Articles

A SIMULATION OF CONVECTIVE DYNAMO IN THE SOLAR CONVECTIVE ENVELOPE: MAINTENANCE OF THE SOLAR-LIKE DIFFERENTIAL ROTATION AND EMERGING FLUX

and

Published 2014 June 11 © 2014. The American Astronomical Society. All rights reserved.
, , Citation Yuhong Fan and Fang Fang 2014 ApJ 789 35 DOI 10.1088/0004-637X/789/1/35

0004-637X/789/1/35

ABSTRACT

We report the results of a magnetohydrodynamic (MHD) simulation of a convective dynamo in a model solar convective envelope driven by the solar radiative diffusive heat flux. The convective dynamo produces a large-scale mean magnetic field that exhibits irregular cyclic behavior with oscillation time scales ranging from about 5 to 15 yr and undergoes irregular polarity reversals. The mean axisymmetric toroidal magnetic field is of opposite signs in the two hemispheres and is concentrated at the bottom of the convection zone. The presence of the magnetic fields is found to play an important role in the self-consistent maintenance of a solar-like differential rotation in the convective dynamo model. Without the magnetic fields, the convective flows drive a differential rotation with a faster rotating polar region. In the midst of magneto-convection, we found the emergence of strong super-equipartition flux bundles at the surface, exhibiting properties that are similar to emerging solar active regions.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Despite the turbulent nature of solar convection, the Sun's large-scale magnetic field exhibits remarkable order and organization such as the 11 yr sunspot cycle (e.g., Maunder 1922) and the Hale's polarity rule of the bipolar active regions (Hale et al. 1919; Hale & Nicholson 1925). In recent years, global fully dynamic three-dimensional (3D) convective dynamo simulations have been making headway in producing the solar-like cyclic behavior of the large-scale magnetic field (e.g., Ghizaru et al. 2010; Racine et al. 2011; Käpylä et al. 2012; Augustson et al. 2013) and the self-consistent formation of buoyant, active region like emerging tubes from dynamo generated strong toroidal fields (Nelson et al. 2011, 2013, 2014). Most of these simulations have differential rotation with cylindrical iso-rotation contours throughout the convection zone (CZ; see review by Charbonneau 2013), and some are considering rotation rate three times the solar rate. In this paper, we present a convective dynamo simulation driven by the solar radiative diffusive heat flux and maintains a differential rotation profile that resembles more closely to the solar differential rotation in the CZ in terms of the pole-equator contrast and the more conical iso-contours of rotation in the mid-latitude region. The convective dynamo produces a large-scale mean magnetic field with irregular cyclic behavior and polarity reversals very similar to a convective dynamo presented in Miesch et al. (2011). We demonstrate in this paper the important role the magnetic fields play in the maintenance of the solar-like differential rotation. We also show the emergence of strong super-equipartition flux tubes near the surface that exhibit some properties similar to emerging solar active regions.

2. THE NUMERICAL MODEL

We solve the following anelastic MHD equations using a finite-difference spherical anelastic MHD code (Fan 2008; Fan et al. 2013):

Equation (1)

Equation (2)

Equation (3)

Equation (4)

Equation (5)

Equation (6)

Equation (7)

In the above, s0(r), p0(r), ρ0(r), T0(r), and ${\bf g} = - g_0 (r) {\hat{\bf r}}$ denote the profiles of entropy, pressure, density, temperature, and the gravitational acceleration of a time-independent reference state of hydrostatic equilibrium and nearly adiabatic stratification, cp is the specific heat capacity at constant pressure, γ is the ratio of specific heats, and v, B, s1, p1, ρ1, and T1 are the velocity, magnetic field, entropy, pressure, density, and temperature to be solved that describe the changes from the reference state. Ω denotes the solid body rotation rate of the Sun and is the rotation rate of the frame of reference, where Ω = 2.7 × 10−6 rad s−1. ${\cal D}$ is the viscous stress tensor: ${\cal D}_{ij} = \rho _0 \nu [ S_{ij} - (2/3)(\nabla \cdot {\bf v}) \delta _{ij} ]$, where ν is the kinematic viscosity, δij is the unit tensor, and Sij is the strain rate tensor given by the following in spherical polar coordinates:

Equation (8)

Equation (9)

Equation (10)

Equation (11)

Equation (12)

Equation (13)

K denotes the thermal diffusivity, and η is the magnetic diffusivity. In Equation (3),

Equation (14)

is the radiative diffusive heat flux, where σs is the Stephan–Boltzman constant, κ is the Rosseland mean opacity.

The simulation domain is a partial spherical shell with r ∈ [ri, ro], spanning from ri = 0.722 Rs at the base of the CZ to ro = 0.971 Rs at about 20 Mm below the photosphere, where Rs is the solar radius, θ ∈ [π/2 − Δθ, π/2 + Δθ] with Δθ = π/3, and ϕ ∈ [0, 2π]. The domain is resolved by a grid with 96 grid points in r, 512 grid points in θ, and 768 grid points in ϕ. J. Christensen-Dalsgaard's (JCD) solar model (Christensen-Dalsgaard et al. 1996) is used for the reference profiles of T0, ρ0, p0, and g0 in the simulation domain. We assumed that s0 = 0 for the reference state. The heating (the last term in Equation (3)) due to the solar radiative diffusive heat flux drives a radial gradient of s1 that drives the convection. We set the thermal diffusivity K = 3 × 1013 cm2 s−1, the viscosity ν = 1012 cm2 s−1, and the magnetic diffusivity η = 1012 cm2 s−1 at the top of the domain, and they all decrease with depth following a $1/ \sqrt{\rho _0}$ profile. The stratification of the domain includes approximately four density scale heights between the top and the bottom, and thus the above diffusivities decrease to K = 4.02 × 1012 cm2 s−1, ν = 1.34 × 1011 cm2 s−1, η = 1.34 × 1011 cm2 s−1 at the bottom of the CZ domain. The rationale for our choice of such depth-dependent diffusivities for the numerical experiments here is that, if the dominant spatial scales of convection decreases with height, it may be expected that more heat is transported by the unresolved scales as one moves toward the top of the simulation domain, and hence the greater diffusivities there. Furthermore, with a low magnetic diffusivity and viscosity in the deep CZ, should buoyant magnetic structures develop, they would be able to better preserve their magnetic buoyancy and rise. Given the above diffusivities, the various diffusive timescales for the simulation are estimated as follows. The viscous and magnetic diffusive timescales (Δr)2/ν and (Δr)2/η range from about 71 yr near the bottom of CZ to about 10 yr near the top, and the thermal diffusive timescale (Δr)2/K ranges from about 2.4 yr near the bottom to about 0.3 yr near the top, where we have used the depth of the CZ domain Δr for the estimate.

We impose ∂s1/∂r = 0 at the bottom and s1 = 0 at the top boundary. We also impose a latitudinal gradient of entropy at the lower boundary:

Equation (15)

where Δsi = 431.4 erg g−1 K−1, corresponding to a pole to equator temperature difference of about 6.8 K, to represent the tachocline induced entropy variation that can break the Taylor–Proudman constraint in the CZ (Rempel 2005). At the two θ boundaries, s1 is assumed symmetric. The velocity boundary condition is non-penetrating and stress-free at the top, bottom, and the two θ boundaries. For the magnetic field we assume perfect conducting walls for the bottom and the θ boundaries and radial field at the top boundary. All quantities are naturally periodic at the ϕ boundaries.

For the initial state, we specify the initial s1 such that its horizontal average: 〈s1t = 0, satisfies

Equation (16)

where Ls is the solar luminosity, Frad is the absolute magnitude of Frad given in Equation (14), and at the lower boundary ri, Ls/4πr2 = Frad. Equation (16) lets the initial thermal conduction together with radiative diffusion completely carry the solar luminosity, which sets up an initial unstable entropy gradient 〈s1t = 0. We start the simulation with a small initial seed magnetic and velocity field and let the magneto-convection evolve to a statistical steady state.

3. RESULTS

3.1. Overview of the Convective Dynamo

Figure 1 shows the magnetic and kinetic energies in the statistically steady convective flows in the simulation domain over a time span of about 74 yr. The total magnetic energy Em maintained by the dynamo is about 10% of the total kinetic energy Ek of the convective envelope. The energy of the azimuthally averaged (mean) magnetic field Em, mean only constitutes a small fraction of Em, oscillating from about 1% to 10% of Em.

Figure 1.

Figure 1. Temporal variation of the total kinetic energy (Ek), total magnetic energy (Em), and the azimuthally averaged mean magnetic field energy (Em, mean) of the statistically steady convective flows in the simulation domain.

Standard image High-resolution image

Figure 2 shows the depth variation of the mean entropy gradient established in the CZ domain in the statistical steady state. The entropy gradient reaches a value of about 4.26 × 10−6 erg−1 K−1 cm−1 at the top boundary at about 0.97 Rs, which is of a similar order of magnitude as the entropy gradient (∼10−5 erg−1 K−1 cm−1) at this depth in the solar model of JCD. Figure 3 shows the various horizontally integrated energy fluxes (normalized to the solar luminosity Ls) through the domain as a function of radius established in the statistical steady state. These are, respectively, the integrated radiative diffusive heat flux (red curve): Lrad = 4πr2Frad, the convective enthalpy flux (black curve):

Equation (17)

the conductive energy flux by thermal diffusion (yellow curve):

Equation (18)

the kinetic energy flux (blue curve):

Equation (19)

the viscous energy flux (black dashed curve):

Equation (20)

the Poynting flux (green curve):

Equation (21)

the resistive energy flux (cyan curve):

Equation (22)

and the sum of all the energy fluxes, Ltot, is shown as the dash–dotted curve in Figure 3. In the above, 〈〉 denotes averaging over the spherical shell surface and time. The gradual decline of Ltot reflects a numerical deviation from exact energy conservation, which is mainly caused by the numerical diffusion of the magnetic field due to the Alfvén wave-upwind scheme used for advancing the induction equation (Fan 2008; Stone & Norman 1992). This numerical dissipation of magnetic energy is not being put back into the thermal energy in the entropy equation and results in a loss of the total energy and hence a loss of about 13% of the total energy flux exiting the domain at the top compared to the total energy flux (Ls) entering the domain from the bottom. We note that the explicit resistive dissipation of the magnetic field (due to η), and both the explicit viscous dissipation (due to ν) and the numerical diffusion of momentum are put into the thermal energy in the entropy equation as resistive and viscous heating to maintain energy conservation. From Figure 3, it can be seen that the enthalpy flux of the resolved convection transports about 66% of the solar luminosity in the middle of the CZ, and due to the high thermal diffusivity K, thermal conduction also transports a substantial fraction of the solar luminosity (about 36% at the middle of the CZ). The kinetic energy flux of the convective flows is downward and peaks at about 16% of the solar luminosity. The energy fluxes due to the Poynting flux (mostly downward), resistive, and viscous transport are all much smaller.

Figure 2.

Figure 2. Depth variation of the mean entropy gradient established in the statistical steady state of the dynamo simulation.

Standard image High-resolution image
Figure 3.

Figure 3. Horizontally integrated energy fluxes due to respectively, radiative diffusion Lrad (red curve), convection Lconv (black curve), thermal conduction Lcond (yellow curve), kinetic energy flux Lkin (blue curve), viscous flux Lvis (dashed line), Poynting flux Lpoyn (green curve), resistive flux Lres (cyan curve), and the sum of all Ltot (dash–dotted black curve), as a function of depth (see the text for the expressions of Lrad, Lconv, Lcond, Lkin, Lvis, Lpoyn, and Lres).

Standard image High-resolution image

Figure 4 shows the depth variation of the peak downflow (solid black curve), and the rms speed vrms (dash–dotted black curve), of the statistical steady convective flows in the domain. Note that in computing vrms, we take out the azimuthally averaged velocity components and only sum up the azimuthally fluctuating parts of the velocity components. Also shown are the corresponding magnetic field strength in equipartition with the peak downflow speed (solid red curve) and the rms speed (dash–dotted red curve). It can be seen that the equipartition field strength Beq corresponding to the peak down flow speed reaches ≈63 kG, while Beq corresponding to the rms speed is ∼10 kG for the deep and mid-convection zone, and decrease to about 5000 G near the top boundary at about r = 0.971 Rs. Following Käpylä et al. (2012), we compute the following non-dimensional numbers characterizing the convective flows. The Reynolds number Re = urmskf ranges from about 130 at the bottom to about 50 at the top, and with a mid-convection zone value of about 128, where kf = 2π/(rori) and urms = ((3/2)〈vr2 + vθ2〉) is the rms velocity averaged over each depth, omitting the contribution from the azimuthal velocity. The Coriolis number CO = (2Ω/urms, allkf) = 1.3, where urms, all = ((3/2)〈vr2 + vθ2〉) with the averaging 〈〉 done for the entire domain. We can compare the values of these non-dimensional numbers with the corresponding ones in Käpylä et al. (2012): Re = 36 and CO = 7.6. It appears the convective flow in our dynamo simulation is moderately more turbulent as characterized by the larger Re, especially in the deeper layers of the CZ. Our Coriolis number CO is significantly lower, indicating that our convective dynamo is operating in a significantly less rotationally dominant regime. If we were to scale their typical rms velocity to be similar to ours urms, all ≈ 100 m s−1, then their CO would imply a significantly more rapidly rotating stellar envelope (with the solar CZ depth) at about five times the solar rotation rate.

Figure 4.

Figure 4. Peak downflow speed, the rms convective speed, and their corresponding equipartition magnetic field strengths as a function of depth.

Standard image High-resolution image

Figure 5(a) shows the latitude-time variation of the mean (azimuthally averaged) toroidal magnetic field at a depth near the bottom of the CZ. The mean toroidal magnetic field tends to be of opposite signs for the two hemispheres, and exhibits an irregular cyclic behavior with oscillations of the field strength on timescales ranging from about 5 to about 15 yr and undergoes irregular sign/polarity reversals. The strongest mean toroidal field is concentrated near the bottom of the CZ (see Figure 5(b)), peaking at about 7 kG. Figure 5(c) shows a shell-slice of Bϕ at a depth near the bottom of the CZ, at a cycle maximum phase indicated by the green line in Figure 5(a). It shows that strong toroidal fields Bϕ of a preferred sign (opposite for the two hemispheres) are concentrated in individual channels or filaments in each hemisphere, reaching a peak field strength of about 30 kG, which exceeds the field strength in equipartition with the local rms convective speed (Beq ≈ 13 kG), but is below the equipartition field strength corresponding to the peak down flow speed (Beq ≈ 63 kG). Thus, these strong field filaments are not passively advected by convective flows, but would be pinned down by the strong down flows if in their paths.

Figure 5.

Figure 5. (a) Latitude-time variation of the mean (azimuthally averaged) toroidal magnetic field at a depth (r = 0.73 Rs) near the bottom of the CZ. (b) Azimuthally averaged toroidal magnetic field distribution in the meridional plane at the time marked by the green line in panel (a). (c) A shell slice of the toroidal magnetic field at a depth (r = 0.73 Rs) near the bottom of the CZ at the same time marked by the green line in panel (a).

Standard image High-resolution image

3.2. Maintenance of the Solar-like Differential Rotation

Figure 6(a) shows the time and azimuthally averaged rotation rate in the convective envelope self-consistently maintained in the convective dynamo simulation. It shows a solar-like differential rotation profile (e.g., Thompson et al. 2003) with a faster rotation rate at the equator than at the polar region by about 30% of the mean rotation rate, and more conical shaped iso-rotation contours in the mid-latitude zone. The time and azimuthally averaged mean meridional flow pattern is shown in Figure 6(b) in terms of the mass flux function f where $\rho _0 \langle {\bf v} \rangle = \nabla \times [(f / r \sin \theta) {\hat{\bf \phi }} ]$. The meridional circulation has a complex multi-cell structure with a counter-clockwise (clockwise) cell pattern in the low latitude region of the northern (southern) hemisphere, i.e., a poleward near-surface flow in the low latitude region.

Figure 6.

Figure 6. Top row panels from left to right show the time and azimuthally averaged mean meridional profile of angular velocity (a), meridional flow mass flux (b), angular momentum flux density in the r direction due to the Reynolds stress (c), the viscous stress (d), and the Maxwell stress (e) resulting from the dynamo simulation. Middle row panels from left to right show the same as those of the top row except for the results from the corresponding HD simulation and there is not a panel for the Maxwell stress. Bottom row panels from left to right show the same as the middle row, except for the results from the HVHD simulation. See the text for the expressions for the various angular momentum flux density RS, VS, and MS.

Standard image High-resolution image

Interestingly, we find that the presence of the magnetic field is necessary for the self-consistent maintenance of the solar-like differential rotation profile in the current parameter regime. We have carried out a hydrodynamic simulation (hereafter referred to as the HD case) which is identical to the present convective dynamo simulation except that the magnetic field is set to zero, and found that a very different differential rotation profile (Figure 6(f)) is established in the statistical steady state of the HD simulation. It shows a significantly larger differential rotation with a faster rotation rate in the polar region than at mid-latitudes and the equator. The iso-rotation contours are also more cylindrical. The meridional flow (Figure 6(g)) shows a more prominent counter-clockwise (clockwise) cell pattern in the northern (southern) hemisphere in the middle depths in the CZ, with much weaker reversed cells in the near surface layer.

To compare the transport of angular momentum in the dynamo and the HD cases, we show in Figures 6(c)–(e) the meridional profile of the angular momentum flux density in the r direction (perpendicular to and away from the rotational axis) due, respectively, to the Reynolds stress of the rotationally influenced convection (panel (c)):

Equation (23)

the viscous stress (panel (d)):

Equation (24)

and the Maxwell stress (panel (e)):

Equation (25)

where 〈〉 denotes time and azimuthal averages and ' denotes the azimuthally varying component. The meridional profiles of the angular momentum flux density RS and VS for the corresponding HD case are shown in Figures 6(h) and (i). We find that there is a significant difference in the angular momentum flux density by the Reynolds stress between the dynamo and HD cases. The RS for the dynamo case shows an overall more outward transport in its meridional distribution compared to that for the corresponding HD case. Near the lower boundary in the mid-latitude range, the presence of the concentrated magnetic field (which helps to damp the convective downflows) results in a more enhanced outward RS flux at the lower boundary layer. The angular momentum flux density MS due to the Maxwell stress in the dynamo case (Figure 6(e)) is found to oppose RS, and its strength is greatest at the lower boundary layer where the magnetic field concentrates. The angular momentum flux density VS simply acts to reduce the differential rotation as expected for both the dynamo and the HD cases. The difference between the dynamo and the HD cases is more clearly seen by evaluating the net angular momentum fluxes in the r direction integrated over individual concentric cylinders of radii r centered on the rotation axis, as shown in Figure 7(a) for the dynamo case and Figure 7(b) for the corresponding HD case. It can be seen that the net angular momentum flux due to RS is outward throughout (except near the top boundary) for the dynamo simulation (black curve in Figure 7(a)), which drives a faster rotation in the outer equatorial region. The net angular momentum flux due to RS is mainly counteracted by the net flux due to MS by the Maxwell stress, with the remaining difference balanced by the significantly smaller net fluxes due to VS and the meridional flow. In contrast, the HD simulation shows a significant inward angular momentum transport due to the Reynolds stress (black curve in Figure 7(b)) across the inner cylinders in the high to mid-latitude region. This drives a faster rotation in the polar region. Thus, it appears that the presence of the magnetic field alters the convective flows such that the resulting Reynolds stress from the convective motions produces a more outward (away from the rotation axis) net transport of the angular momentum needed to drive a solar-like differential rotation. The Rossby number RO = vrms, all/(ΩHp) is about 0.74 for the dynamo case and 0.96 for the HD case, where vrms, all = 125 m s−1 for the dynamo case and 157 m s−1 for the HD case, is the rms velocity (with the azimuthally averaged mean flow velocity taken out) averaged over the entire volume, and Hp is the pressure scale height at the bottom of the CZ. The Rossby number measures the importance of the Coriolis force in the force balance. The lower Rossby number in the dynamo simulation shows that the magnetic fields suppress the convective motions so that they are more rotationally constrained.

Figure 7.

Figure 7. Integrated angular momentum fluxes across cylinders centered on the rotation axis as a function of the radial distance r from the rotation axis for, respectively, the dynamo case (a), the corresponding HD case (b), and the HVHD case (c). The r of the tangent cylinder of the base of the CZ is 0.722 Rs.

Standard image High-resolution image

A recent systematic study by Gastine et al. (2014) of rotating stellar convection considering a wide range of models shows that the differential rotation profile transitions from being solar-like, with a faster rotating equator, to being anti-solar, with a faster polar rotation rate, at a value of about 1 for the Rossby number. This result is found to be quite general, independent of the detailed model setup (presence of a magnetic field, thickness of the convective layer, density stratification). Our HD case with RO = 0.96 appears to be very close to the transition, and a reduction of ∼23% of the overall rms velocity and RO by the presence of the magnetic field in the dynamo case is able to significantly alter the angular momentum transport by the rotationally constrained convection, leading to a transition into the solar-like differential rotation. We note that even though the HD case is quite close to the transition, its anti-solar differential rotation appears to be a stable solution not dependent on the history, i.e., not one of two bistable states (Gastine et al. 2014). We have arrived at the statistically steady HD solution with the anti-solar differential rotation by either starting from the initial setup with a seed velocity field as described in Section 2, or by starting from the statistically steady state dynamo solution (for which the differential rotation is solar-like) and zero out the magnetic field. The solar-like differential rotation obtained in the convective dynamo simulation also appears to be a stable solution that is not dependent on the history. The differential rotation at the pole and equator remain statistically steady without exhibiting any systematic drift for the ≳ 74 yr period (comparable to the maximum viscous timescale near the bottom of CZ) we have run after the dynamo solution has reached a statistical steady state. Also, as we start the dynamo simulation from the initial setup described in Section 2, we find that the convective dynamo goes through an earlier phase of anti-solar differential rotation (for ∼6 yr) before it evolves toward the solar-like differential rotation profile as the mean entropy gradient and the convective energy flux settle down to their statistical steady state. Thus, it appears that the solar-like differential rotation is the preferred stable solution in the dynamo case.

The transition to a solar-like differential rotation can alternatively be achieved in the non-magnetic hydro simulations by simply increasing the viscosity to reduce RO. We have run another hydrodynamic simulation (hereafter referred to as the HVHD case, meaning "high viscosity hydro") where we increase the viscosity ν by 5 times (with the same ρ0−1/2 depth dependence) compared to the HD case (or the dynamo case). The resulting rms velocity of the statistical steady convection reached is vrms, all = 120 m s−1 and the Rossby number RO = 0.71, much closer to those of the dynamo case. The bottom row panels of Figure 6 show, respectively, the resulting differential rotation profile (Figure 6(j)), meridional circulation (Figure 6(k)), the angular momentum flux density in the r direction, RS (Figure 6(l)) due to the Reynolds stress, and VS (Figure 6(m)) due to the viscous stress. The integrated net (outward) angular momentum fluxes across concentric cylinders of radius r centered on the rotation axis is shown in Figure 7(c). It is found that a solar-like differential rotation profile (Figure 6(j)) with faster rotating equator and with a more conical iso-rotation contours in mid-latitude zones is established, although the contrast of rotation rate between the equator and the polar region is larger, about 44% of the mean rotation rate (compared to about 32% in the dynamo case). The mean meridional circulation (Figure 6(k)) shows a counter-clockwise (clockwise) cell pattern in the low-latitude region of the northern (southern) hemisphere, i.e., a poleward near-surface flow in the low latitude region, similar to the dynamo case (Figure 6(b)). We find that the angular momentum transport in the r direction due to the Reynolds stress for the HVHD case is very similar to that for the dynamo case, both in the meridional profile of the flux density (Figure 6(l) compared to Figure 6(c)) as well as in the integrated net flux across the constant r concentric cylinders (Figure 7(c) compared to Figure7(a)). However, this similar outward net angular momentum flux by the Reynolds stress is now balanced almost entirely by the transport due to the viscous stress, with the absence of the Maxwell stress which is the major component that balances the angular momentum flux by the Reynolds stress in the dynamo case. The comparison between the dynamo and the HVHD cases suggests an effective role of enhanced viscosity played by the magnetic fields, which (1) suppresses the large-scale convective motions such that they are more rotationally constrained (lower RO) to produce an outward transport of the angular momentum by the Reynolds stress, necessary to drive a solar-like differential rotation, and (2) takes up the main role to balance the Reynolds stress transport with the Maxwell stress instead of the viscous stress.

Further in Figure 8, we show the various horizontally integrated energy fluxes through the domain for the HD case (upper panel) and the HVHD case (lower panel), in comparison with the energy fluxes shown in Figure 3 for the dynamo case. It can be seen that the dynamo case and the HVHD case show a similar convective energy flux Lconv (reaching about 66% Ls in the dynamo case and about 60% Ls in the HVHD case). In both the dynamo and the HVHD cases, the downward kinetic energy flux Lkin (reaching about 16% Ls in the dynamo and 10% Ls in the HVHD case) is significantly reduced compared to the HD case (reaching about 45% Ls). In fact, in the HD case (upper panel in Figure 8), the downward kinetic energy flux is so large that the outward convective energy flux Lconv exceeds the solar luminosity in the middle of the CZ to counter it. The result here indicates again the similar role played by the magnetic fields and the enhanced viscosity in suppressing the downward convective flows.

Figure 8.

Figure 8. Same as Figure 3, but for the corresponding HD case (upper panel) and the HVHD case (lower panel).

Standard image High-resolution image

3.3. Emerging Flux

In the convective dynamo, the large-scale mean toroidal field as shown in Figure 5(b) is produced by the latitudinal differential rotation shearing a dipolar poloidal mean field. The reason that the mean toroidal field is concentrated towards the bottom of the CZ is mainly due to a downward advective transport of the magnetic energy in the bulk of the CZ, as represented by 〈vr(Bθ2 + Bϕ2)/8π〉 shown in Figure 9(a). This causes the distribution of the magnetic energy (for both the mean field and the small-scale field) to be strongly concentrated toward the bottom (see Figure 9(b)). The decrease with depth of the magnetic diffusivity η would also promote stronger fields toward the bottom but is less important here because of the small magnitude of η and the long diffusive timescale: (Δr)2/η ∼ 10 yr, using the peak η value near the surface and the depth of the CZ domain Δr. The advective timescale for the downward magnetic energy transport across the CZ is Δr/um ∼ 0.5 yr is significantly shorter, where we have used um = 〈vr(Bθ2 + Bϕ2)〉/〈Bθ2 + Bϕ2〉 evaluated at the middle of the CZ as a measure of the transport speed. Thus, the advective transport acts more quickly.

Figure 9.

Figure 9. (a) Advective flux of the horizontal magnetic field energy, and (b) the distribution of horizontal magnetic field energy as a function of depth.

Standard image High-resolution image

In the midst of magneto-convection, we find occasional active regions like flux emergence events in the top layer of the simulation domain. Such an example is shown in Figure 10, where panels (a), (b), (c), and (d) show, respectively, snapshots of Br, Bϕ, vr, and vϕ at a constant r slice at the depth of 30 Mm below the photosphere. The location of the emerging bipolar region is indicated by an arrow in the panels. It is characterized by a diverging bipolar pattern in Br (panel (a)) and the emergence of a strong toroidal field patch reaching a peak field strength of 9800 G (panel (b)) (also see the animation in the online journal). The emerging region corresponds to an up flow region in vr (panel (c)), but the upward velocity is not significantly different from that of other up flow convective cells. The zonal velocity vϕ of the emerging region shows a diverging pattern, and when averaged over the emerging region, is ∼100 m s−1 faster than the mean zonal velocity of that latitude. Figure 10(e) shows the subsurface 3D magnetic field configuration in the convective envelope by showing field lines traced from randomly seeded points throughout the volume. The field lines are colored based on their azimuthal field Bϕ as indicated by the color table. It can be seen that relatively more coherent bundles of strong toroidal flux are embedded in the turbulent magnetic fields. In Figure 10(f), regions of strong field strength where the Alfvén speed (va) exceeds the rms convective velocity (vrms) for the corresponding depth is outlined with the equipartition iso-surfaces (with va/vrms = 1), which are again colored based on the Bϕ value on the iso-surfaces. There is a systematic preference for these strong flux regions to be green or of negative Bϕ (red or of positive Bϕ) in the northern (southern) hemisphere. The arrows in Figures 10(e) and (f) mark the toroidal flux bundle with super-equipartition field strength that gives rise to the emerging region.

Figure 10. Panels (a), (b), (c), and (d) show, respectively, snapshots of Br, Bϕ, vr, and vϕ at a shell slice at the depth of 30 Mm below the photosphere, displayed on the full sphere in Mollweide projection. An animation showing the evolution of Br, Bϕ, vr at the 30 Mm depth, and also Bϕ at a depth near the bottom of the CZ, over a period of about 13 days centered around the time instant shown in this figure is also available in the online journal. Panels (e) and (f) show, respectively, 3D views of the magnetic field lines and the equipartition field iso-surfaces of va/vrms = 1 with va being the Alfvén speed and vrms being the rms convective velocity for the corresponding depth.

(An animation and a color version of this figure are available in the online journal.)

Video Standard image High-resolution image

Figure 11 shows a more zoomed in view of the thermodynamic properties of the emerging region at the same depth as that shown in the upper four panels of Figure 10. We see that there is a systematic reduction of density (i.e., buoyant; see Figure 11(b)) and pressure (Figure 11(d)) in the emerging region compared to the surrounding, although the reduction magnitude is rather moderate compared to the fluctuations seen in strong downflow lanes and in strong vertical flux tubes in the downflow lanes. The temperature change in the emerging region compared to the surrounding is smaller, partly due to the large thermal conduction. Averaged over the emerging region (area enclosed in the yellow contour in Figure 11(a), 〈ρ10〉 ≈ −1.6 × 10−5, 〈p1/p0〉 ≈ −1.1 × 10−5, and 〈T1/T0〉 ≈ 0.5 × 10−5. It can be seen that the temperature change is relatively small compared to the density and pressure change in the emerging region. This suggests that the buoyancy or reduction in density is mainly due to the reduction in gas pressure provided largely by the presence of the magnetic pressure, instead of mainly due to an increase in temperature. In other words, the buoyancy contribution is more from the magnetic buoyancy than the thermal buoyancy.

Figure 11.

Figure 11. Zoomed in view centered on the emerging region on the horizontal surface at the same depth (30 Mm) as shown in Figure 10 with (a) showing Bϕ and a yellow contour marking the emerging region where Bϕ < −5000 G and the ratio of the Alfvén speed over the rms convective speed va/vrms > 1, (b) showing density change ρ1 relative to the reference state ρ0 at that height, (c) showing temperature change T1 relative to the reference state T0, and (d) showing pressure change p1 relative to the reference state p0.

Standard image High-resolution image

As can be seen from Figure 10(e), the emerging flux region is the apex of a roughly east-west oriented (toroidal) super-equipartition flux bundle which remain relatively coherent for some distance, before the two ends connect in complex ways to other flux systems. The following end of the coherent flux bundle extends into the middle of the CZ. The fact that the emerging flux has a prograde zonal speed of ∼100 m s−1 relative to the mean zonal speed of the latitude indicates that it is not a toroidal flux tube rising in isolation from the bottom of the CZ. If it were, it would have a retrograde flow due to angular momentum conservation as is found in many previous studies of isolated rising flux tubes in the rotating solar CZ (e.g., Caligari et al. 1995; Fan 2008, 2009). The emerging flux bundle must have well mixed with the local plasma through reconnections, and is continually sheared and amplified by the differential rotation and the local flows against resistive dissipation. The sequence of images in Figure 12 show the sub-surface development of the super-equipartition emerging flux bundle (marked by the arrow) over a nine-day period prior to the time of the flux emergence event shown in Figure 10. It shows that local shear in the upper CZ contributes significantly to the development of the emerging flux bundle. The left column images show iso-volumes of super-equipartition fields with the surface of the volume colored by Bϕ. The right column images show representative field lines traced from the iso-volume corresponding to the emerging flux bundle. It can be seen that a segment of a super-equipartition flux bundle in the middle of the CZ is sheared and stretched in the prograde direction into a hairpin turn with the upper side of the hairpin forming the emerging flux bundle reaching the top boundary.

Figure 12.

Figure 12. Left column images show iso-volumes of super-equipartition fields (where va/vrms > 1) for a period of 9 days prior to the time of the flux emergence event shown in Figure 10, with the arrow marking the evolution of the emerging flux bundle that produces the flux emergence event shown in Figure 10. Right column images show representative field lines traced from points in the iso-volume corresponding to the emerging flux bundle.

Standard image High-resolution image

We have done a statistical study of the super-equipartition emerging fields. For a time period of about one year centered at the cycle maximum phase (green line in Figure 5(a)) and at an interval of 12 hr, we find in the shell slice at 30 Mm depth all the area where the emerging horizontal field exceeds $\sqrt{2}$ times the field strength in equipartition with the rms convective velocity of that depth. For each pixel (or grid point) of the selected emerging field area, we compute a tilt angle of the horizontal field vector based on the local Bϕ and Bθ. The resulting tilt angle distribution of all the pixels is shown in Figure 13. The quadrant of the tilt angle is such that, if the sign of the azimuthal field Bϕ is consistent with Hale's polarity rule of the cycle, i.e., negative (positive) in the northern (southern) hemisphere, then the tilt angle falls in quadrants I and IV. If the horizontal field vector is tilted clockwise (anti-clockwise) from the cycle preferred azimuthal field direction in the northern (southern) hemisphere by an acute angle, consistent with the mean tilt of solar active regions, then the tilt angle falls in quadrant I. We find from Figure 13 that there is a preference for Hale's polarity rule for the emerging azimuthal field by a ratio of 2.4 to 1 in the area. For those pixels satisfying Hale's rule, the mean tilt angle is 7fdg5, with an estimated uncertainty of 1fdg6. Thus, the super-equipartition emerging fields have a statistically significant mean tilt similar to the active region mean tilt (e.g., Stenflo & Kosovichev 2012).

Figure 13.

Figure 13. Distribution of the tilt angles of the horizontal field vectors in strong emerging field areas at 30 Mm depth. See the text about the tilt angle quadrants I, II, III, and IV.

Standard image High-resolution image

4. DISCUSSIONS

We have presented a 3D convective dynamo driven by the solar radiative diffusive heat flux in the solar CZ, and with a latitudinal gradient of entropy imposed at the bottom, representing the tachocline-induced thermal variation that can break the Taylor–Proudman constraint in the CZ (Rempel 2005). The convective dynamo produces a large-scale mean field that undergoes irregular cycles and polarity reversals, and self-consistently maintains a solar-like differential rotation with faster rotation at the equator than at the polar region by about 30% and more conical iso-rotation contours in mid-latitudes (e.g., Thompson et al. 2003).

The irregular cyclic behavior of the mean field in our model differs from those in the literature (e.g., Ghizaru et al. 2010; Racine et al. 2011; Käpylä et al. 2012). By comparing the Reynolds number Re and the Coriolis number CO (as defined in Käpylä et al. 2012) achieved in our dynamo run with those of the cyclic dynamo simulation presented in Käpylä et al. (2012) (see Section 3.1), we find that their convective dynamo is operating in a significantly more rotationally constrained regime, with their CO being about five times ours. Our convective flows appear to be only moderately more turbulent compared to theirs as reflected in the similar order of magnitude for the Re values. If we consider our convective rms velocity to be similar to theirs, then their dynamo would be effectively operating in a stellar envelope that is rotating at about five times the solar rotation rate. This is probably the main reason we obtain a very different mean field dynamo behavior compared to that of Käpylä et al. (2012). Our irregularly cycling convective dynamo model also differs significantly in many ways from the cyclic convective dynamo model described in Ghizaru et al. (2010); Racine et al. (2011). Their dynamo model used an implicit large eddy code with no explicit viscosity, magnetic diffusion, and thermal diffusion. However, our convective dynamo appears to be operating in a more turbulent regime if we compare the convective downflow speed obtained: about 25 m s−1 at r = 0.954 Rs in (Ghizaru et al. 2010) versus our ∼300 m s−1 at the same depth. Given that both models use the solar rotation rate for the convective envelope, this suggests that their dynamo model is also operating in a significantly more rotationally constrained regime with a significantly lower Rossby number compared to ours. The Newtonian cooling treatment of the entropy equation used in Ghizaru et al. (2010); Racine et al. (2011) is very different from our treatment of the energy transport which forces the solar luminosity through the convective domain. Furthermore, their model includes a sub-adiabatically stratified overshoot layer at the bottom of the CZ which our model does not have. All these contribute to the significant differences in the resulting dynamo behavior.

In both Käpylä et al. (2012) and Racine et al. (2011), because of the significantly higher CO or lower RO, the convective flows in their corresponding hydro cases in the absence of the magnetic fields are already driving a solar-like differential rotation, and the addition of the magnetic fields in their dynamo cases appear to mainly reduce the differential rotation (Charbonneau 2013). On the other hand, for our convective dynamo in a significantly less rotationally constrained regime with RO closer to 1, the presence of the magnetic field is found to play an important role for the maintenance of the solar-like differential rotation, without which a faster rotating polar region results, as is shown with the corresponding HD simulation. A solar-like differential rotation profile can alternatively be achieved in the hydro case by increasing the viscosity as shown in the HVHD simulation. Comparing between the dynamo case and the HVHD case indicates that in several aspects the magnetic field plays an effective role of enhanced viscosity to (1) suppress the large scale convective motions such that they become more rotationally constrained to produce an outward transport of angular momentum by the Reynolds stress needed to drive a solar-like differential rotation, (2) take up the main role of balancing the Reynolds stress transport with the Maxwell stress transport instead of the viscous stress under low viscosity conditions, and (3) reduce the downward kinetic energy flux. Our resulting differential rotation self-consistently maintained in the convective dynamo simulation is in fairly good agreement with the observed solar differential rotation in both the pole-equator contrast and also the more conical shaped iso-rotation contours in the mid-latitude zone. The more conical iso-rotation contours are achieved by the latitudinal gradient of entropy imposed at the lower boundary as has been described in Rempel (2005); Miesch et al. (2006). The latitudinal gradient of entropy is spread into the bulk of the CZ due to the large thermal diffusivity, and this latitudinal gradient in the CZ provides the necessary balance in the ϕ component of the vorticity equation to allow for a non-cylindrical differential rotation to be established (Rempel 2005). Thus, the role of the imposed latitudinal entropy gradient is to change the shape of the iso-rotation contours from cylindrical to more conical. It is not the reason for the change of differential rotation from anti-solar (fast rotating pole) to solar-like (faster equator) between the HD and the magnetic cases, both of which have the same latitudinal entropy gradient imposed. That change is brought about by the change in the direction of Reynolds stress transport of the angular momentum.

We also note here the possible effect of the small deviation from total energy conservation in our dynamo simulation as seen in Figure 3, where the total energy flux Ltot exiting the domain at the top is about 13% less than the input solar luminosity. As pointed out in Section 3.1, this is due to the loss of the magnetic energy dissipated by the numerical diffusion, which is not put back into the thermal energy and hence does not have to be carried out by thermal conduction at the top. This would mean a slightly weakened driving of convection where a reduced solar luminosity (reduced by up to about 13% near the top) is forced through the domain. The heat fluxes by thermal conduction and convection would have been slightly higher if the energy conservation were strictly adhered to. This, however, does not affect our finding that the magnetic field takes up the role of an enhanced viscosity, and if it damps the large-scale convective motions to the level similar to that of the HVHD case (with convection carrying roughly 66% of the solar luminosity) a solar-like differential rotation can be achieved by the resulting Reynolds stress. The larger question that remains is how to maintain the solar differential rotation with a convection that can transport nearly 100% of the solar luminosity without resorting to an ad hoc thermal diffusion (conduction) that carries a substantial fraction (about 36% at the middle of the CZ) of the solar luminosity.

In our dynamo simulation the large-scale mean toroidal field, antisymmetric with respect to the equator, is concentrated at the bottom of the CZ, unlike many of the recent convective dynamo simulations with cylindrical iso-rotation contours and significantly faster (than solar) rotation rates (e.g., Nelson et al. 2013; Augustson et al. 2013), where strong wreath-like toroidal field structures are present in the equatorial region of the bulk of the CZ. In the 3D magnetic field of the present simulation (Figure 10), occasional more coherent toroidal flux bundles of super-equipartition field strength are embedded in the turbulent small scale fields, as discussed in Section 3.3. Some of these super-equipartition flux bundles rise to the surface to produce active region like flux emergence events. Although these emerging flux bundles show significant magnetic buoyancy, they are not flux bundles rising in isolation from the bottom of the CZ, but are a product of continued reconnection and shear amplification by local flows in the CZ. There is a preference for the azimuthal field in the strong emerging field regions to conform to Hale's polarity rule by a ratio of 2.4 to 1 in area, and a statistical significant mean tilt angle of 7fdg5 ± 1fdg6 for the emerging horizontal fields, consistent with the active region mean tilt. However, the violation from Hale's rule is far greater than that of solar active regions. This is because of the very weak mean field component (∼5% of the total magnetic energy) in the current convective dynamo. It is very likely that our dynamo model is still significantly over-estimating the giant-cell convective speed (Hanasoge et al. 2012) and the Sun's dynamo is operating in a significantly more rotationally constrained regime. This is indicated by the more rotationally constrained convective dynamo models of Käpylä et al. (2012); Augustson et al. (2013), which are able to achieve a more regular, solar-cycle-like cyclic behavior and stronger mean field by effectively increasing the rotation rate. With a significantly lower magnetic diffusion (than have been achieved by current global convective dynamo models), the solar dynamo may be operating in a significantly stronger field regime with a much more suppressed giant-cell convection and lower Rossby number. This may lead to a stronger Reynolds stress transport of angular momentum that needs to be balanced by a stronger Maxwell's stress and hence the generation of a stronger mean field. However, the question remains how such suppressed convective flows transport the solar luminosity through the CZ, although the convective energy transport in the more magnetic-buoyancy-dominated regime may be quite different. Furthermore, the inclusion of an overshoot layer below the base of the CZ may be important for the operation of the solar cycle dynamo. The convective dynamo model of Ghizaru et al. (2010); Racine et al. (2011) has achieved a much stronger large-scale mean field component, with the mean field magnetic energy comparable to that of the small-scale magnetic energy during cycle maxima (compared to the 10% in our model), by allowing penetration and storage of the strong mean field in the stable overshoot layer.

We thank Matthias Rempel and Piyali Chatterjee for discussions and helpful comments on the paper, and Doug Braun for helpful suggestions on error analysis. We also thank the anonymous referees for helpful comments. This work is supported by the NASA LWSCSW grant NNX13AG04A and NASA HSR grant NNX10AB81G to NCAR. NCAR is sponsored by the National Science Foundation. F.F. is supported by the Advanced Study Program of NCAR. The numerical simulations were carried out on the Pleiades supercomputer at the NASA Advanced Supercomputing Division under project GIDs s1106, s0925.

Please wait… references are loading.
10.1088/0004-637X/789/1/35