because it di- reduces and the agreement with our predictions improves. rectly measures whether or not the convection is efficient Note that we tend to overestimate the convection speed and this is a critical factor in determining whether or not relative to the simulations. This is just because we have set our predictions are applicable. all dimensionless constants to unity in our theory. In practice The upper panel of Fig.7 shows the same simulations we can absorb the apparent constant offset into our predic- coloured by P´eclet number4. There is a clear trend towards tions by setting the prefactor λ ≈ 0.2 in our relations. less shear with lower Pe. In the lower panel we retain only 3c Finally, while we have focused on simulations in spheri- those simulations with Pe > 102. Doing so significantly re- cal geometries, our results also agree with the cartesian sim- duces the scatter and improves the agreement with our pre- ulations of Barker et al.(2014), who find, as we do, that dictions. 3 ∝ Ω−1/5 at fixed heat flux. Some of the remaining scatter is due to density strati- c fication. In particular, the different tracks that can be seen in the data of Gastine & Wicht(2012) in Fig.6 correspond to different degrees of density stratification. The strength of this effect decreases with increasing P´eclet number. This 4.2 MHD Simulations could be because, as the density stratification increases, While hydrodynamic systems provide a helpful test of our the velocity required to carry the heat flux becomes large arguments, the most common astrophysical case is the near- near the top of the domain and small near the bottom. So ideal MHD limit. Fig. 10 shows our predicted differential when the mean Pe remains constant, increasing stratification rotation scaling along with the findings of various MHD sim- means that an increasingly large proportion of the domain ulations of convection. As before we include both anelastic has a low Pe. As such more work may be required to probe and fully compressible simulations. We employ the same con- whether this effect holds at the large Pe > 106 that apply in ventions that filled shapes denote spherical domains, open the Sun for instance. shapes denote shellular domains, squares indicate radial If there is an effect of density stratification at large shear, circles indicate latitudinal shear and triangles indicate Pe, it could be due to the compressibility torque described root-mean square of the shear integrated over the domain. by Glatzmaier et al.(2009). Such effects are potentially very physically important if they hold in the astrophysically rel- There is much less scatter between simulations per- formed by different groups than in the case of hydrodynamic evant high-Pe limit particularly because some systems, such as red giants, exhibit extreme degrees of density stratifica- simulations. This is largely because most of our data come tion in their convection zones. from simulations performed with the ASH software instru- ment and that these data have been reported in a relatively Focusing just on the high-Pe number data shown in the uniform way. lower panel of Fig.7, we see an overall trend for the relative There is also less scatter between the simulations in each individual work. We believe this is partly because these 4 For details of how this was calculated for each simulation see simulations were performed using a smaller set of software AppendixA. instruments and partly because they used a more uniform

MNRAS 000,1–27 (2020) 10 Adam S. Jermyn

Convective Velocity (Hydrodynamic) Convective Velocity (Hydrodynamic) 105 0 100 10 104

1 | 10 103 N | e h

1 P 10 / c 2 v 10 2

| 10 N |

h 1 / 10 c v 2 10 10 3 100 10 2 10 1 100 101 102 103 /|N|

10 3 Pe > 102 105 100 4 10 2 10 1 100 101 102 103 10 /|N| 1 | 10 103 N

Prediction Brown et al. (2008) | e h P / Gastine et al. (2012) Brun et al. (2009) c 2 v 10 Gastine et al. (2013) Brun et al. (2017) 10 2 Browning et al. (2004) Featherstone et al. (2015) Augustson et al. (2012) Mabuchi et al. (2015) 101 Augustson et al. (2013) Guerrero et al. (2013) 10 3 Augustson et al. (2016) Käpylä et al. (2011) 100 Matt et al. (2011) 10 2 10 1 100 101 102 103 /|N|

Figure 8. The normalized convection speed 3c/h |N | is shown Pe > 103 105 alongside our predictions as a function of Ω/|N | for a variety 100 of hydrodynamic convection simulations. Filled shapes indicate simulations performed on spherical domains while open shapes 104 indicate those performed in shellular domaints. Triangles denote 10 1 | 103 outputs which were averaged over the simulation domain. Our N | e h P /

prediction is constant for slow rotation (Ω < |N |) and scales as c 2 −1 v 10 (Ω/|N |) for rapid rotation (Ω > |N |). 10 2

101

10 3 100 2 1 0 1 2 3 set of dimensionless parameters5. We can see the latter in a 10 10 10 10 10 10 /|N| plot coloured by P´eclet number, shown in the upper panel of Fig. 11. There are far fewer simulations at very low Pe than in the hydrodynamic sample and this results in less Figure 9. The normalized convection speed 3c/h |N | is shown scatter. When we restrict the sample to just Pe > 102 (lower alongside our predictions as a function of Ω/|N | for a variety of panel) the scatter goes away almost entirely and the remain- hydrodynamic convection simulations, coloured by P´eclet number ing simulations lie nicely on our predicted slope, albeit out Pe. The upper panel shows data from all simulations for which we to slower rotation rates than we expect. could compute Pe, the lower shows just those simulations with Pe > 102, and the lower panel shows just those with Pe > 103. Our Overall we see a steeper slope than our predicted scal- prediction is constant for slow rotation (Ω < |N |) and scales as ing in the rapidly rotating limit. This conclusion is driven (Ω/|N |)−1 for rapid rotation (Ω > |N |). mostly by the results of Yadav et al.(2013a) and Augustson et al.(2016), which suggest that |R∇ ln Ω| ∝ Ω−2 rather than −1 our predicted Ω . This is consistent with the power-law fits flux7, suggesting that |R∇ ln Ω| ∝ Ω−1.328 , Ω−1.449 , Ω−1.210 provided by Yadav et al.(2013a) for the zonal (azimuthal) or Ω−1.2911 . Yadav et al.(2013b) similarly find a scaling of 6 and non-zonal (meridional) . We may trans- Ω−1.35. late into predictions for the differential rotation at fixed heat Such steep power laws appears inconsistent with the ob- servations shown in Figs.3 and4 so we remain uncertain as to whether the slope in nature is truly steeper than our pre-

5 7 ∗ −3 The greater uniformity of dimensionless parameters in these |R∇ ln Ω| ≈ Rozonal. Using RaQ ∝ Ω we then obtain the scaling simulations is likely because the simulation must be fairly tur- for |R∇ ln Ω|. bulent (advection-dominated) to see a dynamo in the first place, 8 Using their fit for dipolar dynamos with no Pm dependence. which effectively sets a floor on the P´eclet number. 9 Using their fit for dipolar dynamos with a Pm dependence. 6 See also Davidson(2013) for more discussion of these scaling 10 Using their fit for multipolar dynamos with no Pm dependence. relations. 11 Using their fit for multipolar dynamos with a Pm dependence.

MNRAS 000,1–27 (2020) Convective Differential Rotation II 11

MHD Simulations MHD Simulations 104 101 101 100 103 100 10 1 | 10 2 n

2 e l 10 1 P 10 3 R

| 10

10 4 2 1 | 10 10 5

n 10 l

R 6 | 10 10 3 100 10 2 100 102 104 106 /|N| 10 4 Pe > 102 104 10 5 101

100 103 10 6 10 1

2 0 2 4 6 | 10 10 10 10 10 10 2 n

2 e l 10 /|N| P 3 R

| 10

10 4 101 Prediction Mabuchi et al. (2015) 10 5 Augustson et al. (2013) Varela et al. (2016) Augustson et al. (2016) Yadav et al. (2013) 10 6 100 Brun et al. (2005) Soderlund et al. (2014) 10 2 100 102 104 106 /|N|

Figure 10. The relative differential rotation |R∇ ln Ω| is shown Figure 11. The relative differential rotation |R∇ ln Ω| is shown alongside our predictions as a function of Ω/|N | for a variety of alongside our predictions as a function of Ω/|N | for a variety of MHD convection simulations. Circles indicate latitudinal shear MHD convection simulations, coloured by P´eclet number Pe. The and triangles indicate root-mean-squared shear integrated over upper panel shows data from all simulations for which we could the domain. Filled shapes indicate simulations performed on compute Pe while the lower shows just those simulations with spherical domains while open shapes indicate those performed in Pe > 102. Our prediction is constant for slow rotation (Ω < |N |) shellular domaints. Our prediction is constant for slow rotation and scales as (Ω/|N |)−1 for rapid rotation (Ω > |N |). (Ω < |N |) and scales as (Ω/|N |)−1 for rapid rotation (Ω > |N |). diction. However, if this is indeed what happens it points to a relative to rotation on average, deep in the core where con- difficulty in our analysis. In Section 7 of Paper I we assumed vection is slow we might still be in the rapidly rotating limit. that each of the shear, circulation and baroclinicity are as A further test of our predictions comes from the con- large as allowed by the conditions of heat and momentum vective velocities. Fig. 12 shows our predictions for the ratio balance. While it must be that in each equation at least one 3c/h|N| alongside the results of the subset of these simula- of these saturates its bounds, it need not be the case that tions for which we were able to calculate this ratio. they all do. If they do not then our predictions in Tables1 We see that the data generally exhibit the same trend and2 are really upper bounds, only one of which must be of a plateau towards slow rotation and a decline towards saturated in any given scenario. So for instance it could be rapid rotation. Most of the scatter in these results is directly that the cylindrical vertical differential rotation ∂zΩ satu- attributable to not being in the limit of efficient convection. − rates our bound and scales like Ω 2, similar to what we see We can see this in the lower panel where we have filtered in Fig. 10. In that case there is no formal requirement that for Pe > 102 and see significantly better agreement with our the radial shear ∂RΩ also saturates its bound, in which case predicted slope and plateau. −2 the shear only needs to be as large as ∂zΩ ∝ Ω . Other authors have found similar, though typically In the slowly rotating regime the simulations of Varela weaker, scalings of MHD convection speed at constant heat −1/2 −0.23 et al.(2016) hint at a plateau but those of Augustson et al. flux. We predict 3c ∝ Ω , while others find Ω to (2016) instead show a continuing increase of relative shear Ω−0.29 (Christensen & Aubert 2006)12, Ω−0.41 (Yadav et al. with decreasing rotation rate. This could indicate that we have misplaced the break point between slow and rapid ro- tation. On the other hand, all of the simulations which show this trend were performed on a spherical domain and so it could be that, even though the convective turnover is fast 12 See their equations (30) and (31).

MNRAS 000,1–27 (2020) 12 Adam S. Jermyn

Convective Velocity (MHD) MHD Simulations 105 0 10 106

104

1 10 3

| 10 4

N 10 | e h P / c v 102 10 2

1 2 2 10 c 10 v / 2 A 10 3 v 100 10 2 10 1 100 101 102 103 /|N| 100

Pe > 102 105 100

10 2 104

10 1 3 | 10 10 2 10 1 100 101 102 103 N | e h P

/ /|N| c v 102 2 10 Prediction Mabuchi et al. (2015) Augustson et al. (2013) Varela et al. (2016) 101 Augustson et al. (2016) Yadav et al. (2013) Brun et al. (2005) 10 3 100 10 2 10 1 100 101 102 103 /|N|

3 / | | 32 /32 Figure 12. The normalized convection speed c h N is shown Figure 13. The ratio A c of magnetic energy to convective alongside our predictions as a function of Ω/|N | for a variety of kinetic energy is shown alongside our predictions as a function of MHD convection simulations, coloured by P´ecletnumber Pe. Our Ω/|N | for a variety of MHD convection simulations. Filled shapes prediction is constant for slow rotation (Ω < |N |) and scales as indicate simulations performed on spherical domains while open (Ω/|N |)−1 for rapid rotation (Ω > |N |). shapes indicate those performed in shellular domaints. All of these data were averaged over the simulation domain. Our prediction is constant for slow rotation (Ω < |N |) and scales as Ω/|N | ≈ 2013a)13 and Ω−0.32 to Ω−0.47 (Aubert et al. 2017)14. This (Ω/|N |)2 for rapid rotation (Ω > |N |). weaker scaling could be a result of the P´eclet number de- pendence we have noted, though of course it could also be that our theory predicts a steeper scaling law than actually occurs. Finally we examine the ratio of magnetic to kinetic en- ergy, shown as a function of Ω/|N| in Fig. 13. Our predic- tions are shown as a solid line and the results of simula- tions are shown as a variety of shapes, following the previ- ous convention that solid shapes indicate spherical domains, open shapes indicate shellular domains, and triangles de- note measurements averaged over the simulation domain. Note that our predictions for this energy ratio are not new: they are identical to those made by several authors includ- ing Starchenko & Jones(2002) and Augustson et al.(2016). There is significant scatter between simulations per- formed by different groups but within the sets of simulations with the most dynamic range in Ω/|N| we see a plateau to- Our predicted scalings are also similar to those that wards slow rotation and scaling similar to that of our pre- found by several groups in fits to their own numerical sim- dictions for rapid rotation. ulations. For instance at fixed heat flux we predict that the 1/4 magnetic field strength scales as B ∝ 3A ∝ Ω and in simu- lations this scaling has been found to be Ω−0.02 (Christensen 15 −0.11 16 0 13 From their multipolar fit with magnetic de- & Aubert 2006) , Ω (Yadav et al. 2013a) , Ω (Yadav − pendence. et al. 2013a)17, and between Ω 0.02 and Ω0.19 (Aubert et al. 14 See their figure 11(a). Their  is proportional to Ω−3. 2017)18.

MNRAS 000,1–27 (2020) Convective Differential Rotation II 13

Baroclinic Angle both slow and rapid rotation, in agreement with the argu- ments of Balbus & Schaan(2012). The scaling we find sug- 100 gests that the baroclinic angle in convection zones is indeed driven by rotation-induced anisotropy in the convective heat 10 1 flux (Jermyn et al. 2018b).

10 2 5 FUTURE TESTS 10 3 We have tested our predictions under many different circum- stances but there are further tests that would be fruitful in 10 4 the future. First, while we have tested our predictions for the latitudinal and spherical radial shear, we have not made a 10 5 quantitative test of our prediction that the cylindrical verti- cal shear ∂zΩ is suppressed relative to the radial component. This is seen in simulations (see e.g. Mabuchi et al. 2015) but 10 6 so far as we are aware it has not been quantified. It ought 10 2 100 102 104 106 to be straightforward to add ∂ Ω and ∂ Ω to the output of /|N| R z future simulations or even to run previous simulations again Prediction Augustson et al. (2013) with these additional outputs. Brun et al. (2017) Matt et al. (2011) Secondly, it would be good to test our predictions for the Augustson et al. (2012) Raynaud et al. (2018) meridional circulation velocities ur and uθ . There have been measurements of the meridional circulation in the Sun (Zhao et al. 2013; Rajaguru & Antia 2015; Schad & Roth 2020) but there are still significant disagreements between the different inversion techniques which make a direct comparison to our Figure 14. The angle ξ between the pressure gradient and the theory challenging. On the other hand, such a comparison entropy gradient is shown for a variety of simulations. Our pre- 2 could be done with simulations by reporting the time- and diction scales as (Ω/|N |0) for slow rotation (Ω < |N |0) and is azimuthal-average of the meridional velocity field. As far as constant for rapid rotation (Ω > |N |0). we are aware this has not been reported 19 but we see no particular barrier to doing so. 4.3 Baroclinicity More broadly, further tests from simulations primarily face the challenge of reaching high enough P´eclet numbers to As described in appendixA, we have inferred ξ from the be applicable to efficient convection. The primary limitation reported results of various hydrodynamic and MHD simu- to this is the availability of computing resources, so we hope lations. Our prediction is the same for both kinds of sim- that more simulations will be performed at Pe ≈ 103 to 105 ulations and so we analyze them together. Some authors in the near future. reported the temperature difference between different lati- Finally, the most useful observational comparisons are tudes in a model while others reported latitudinal gradients those that probe the most rapidly-rotating systems. These of closely related quantities like the . The are the systems which are best able to test the slopes we have results of our inference and standardization are shown in predicted and to thereby either verify or falsify our theories. Fig. 14. In this regard data such as the differential rotation mea- We find reasonable agreement with our predicted Ω2 surements of Davenport et al.(2015) are extremely useful scaling for slow rotators and, as expected, we see ξ satu- because, while they report observations of a single system, rating in the limit of rapid rotation. There is considerable it is well-characterized and extremely rapidly rotating, par- scatter among simulations in the saturated regime and even ticularly relative to its slow convective turnover time. This considerable disagreement amongst simulations performed allows it to more strongly constrain the scaling of differential by the same group with the same code. This may reflect a rotation. sensitivity of the saturated ξ to the geometry or other di- mensionless parameters in the simulation. These data support our conclusion in Part I that the thermal wind term is one of the dominant terms in both 6 PRIOR PREDICTIONS the hydrodynamic and MHD regimes and in the limits of In Part I we made predictions for the scaling of the differen- tial rotation, meridional circulation, magnetic field and baro- 15 ∗ See their equation (33). Note that Lo ∝ 3A/Ω and their RaQ ∝ clinic angle. We are aware that two of our predicted scalings −3 ∗ ( 1−0.34×3 Ω , such that Lo ∝ (RaQ ) 0.34) means that 3A ∝ Ω . 16 From their dipolar fit with no de- pendence. 19 Some authors (e.g. Augustson et al. 2012) report the root-mean 17 From their dipolar fit with magnetic Prandtl number depen- square of the meridional circulation. Despite the similar name dence. this is not the quantity we predict. By squaring before taking 18 See their figure 11(a). Their  is proportional to Ω−3 and their a time-average they are measuring the strength of axisymmetric λ ∝ BΩ−1. convective motions rather than the slow mean meridional flow.

MNRAS 000,1–27 (2020) 14 Adam S. Jermyn have been suggested previously. These are that for the mag- cations. We have predicted that the relative shear is of or- netic field (Starchenko & Jones 2002) and that of convection der unity for slowly rotating convection zones and falls off speed with rotation rate the hydrodynamic limit (Stevenson quickly once the rotation becomes faster than the convec- 1979). The remaining scalings are new. tive turnover frequency. In particular, the relative shear per However, there have been many previous predictions of unit log r is at most of order |N|/Ω in the rapidly-rotating the same quantities and it is worth comparing our predic- limit. A consequence of this is that the shear is approxi- tions with these. Our prediction that |R∇ ln Ω| is a constant mately bounded by the Brunt-V¨ ¨ais¨al¨a frequency, which is of of order unity for slowly-rotating systems is consistent with order the convective turnover frequency. This is because if predictions based on the Λ effect (e.g.K uker¨ et al. 2019, Ω > |N| then the absolute shear per unit log r is |N| while if and references therein). Gilman & Foukal(1979) predicted Ω < |N| it is Ω, which is less than Brunt-V¨ ¨ais¨al¨a frequency that for slowly-rotating systems the angular momentum of |N|. Hence we expect the maximum absolute shear across a fluid parcels is conserved in convection and so that Ω ∝ r−2, convection zone to be or |R∇ ln Ω| = 2, consistent with our predictions. Kissin & ¹ Thompson(2015a) make the same prediction of Ω ∝ r−2 ∆Ωmax ≈ |N|d ln r. (15) for slowly-rotating systems and provide two possible scal- convection zone ings for rapidly-rotating ones, |R∇ ln Ω| = α, 1 < α < 3/2 Note that this upper bound is robust even if, as suggested and Ω ≈ Ω0 + |N| for some suitable choice of Ω0. Both of by Fig. 10, the differential rotation falls off faster with Ω these are inconsistent with our predictions, particularly the than we have predicted. latter, which introduces a new characteristic frequency-scale Because, in the MHD regime, |N| deviates only weakly Ω0 and predicts a characteristic length-scale for the shear of from |N|0 we may estimate the right-hand side of equa- −1 (∂r ln |N|) ≈ h. tion (15) using |N|0 from stellar models. For the red giants −6 −1 Our prediction for the scaling of the convection speed studied by Deheuvels et al.(2015) ∆Ωmax ≈ 4×10 s . This 2 −1 (3c ≈ h|N| Ω ) is based on the theory of Stevenson(1979), is on the order of what Cantiello et al.(2014) and Kissin and has been suggested subsequently along various differ- & Thompson(2015b) suggest is needed to match asteroseis- ent lines by Starchenko & Jones(2002), Christensen(2010) mic inferences of core-envelope shear in red giants, so an or- and Barker et al.(2014). In the hydrodynamic case, our pre- der unity pre-factor in equation (15) would suffice to bring diction for the scaling of |N| with Ω matches that of Barker our bound into agreement with their calculations, and it et al.(2014) as well. We previously predicted a somewhat seems possible that most or all of the differential rotation in 3/2 −1/2 different scaling (Jermyn et al. 2018a) of 3c ≈ h|N| Ω such stars is in their convective envelopes. However, because because, unlike Stevenson(1979), Jermyn et al.(2018a) did equation (15) is an upper bound, it remains possible that a not impose the lower bound on the vertical wavenumber. substantial component lies in the radiative zone (Fuller et al. Such a lower bound is physically motivated for stars by the 2019) and some astereoseismic evidence indicates that there finite scale height, so we favour the scaling of Stevenson is more shear in the radiative zones of red giants than in (1979) here. their convection zones (Mauro et al. 2018). Our prediction for the scaling of the the magnetic field A further prediction is that rapidly-rotating convecting 3 ≈ | | ≈ | |3/4Ω1/4 stars ought to exhibit significant equator to pole temper- with rotation rate ( A h N h N 0 ) is based on the relationship between heat flux and magnetic field by Chris- ature differences. These differences are due to the Coriolis > effect acting on convective motions and not the von Zeipel tensen & Aubert(2006) and the bound that if 3A ∼ h|N| then the magnetic field begins to inhibit the linear convective in- effect. This has been seen in simulations (Raynaud et al. stability (Gough & Tayler 1966). This prediction agrees with 2018) and observations (Che et al. 2011), and our scaling the theory of Stevenson(1979), and Starchenko & Jones relations complement these by providing a theoretical curve with which to interpret the observations. In particular, in (2002) predict both the same scaling for 3A and for |N|. However, a variety of other scaling laws have been proposed. the usual notation, These are summarized in Section 4.2. They generally agree T ∝ gβ , (16) that, at fixed |N|, the magnetic field strength varies as a eff eff small power of Ω ranging from −0.02 to 0.19, all of which is where geff is the effective acceleration due to gravity and cen- somewhat smaller than we have predicted. trifugal effects and β = 1/4 is the von Zeipel exponent (von Zeipel 1924). Differentiating this with respect to latitude yields

7 DISCUSSION ∂ ln T ∂ ln g Ω2r cos θ eff = β eff ≈ β . (17) Up to this point we have focused on testing our predictions ∂θ ∂θ g against observations and simulations. We have found general Our findings (equation A108) suggest instead that agreement on the existence and location of plateaus, as well as on the trends away from plateaus. Importantly, in sec- ∂ ln T r|N|2 eff = ξ . (18) tion4, we found that simulations show a strong dependence ∂θ geff on P´eclet number and that they only converge to our predic- 2 2 tions as Pe becomes large in the limit of efficient convection. For slowly-rotating systems, ξ ∝ Ω /|N| and this becomes Because this is the most common limit in astrophysical sys- ∂ ln T Ω2r tems it matters that simulations attempting to reproduce eff ∝ , (19) ∂θ g observed differential rotation operate in this regime. eff We now turn to exploring broader astrophysical impli- which is equivalent to the von Zeipel result, possibly with

MNRAS 000,1–27 (2020) Convective Differential Rotation II 15 a different β. For rapidly-rotating systems with Ω > |N|, ξ DATA AVAILABILITY saturates and Details of how the data from various sources were obtained, standardized and processed are provided in AppendixA, ∂ ln T r|N|2 eff ≈ . (20) and all scripts and files needed to reproduce this analysis are ∂θ g eff available on https://doi.org/10.5281/zenodo.3992227. This saturation could be why observations of late-type stars have inferred β smaller than 1/4: if the systems enter the saturated regime then a fit to the unsaturated scaling would REFERENCES tend to estimate a lower β (Che et al. 2011). Abdul-Masih M., et al., 2016, AJ, 151, 101 For extremely rapid rotators we also predict that con- Aerts C., Mathis S., Rogers T. M., 2019, Annual Review of As- vection is suppressed. In stars this leads to an increase in tronomy and Astrophysics, 57, 35 the Brunt-V¨ ¨ais¨al¨a frequency N and thence a steepening of Altrock R. C., Canfield R. C., 1972, Sol. Phys., 23, 257 the entropy gradient to carry the nuclear luminosity. This Ammler-von Eiff M., Reiners A., 2012, A&A, 542, A116 is likely to be difficult to observe directly in measurements Antia H. M., private communication, 2016 such as the stellar radius or effective temperature because Antia H. M., Basu S., 2001, ApJ, 559, L67 the effect is even smaller than centrifugal expansion (see ap- Antia H. M., Basu S., Chitre S. M., 2008, ApJ, 681, 680 pendixB). On the other hand, it could be observed in astero- Aubert J., Gastine T., Fournier A., 2017, Journal of Me- chanics, 813, 558 seismic measurements. The cores of massive stars likely emit Augustson K. C., Brown B. P., Brun A. S., Miesch M. S., Toomre internal gravity waves, with a peak frequency thought to be J., 2012, ApJ, 756, 169 proportional to the Brunt-V¨ ¨ais¨al¨a frequency (Rogers et al. Augustson K. C., Brun A. S., Toomre J., 2013, The Astrophysical 2013; Couston et al. 2018). Moreover convection in these Journal, 777, 153 stars is so slow that nearly all of them are in the rapidly ro- Augustson K. C., Brun A. S., Toomre J., 2016, ApJ, 829, 92 tating limit. Hence if our predictions hold then there ought Aurnou J., Heimpel M., Wicht J., 2007, Icarus, 190, 110 to be a systematic shift in the peak frequency of these waves Bailey J. E., et al., 2015, Nature, 517, 56 as a function of rotation that scales as Ω1/4. Balbus S. A., Schaan E., 2012, MNRAS, 426, 1546 Finally, note that in the very rapid limit we expect the Balbus S. A., Latter H., Weiss N., 2012, MNRAS, 420, 2457 Balona L. A., Abedigamba O. P., 2016, MNRAS, 461, 497 Taylor-Proudman constraint to hold. That is we predict that Barker A. J., Dempsey A. M., Lithwick Y., 2014, ApJ, 791, 13 the star forms nested cylinders of constant rotation because Bazot M., Benomar O., Christensen-Dalsgaard J., Gizon L., |∂zΩ|  |∇Ω|. We could not find data to probe this predic- Hanasoge S., Nielsen M., Petit P., Sreenivasan K. R., 2019, tion quantitatively but essentially all of the simulations we A&A, 623, A125 examined show this feature in the rapidly rotating limit. Beck P. G., et al., 2012, Nature, 481, 55 Benomar O., et al., 2018, Science, 361, 1231 Bonanno A., Fr¨ohlich H.-E., Karoff C., Lund M. N., Corsaro E., Frasca A., 2014, A&A, 569, A113 Borkovits T., Hajdu T., Sztakovics J., Rappaport S., Levine A., B´ır´oI. B., Klagyivik P., 2016, MNRAS, 455, 4136 Brown B. P., Browning M. K., Brun A. S., Miesch M. S., Toomre ACKNOWLEDGEMENTS J., 2008, ApJ, 689, 1354 Browning M. K., Brun A. S., Toomre J., 2004, ApJ, 601, 512 The Flatiron Institute is supported by the Simons Founda- Brun A. S., Palacios A., 2009, ApJ, 702, 1078 tion. ASJ thanks the Gordon and Betty Moore Foundation Brun A. S., Browning M. K., Toomre J., 2005, The Astrophysical (Grant GBMF7392) and the National Science Foundation Journal, 629, 461 (Grant No. NSF PHY-1748958) for supporting this work. Brun A. S., Antia H. M., Chitre S. M., 2010, A&A, 510, A33 ASJ also acknowledges financial support from a UK Mar- Brun A. S., et al., 2017, ApJ, 836, 192 shall Scholarship as well as from the IOA, ENS and CEBS Buchler J. R., Yueh W. R., 1976, ApJ, 210, 440 to work at ENS Paris and CEBS in Mumbai. PL acknowl- Cantiello M., Braithwaite J., 2019, ApJ, 883, 106 Cantiello M., Mankovich C., Bildsten L., Christensen-Dalsgaard edges travel support from the french PNPS (Programme Na- J., Paxton B., 2014, ApJ, 788, 93 tional de Physique Stellaire) and from CEBS. CAT thanks Cassisi S., Potekhin A. Y., Pietrinferni A., Catelan M., Salaris Churchill College for his fellowship. SMC is grateful to the M., 2007, ApJ, 661, 1094 IOA for support and hospitality and thanks the Cambridge- Chaboyer B., Zahn J.-P., 1992, Astronomy and Astrophysics, 253, Hamied exchange program for financial support. ASJ, SMC 173 and PL thank Bhooshan Paradkar for productive conversa- Che X., et al., 2011, ApJ, 732, 68 tions related to this work. The authors thank Jim Fuller, Christensen U. R., 2010, Space Sci. Rev., 152, 565 Douglas Gough, Steven Balbus, Chris Thompson for com- Christensen U. R., Aubert J., 2006, Geophysical Journal Interna- ments on this manuscript. We are also thankful to the anony- tional, 166, 97 mous referee, whose suggestions led us to include more data Chugunov A. I., Dewitt H. E., Yakovlev D. G., 2007, Phys. Rev. D, 76, 025028 as well as the sensitivity analysis in AppendicesC andD. Conroy K. E., et al., 2014, PASP, 126, 914 ASJ is grateful to Lars Bildsten for helpful suggestions re- Couston L.-A., Lecoanet D., Favier B., Le Bars M., 2018, Journal garding the data analysis, to Frank Timmes and Matteo of , 854, R3 Cantiello for suggestions regarding the presentation, and to Cox J. P., Giuli R. T., 1968, Principles of Daniel Lecoanet, Evan Anders and Matthew Browning for Cyburt R. H., et al., 2010, ApJS, 189, 240 comments regarding the interpretation of simulations. Davenport J. R. A., Hebb L., Hawley S. L., 2015, ApJ, 806, 212

MNRAS 000,1–27 (2020) 16 Adam S. Jermyn

Davidson P. A., 2013, Geophysical Journal International, 195, 67 Langanke K., Mart´ınez-Pinedo G., 2000, Nuclear Physics A, 673, Deheuvels S., Ballot J., Beck P. G., Mosser B., Østensen R., Gar- 481 c´ıaR. A., Goupil M. J., 2015, A&A, 580, A96 Liu J., Goldreich P. M., Stevenson D. J., 2008, Icarus, 196, 653 Donahue R. A., Saar S. H., Baliunas S. L., 1996, ApJ, 466, 384 Lurie J. C., et al., 2017, AJ, 154, 250 Donati J. F., Collier Cameron A., 1997, MNRAS, 291, 1 Mabuchi J., Masada Y., Kageyama A., 2015, ApJ, 806, 10 Donati J.-F., et al., 2008, MNRAS, 390, 545 Maeder A., 1998, in Howarth I., ed., Astronomical Society of the Eddington A. S., 1929, MNRAS, 90, 54 Pacific Conference Series Vol. 131, Properties of Hot Lumi- Featherstone N. A., Miesch M. S., 2015, The Astrophysical Jour- nous Stars. p. 85 nal, 804, 67 Masmoudi N., 2007, Journal des Mathematiques Pures et Ap- Ferguson J. W., Alexander D. R., Allard F., Barman T., Bodnarik pliquees, 88, 230 J. G., Hauschildt P. H., Heffner-Wong A., Tamanai A., 2005, MatijeviˇcG., PrˇsaA., Orosz J. A., Welsh W. F., Bloemen S., ApJ, 623, 585 Barclay T., 2012, AJ, 143, 123 Frasca A., Fr¨ohlich H.-E., Bonanno A., Catanzaro G., Biazzo K., Matt S. P., Do Cao O., Brown B. P., Brun A. S., 2011, As- Molenda-Zakowicz˙ J., 2011, A&A, 532, A81 tronomische Nachrichten, 332, 897 Fuller G. M., Fowler W. A., Newman M. J., 1985, ApJ, 293, 1 Mauro M. P. D., Ventura R., Corsaro E., Moura B. L. D., 2018, Fuller J., Piro A. L., Jermyn A. S., 2019, MNRAS, 485, 3661 The Astrophysical Journal, 862, 9 Gastine T., Wicht J., 2012, Icarus, 219, 428 Miesch M. S., Toomre J., 2009, Annual Review of Fluid Mechan- Gastine T., Wicht J., Aurnou J. M., 2013, Icarus, 225, 156 ics, 41, 317 Gastine T., Yadav R. K., Morin J., Reiners A., Wicht J., 2014, Nielsen M. B., Schunker H., Gizon L., Schou J., Ball W. H., 2017, MNRAS, 438, L76 A&A, 603, A6 Gilliland R. L., et al., 2010, Publications of the Astronomical Oda T., Hino M., Muto K., Takahara M., Sato K., 1994, Atomic Society of the Pacific, 122, 131 Data and Nuclear Data Tables, 56, 231 Gilman P. A., 1977, Geophysical and Astrophysical Fluid Dynam- Ouazzani R. M., Marques J. P., Goupil M. J., Christophe S., ics, 8, 93 Antoci V., Salmon S. J. A. J., Ballot J., 2019, A&A, 626, Gilman P. A., 1979, ApJ, 231, 284 A121 Gilman P. A., 1980, in Gray D. F., Linsky J. L., eds, Lecture Notes Paxton B., Bildsten L., Dotter A., Herwig F., Lesaffre P., Timmes in Physics, Berlin Springer Verlag Vol. 114, IAU Colloq. 51: F., 2011, ApJS, 192, 3 Stellar Turbulence. pp 19–37, doi:10.1007/3-540-09737-6 7 Paxton B., et al., 2013, ApJS, 208, 4 Gilman P. A., Foukal P. V., 1979, ApJ, 229, 1179 Paxton B., et al., 2015, ApJS, 220, 15 Glatzmaier G., Evonuk M., Rogers T., 2009, Geophysical and Paxton B., et al., 2018, ApJS, 234, 34 Astrophysical , 103, 31 Paxton B., et al., 2019, The Astrophysical Journal Supplement Gough D. O., Tayler R. J., 1966, MNRAS, 133, 85 Series, 243, 10 Guerrero G., Smolarkiewicz P. K., Kosovichev A. G., Mansour Pols O. R., Tout C. A., Eggleton P. P., Han Z., 1995, MNRAS, N. N., 2013, ApJ, 779, 176 274, 964 Guillot T., et al., 2018, Nature, 555, 227 Potekhin A. Y., Chabrier G., 2010, Contributions to Hermes J. J., et al., 2017, ApJS, 232, 23 Physics, 50, 82 Iglesias C. A., Rogers F. J., 1993, ApJ, 412, 752 Rajaguru S. P., Antia H. M., 2015, ApJ, 813, 114 Iglesias C. A., Rogers F. J., 1996, ApJ, 464, 943 Rappaport S., Deck K., Levine A., Borkovits T., Carter J., El Ingersoll A. P., et al., 2004, Dynamics of Jupiter’s . Mellah I., Sanchis-Ojeda R., Kalomeni B., 2013, ApJ, 768, 33 pp 105–128 Rast M. P., Ortiz A., Meisner R. W., 2008, The Astrophysical Itoh N., Hayashi H., Nishikawa A., Kohyama Y., 1996, ApJS, 102, Journal, 673, 1209 411 Raynaud R., Rieutord M., Petitdemange L., Gastine T., Putigny Jermyn A. S., Lesaffre P., Tout C. A., Chitre S. M., 2018a, MN- B., 2018, A&A, 609, A124 RAS, 476, 646 Reiners A., Schmitt J. H. M. M., 2003, A&A, 412, 813 Jermyn A. S., Tout C. A., Chitre S. M., 2018b, MNRAS, 480, Reinhold T., Gizon L., 2015, A&A, 583, A65 5427 Ricker G. R., et al., 2015, Journal of Astronomical Telescopes, Jermyn A. S., Chitre S. M., Lesaffre P., Tout C. A., 2020 Instruments, and Systems, 1, 014003 Jones C. A., Boronski P., Brun A. S., Glatzmaier G. A., Gastine Rogers T. M., 2015, ApJ, 815, L30 T., Miesch M. S., Wicht J., 2011, Icarus, 216, 120 Rogers F. J., Nayfonov A., 2002, ApJ, 576, 1064 Joyce M., Chaboyer B., 2018, ApJ, 856, 10 Rogers T. M., Lin D. N. C., McElwaine J. N., Lau H. H. B., 2013, K¨apyl¨a P., 2011, Astronomische Nachrichten, 332, 43 The Astrophysical Journal, 772, 21 K¨apyl¨a P. J., Mantere M. J., Guerrero G., Brandenburg A., Chat- Roxburgh I. W., COROT Team 1999, in Gimenez A., Guinan terjee P., 2011, A&A, 531, A162 E. F., Montesinos B., eds, Astronomical Society of the Pacific Kaspi Y., Flierl G. R., Showman A. P., 2009, Icarus, 202, 525 Conference Series Vol. 173, Stellar Structure: Theory and Test Kaspi Y., et al., 2018, Nature, 555, 223 of Connective Energy Transport. p. 357 Kiraga M., Stepien K., 2007, Acta Astron., 57, 149 Saumon D., Chabrier G., van Horn H. M., 1995, ApJS, 99, 713 Kirk R. L., Stevenson D. J., 1987, ApJ, 316, 836 Schad A., Roth M., 2020, ApJ, 890, 32 Kissin Y., Thompson C., 2015a, ApJ, 808, 35 Schou J., et al., 1998, ApJ, 505, 390 Kissin Y., Thompson C., 2015b, The Astrophysical Journal, 808, Soderlund K. M., Heimpel M. H., King E. M., Aurnou J. M., 35 2013, Icarus, 224, 97 Kitchatinov L. L., Ruediger G., 1995, A&A, 299, 446 Spruit H. C., 2002, A&A, 381, 923 Klion H., Quataert E., 2017, Monthly Notices of the Royal As- Stancliffe R. J., Fossati L., Passy J. C., Schneider F. R. N., 2015, tronomical Society: Letters, 464, L16 A&A, 575, A117 Kuker¨ M., Rudiger¨ G., Olah K., Strassmeier K. G., 2019, A&A, Starchenko S. V., Jones C. A., 2002, Icarus, 157, 426 622, A40 Stevenson D. J., 1979, Geophysical and Astrophysical Fluid Dy- Kushnir D., Zaldarriaga M., Kollmeier J. A., Waldman R., 2016, namics, 12, 139 MNRAS, 462, 844 Takeda Y., UeNo S., 2017, Sol. Phys., 292, 123 LaCourse D. M., et al., 2015, MNRAS, 452, 3561 The LIGO Scientific Collaboration et al., 2018, arXiv e-prints,

MNRAS 000,1–27 (2020) Convective Differential Rotation II 17

Thompson M. J., 1991, Helioseismic Inversion. p. 61, reported shear Ω1 is 2/15 times the difference between the doi:10.1007/3-540-54420-8 52 pole and the equator. There are π/2 radians between the Timmes F. X., Swesty F. D., 2000, ApJS, 126, 501 pole and the equator so we approximate Varela J., Strugarek A., Brun A. S., 2016, Advances in Space Research, 58, 1507 2  15  |R∇Ω| ≈ |∂ Ω| ≈ Ω . (A1) Vidotto A. A., et al., 2014, Monthly Notices of the Royal Astro- θ π 2 1 nomical Society, 441, 2361 Vilhu O., 1984, A&A, 133, 117 Uncertainties in both the shear and mean rotation rate Wright N. J., Drake J. J., 2016, Nature, 535, 526 were reported. These were propagated into the relative shear Wright N. J., Drake J. J., Mamajek E. E., Henry G. W., 2011, by the formula ApJ, 743, 48 uv Yadav R. K., Gastine T., Christensen U. R., 2013a, Icarus, 225, t  2 Õ ∂µ 2 185 dµ = dxj , (A2) ∂xj Yadav R. K., Gastine T., Christensen U. R., Duarte L. D. V., j 2013b, ApJ, 774, 6 where µ is a function of interest and x are variables on Yadav R. K., Gastine T., Christensen U. R., Duarte L. D. V., j Reiners A., 2016, Geophysical Journal International, 204, which it depends. This amounts to the assumption that er- 1120 rors are uncorrelated and small relative to the scale over Zackay B., Venumadhav T., Dai L., Roulet J., Zaldarriaga M., which higher-order derivatives are relevant. 2019, Phys. Rev. D, 100, 023007 Stellar models were constructed which matched the Zhao J., Bogart R. S., Kosovichev A. G., Duvall T. L. J., Hartlep ages, metallicities and masses provided in their table S2. The T., 2013, ApJ, 774, L29 average |N|0 (Brunt-V¨ ¨ais¨al¨a frequency) in the convection von Zeipel H., 1924, MNRAS, 84, 665 zone of each model was calculated following equation (7).

APPENDIX A: DATA PROCESSING A1.2 Bazot Observations Here we explain the way in which data were processed to Data were taken from tables 1 and A3 of Bazot et al.(2019) produce our figures. All scripts and inlists as well as auxiliary who inferred differential rotation through asteroseismology data files used in this data processing are available on https: for several solar-type stars and reported in the form of the //doi.org/10.5281/zenodo.3992227. latitudinal shear and mean rotation rates. The reported In many cases the data processing required stellar shear Ω1 is 2/15 times the difference between the pole and models. These were produced using revision 11701 of the the equator. There are π/2 radians between the pole and the Modules for Experiments in Stellar Astrophysics MESA. equator so we approximate The MESA equation of state (EOS) is a blend of the 2  15  OPAL (Rogers & Nayfonov 2002), SCVH (Saumon et al. |R∇Ω| ≈ |∂ Ω| ≈ Ω . (A3) θ π 2 1 1995), PTEH (Pols et al. 1995), HELM (Timmes & Swesty 2000) and PC (Potekhin & Chabrier 2010) equations of Uncertainties in both the shear and mean rotation rate state. Radiative opacities are primarily from OPAL (Igle- were reported for 16 Cyg A. For 16 Cyg B a weighted average sias & Rogers 1993, 1996), with low-temperature data ac- of the three modes was used following the recommendation cording to Ferguson et al.(2005) and the high-temperature, in the caption of table 1. The uncertainty of this average Compton-scattering dominated regime by Buchler & Yueh was computed as the weighted root-mean-squared difference (1976). Electron conduction opacities are according to Cas- between the three modes and the average. sisi et al.(2007). Nuclear reaction rates are from JINA REA- Uncertainties were propagated into the relative shear CLIB (Cyburt et al. 2010) plus additional tabulated weak equation (A2). Stellar models were constructed which reaction rates (Fuller et al. 1985; Oda et al. 1994; Langanke matched the ages, metallicities and masses provided in their & Mart´ınez-Pinedo 2000). Electron screening is included via table A3, and the average |N|0 in the convection zone of each the prescription of Chugunov et al.(2007). Thermal neutrino model was calculated according to equation (7). loss rates are according to Itoh et al.(1996). Mixing length theory was implemented following Cox & Giuli(1968) with mixing length parameter α = 2. Models were created on the A1.3 Ammler-von Eiff Observations pre- and evolved from there. All other param- Data were taken from table 2 of Ammler-von Eiff & Reiners eters were set to their defaults. (2012). The latitudinal differential rotation was spectroscop- ically inferred for several√ A to F stars and reported in the A1 Latitudinal Shear form of the coefficient η/ sin i, where η is the pole to equa- tor shear divided by the equatorial angular velocity. For each We describe the observations. object our hypothesis is that the latitudinal differential ro- tation arises owing to the surface convection zone. So we A1.1 Benomar Observations identify 2 Data were taken from tables S2 and S3 of Benomar et al. |R∇Ω| ≈ |∂θ Ω| ≈ |Ωpole − Ωequator|. (A4) (2018), who inferred differential rotation through asteroseis- π mology for several solar-type stars and reported it in the Stars without a surface convection zone were filtered form of the latitudinal shear and mean rotation rates. The out of the sample. Many systems were consistent with zero

MNRAS 000,1–27 (2020) 18 Adam S. Jermyn shear to within the error bars. So we only include those with to calculate Ω = 3/R and Ω/|N|avg. Objects hotter than the at least a 1σ detection of shear. hottest model in the grid or cooler than the coolest model The rotation rate is reported as 3 sin i, where 3 is the in the grid were excluded from the sample. surface rotation velocity. Inclinations were not reported. So rotation speeds are considered to be lower bounds and shear measurements to be upper bounds. A1.8 Lurie Observations Eighty stellar models were constructed with metal- licity Z = 0.02 and spaced uniformly in mass from Data were taken from Lurie et al.(2017), who report differ- 0.2 M to 1.4 M . The models were evolved to an age of ential rotation inferred from star spot observations of Kepler −2.5 (M/M ) GYr or 10 GYr, whichever was smaller. For each eclipsing binary systems. They infer the parameter η by object the model which best-matched the reported T in eff Pmax − Pmin their table 5 was used to compute Ω = 3/R and |N|avg. η = , (A5) Pmax

where the periods Pmin and Pmax are the first minimum and A1.4 Frasca Observations maximum in the periodogram of the lightcurve of the sys- tem. They report these periods and so we compute η using Data were taken from Frasca et al.(2011). The differential equation (A5) and translate this into a shear using equa- rotation was inferred from star spot observations of the K tion (A4). star KIC 8429280 and reported in the form of the pole to Triple-star systems were excluded by means of the cata- equator shear, which we treat with equation (A4). The rota- logues of Rappaport et al.(2013) and Borkovits et al.(2016) tion rate is reported as 3 sin i, where 3 is the surface rotation on the grounds that the inference of Lurie et al.(2017) is velocity. The inclination is also reported and so we use it based on the assumption of a binary system. We further ex- to compute 3. We employ the same grid of main-sequence clude the false positive systems described by Borkovits et al. stellar models from appendix A1.3 to compute |N| and R avg (2016) because these have unusual light curves which could and use these to calculate Ω = 3/R and Ω/|N| . avg interfere with the inference of rotation periods. Objects were matched by effective temperature to the A1.5 Bonanno Observations grid of main-sequence stellar models from appendix A1.3, from which we computed |N|avg and Ω/|N|avg. The effective Data were taken from Bonanno et al.(2014) who report dif- of these objects were obtained from the Kepler ferential rotation inferred from star spot observations of the Eclipsing Binary catalog (Matijeviˇcet al. 2012; Conroy et al. late F star KIC 5955122. The differential rotation was re- 2014; LaCourse et al. 2015; Abdul-Masih et al. 2016). Ob- ported in the form of a pole to equator shear, which we treat jects with no temperature in the catalogue were excluded. with equation (A4). We normalise this to the mean rotation Objects hotter than the hottest model in the grid or cooler rate and take the result to be approximately |R∇Ω|/Ω.A than the coolest model in the grid were excluded from the stellar model was constructed which matched the age, metal- sample. licity and mass reported for KIC 5955122 and we used this to compute |N|avg and Ω/|N|avg. A1.9 Davenport Observations A1.6 Donati Observations Data were taken from Davenport et al.(2015), who re- Data were taken from Donati et al.(2008). The differential port differential rotation inferred from star spot observa- rotation was inferred for a sample of M-dwarfs from spec- tions of the M-dwarf GJ 1243. The spots reveal a shear of − tropolarimetry and reported as the surface equator to pole 0.0047 rad d 1. After accounting for spot latitude, they report − difference in Ω normalised against that of the equator. We a pole-equator shear of 0.012 ± 0.002 rad d 1. The rotation treat this with equation (A4). period of the is reported as 0.592596 ± 0.00021 d. We Donati et al.(2008) report the Rossby number in the convert this into η and then use equation (A4) to obtain a form of τcProtation, where τc = 1/|N|0 (Gilman 1980) was de- relative shear. Uncertainties were propagated along the way termined from an empirical calibration to non-rotating stel- with equation (A2). The Rossby number is also reported, as lar models (Kiraga & Stepien 2007) and Protation is the aver- by Donati et al.(2008) and we handle it in the same manner age spin period of the star. So we take Ω ≈ 2πRo−1. as in appendix A1.6. |N |0

A1.7 Reinhold Observations A2 Radial Shear Data were taken from Reinhold & Gizon(2015), who report We describe the observations. differential rotation inferred from star spot observations of Kepler main-sequence and subgiant stars. Stars which were flagged as having highly stable periods were excluded as pos- A2.1 Deheuvels Observations sible binary or pulsator systems. The differential√ rotation was reported in the form of the Data were taken from Deheuvels et al.(2015). The differen- coefficient η/ sin i, which we handle as in appendix A1.4. tial rotation was seismically inferred for several red giants We employ the same grid of main-sequence stellar models and reported in the form of the difference between the core from appendix A1.3 to compute |N|avg and R and use these and envelope rotation rates divided by the envelope rotation

MNRAS 000,1–27 (2020) Convective Differential Rotation II 19 rate. We take this to equal Browning et al.(2004) report the

Ωcore ¹ 3˜L ln ≈ |R∇ ln Ω|d ln R, (A6) Re = , (A12) Ωenvelope convection zone ν on the assumption that most of the differential rotation is where 3˜ is the root-mean square of the convective velocity, developed in the convection zone. We thereby infer L is the radial extent of the convective region and ν is the microscopic viscosity. They also report the Prandtl number Ω ¹ −1 |R∇ ln Ω| ≈ ln core d ln R . (A7) ν avg Ω Pr = = 0.25 (A13) envelope convection zone α Stellar models were constructed with MESA which and the Rayleigh number matched the masses and radii in table 2 of Deheuvels g∆sL3 ∂ρ et al.(2015). These were used to compute |N|avg and Ra = , (A14) ∫ ρνα ∂s convection zone d ln R. where α is the thermal diffusivity, ρ is the density, s is the entropy, ∆s is the entropy change across the domain and A2.2 Klion Observations g is the acceleration owing to gravity. We assume that the Data were taken from the text of Klion & Quataert(2017) partial derivative is taken at constant pressure, for the red giant Kepler-56. The differential rotation was  ∂ ln ρ  γ − 1 seismically inferred and reported in the form of a power-law = . (A15) ∂s P γ Ω(r) ∝ rβ (A8) With this information we obtain the convective velocity by with β ≈ 1, so we take |R∇Ω|/Ω ≈ 1. The s 3 (γ − 1)Pr is provided in the text as 74 ± 3d. A solar-metallicity stellar c ≈ Re (A16) model was constructed with MESA which matched the mass h|N| γRa and radius reported for Kepler 56. This was used to compute and the P´eclet number as |N|avg. √ Pe = RePr ≈ RaPr. (A17)

A2.3 Nielsen Observations The Rossby number actually realised in the flow is also reported. This they define as Data were taken from table 1 of Nielsen et al.(2017). The ωφ,convective differential rotation was seismically inferred for several solar- Ro = , (A18) like stars and reported as envelope and interior angular fre- 2Ω quencies. We then approximate the shear as where ωφ,convective ≈ 3c/h is the root-mean square of the vor- ticity of the convective flow. So we find 1  max(Ωcore, Ωenvelope)  |R∇Ω| ≈ R∂ Ω ≈ − 1 , (A9)   R Ω 1 3c ∆ ln R min(Ωcore, Ωenvelope) ≈ , (A19) |N| 2Ro h|N| where ∆ ln R was computed as where the term in parentheses is given by equation (A16). ¹ dr ∆ ln R = (A10) convection zone max(r, h) as we have done in equation (7). Uncertainties were prop- A3.2 Augustson 2012 Simulations agated via equation (A2). Objects were matched by effec- These data come from three-dimensional anelastic hydrody- tive temperature to the grid of main-sequence stellar mod- namic simulations of convection in a rotating spherical shell els from appendix A1.3, from which we computed |N|avg and domain (Augustson et al. 2012). The differential rotation is Ω/|N|avg. reported as the difference in angular velocity between the equator and a latitude of π/3, normalized to the equatorial A3 Hydrodynamic Simulations angular velocity, so we process this as equation (A11). The Rossby number actually realised in the flow is also reported, We describe how we extract the relevant data from various as equation (A18), so we handle this as before. hydrodynamic simulations. The temperature difference ∆T between the equator and a latitude of π/3, averaged over depth, is reported. To obtain the baroclinic angle ξ from ∆T we write A3.1 Browning 2004 Simulations ∂θ s These data come from three-dimensional anelastic hydro- ξ ≈ . (A20) r∂r s dynamic simulations of convection in a rotating spherical domain (Browning et al. 2004). The differential rotation is This is valid in the limit of a radial pressure gradient and reported as the difference in angular velocity between the a mostly-radial entropy gradient, so that ξ is the small pro- equator and a latitude of π/3, normalized to the equatorial jection of ∇s/|∇s| along eθ . Augustson et al.(2012) define angular velocity, so we write ∆SgL3 Ra = , (A21) 3 Ωequator − Ωπ/3 cpνα |R∇ ln Ω| ≈ . (A11) π Ωequator where Ra is the Rayleigh number, ∆S is the radial change

MNRAS 000,1–27 (2020) 20 Adam S. Jermyn

in dimensionful entropy across the simulation domain, L the A3.3 Matt 2011 Simulations vertical size of the domain, g is the acceleration owing to These data come from three-dimensional anelastic hydro- gravity, c is the specific heat at constant pressure, ν is the P dynamic simulations of convection in a rotating spherical viscosity and α is the thermal diffusivity. shell domain (Matt et al. 2011). The differential rotation is The ratio ∆S/c is related to the dimensionless entropy P reported as the difference in angular velocity between the change ∆s by equator and a latitude of π/3, normalized to the equato- ∆S  γ  rial angular velocity, so we use equation (A11) to compute ∆s = . (A22) cP γ − 1 |R∇ ln Ω|. The Rossby number actually realised in the flow is also reported, so we follow the procedure we used in ap- Identifying pendix A3.1 to calculate Ω/|N|. ∆s ≈ L∂r s, (A23) The temperature difference between the equator and a latitude of π/3 is also reported. We follow the same proce- we then obtain dure as in appendix A3.2 to compute ξ from this, except να  γ  that we compute r∂ s ≈ Ra . (A24) r 4 γ − 1 gL GM g = , (A35) They further let the Prandtl number Pr be ν/α and the Tay- r2 lor number Ta be 4Ω2L4/ν2. With these we obtain where G is Newton’s constant and M is the mass of the star γ  Ra   Ω2r  because Matt et al.(2011) do not report g but do report M r∂ s ≈ 4 . (A25) r γ − 1 Pr Ta g and R. We compute the ratio 3c/h|N| with the Rossby, Prandtl, So Taylor and Rayleigh numbers as in appendix A3.2. Finally, γ − 1  Pr Ta   g  we obtain the P´eclet number from equation (A17). ξ ≈ ∂θ s. (A26) 4γ Ra Ω2r Assuming that isobars are nearly spherical, we see that

∂s A3.4 Brown 2008 Simulations ∂θ s ≈ ∂θ ln T . (A27) ∂ ln T P These data come from simulations of three-dimensional With equation (2) we find anelastic hydrodynamic convection in rotating spherical shells, reported in tables 1 and 2 of Brown et al.(2008).

∂s γ The differential rotation is reported as the kinetic energy of = , (A28) ∂ ln T P γ − 1 differential rotation, defined to be the mass-weighted aver- so age of the squared difference between the rotation rate and the mean rotation rate. That is, they report 1  Pr Ta   g  ξ ≈ ∂θ ln T (A29) 4 Ra Ω2r  2 DRKE ≈ ρR2h Ω − Ω i, (A36) Augustson et al.(2012) provide g, R, Pr, Ta and Ra. They also provide a quantity Ω/Ω , which we take to be the mean where h...i denotes the mass-weighted average and Ω¯ is the angular velocity of the simulation normalized by the typical mean rotation rate. They further report the convective ki- − − solar angular velocity of 4.3 × 10 7cycles s 1. We then esti- netic energy mate 2 3 ∆T CKE ≈ ρ3c . (A37) ∂θ ln T ≈ , (A30) π Tbase CZ To ensure consistent normalization we estimate 3c from their where we use the temperature Tbase CZ at the base of the reported Rossby number as convection zone because the temperature variation is domi- nated by that region. 3c ≈ 2ΩRo, (A38) We are also interested in the convection speed normal- so that the shear is ized by 3c/h|N|. To obtain this we write r r 3 3 r γ 3c DRKE DRKE c ≈ c (A31) |R∇Ω| ≈ ≈ 2ΩRo , (A39) h|N| h (γ − 1)|g∂r s| R CKE CKE 3 r γr where Ro is the Rossby number actually realised in the flow ≈ c (A32) h (γ − 1)g|r∂r s| and of the form of equation (A18). We calculate Ω/|N| from Ro as in appendix A3.1 3c tv γ ≈ (A33) We compute the ratio 3 /h|N| with the Rossby, Prandtl, 2Ωh  Ra  c Pr Ta Taylor and Rayleigh numbers as in appendix A3.2. The one r difference is that the Rayleigh number is reported as in equa- Pr Ta ≈ Ro γ . (A34) tion (A14). So we must divide it by (γ − 1)/γ before using Ra equation (A34). As before, we obtain the P´eclet number from Finally, we obtain the P´eclet number by equation (A17). equation (A17).

MNRAS 000,1–27 (2020) Convective Differential Rotation II 21

A3.5 Brun 2009 Simulations To compute 3c/h|N| we use the Rossby number actually realised in the flow. With this we find These data come from three-dimensional anelastic hydro- 3 √ dynamic simulations of convection in a rotating spherical c ≈ 2Ro R∗. (A42) domain reported by Brun & Palacios(2009). The differen- h|N| tial rotation is reported as the mass-weighted average of the We compute the Reynolds number as Re = 2ΩLRo/ν squared difference between the rotation rate and the mean from the given Ω, L and viscosity ν. We then compute the rotation rate. We translate this into a shear as we did in P´eclet number from equation (A17). appendix A3.4. The Rossby number actually realised in the flow is reported in the form of equation (A18) and we com- pute Ω/|N| from this as in appendix A3.1. A3.8 Augustson 2013 & 2016 Simulations We compute the ratio 3c/h|N| in the same way as we did for the data from Augustson et al.(2012), using the These data come from anelastic three-dimensional anelas- Rossby, Prandtl, Taylor and Rayleigh numbers. Two differ- tic hydrodynamical simulations of rotating convection in ences are that Brun & Palacios(2009) used a Prandtl num- a spherical domain (Augustson et al. 2016) and spherical ber of 1 and that they report the Rayleigh number as in shells (Augustson et al. 2013). The differential rotation is equation (A14). So we must divide Ra by (γ −1)/γ before us- reported as the mass-weighted average of the squared differ- ing equation (A34). They also do not report γ so we assume ence between the rotation rate and the mean rotation rate. that it is 5/3. As before, we obtain the P´eclet number from We translate this into a shear as we did in appendix A3.4. equation (A17). The Rossby number actually realised in the flow is reported in the form of equation (A18) and we compute Ω/|N| from this as in appendix A3.1. A3.6 Brun 2017 Simulations The temperature difference between the equator and a latitude of π/3 is also reported by Augustson et al.(2013). These data come from three-dimensional anelastic hydro- We follow the same procedure as in appendix A3.2 to com- dynamic simulations of convection in a rotating spherical pute ξ from this. We likewise compute the ratio 3 /h|N| in domain (Brun et al. 2017). The differential rotation is re- c the same way as we did in appendix A3.2. We obtain the ported as the mass-weighted average of the squared differ- P´eclet number from equation (A17). ence between the rotation rate and the mean rotation rate. We translate this into a shear as we did in appendix A3.4. The Rossby number actually realised in the flow is reported A3.9 Guerrero Simulations in the form of equation (A18) and we compute Ω/|N| from this as in appendix A3.1. These data come from simulations of three-dimensional The temperature difference between the equator and a anelastic convection in rotating spherical shells, reported in latitude of π/3 is also reported. We follow the same proce- table 1 of Guerrero et al.(2013). The reported differential dure as in appendix A3.2 to compute ξ from this, except rotation rate is the surface equator-pole difference in Ω nor- that we compute g as in appendix A3.3. malised to that of the equator. Because this spans π/2 ra- We compute the ratio 3c/h|N| in the same way as we did dians we divide by π/2 to give it the same units as ∂θ Ω/Ω. in appendix A3.5. As before, we obtain the P´eclet number We take this as to be approximately |R∇Ω|/Ω. from equation (A17). They report R∗ as in equation (A40). We compute Ω/|N| from this as in appendix A3.7. They also report the Rossby number in the form A3.7 Featherstone 2015 v˜ These data come from three-dimensional anelastic hydrody- Ro = , (A43) 2Ω¯ L namic simulations of convection in rotating spherical shells, reported by Featherstone & Miesch(2015). The differen- where v˜ is the root-mean square of the fluctuating velocity, ¯ tial rotation is reported as the mass-weighted average of the L is the vertical size of the domain and Ω is the rotation squared difference between the rotation rate and the mean rate of the frame with zero total angular momentum. We rotation rate. We translate this into a shear as we did in use this to find 3 √ appendix A3.4. The Rossby number actually realised in the c ≈ 2Ro R∗. (A44) flow is reported in the form of equation (A18), and we com- h|N| Ω/| | pute N from this as in appendix A3.1. We were unable to determine the P´eclet number in these Featherstone & Miesch(2015) report a modified simulations because the EULAG software instrument they Rayleigh number employ does not use explicit diffusivities and so makes it g|∂ S| difficult to determine precise values for dimensionless pa- R∗ ≡ r , (A40) 2 cPΩ rameters such as Pe. where these quantities are all as defined previously. Using equations (2), (1) and (A22) we find A3.10 Mabuchi Simulations (HD) |N|2 R∗ = , (A41) These data come from three-dimensional fully compressible 2 Ω hydrodynamic simulations of rotating convection in spher- from which we compute Ω/|N|. ical shells, reported by Mabuchi et al.(2015). Data were

MNRAS 000,1–27 (2020) 22 Adam S. Jermyn taken from their table 1. The reported differential rotation from which we compute Ω/|N|. We further obtain is in the form     3c 2 Ro Ω − Ω¯ = . (A55) α ≡ , (A45) h|N| π Roconv Ω¯ We calculate the P´eclet number from equation (A17). where Ω is evaluated at the surface of the star on the equator. The convective Rossby number is reported as A3.12 Gastine 2012 Simulations 2 Ra Roconv ≡ , (A46) Ta Pr These data come from three-dimensional hydrodynamic sim- where ulations of anelastic convection in rotating spherical shells, reported by Gastine & Wicht(2012).  2Ω¯ d 2 Ta ≡ , (A47) The polytropic index is reported as m = 2, so that γ = ν 3/2. The same was used in each of these d is the vertical size of the domain and simulations, and reported as 4   ν gd ∂r s Ek = = 10−4, (A56) Ra ≡ − . (A48) 2 νξ cP Ωframed With these definitions, the convective Rossby number sim- where Ωframe is the angular velocity of the simulation refer- plifies to ence frame, d is the vertical size of the domain and ν is the microscopic viscosity. | |2 2 N The Prandtl number was also held constant and re- Roconv = , (A49) 4Ω2 ported as from which we compute Ω/|N|. ν Pr ≡ = 1, (A57) The Rossby number actually realised in the flow is also α reported as where α is the microscopic thermal diffusivity. Finally, the π3 Ro = c , (A50) aspect ratio η was held constant and reported as Ωd rinner where d is the radial extent of the convection zone. From η = . (A58) router this we obtain     This is related to the vertical size d of the domain by 3c 2 Ro = . (A51) d h|N| π Roconv η = 1 − . (A59) router We obtain the P´eclet number from equation (A17). We calculate the Reynolds number using the height of the un- The Rayleigh number is reported as − stable layer (0.3), the thermal diffusivity (3.9×10 4) and the gd3∆S reported root-mean-square velocities, all in code units. Ra ≡ , (A60) cpνα

where cp is the specific heat at constant pressure, g is the A3.11 K¨apyl¨a Simulations acceleration due to gravity and ∆S is the difference in specific dimensionful entropy across the domain. The ratio ∆S/c These data come from simulations of three-dimensional fully p may be written as compressible hydrodynamical convection in rotating spher- ical shells (K¨apyl¨a et al. 2011). Data were taken from their ∆S 1 γ − 1 = ln p − ln ρ = ∆s, (A61) table 1. The reported differential rotation rate is the surface cp γ γ difference of Ω over 2π/3 radians normalised to that of the where s is the dimensionless entropy defined in equation (2). equator. We thus divide by 2π/3 to estimate ∂ Ω/Ω and take θ So this as to be approximately |R∇Ω|/Ω. The Rayleigh number is reported as in equation (A48). (γ − 1)gd3∆s Ra = . (A62) The Coriolis number was reported as Ro−1, with Ro defined ναγ by equation (A50). The Prandtl number is reported as well. Approximating the entropy gradient as radial, we write Finally, the Reynolds number is reported as ∂s v˜d g∆s ≈ gd ≈ γd|N|2, (A63) Re = , (A52) ∂r 2πν so that where d is the vertical size of the domain and ν˜ is the vis- (γ − 1)d4|N|2 cosity. Ra ≈ (A64) With these pieces we can write the , de- να fined by equation (A47), as and, with these pieces, we may write the ratio 2 2 s Ta = (4π CoRe) (A53) Ω (γ − 1)Pr ≈ . (A65) and find |N| Ek2Ra Ra |N|2 Ro2 = = , (A54) The ratio of the axisymmetric azimuthal kinetic energy conv 2 Ta Pr 4Ω in the frame rotating at Ωframe to the total kinetic energy

MNRAS 000,1–27 (2020) Convective Differential Rotation II 23 was reported. In the language of Brown et al.(2008) this We compute the ratio of the convection speed to h|N| ratio is with equation (A64) to be DRKE s , (A66) 3 3 n Re2(γ − 1)Pr DRKE + CKE c ≈ c ≈ n , (A75) h|N| d|N| 2 from which we can compute DRKE/CKE. Ek Ra The Rossby number was also reported as where n is defined by equation (A72) and Re is the Reynolds ! number of fluctuations in the flow. ν h32i1/2 h32i1/2 Ro ≡ EkRe = = , (A67) We obtain the P´eclet number from equation (A17). Here Ωd2 νd Ωd we use the fluctuating Reynolds number, because the other Reynolds numbers reported by Gastine & Wicht(2012) are where we have inferred the definition of Re from the text broken down into specific spatial components. surrounding their equation (26) and h32i1/2 is the time- and volume-averaged root-mean square of the velocity. Putting these two numbers together we find r A3.14 Aurnou Simulations h32i1/2 DRKE |R∇ ln Ω| ≈ (A68) ΩR CKE These data are from simulations of three-dimensional hy- r drodynamic Boussinesq convection in rotating spherical d DRKE ≈ (A69) shells (Aurnou et al. 2007). They appear in the second sum- routerRo CKE mary figure of Gastine et al.(2014) and were extracted with r DRKE automated graphic data extraction software to provide α as ≈ (1 − η) . (A70) CKE in equation (A45). Points with α < 0.02 were removed from the sample because the plot was scaled linearly and below We computed the ratio of the convection speed to h|N| this limit the figure resolution was insufficient to ensure ac- as curacy. To measure the rotation we note that α is provided r 3 3 n  Ω   3  Ω DRKE as a function of the convective Rossby number, which Gas- c ≈ c = n c = n Ro 1 − , (A71) h|N| d|N| |N| Ωd |N| CKE tine et al.(2014) define where the final square root factor corrects the Rossby num- s Ek2Ra ber for the fraction of energy actually in convection, as op- Ro = . (A76) c Pr posed to differential rotation, and Inspection of equations (2-4) of (Aurnou et al. 2007) reveals  Nρ  n ≡ max 1, (A72) that this quantity is just |N|/Ω, from which we compute γ Ω/|N|. is the number of pressure scale heights in the domain, which We obtain the P´eclet number from equations (A17) we cut off at unity because that is the relevant mixing length and (A73). To find the Ekman number we note that the in the limit of no density stratification. Here Nρ is the re- quantity Ra∗ reported by Gastine et al.(2014) is not just ported number of density scale heights in the domain. We the convective Rossby number, but rather obtain the P´eclet number using the relation Pr − Ra∗ = Ra . (A77) Re = Ek 1Ro, (A73) Ek2 where Ek is the Eckman number, and use this Reynolds num- From this we can solve for Ek and thence for Pe. ber in equation (A17).

A3.15 Kaspi Simulations A3.13 Gastine 2013 Simulations These data are from three-dimensional anelastic general cir- These data are from three-dimensional hydrodynamic simu- culation simulations over a thick spherical shell (Kaspi et al. lations of anelastic convection in rotating spherical shells, re- 2009). They appear in the second summary figure of Gas- ported by Gastine et al.(2013). The Prandtl number, aspect tine et al.(2014) and were extracted with automated graphic ratio, domain size, poltropic indices are the same as those in data extraction software providing the quantity α given in appendix A3.12. The Rayleigh number and Eckman number equation (A45). These were extracted and processed as in are reported in the same manner but vary from simulation appendix A3.14. to simulation. We compute the Rossby number from these in the same way as in appendix A3.12. Gastine et al.(2013) report the time-averaged Rossby A3.16 Gilman Simulations number at the surface on the equator (Roe). We interpret this directly as a measure of differential rotation, so that These data come from simulations of three-dimensional Boussinesq convection in rotating spherical shells reported Ωequator,surface − Ωframe |R∇ ln Ω| = Ro ≡ , (A74) in Gilman(1977) and Gilman(1979). The results appear in e Ω frame the second summary figure in Gastine et al.(2014). These where Ωframe is defined as in appendix A3.12. were extracted and processed as in appendix A3.14.

MNRAS 000,1–27 (2020) 24 Adam S. Jermyn

A3.17 Rogers Simulations The polytropic index n is not specified, so we assume n = 3/2. The pressure scale height is then related to the These data are from three-dimensional hydrodynamic simu- density scale height by a factor of γ = 5/3, so lations of anelastic convection in a spherical domain reported by Rogers(2015). The results were taken from their table. h γhρ = , (A87) The table lists 3c for the two extremal convective forcing r r functions and so we linearly interpolate these as where hρ is the density scale height. The average density −1 3c = [2.9 + (q − 1.5) ∗ (4.5 − 2.9)] km s , (A78) scale height in the simulation domain is just the size of the domain divided by the number of scale heights it contains, where q equals their Q¯/3c and is the convective forcing. We take the scale height to be the size of their convection zone, so which is 0.3R . Using the Ω values in their table we obtain h γ∆r γ ≈ = (1 − χ). (A88) 3c/hΩ. We then employ the scaling in table2 to write r Nρr Nρ s Ω Ωh Inserting this into equation (A86) we find = (A79) |N| 3 c γ ξ ≈ (1 − χ)∂θ ln Nu. (A89) when Ω > |N| and Nρ Ω Ωh = (A80) We then estimate |N| 3 c 2 Nuequator − Nupole when Ω < |N|. Note that this definition assumes that our ∂θ ln Nu ≈ 2 (A90) π Nu + Nu scaling for the convective velocity holds, and so we cannot equator pole use these data to test that scaling. and with equation (A89) obtain ξ. The differential rotation is reported as the ratio of the The table contains the Ekman number core-to-envelope angular velocity. We estimate the shear ν from this with equation (A9). Ek = . (A91) 2 To obtain the Reynolds number we use their reported Ωframed 13 2 −1 viscosity ν = 4 × 10 cm s and let The Rayleigh number is reported as 3c h Re = . (A81) gdr2∆s ν Ra = , (A92) να We then determine the P´eclet number through equa- tion (A17) using the Prandtl number where we have used α = κcp to convert from their notation ν to ours. Using d/r = 1 − χ we see that this Rayleigh number Pr = , (A82) −1 α differs from that of equation (A62) by a factor of γ(γ−1) (1− χ)−2, so we may use equation (A65) to find where the α = 5 × 1011cm2 s−1. s Ω γPr ≈ . (A93) A3.18 Raynaud Simulations |N| Ek2(1 − χ)2Ra These data are from simulations of three-dimensional anelas- tic hydrodynamic convection in rotating spherical shells, re- ported in table B.1 of Raynaud et al.(2018). The table con- A4 MHD Simulations tains the number of density scale heights in the simulation We describe how we extract the relevant data from various domain (Nρ), the ratio of the outer radius of the domain to the inner radius (χ) and the Nusselt number (Nu) at both MHD simulations. the equator and pole. The Nusselt number is proportional to the radial entropy gradient. The latitudinal derivative of this is then A4.1 Soderlund Simulations ∂θ Nu ∝ ∂r ∂θ s, (A83) These data are from simulations of three-dimensional so Boussinesq MHD convection in rotating spherical shells of varying thickness (Soderlund et al. 2013). These appear in ∂r ∂θ s ∂ ln Nu = . (A84) the second summary figure in Gastine et al.(2014) and were θ ∂ s r extracted with automated graphic data extraction software, If we estimate that the radial derivative in the numerator to providing the quantity α given in equation (A45). Points produce a factor of order the pressure scale height then with α < 0.02 were removed from the sample because the ∂θ s plot was scaled linearly and below this limit the figure res- ∂θ ln Nu ≈ . (A85) h∂r s olution was insufficient to ensure accuracy. To measure the rotation we note that α is provided By equation (A20) we then obtain as a function of the convective Rossby number, defined as h |N|/Ω. We obtain the P´eclet number from equations (A17) ξ ≈ ∂ ln Nu. (A86) r θ and (A73).

MNRAS 000,1–27 (2020) Convective Differential Rotation II 25

A4.2 Mabuchi Simulations (MHD) reported in tables 1 and 2 of Varela et al.(2016). The differ- ential rotation is reported as the difference in angular veloc- These data are from three-dimensional fully compressible ity between the equator and a latitude of π/3, normalized to MHD simulations of rotating convection in spherical shells, the equatorial angular velocity, so we follow equation (A11) reported by Mabuchi et al.(2015). Data were taken from to compute |R∇ ln Ω|. their table 1. The reported differential rotation is in the The Rossby number actually realised in the flow is re- form of α as defined by equation (A45). The data analysis ported in the form of equation (A18). We calculate Ω/|N| was performed in the same way as for the hydrodynamic from this as in appendix A3.1. The magnetic energy and case in appendix A3.10. Mabuchi et al.(2015) also report the convective kinetic energy are also reported and used to the magnetic energy and the kinetic energy in the meridional calculate 32 /32. plane and we take the ratio of these to be 32 /32. A c A c We compute the ratio 3 /h|N| in the same way as we We obtain the P´eclet number from equation (A17). We c did in appendix A3.2, using the Rossby, Prandtl, Taylor and calculate the Reynolds number using the height of the un- Rayleigh numbers. The one difference is that the Rayleigh stable layer (0.3), the thermal diffusivity (3.9 × 10−4), and number is reported as in equation (A14), so we must divide the reported root-mean-square velocities, all in code units. it by (γ − 1)/γ before using equation (A34). We obtain the P´eclet number from equation (A17). A4.3 Brun 2005 Simulations

These data are from three-dimensional anelastic MHD simu- A4.6 Yadav 2013 Simulations lations of convection in a rotating spherical domain, reported by Brun et al.(2005). The differential rotation is reported as These data are from Boussinesq three-dimensional MHD the mass-weighted average of the squared difference between simulations of rotation convection in a spherical shell do- the rotation rate and the mean rotation rate. We translate main, reported in table A.2 of Yadav et al.(2013a). The this into a shear as we did for the data of Brown et al.(2008), Rayleigh number is reported as taking into account the scaling of 3c in MHD systems given gd3∆T d ln V in Table1. Ra = , (A95) αν dT The Rossby number actually realised in the flow is also reported. This they define as where d ln V/dT is the thermal expansivity. The convention given in their equation (1) shows that the modified Rayleigh 3 Ro = c . (A94) number 2ΩR RaEk2 For this calculation we take h ≈ R because the average depth Ra∗ = (A96) is comparable to the radius. We then calculate Ω/|N| from Pr this as in appendix A3.1. equals |N|2/Ω2, so we compute this from the given Ra, Ek The magnetic energy and the convective kinetic energy and Pr and thereby obtain Ω/|N|. 32 /32 are also reported. We use this to calculate A c . We com- Yadav et al.(2013a) provide the Rossby number num- pute the ratio 3c/h|N| in the same way as we did in ap- ber, from which we obtain the Reynolds number pendix A3.1. The one difference is that Brun et al.(2005) −1 do not report the adiabatic exponent γ, so we assume it is Re = RoEk . (A97) 5/3. We then find Brown et al.(2008) report the P´eclet number so we do 3 Reν not need to compute it. c = . (A98) h|N| dh|N| 2 A4.4 Augustson 2013 & 2016 Simulations Letting h ≈ d and using ν = Ωd Ek we find s r These data come from anelastic three-dimensional MHD 3 ReEkΩ Pr Pr c = = ReEk = Re . (A99) simulations of rotating convection in a spherical do- h|N| |N| RaEk2 Ra main (Augustson et al. 2013) and spherical shells (August- son et al. 2016). The differential rotation is reported as the We use their Ro` in place of Ro because Yadav et al.(2013a) mass-weighted average of the squared difference between the report the former and say that it is more reflective of the rotation rate and the mean rotation rate. We translate this ratio of inertial to Coriolis forces than the latter. The zonal into a shear as we did in appendix A3.4, taking into account Rossby number is also provided, and we interpret this as the scaling of 3c in MHD systems given in Table1. The data giving analysis was performed in the same way as for the hydrody- |R∇ ln Ω| ≈ Ro . (A100) namic case. The magnetic energy and the convective kinetic zonal 32 /32 energy are also reported. We use this to calculate A c . We We next compute the P´eclet number as obtain the P´eclet number from equation (A17). Pe = RePr. (A101) Finally, the Lorentz number is reported as A4.5 Varela 2016 Simulations B2 These data are from anelastic three-dimensional MHD sim- Lo ≡ h i1/2, (A102) ulations of rotating convection in a spherical shell domain, 4πρΩ2d2

MNRAS 000,1–27 (2020) 26 Adam S. Jermyn where h...i denotes a volume average and d is the domain A6 Jupiter Data thickness. From this we obtain The profile of differential rotation in the surface layers of 2 2 2 2 Jupiter was taken to be that given by Kaspi et al.(2018). 3A = Lo Ω d , (A103) This yields the variation of the characteristic zonal flow rate so with depth but does not provide the shear itself. To com- pute this we first note that roughly 30 per-cent of Jupiter’s 32 A ≈ Lo2Ro−2. (A104) surface is covered in latitudinal bands with velocities of or- 2 ` −1 3c der 100 km s , while the remaining latitudes contain bands with velocities of order 25km s−1 (Kaspi et al. 2018). The former are also approximately twice as wide per band as the latter. There are approximately 15 bands in total, and so we A5 Solar Data estimate the typical shear at the surface to be A helioseismic inversion of the solar profile was obtained 15  1  from Antia(2016). It corresponds to that appearing in An- |R∇Ω| ≈ 0.3 × 25km s−1 + × 0.7 × 100km s−1 , (A110) πR 2 tia et al.(2008). This includes Ω as a function of position J 9 throughout the convective envelope as well as in the upper where RJ ≈ 7×10 cm is the radius of Jupiter. We then assume portions of the radiative envelope. This profile was then su- that the band velocities scale with depth following the profile persampled on to a grid running from 0.5R to R in the   r−a−H   radial direction, with 100 uniformly spaced points, and from  1 + tanh ∆H  3 ∝ ( − ) (r−α)/H  0 to π in the latitude, with 70 uniformly spaced points. By φ, band  1 α e + α    (A111)  1 + tanh H  applying a differentiating Gaussian filter with width equal  ∆H  to five grid points in each of the radial and latitudinal direc-   found by Kaspi et al.(2018), where α, H, a and ∆H are given tions we computed R∇Ω/Ω from this profile. We then aver- therein. The shear then scales as aged each component of this over latitude, weighted by sin θ, 3φ, band and took the square root of the result to produce a measure |R∇Ω| ∝ . (A112) of the mean radial and latitudinal differential rotation at r each radial slice. We further assume that the mean rotation period remains The radial profile of |N| was obtained from a MESA model 9.92hr (Kaspi et al. 2018). of a 1 M star with metallicity Z = 0.02 at an age of 4.6 Gyr. For regions deeper than 3000 km we bound the band ve- This was used to compute Ω/|N| everywhere in the solar locity above by the inferred 6 m s−1 (Guillot et al. 2018) and convection zone. We then averaged Ω2/|N|2 over latitudes, use the band structure above to compute the shear. Fur- weighting by sin θ, and computed the square root to deter- thermore we treat the number of bands as varying linearly mine the mean Ω/|N| in each radial slice. in radial coordinate between 1 near the centre of the planet Altrock & Canfield(1972) provide an upper bound on and 15 at the surface. This is an approximation of the cylin- the temperature difference between the solar pole and equa- drical nature that Kaspi et al.(2018) assumed for Jupiter’s tor. Following equation (A20) we write differential rotation. Following this assumption we consider this estimate to be for both the spherical-radial and latitu- ∂ s ξ ≈ θ . (A105) dinal shears. r∂ s r All that remains is to compute |N|0. To do this we We then note that, from equation (1), ran a MESA model of a Jupiter-mass planet irradiated by 12500 erg cm−2 s−1 at a pressure of 1 bar. This produced a γ |N|2 model within 3 per cent of Jupiter’s radius and an effective ∂ s ≈ (A106) r γ − 1 g temperature of 136 K, close to the Jovian 1 bar temperature of 165 K (Ingersoll et al. 2004). To ensure efficient convection so we exclude the and the top scale height of the g(γ − 1) convection zone from our analysis. The resulting profile of ξ ≈ ∂θ s (A107) γr|N|2 |N|0 is quite close to what the order of magnitude estimate  1/3 and inserting equation (A27) we find F |N|0 ≈ (A113) ρz3 g ξ ≈ ∂θ ln T. (A108) r|N|2 would suggest. Here F is the heat flux, ρ is the density and z is depth from the surface. Because the temperature measurement is at the surface we compute ξ using |N|2, g and r at the surface, estimating APPENDIX B: EXPANSION  2  Tpole − Tequator ∂θ ln T ≈ . (A109) π Teff Using equations (2) and (1) we may relate the increase in |N| to a change in the pressure and density gradients. For a The factor of 2/π accounts for the number of radians in- radial gravity field volved in estimating the latitudinal derivative from a finite difference. To plot this we take the rotation rate of the Sun 2 1 N = − g (∂r ln P − γ∂r ln ρ) . (B1) to be its mean rate and compute Ω/|N| for the surface |N|. γ

MNRAS 000,1–27 (2020) Convective Differential Rotation II 27

Latitudinal Spot Observations show the results of a variety of different choices on the shape of the latitudinal shear data from Lurie et al.(2017). In par- 100 Prediction (MHD) ticular, we show weighting by log r, r, and r2 and averaging logr-log|N|0 in log |N|0, as well as weighting by log r and averaging in r-log|N|0 p 2 10 1 log |N| , |N| , |N| , and |N| . Different choices of averaging 2 0 0 0 0 r -log|N|0 and weighting produce very different overall scales for |N|0, logr-|N|0 by up to a factor of 102, but do a good job at preserving the | 10 2 logr- |N|0 shape of the data (i.e. slope of shear versus Ω). Because we n l 2 logr-|N|0

| are primarily comparing our theory with the shape of the

10 3 data, and admit uncertainty as to the transition between the slowly-rotating and rapidly-rotating limits, the precise choice of averaging therefore makes only a small difference 10 4 to our comparisons.

10 2 100 102 104 106

/|N|0 APPENDIX D: SENSITIVTY TO STELLAR MODELLING

Figure C1. The relative latitudinal shear |∂θ ln Ω| is shown as In addition to the systematic uncertainty in |N|0 arising from a function of Ω/|N |0 for observed convecting stars inferred from uncertainties in averaging and weighting, there are uncer- star spot measurements by Lurie et al.(2017) along with our tainties involved in matching stellar models to the observed predictions. The same data are shown with different colours rep- characteristics of stars. In matching main-sequence models resenting different averaging schemes for |N |0 which weight by to observations we have relied on the observed effective tem- different factors in radius and average different functions of |N |0. peratures, and for giants with asteroseismic observations we have directly used the inferred masses, ages and metallici- Inserting equation (3) and rearranging, we find ties. These matching exercises are subject to several uncer- d ln ρ 1 hN2 tainties. First, different stellar models constructed with dif- ≈ − . (B2) d ln p γ g ferent software instruments typically disagree on observable properties at the 1 − 5 per-cent level even with identical in- For Ω  |N| in the MHD limit the density gradient changes 0 put physics (Stancliffe et al. 2015). Next, input physics like by an amount opacities are uncertain at the 10 − 20 per-cent level (Bailey 2 !  d ln ρ  h h Ω2/7 hN et al. 2015). In addition different choices of age on the main- δ ≈ (|N|2 − |N|2) ≈ − |N|2 ≈ − 0 . 0 / sequence, or of Helium abundance, can also lead to changes d ln p g g |N|2 7 g 0 in effective temperature on this level. (B3) Most importantly, though, 1D stellar models rely on If this change in the density gradient results in a change in mixing length theory, which is a highly simplified prescrip- the mean density of the convection zone of the same order, tion for convection. One known source of uncertainty in this the radius of the star is changed by an amount theory is the mixing length parameter αMLT, which is seen to vary by as much as 20 per-cent between stars which differ 2/7 2 ! δR d  d ln ρ  d Ω hN by 0.2M (Joyce & Chaboyer 2018). Naively extrapolating ≈ − cz δ ≈ cz 0 , (B4) R R d ln p R | |2/7 g this over the range of stellar masses we have considered, N 0 this perhaps leads to a factor of 2 uncertainty in αMLT. In −2/3 where dcz is the depth of the convection zone. limit of efficient convection, |N| ∝ α (see e.g. Cantiello & 0 MLT√ √ By comparison the bloating owing to centrifugal effects Braithwaite 2019, and note that |N| ∝ ∇ − ∇ ∝ ∇ − ∇0), is 0 ad so the dominant source of uncertainty in |N|0 would seem to δR Ω2R be that in the mixing length, and more broadly in using such ≈ , (B5) R g a simplified theory for convection in the first place. which is larger by a factor This paper has been typeset from a TEX/LATEX file prepared by 2 12/7 (δR/R)centrifugal R  Ω  the author. ≈  1. (B6) (δR/R)convective hdcz |N|0

APPENDIX C: SENSITIVITY TO |N| AVERAGING As described in AppendixA and Section 3.1, stellar mod- els produce profiles of |N|0 rather than a single number. To compute an average |N|0 from such a model for compari- son, we have had to make a choice (equation (7)) of how to weight different parts of the stellar model. In Fig. C1 we

MNRAS 000,1–27 (2020)