Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 Phil. Trans. R. Soc. A (2012) 370, 3049–3069 doi:10.1098/rsta.2011.0507 R EVIEW The solar dynamo B Y M ARK S. M IESCH* High Altitude Observatory, National Center for Atmospheric Research, 3080 Center Green Drive, Boulder, CO 80301-2252, USA The origins of solar magnetism lie below the visible surface of the Sun, in the highly turbulent convection zone. Turbulent convection operates in conjunction with rotational shear, global circulations and intricate boundary layers to produce the rich diversity of magnetic activity we observe. Here, we review recent insights into the operation of the solar dynamo obtained from solar and stellar observations and numerical models. Keywords: Sun interior; Sun activity; solar dynamo; magnetohydrodynamics 1. Introduction The year 2008 marked the 100th anniversary of George Ellery Hale’s monumental detection of magnetic fields in sunspots by means of the Zeeman effect [1]. It was the first demonstration of magnetism in an astronomical object and the discovery transformed not only solar physics but all of astrophysics. The 11 year sunspot cycle, discovered 65 years earlier by Heinreich Schwabe [2], was soon seen in a new light (although not without controversy), as a remarkable example of hydromagnetic self-organization on a colossal scale, whereby bulk motions of a turbulent plasma produce organized, global patterns of magnetic activity. Hale himself appreciated the implications, drawing parallels between sunspots and laboratory experiments of magnetic fields generated by rotating electrically charged discs. Astrophysical dynamo theory was born (or at least reborn, because prior links between geomagnetic storms and solar activity had already provided inklings of solar magnetism). Modern observations reveal the handiwork of the solar dynamo with marked clarity. Filamentary magnetic structures such as coronal loops and streamers permeate the solar atmosphere, and the release of magnetic energy accelerates the solar wind and powers eruptive events such as flares and coronal mass ejections. These in turn shape the heliosphere and regulate the space environment of planets. Solar and stellar variability and rotational evolution are intimately linked with magnetism. Yet, the sunspot cycle still stands as one of the most fundamental and enigmatic challenges of solar physics. In this paper, we provide a brief overview of solar dynamo theory as it now stands, focusing on fundamental building blocks and active frontiers. In order to *[email protected] One contribution of 11 to a Theme Issue ‘Astrophysical processes on the Sun’. 3049 This journal is © 2012 The Royal Society Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 3050 M. S. Miesch set the stage for the subsequent discussion, we give a brief overview of solar magnetism in §2. We then address the nature of convective dynamos in §3, highlighting the distinction between small-scale and large-scale dynamo action and its implications for stars and the Sun in particular. In §4, we focus on the solar activity cycle, highlighting current theoretical paradigms and open questions. Section 5 is a summary and a look to the future. 2. Solar magnetism (a) The solar cycle The cyclic nature of the global solar magnetic field is most clearly demonstrated in the structure and evolution of the mean field threading through the solar surface. Figure 1a shows the radial component of the magnetic field in the solar photosphere inferred from Zeeman spectroscopy as a function of latitude and time. A longitudinal average is approximated by combining measurements near the centre of the disc spanning one mean rotation period, known as a Carrington rotation. The low-latitude behaviour seen in figure 1a reflects the well-known solar butterfly diagram, prominent also in sunspot observations [4–6]. At the beginning of a cycle, bipolar active regions appear at latitudes of roughly ±30◦ , signifying the emergence of coherent subsurface flux structures. As the cycle proceeds, these active regions disperse and new flux emerges at progressively lower latitudes, reaching the equator in approximately 11 years.1 As the active bands approach the equator, the emergence of new active regions largely ceases, as signified for example by a minimum in the fraction of the surface covered by spots (figure 1b). This is known as solar minimum, when the large-scale structure of the magnetic field in the photosphere and overlying corona is relatively simple,2 dominated by polar caps of opposite polarity (e.g. 1996–1997 in figure 1a). Much of the magnetic energy is in the dipole component, but higher order multi-poles also contribute significantly [7]. This was particularly evident during the recent extended minimum in 2008–2009 when the dipole contribution dropped unusually low relative to higher order contributions (J. Luhmann 2010, unpublished data). The appearance of active regions at mid-latitudes signifies the beginning of a new 11 year cycle. The polarity of the polar fields during solar minimum reverses each 11 year sunspot cycle; so the net period of the magnetic activity cycle is approximately 22 years. Regions of positive and negative polarity within an active region are oriented approximately along an east–west axis, although the trailing spots (westward edge, following the sense of rotation) are systematically displaced poleward 1 Individual active regions do not exhibit latitudinal propagation until after they fragment and disperse. 2 The photospheric magnetic field at solar minimum exhibits many small-scale, short-lived features that cover most of the surface and dominate the magnetic energy. Thus, the magnetic topology may actually be more complex at solar minimum relative to solar maximum in the sense of more photospheric null points, separators and separatrix surfaces. Still, the large-scale magnetic structure is relatively simple in terms of the relative prominence of low-order, nearly axisymmetric multi-pole moments. Phil. Trans. R. Soc. A (2012) Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 3051 Review. The solar dynamo (a) 90 N 30 N EQ 30 S 90 S Mg II index spot area (b) 8000 6000 4000 2000 0 (c) 0.29 0.28 0.27 0.26 1975 1980 1985 1990 1995 2000 2005 2010 date Figure 1. (a) Vertical magnetic flux through the solar photosphere as a function of latitude and time, averaged over longitude (courtesy of David Hathaway). Blue and yellow regions denote inward and outward polarities. The field strength is inferred from Zeeman measurements near the centre of the disc assuming a radial field orientation, and longitudinal averages are constructed from synoptic maps spanning one Carrington rotation (data compiled from the Vacuum Tower Telescope and Synoptic Optical Long-term Investigations of the Sun (SOLIS) instruments at NSO/Kitt Peak and the Michelson Doppler Imager (MDI) instrument onboard the SOlar and Heliospheric Observatory (SOHO)). (b) The area of the solar disc covered by sunspots as a function of time, in units of millionths of a hemisphere (courtesy of David Hathaway; data compiled from Royal Greenwich Observatory, US Air Force and NOAA facilities). (c) Mg II index (non-dimensional; see text) as a function of time (courtesy of Marty Snow; data compiled from the UARS, ERS-2, SORCE and additional NOAA missions; for further information, see Viereck et al. [3]). relative to the leading (eastward edge) polarity. The tilt of the axis increases with latitude, known as Joy’s Law, and is thought to arise through the influence of the Coriolis force on buoyantly rising toroidal flux structures [8]. Furthermore, the sense of the leading and trailing polarities in active regions reverses sign across the equator, known as Hale’s polarity law, implying antisymmetry in the subsurface toroidal fields that give rise to them. The emergent flux structures that form photospheric active regions apparently decouple from deeper seated dynamics within a few days [9], fragmenting and dispersing over the course of several weeks. The dispersal and subsequent transport of the residual flux from active regions is well captured by quasitwo-dimensional models that incorporate turbulent diffusion from convective motions as well as advection by a time-varying, axisymmetric, poleward flow Phil. Trans. R. Soc. A (2012) Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 3052 M. S. Miesch of order 10–30 m s−1 [10,11]. Such a flow has been detected in the solar surface layers by multiple independent techniques, including Doppler measurements, local helioseismic inversions and correlation tracking [12]. The combined effect of Joy’s Law, active region fragmentation and surface transport by convection and by meridional flow induces a net global, mean poloidal field by means of a phenomenon known as the Babcock–Leighton (BL) mechanism [13,14]. Trailing residual flux from active regions is transported poleward, where it interacts with pre-existing polar fields of the opposite polarity (figure 1a). This process apparently accounts for the cyclic polarity reversals in the polar caps, although other mechanisms may also contribute to global poloidal field generation (§4). Although sunspots and active regions are the most familiar manifestations of cyclic solar activity, corresponding variability is observed in many other diagnostics, such as solar irradiance, the frequency of eruptive events (coronal mass ejections and flares) and geomagnetic activity induced by solar forcing [15,16]. An example is shown in figure 1c, which traces what is known as the Mg II index, quantifying the relative emission in the core and wing of the Mg II spectral line. This Mg II emission originates in the chromosphere, and the core-to-wing ratio is observed to peak above active regions and supergranular downflow lanes where the photospheric magnetic flux is predominantly vertical. The Mg II index is commonly taken as a proxy for solar and stellar magnetic activity [17]. Although continuous multi-wavelength monitoring of solar irradiance variability and related magnetic activity is limited to the past three to five decades, available data suggest that similar variations in solar magnetic activity have been operating unabated for at least 10 000 years, with extended periods of relatively low and high activity known as grand minima and grand maxima [5,18]. Continuous sunspot records date back to the early seventeenth century, and earlier variations in solar activity can be inferred from cosmogenic chemical isotopes found in tree rings and ice cores. The data are consistent with the notion that the 22 year activity cycle has been operating continually for at least several millennia, although the most ancient isotope data often have insufficient temporal resolution to demonstrate this conclusively. Longer term variations such as grand minima and maxima appear relatively random. There is some possible evidence for quasi-periodic behaviour on time scales of 80–2000 years, but none as prominent or regular as the primary 11 year sunspot cycle [18]. (b) Order amid chaos Although cyclic activity has long been regarded as the most important and formidable challenge of solar dynamo theory, a closer look at solar magnetism reveals that this striking regularity exists amid a turbulent maelstrom of smallscale magnetic fluctuations. High-resolution Zeeman measurements of the solar photosphere reveal rapidly varying magnetic structure on the smallest observed scales of order 100 km, while more sophisticated measurements based on full Stokes polarimetry and the Hanle effect suggest that there is yet more magnetic energy on even smaller scales that are beyond current resolution limits [19–22]. This small-scale flux component replenishes itself daily and has been referred to as the Sun’s magnetic carpet [23,24]. Phil. Trans. R. Soc. A (2012) Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 Review. The solar dynamo 3053 Photospheric magnetograms (vertical field maps inferred from Zeeman measurements) reveal patches of magnetic flux spanning a vast range of scales (figure 2a,b). By applying clumping algorithms to cross-calibrated data from the SOlar and Heliospheric Observatory (SOHO)/Michelson Doppler Imager (MDI) and Hinode/Solar Optical Telescope (SOT), Parnell et al. [25] have demonstrated that the flux distribution is self-similar, following a power law extending from the smallest resolved features (area ∼1014 cm2 , flux ∼1016 Mx) all the way to sunspots and active regions (area ∼2 × 1019 cm2 , flux ∼1022 –1023 Mx). The distribution persists at solar minimum and solar maximum, the main difference being the presence of active regions in the latter, which extends the distribution to higher flux values (figure 2c). The best-fit power-law exponent for the data in figure 2c is −1.85. Careful scrutiny of time series suggests that the distribution of emerging flux is even steeper, following a power law with a best-fit exponent of −2.74 [26]. Presumably, it is this emergent flux spectrum that characterizes the dynamo, while the shallower spectrum shown in figure 2c may reflect the coalescence of flux elements through passive advection by the granular flow or through dynamical feedback mechanisms such as convective collapse [27]. Indeed, reprocessing of surface flux through coalescence and/or fragmentation can readily yield power laws similar to observational inferences [28]. Still, the power-law behaviour of emergent bipoles and the persistence of power-law behaviour throughout the solar cycle suggests that the scale similarity may be inherent in the field generation and not just due to surface processing of emergent flux. The distribution function f (F) shown in figure 2c is defined such that the total flux per unit area from flux concentrations with values between F1 and F2 is given by F (F1 , F2 ) = F2 Ff (F) dF Mx cm−2 . (2.1) F1 The emergent flux spectrum obtained by Thornton & Parnell [26] is defined similarly, but the integrated flux F (F1 , F2 ) now has units of Mx cm−2 d−1 . Thus, a power-law distribution f (F) ∝ F−a with a > 2 indicates that the unsigned flux emerging on small scales dominates over that emerging on large scales. In particular, the rate of flux emergence in structures smaller than 1020 Mx exceeds that in active regions by nearly three orders of magnitude [26]. This estimate must be regarded as a lower limit, since much of the small-scale flux is likely to be unresolved, as suggested by Hanle measurements [19]. Furthermore, if the mean-field strength in small and large flux concentrations is comparable (justified as a first approximation by the common saturation level of figure 2a,b), this implies that most of the magnetic energy produced by the dynamo is likewise in small-scale, turbulent fluctuations. Thus, the solar dynamo does not just produce the solar cycle. Rather, it produces a continuous spectrum of dynamically significant fields extending from the global dipole moment to sunspots to turbulent fluctuations on the smallest sub-granule scales we can currently resolve. The common power law across this vast range of scales suggests that the sunspot cycle and the magnetic carpet are dynamically linked. Although this does not necessarily rule out meaningful models targeted at specific aspects of the solar dynamo, it is wise to keep in mind that the multi-faceted manifestations of solar magnetism have a common origin. Phil. Trans. R. Soc. A (2012) Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 3054 M. S. Miesch (b) (c) frequency (× 10–46 Mx–1 cm–2) (a) 1012 1010 SOD/NFI 06/2007 MDI/FD 12/2001 MDI/FD 12/2007 108 106 104 102 100 1016 1017 1018 1019 1020 1021 1022 1023 flux (Mx) Figure 2. The distribution of vertical magnetic flux in the solar photosphere (from Parnell et al. [25]). Full-disc (FD) observations from (a) the Michelson Doppler Imager (MDI) instrument on the SOlar and Heliospheric Observatory (SOHO) combined with higher resolution observations from (b) the Solar Optical Telescope (SOT) instrument on Hinode reveal a power-law distribution (c) spanning five decades of flux. White and black in (a,b) indicate outward and inward fluxes with saturation levels of 200 Mx cm−2 . The Hinode/SOT field of view is indicated by the smaller of the two squares in (a). (c) Includes SOT and MDI FD measurements at solar minimum (2007) as well as MDI FD measurements at solar maximum (2001), with colours as indicated. The dashed line indicates a power law with an index of −1.85. (c) Convection, rotation and magnetic activity It has been realized since the formative years of solar dynamo theory a half century ago that the key ingredient in producing organized patterns of magnetic activity from disorganized convective motions is rotation [29–31]. Rotation induces anisotropic momentum and energy transport, which in turn establishes differential rotation and global meridional circulations (MCs) that amplify and transport magnetic flux [32]. Rotation also gives rise to helical flows and fields that can promote hydromagnetic self-organization by coupling large and small scales. There are many subtleties in how such processes proceed in practice (reviewed in §3b) and there are still many uncertainties as to what physical mechanisms are responsible for establishing cyclic solar activity (reviewed in §4). Yet, it is still fair to say that, to the best of our current understanding, if the Sun were not rotating, it would have a magnetic carpet but not a magnetic cycle. In the past few decades, helioseismology has revolutionized solar dynamo theory by mapping out the solar internal rotation profile throughout most of the convection zone (figure 3a). The solar envelope exhibits prominent rotational shear, with the angular velocity U decreasing monotonically by about 30 per cent from the equator to latitudes of 60–70◦ (higher latitude inversions are unreliable owing to the insensitivity of global acoustic modes to rotation near the rotation axis). This latitudinal shear persists throughout the convection zone, with radial shear largely confined to upper and lower boundary layers. The lower boundary layer is known as the tachocline and represents a transition from latitudinal shear in the convection zone to nearly uniform rotation in the radiative zone. At the top of the convection zone is the near-surface shear layer (NSSL), where the rotation rate increases inward by about 2–3% from the photosphere to r ∼ 0.95R. In this Phil. Trans. R. Soc. A (2012) Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 3055 Review. The solar dynamo 450 (b) 400 375 350 325 300 latitude 425 2 60 40 20 0 –20 –40 –60 dW/2p (nHz) (a) 1 0 –1 –2 1996 1998 2000 2002 2004 2006 2008 date (yr) Figure 3. The solar internal rotation. (a) Angular velocity (nHz) as a function of latitude and radius inferred from SOHO/MDI observations using a subtractive optimally localized averages (SOLA) inversion technique [33]. The dotted line indicates the base of the convection zone and white regions near the rotation axis indicate a lack of reliable data. (b) Angular velocity variations relative to a temporal mean versus latitude and time at a depth of 0.99 R, based on MDI and Global Oscillations Network Group (GONG) data (from Howe et al. [34]). Solid contours indicate photospheric magnetic activity based on magnetograms from the SOLIS instrument on Kitt Peak. paper, we focus on the dynamo implications of this differential rotation profile, which will be addressed in §§3b and 4. For a discussion of how mean flows are established and maintained, see Miesch & Toomre [32]. It is often assumed that the solar dynamo operates in a kinematic regime such that fields are passively amplified and transported by the flow. Because this assumption helps us to justify the concept of the a-effect (§3b), it underlies virtually all dynamo models of the solar cycle. Yet, it is also widely realized that the kinematic assumption cannot be strictly valid for the Sun. In the kinematic limit, the magnetohydrodynamic (MHD) induction equation is linear and dynamo action can be regarded as a linear instability with an exponential growth rate. Because the amplitude of the solar cycle does not appear to be growing exponentially, this implies that there must be some nonlinear feedback between fields and flows that limits it. Dynamo saturation, such as magnetic self-organization, is a subtle issue that we will address briefly in §3b. Here, we merely highlight the torsional oscillations, which represent the most compelling evidence we have that systematic feedbacks between large-scale fields and flows do occur. These are alternating bands of faster and slower rotation that appear and propagate in latitude in conjunction with the solar activity cycle (figure 3b). In particular, there is both a low-latitude branch that propagates equatorwards along with the bands of magnetic activity as well as a high-latitude branch that propagates poleward on a comparable time scale. The torsional oscillations are clearly associated with magnetism but their significance, if any, with regard to the operation or saturation of the dynamo is unclear. With an amplitude of less than 0.5 per cent, they do not appear to be large enough to reflect dynamo saturation through the suppression of rotational shear (although MHD convection simulations and non-kinematic mean-field models do exhibit a reduction in rotational shear relative to non-magnetic models, providing some evidence for dynamo saturation via the so-called U-quenching; Phil. Trans. R. Soc. A (2012) Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 3056 M. S. Miesch [35–37]). Current models generally regard them as a passive by-product of cyclic activity, although it is worth noting that they appear each cycle before the emergence of sunspots on the surface (figure 3c). The poleward branch is typically attributed to the mean Lorentz force or turbulent diffusion responding to cyclic lower latitude forcing, but the meridional structure of the low-latitude branch suggests that it may be thermally driven, induced by enhanced radiative losses in active bands3 [38]. It is clear from stellar observations that convection and rotation play a central role in magnetic activity. Late-type stars with convective envelopes such as the Sun generally exhibit similar signatures of magnetic activity, including chromospheric and coronal emission and photometric variations indicative of spots [17]. Although unresolved, the most active stars also exhibit detectable Zeeman splittings [39]. Inferred field strengths scale roughly as the square of the rotation rate, B̃ ∝ U2 , although precise estimates for the power-law index range widely from 1.2 to 2.8 [40–42]. The increase in B̃ is likely to be due in part to an increase in rotational shear DU, which also increases with U, although precisely how is still being debated. Various studies have suggested power-law relationships such that DU ∝ Un but estimates for the index n range from 0.1 to 0.7 [43–45]. Such scaling uncertainties for B̃ and DU are most probably owing to heterogeneous sampling but they may also reflect intrinsic variability owing to sensitivities in the underlying nonlinear dynamics [42]. Scaling relationships for both B̃ and DU are made somewhat tighter when U is replaced by the inverse Rossby number tc U (particularly for heterogeneous samples), where tc is a convective turnover time obtained from theoretical models or empirical fitting. Scaling relationships for both B̃ and DU saturate beyond a threshold value that depends weakly on stellar type [41,42]. For solar-like stars, the threshold is about 10U , beyond which B̃ becomes insensitive to U and DU appears to be suppressed, presumably by the Lorentz force (thus providing evidence for U-quenching). In this saturated regime, the dynamo appears to operate at peak efficiency, with the mean-field strength B̃ determined by the stellar luminosity [46]. Many late-type stars certainly exhibit magnetic activity cycles but, despite much literature on the subject, there are few well-established results. Data are scarce, largely because observations spanning multiple years or even decades are needed to verify cycle periods and amplitudes. There is some evidence that cycle periods for solar-type stars with moderate rotation rates tend to decrease somewhat with U but there is much scatter and there may be multiple branches signifying distinct dynamo regimes. We refer the interested reader to the reviews by Rempel [47], Saar [42] and Lanza [48]. Still, it is clear that the interplay between convection and rotation lies at the root of solar and stellar magnetic activity, so we turn now to discuss the nature of convective dynamos. 3 Sunspots are famously dark relative to their surroundings but their appearance on the solar disc is accompanied by bright, more weakly magnetized regions known as faculae. Facular brightenings generally overwhelm sunspot darkenings; so the Sun is roughly 0.1 per cent brighter at solar maximum [15]. Phil. Trans. R. Soc. A (2012) Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 Review. The solar dynamo 3057 3. Convective dynamos The starting point for any solar dynamo model is the magnetic induction equation under the MHD approximation vB = V(v × B − hV × B), (3.1) vt where v and B are the fluid velocity and magnetic field and h is the magnetic diffusivity of the plasma. The (global-scale) ratio of the advection term to the diffusive term on the right-hand side (RHS) is the magnetic Reynolds number Rm = UL/h, where U and L are characteristic velocity and length scales, respectively. (a) Small-scale dynamos The best understood class of turbulent dynamos are isotropic, homogeneous, incompressible flows with large magnetic Prandtl numbers Pm = n/h 1, where n is the kinematic viscosity. Here, the kinetic energy spectrum scales with the spatial wavenumber k as E(k) ∝ k −p within an inertial range that extends approximately from the energy injection scale k0 to the viscous dissipation scale kn = Req k0 k0 . Here, Re = UL/n is the Reynolds number and q = 2/(p + 1). For Kolmogorov turbulence, p = 5/3 and q = 3/4. For such a flow, the magnetic field is most efficiently generated on small scales, below the viscous dissipation scale k > kn , where the velocity field is spatially smooth but temporally random [49,50]. Quasi-laminar (yet chaotic) stretching generates folded field topologies such that variations in the direction of B occur on a length scale comparable to the size of eddies e , whereas −1/2 e [49]. The variations perpendicular to B occur on a resistive scale r ∼ Rm −1/2 e . resulting magnetic energy peaks near the ohmic dissipation scale r ∼ Rm From equation (3.1), we anticipate that the kinematic dynamo growth rate should scale with the strength of the shear kuk ∼ k (3−p)/2 , where uk2 = kE(k). For a general velocity field with p < 3, this implies that the smallest eddies are most efficient at amplifying field because they have the shortest turnover times. Thus, e ∼ n and r ∼ Pm1/2 n where n = 2p/kn . This picture breaks down in stars where Pm < 1. Here, the ohmic dissipation scale lies within the inertial range such that r > n . There are no scales that appear spatially smooth from the perspective of the magnetic field. Small eddies still have the fastest turnover times but ohmic diffusion will suppress field amplification on scales smaller than r . Thus, we might still expect magnetic energy generation to be most efficient near the ohmic dissipation scale r but now quasi-laminar stretching by larger eddies > r must compete against the disruptive effects of turbulent mixing from smaller eddies < r . Dynamo action is more difficult to sustain and depends sensitively on the slope of the spectrum p near r ; steeper slopes (larger p) facilitate field generation [50]. This is most readily demonstrated in numerical simulations where Rm can be held fixed while Pm is varied by changing the value of n. As n is decreased below h such that Pm passes through unity, there is a sharp rise in the critical value of Rm needed to sustain dynamo action4 [51,52]. 4 The value of Rm in stars is sufficiently large that supercriticality is ensured, despite the small value of Pm . Phil. Trans. R. Soc. A (2012) Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 3058 M. S. Miesch Despite the subtleties at low Pm , it is clear that isotropic, homogeneous, turbulent dynamos tend to generate intermittent magnetic fields on scales much smaller than the typical scale of the velocity field LB LV . In short, turbulent flows beget turbulent fields. Such dynamos have been referred to as small-scale dynamos in order to distinguish them from large-scale dynamos, which may generate magnetic fields on scales much larger than the velocity field LB LV . Small-scale dynamo action has been proposed to account for the intricate flux distribution of the magnetic carpet [21,53,54]. This would then imply that the magnetic energy in the photosphere may well peak on scales that are currently unresolved, possibly extending down to ohmic dissipation scales of order 100 m. (b) Large-scale dynamos: from fluctuations to flux The principal agents of self-organization in turbulent dynamos are helicity and shear. Although the significance of these agents was originally appreciated within the context of the kinematic mean-field theory, they continue to take centre stage as three-dimensional MHD simulations become increasingly important tools at the leading edge of dynamo research. Chaotic stretching is still required to generate magnetic energy but helicity and shear—coupled with the geometry, boundary conditions and other sources of anisotropy and instability such as density stratification and buoyancy—play a dominant role in shaping the large-scale magnetic topology. The pioneers of mean-field dynamo theory emphasized the significance of the kinetic helicity, defined as Hk = u · vV , where u = V × v is the fluid vorticity and V denotes a volume integral [29–31]. The combined influence of rotation and density stratification induces a helicity density u · v, which reflects the presence of spiralling flows that break anti-dynamo constraints associated with two dimensionality and mirror symmetry. More recently, the focus has shifted to the magnetic helicity Hm = A · BV , where A is the vector magnetic potential, defined such that B = V × A. Magnetic helicity is produced by means of kinetic helicity through magnetic induction but has its own particular significance, largely because Hm , unlike Hk , is an ideal invariant of the MHD equations. However, because A is undefined within the addition of an arbitrary scalar gradient VL (the gauge), Hm is not, in general, uniquely defined for a finite spatial domain. This ambiguity can be resolved by defining a gauge-invariant relative helicity with respect to a reference field Bp as follows [55]: HR = (A + Ap ) · (B − Bp ) dV , (3.2) V where V is the volume of the domain in question and Bp = V × Ap . The reference field is generally taken to be potential (V × Bp = 0) with a Coulomb gauge (V · Ap = 0). Furthermore, it is defined such that the component normal to the surface S that encloses V is the same as the actual field: Bp · n̂ = B · n̂ on S, which specifies Bp and Ap uniquely. If no field lines leave the domain, i.e. B · n̂ = 0 on S, then Ap = Bp = 0 and HR = Hm . Phil. Trans. R. Soc. A (2012) Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 Review. The solar dynamo 3059 If we take S to be the spherical surface of the Sun, or a single hemisphere, then the evolution equation for HR , obtained from equations (3.2) and (3.1), is given by [55] dHR 8p 4ph =− J · dS, (3.3) hJ · B dV − Ap × v × B + dt c V c S where c is the speed of light, J = c(4p)−1 V × B is the current density and we have used cgs units throughout. Equation (3.3) is gauge invariant (with respect to A). If h is constant in V then the first term on the RHS is equal to −8phc −1 Hc , where Hc = J · BV is the current helicity. Thus, the rate of change of HR is determined by ohmic dissipation, ∝ Hc , and the helicity flux through the surface S, given by the integrand of the second term on the RHS of (3.3). If the field is entirely contained within V (B · n̂ = 0 on S) and if the fluid is perfectly conducting (h = 0), then the magnetic helicity is conserved (HR = Hm , dHm /dt = 0). To demonstrate the significance of helicity and shear to the generation of largescale fields, we consider the longitudinally averaged induction equation (3.1) vB (3.4) = lBm · VU + V[vm × B + E − hV × B], vt where l = r sin q is the cylindrical radius and E = v × B . Throughout this paper, angular brackets denote averages over longitude (not to be confused with V , which indicates a volume integral) and primes indicate fluctuations about the mean, e.g. v = v − v. The subscript m denotes components in the meridional plane, e.g. Bm = Br r̂ + Bq q̂. With regard to the solar activity cycle, the primary challenge of solar dynamo theory is to determine the nature of the turbulent electromotive force (emf) E. Correlations between fluctuating fields and flows are generally attributed to convective motions but they may also arise from other phenomena, such as magnetic buoyancy (MB) or magneto-shear instabilities [4,6]. Explicit expressions for E can only be obtained for idealized scenarios, the best known being kinematic mean-field dynamo theory in which the field is passively advected by the flow (the Lorentz force is neglected) and it is assumed that characteristic length and time scales of the fluctuations are small relative to the mean. Under these conditions, the turbulent emf is linear in B and its components can be expressed (in Cartesian coordinates) as a converging infinite series [4,6] vBj + ··· . (3.5) Ei = aij Bj + bijk vxk Under the additional assumptions of quasi-isotropy (antisymmetric under mirror reflections but otherwise isotropic), and homogeneity, the a and b tensors reduce to scalar coefficients proportional to the kinetic helicity and the kinetic energy of the fluctuating motions, respectively. The a-effect is inherently helical. The generation of magnetic energy is associated with a transfer of magnetic helicity from fluctuating to mean fields [56]. In strictly kinematic models, this transfer is implicit (fluctuating fields never appear explicitly), but many mean-field models incorporate Lorentzforce feedbacks through slight adjustments to the basic mean-field paradigm [57]. To lowest order, Lorentz force feedbacks modify the a-effect by contributing an extra term such that a = ak + am = −(t/3)(Hk − (rc)−1 Hc ), where Hk and Hc are Phil. Trans. R. Soc. A (2012) Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 3060 M. S. Miesch the kinetic and current helicity associated with the fluctuating motions. A meanfield analogue of equation (3.3), accounting for the conservation of magnetic helicity in the absence of ohmic dissipation, can be used to derive an expression for am , the result being [32,57,58] ak , (3.6) a= 2 ) 1 + Rm B2 V /(Beq 2 = 4pru 2 and Rm = u 2 (3th)−1 are, respectively, the equipartition field where Beq strength and the magnetic Reynolds number associated with the turbulent flow. In short, Lorentz force feedbacks lead to a marked suppression of the turbulent a-effect for large Rm that can be attributed to the near conservation of magnetic helicity. The generation of helical mean flows gives rise to a build-up of small-scale helicity of the opposite sense that can efficiently quench the dynamo, unless it is either dissipated or removed from the domain by the helicity flux in equation (3.3). This conclusion is not restricted to the mean-field theory. In fact, the first derivation of am was based on an eddy-damped, quasi-normal Markovian (EDQNM) turbulence model that did not adopt the kinematic and scaleseparation assumptions required by traditional mean-field theory [59]. Results exhibited a self-similar inverse cascade of magnetic helicity from small to large scales,5 which was later demonstrated in idealized numerical simulations of MHD turbulence [60,61]. However, MHD convection simulations have demonstrated that rotation alone is not sufficient to promote the generation of large-scale fields by means of an inverse cascade of magnetic helicity; rather, results are sensitive to boundary conditions, geometry and the strength and orientation of rotational shear (reviewed by Tobias et al. [62] and Brandenburg [63]). Whatever the nature of the initial saturation on a dynamical time scale, the helicity will continue to evolve on a diffusive time scale according to equation (3.3). In a closed system, a steady state can only be reached if the current helicity vanishes J · B = 0. A helical mean field can still persist, provided the helicity of the opposite sign prevails at small scales such that J · B = −J · B . If J ∼ kB, where k is a characteristic wavenumber (k0 for the mean and k for the fluctuations), then saturation occurs when B2 ∼ k /k0 (B )2 . In other words, energy in the mean fields will grow slowly until it exceeds small-scale fields by the ratio of length scales. This was first derived and demonstrated in numerical simulations of MHD turbulence by Brandenburg [60]. The first term in equation (3.4) is the well-known U-effect whereby mean toroidal field Bf is generated from mean poloidal field Bm and amplified by means of rotational shear. The prominence of sunspots and the large amount of flux contained in toroidal relative to poloidal fields suggests that the U-effect must play an essential role in cyclic solar activity (§4). Rotational shear can also influence the topology of large-scale fields by inducing a flux of smallscale magnetic helicity along isorotation surfaces that can potentially alleviate dynamical a-quenching [57,64]. Global numerical simulations of convective dynamos verify that rotation promotes and regulates dynamo action by means of helicity and shear. Figure 4 illustrates a simulation of a young solar-type star rotating at five times the solar 5 As opposed to the forward cascade of energy from large to small scales. Phil. Trans. R. Soc. A (2012) Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 Review. The solar dynamo 3061 rate [65]. Enhanced convective Reynolds stresses produce prominent differential rotation, which in turn generate strong toroidal fields (approx. 10 kG) in confined bands with opposite polarity in the Northern and Southern Hemispheres, as inferred from solar observations (figure 1). These toroidal flux structures account for over half of the total magnetic energy of the dynamo and persist amid the intense turbulence of the convection zone. This is in contrast to the simulation at the solar rotation rate where the kinetic energy of the convection exceeds that in the differential rotation and turbulent pumping effectively confines such structures to the base of the convection zone and tachocline [66–68]. Simulations such as this often exhibit quasi-cyclic polarity reversals, as demonstrated in figure 4d for a fast rotator. By using sophisticated numerical methods featuring implicit time stepping and minimal numerical dissipation, Ghizaru et al. [68] (see [37]) have recently achieved the first global MHD simulation of solar convection to exhibit cyclic polarity reversals on a decadal time scale (figure 5). Understanding the origin of such behaviour requires a closer look at the fundamental physical mechanisms that give rise to magnetic activity cycles in simulations and stars, which we now explore (see §4). 4. Origins of cyclic solar activity Dynamo models of the solar cycle are generally based on the longitudinally averaged induction equation (3.4), with some theoretically or empirically motivated parametrization for the turbulent emf E. Nearly all recent models invoke an emf of the form E = P(r, q : B) + g × B + ht J. (4.1) The first term on the RHS is the source term for the mean poloidal field. The simplest representation is a quasi-isotropic, homogeneous, kinematic a-effect,6 as described in §3b: P(r, q : B) = ak Bf f̂. However, models often employ a nonlinear quenching formulation such as in equation (3.6) in order to limit the dynamo amplitude. Furthermore, many current models employ a nonlocal source term such that the value of P(r, q : B) near the solar surface is related to the integrated toroidal flux near the base of the convection zone [35,69,70]. This is intended to implicitly capture the mean magnetic induction arising from the buoyant destabilization, rise, emergence and dispersal of the toroidal flux structures that give rise to active regions; namely, the BL mechanism discussed in §2a. Although the BL mechanism is phenomenologically distinct from the turbulent a-effect, it is functionally similar in that Ef is linearly proportional to the mean toroidal field Bf , apart from quenching and non-locality. The second and third terms in equation (4.1) are transport terms and may be related directly to the kinematic expansion in equation (3.5). The g term arises from the antisymmetric off-diagonal components of the a tensor and reflects magnetic pumping by convective motions. Again, in terms of magnetic induction, this is phenomenologically distinct but functionally equivalent to a mean MC vm ; 6 Strictly speaking, the a appearing here is the symmetric part of the full a tensor appearing in k equation (3.5). The antisymmetric part reflects transport and is expressed here separately as the magnetic pumping vector g, which may be regarded as a pseudo-flow. Phil. Trans. R. Soc. A (2012) Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 3062 (b) (c) (d) latitude (°) (a) M. S. Miesch 50 0 –50 6 8 10 12 14 16 18 time (yr) Figure 4. Convective dynamo simulation of a young, rapidly rotating, solar-like star (rotation period 5.6 days; from Brown et al. [65]). (a) Radial velocity near the top of the convection zone in a Molleweide projection (yellow upflow, blue/black downflow; saturation range ±70 m s−1 ). (b) Timeaveraged angular velocity (white faster, blue slower; range 1980–2100 nHz) and (c) toroidal magnetic field (red eastward, blue westward; range ±10 kG), both averaged over longitude. (d) Mean toroidal field in the lower convection zone as a function of latitude and time, illustrating polarity reversals on a time scale of roughly 4 years (colour table as in c). Poleward propagating magnetic features are accompanied by torsional oscillations (cf. §2c). (b) 90 60 30 0 –30 –60 –90 latitude (°) (a) 0 50 100 150 200 time (yr) 250 300 Figure 5. Magnetic cycles in a solar convection simulation (from Racine et al. [37]; see also Ghizaru et al. [68]). (a) Snapshot of longitudinal field Bf near the base of the convection zone and (b) longitudinally averaged toroidal field Bf as a function of latitude and time. Yellow/orange and green/blue regions indicate eastward and westward fields, respectively, with a saturation range of (a) ±2 kG and (b) ±4 kG. Note the regularity of reversals and the slight equatorwards propagation. cf. equation (3.4). There is also a zonal component to g which can transport poloidal magnetic flux in longitude [37,71]. The turbulent diffusion ht is usually assumed to be isotropic, although this is not strictly consistent with a non-zero g. As with the a-effect, mean-field models are increasingly incorporating implicit Lorentz-force feedbacks by means of h-quenching [72,73]. In most solar dynamo models, the mean toroidal field is generated by the U-effect (see §3b) near the base of the convection zone—region III in figure 6a (for a notable exception focusing on the NSSL, see [74]). Although one may expect that the strong radial shear in the tachocline is responsible, many recent mean-field models find that the latitudinal shear in the lower convection zone is more efficient at generating the global, equatorially antisymmetric toroidal bands that give rise to sunspots [35,75,76]. By contrast, the principal mechanisms and even the location of the poloidal source term P are currently a matter of debate. Photospheric measurements suggest that the flux emerging in active regions is more than sufficient to account for the magnetic polarity reversal in the polar caps, as suggested by figure 1a. The flux in a single active region is of order Phil. Trans. R. Soc. A (2012) Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 3063 Review. The solar dynamo (a) (b) MC 80 latitude CT MB 60 40 20 0 III II I 5 10 time (yr) 15 Figure 6. (a) Schematic illustrating distinct physical regimes in the solar envelope including (I) the near-surface shear layer in the upper convection zone, (II) the bulk of the convection zone and (III) the tachocline, overshoot region and lower convection zone (from Miesch & Toomre [32]). Coupling mechanisms between the regions are also highlighted, including the meridional circulation (MC), convective transport (CT) and magnetic buoyancy (MB). (b) A representative example of a Babcock–Leighton flux-transport solar dynamo model from Rempel [35]. Contours indicate the mean toroidal field near the base of the convection zone (solid lines eastward, dotted lines westward; range ±12.8 kG) and the image represents the mean radial field Br near the surface (red outward, blue inward; range ±100 G), both plotted versus latitude and time. 1022 –1023 Mx, with about 1024 –1025 Mx emerging over the course of an 11 year sunspot cycle [77]. If the average magnetic field strength poleward of ±70◦ latitude is roughly 10 G, then a polarity reversal would require a flux of order 4 × 1022 Mx. Thus, even if much of the flux emerging in active regions is lost through magnetic reconnection or vertical transport of magnetic loops, the BL mechanism can account for reversals of the surface dipole moment—threading region I in figure 6a. Yet, can the redistribution of active region flux in the photospheric boundary layer really determine the global dipole moment of the entire solar envelope? The flux in photospheric active regions (region I) and the associated magnetic energy is likely to be a small fraction of the total unsigned flux and energy in the remainder of the convective zone (regions II and III), where most of the mass and kinetic energy of the solar envelope reside. Photospheric observations suggest that active regions make up only a small fraction of the magnetic flux and energy in the photosphere, let alone the entire convection zone (§2b). Furthermore, cosmogenic isotope measurements suggest that the 22 year magnetic activity cycle persisted throughout the Maunder minimum, a 70 year interval in the seventeenth century when there were very few sunspots [78]. Convection simulations exhibit cyclic magnetic activity without capturing the dynamics of flux emergence and dispersal that underlie the BL mechanism (figures 4 and 5; although aspects of these simulations do not match solar observations). Amplitude and parity issues have also motivated several authors to advocate for additional deep-seated sources of poloidal field in BL dynamo models, including convection and tachocline instabilities [79]. Still, the BL mechanism is undeniably operating and plays a prominent role in most recent solar dynamo models. Cyclic behaviour in the Sun is intimately linked with propagation; activity bands migrate equatorwards, and residual vertical flux migrates poleward during the course of each cycle, as evident in figure 1a. Within the MHD framework, there Phil. Trans. R. Soc. A (2012) Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 3064 M. S. Miesch are essentially three physical mechanisms that can give rise to the propagation of mean fields and can thus establish cyclic variability: (A) a spatial offset between poloidal and toroidal sources (i.e. a dynamo wave), (B) advection by the mean MC, and (C) transport by convection or other non-axisymmetric motions (magnetic pumping, turbulent diffusion, etc.). These are not mutually exclusive; multiple mechanisms can operate simultaneously and some phenomena, such as MB instabilities, can contribute to all three. The earliest solar dynamo models were of type (A). Cyclic behaviour is attributed to a phase shift between the a-effect and the U-effect that admits travelling wave solutions to the linear, kinematic induction equation (3.4) [6,30]. The propagation speed and the growth rate of the eigenmodes are proportional to (a|VU|)1/2 , with waves travelling along isorotation surfaces. The natural location for such a dynamo wave is the low-latitude tachocline, where the nearly radial orientation of VU and the sign of a (∝ −Hk ) are favourable for equatorwards propagation. This, together with the need to avoid dynamical a-quenching, motivated interface dynamo models [80,81] whereby the a-effect occurs in the convection zone (region II) while the U-effect occurs in the tachocline (region III). Cyclic behaviour still hinges on a dynamo wave but now the cycle period depends also on the coupling mechanism between regions II and III, usually taken to be turbulent diffusion. Most recent solar dynamo models advocate mechanism (B), whereby an equatorwards circulation of 1–3 m s−1 near the base of the convection zone (region III) accounts for the equatorwards migration of active bands as manifested in the solar butterfly diagram and largely determines the period of the solar cycle [35,69,73,75,76,79]. These are known as flux-transport (FT) dynamo models, most of which attribute poloidal field generation to the BL mechanism. Furthermore, many BL–FT dynamo models operate in a socalled advection-dominated regime whereby the coupling between regions I and III is achieved by the MC. Such models generally employ single-celled circulation profiles as illustrated in figure 6a, but recent results by Dikpati et al. [82] based on modelling and photospheric observations suggest that the presence of a high-latitude counter-cell may account for why cycles are typically shorter (approx. 11 years) than the most recent one (approx. 13 years). An equatorwards circulation of several m s−1 at the base of the convection zone is consistent with global convection simulations, but estimates of the efficiency of convective transport (CT) are about an order of magnitude larger than typical circulation speeds, calling into question the advection-dominated regime [67,83,84]. Magnetic pumping or anisotropic magnetic diffusivity, mechanism (C), can in principle induce cyclic activity in mean-field models but in practice it generally operates in conjunction with other processes. For example, BL models can be constructed in which turbulent diffusion or magnetic pumping account for the coupling between regions I and III [85–87]. The challenge here is to account for the relative inefficiency of CT; if coupling were to occur on a turnover time scale (approx. 1 month), cycles would be too short. The picture emerging from global convection simulations is only now beginning to take shape. For the simulation shown in figure 4, cyclic activity appears to be primarily owing to a nonlinear dynamo wave (A), even though E is not well represented by a simple a-effect. Poleward circulations associated with polar Phil. Trans. R. Soc. A (2012) Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 Review. The solar dynamo 3065 slip instabilities, and torsional oscillations are also present and may play a role in polarity reversals. The simulation shown in figure 5 appears to be more amenable to a mean-field treatment and shows evidence for all three mechanisms for cyclic variability: (A) spatial offsets between toroidal and poloidal sources, (B) an equatorwards circulation of approximately 0.2 m s−1 near the base of the convection zone, and (C) equatorwards turbulent pumping of a similar amplitude, approximately 0.2 m s−1 [37]. 5. Summary and outlook Kinematic, axisymmetric mean-field models are still the workhorse of solar dynamo theory, but verifying their underlying assumptions and accounting for the full richness of solar magnetic activity will require more sophisticated threedimensional MHD simulations. A major challenge for the future will be to link the cyclic global dynamo and the magnetic carpet into a unified theoretical paradigm. Three-dimensional dynamo simulations must also be linked to simulations and observations of flux emergence. Current dynamo models take the strength of the mean toroidal field near the base of the convection zone as a proxy for the sunspot number (figure 6b). However, there is no guarantee that this relationship is linear and the details of how it unfolds may have important implications for the operation of the dynamo as well as solar cycle predictability and the space weather implications of emergent flux. The solar cycle itself still harbours many mysteries. What processes dominate the generation of large-scale poloidal fields as encapsulated in the essential but enigmatic turbulent emf? Can the BL mechanism overwhelm the turbulent a-effect? Might MHD instabilities contribute to mean-field generation? What regulates the cycle period and amplitude? Can a weak but persistent meridional flow prevail over turbulent transport? Does the spectral and spatial flux of magnetic helicity play a crucial role in determining the large-scale magnetic topology and dynamo saturation? Is flux emergence merely a by-product of dynamo action or is it essential for the establishment of cyclic activity? How does the Sun fit within the context of other stars? Inevitably, answers can only come from diligent observation and innovative modelling. Thanks to continuing advances in instrumentation and high-performance computing, the future is bright. We are grateful to Paul Charbonneau, Guiliana De Toma, David Hathaway, Rachael Howe, Clare Parnell, Étienne Racine, Matthias Rempel and Marty Snow for help in preparing and interpreting figures. We also thank Paul Charbonneau, Matthias Rempel, Janet Luhmann and the two anonymous referees for discussions (referees excluded) and comments on the manuscript. Research support is provided through NASA grants NNH09AK14I and NNX08AI57G and computing resources were provided by NASA HEC (Pleiades) and NSF PACI (NCSA, TACC, PSC and SDSC). The National Center for Atmospheric Research is sponsored by the National Science Foundation. References 1 Hale, G. E. 1908 On the probable existence of a magnetic field in sun-spots. Astrophys. J. 28, 315–343. (doi:10.1086/141602) 2 Schwabe, S. H. 1844 Sonnen–Beobachtungen im Jahre 1843. Astron. Nachr. 21, 234–235. (doi:10.1002/asna.18440211505) Phil. Trans. R. Soc. A (2012) Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 3066 M. S. Miesch 3 Viereck, R. A., Floyd, L. E., Crane, P. C., Woods, T. N., Knapp, B. G., Rottman, G., Weber, M., Puga, L. C. & DeLand, M. T. 2004 A composite Mg II index spanning from 1978 to 2003. Space Weather 2, S10005. (doi:10.1029/2004SW000084) 4 Ossendrijver, M. 2003 The solar dynamo. Astron. Astrophys. Rev. 11, 287–367. (doi:10.1007/s00159-003-0019-3) 5 Solanki, S. K., Inhester, B. & Schüssler, M. 2006 The solar magnetic field. Rep. Prog. Phys. 69, 563–668. (doi:10.1088/0034-4885/69/3/R02) 6 Charbonneau, P. 2010 Dynamo models of the solar cycle. Living Rev. Solar Phys. 7, 3. See http://www.livingreviews.org/lrsp-2010-3. 7 Makarov, V. I., Tlatov, A. G., Callebaut, D. K., Obridko, V. N. & Shelting, B. D. 2001 Large-scale magnetic field and sunspot cycles. Solar Phys. 198, 409–421. (doi:10.1023/A:1005249531228) 8 Fan, Y. 2009 Magnetic fields in the solar convection zone. Living Rev. Solar Phys. 6, 4. See http://www.livingreviews.org/lrsp-2009-4. 9 Schüssler, M. & Rempel, M. 2005 The dynamical disconnection of sunspots from their magnetic roots. Astron. Astrophys. 441, 337–346. (doi:10.1051/0004-6361:20052962) 10 Baumann, I., Schmitt, D. & Schüssler, M. 2006 A necessary extension of the surface flux transport model. Astron. Astrophys. 446, 307–314. 11 Schrijver, C. J. & Liu, Y. 2008 The global solar magnetic field through a full sunspot cycle: observations and model results. Solar Phys. 252, 19–31. (doi:10.1007/s11207-008-9240-6) 12 Ulrich, R. K. 2010 Solar meridional circulation from doppler shifts of the FeI line at 5250 Å as measured by the 150-foot solar tower telescope at the Mt. Wilson observatory. Astrophys. J. 725, 658–669. (doi:10.1088/0004-637X/725/1/658) 13 Babcock, H. W. 1961 The topology of the sun’s magnetic field and the 22-year cycle. Astrophys. J. 133, 572–587. (doi:10.1086/147060) 14 Leighton, R. B. 1964. Transport of magnetic fields on the sun. Astrophys. J. 140, 1547–1562. (doi:10.1086/148058) 15 Frölich, C. & Lean, J. 2004. Solar radiative output and its variability: evidence and mechanisms. Astron. Astrophys. Rev. 12, 273–320. (doi:10.1007/s00159-004-0024-1) 16 Hathaway, D. H. 2010 The solar cycle. Living Rev. Solar Phys. 7, 1. See http://www. livingreviews.org/lrsp-2010-1. 17 Schrijver, C. J. & Zwaan. 2000 Solar and stellar magnetic activity. Cambridge, UK: Cambridge University Press. 18 Usoskin, I. G. 2008 A history of solar activity over millennia. Living Rev. Solar Phys. 5, 3. See http://www.livingreviews.org/lrsp-2008-3. 19 Trujillo Bueno, J., Shchukina, N. & Asensio Ramos, A. 2004 A substantial amount of hidden magnetic energy in the quiet sun. Nature 430, 326–329. (doi:10.1038/nature02669) 20 Lites, B. W. et al. 2008 The horizontal magnetic flux of the quiet-sun internetwork as observed with the HINODE spectro-polarimeter. Astrophys. J. 672, 1237–1253. (doi:10.1086/522922) 21 Pietarila Graham, J., Danilovic, S. & Schüssler, M. 2009 Turbulent magnetic fields in the quiet sun: implications of HINODE observations and small-scale dynamo simulations. Astrophys. J. 693, 1728–1735. (doi:10.1088/0004-637X/693/2/1728) 22 Solanki, S. K. et al. 2010 SUNRISE: instrument, mission, data, and first results. Astrophys. J. Lett. 723, L127–L133. (doi:10.1088/2041-8205/723/2/L127) 23 Title, A. M. & Schrijver, C. J. 1998 The sun’s magnetic carpet. In Cool stars, stellar systems and the sun (eds R. A. Donahue & J. A. Bookbinder), pp. 345–358. ASP Conference Series, vol. 154. San Francisco, CA: Astronomical Society of the Pacific. 24 Hagenaar, H. J., Schrijver, C. J. & Title, A. M. 2003 The properties of small magnetic regions on the solar surface and the implications for the solar dynamo(s). Astrophys. J. 584, 1107–1119. (doi:10.1086/345792) 25 Parnell, C. E., DeForest, C. E., Hagenaar, H. J., Johnston, B. A., Lamb, D. A. & Welsch, B. T. 2009 A power-law distribution of solar magnetic fields over more than five decades in flux. Astrophys. J. 698, 75–82. (doi:10.1088/0004-637X/698/1/75) 26 Thornton, L. M. & Parnell, C. E. 2011 Small-scale flux emergence observed using HINODE/SOT. Solar Phys. 269, 13–40. (doi:10.1007/s11207-010-9656-7) Phil. Trans. R. Soc. A (2012) Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 Review. The solar dynamo 3067 27 Parker, E. N. 1978 Hydraulic concentration of magnetic fields in the solar photosphere. VI. Adiabatic cooling and concentration in downdrafts. Astrophys. J. 221, 368–377. (doi:10.1086/156035) 28 Crouch, A. D., Charbonneau, P. & Thibault, K. 2001 Supergranulation as an emergent length scale. Astrophys. J. 662, 715–729. (doi:10.1086/515564) 29 Parker, E. N. 1955 Hydromagnetic dynamo models. Astrophys. J. 122, 293–314. (doi:10.1086/ 146087) 30 Moffatt, H. K. 1978 Magnetic field generation in electrially conducting fluids. Cambridge, UK: Cambridge University Press. 31 Krause, F. & Rädler, K.-H. 1980 Mean-field magnetohydrodynamics and dynamo theory. Oxford, UK: Pergamon Press. 32 Miesch, M. S. & Toomre, J. 2009 Turbulence, magnetism, and shear in stellar interiors. Annu. Rev. Fluid Mech. 41, 317–345. (doi:10.1146/annurev.fluid.010908.165215) 33 Schou, J. et al. 1998 Helioseismic studies of differential rotation in the solar envelope by the solar oscillations investigation using the Michelson Doppler imager. Astrophys. J. 505, 390–417. (doi:10.1086/306146) 34 Howe, R., Christensen-Dalsgaard, J., Hill, F., Komm, R., Schou, J. & Thompson, M. J. 2009 A note on the torsional oscillation at solar minimum. Astrophys. J. Lett. 701, L87–L90. (doi:10.1088/0004-637X/701/2/L87) 35 Rempel, M. 2006 Flux-transport dynamos with Lorentz force feedback on differential rotation and meridional flow: saturation mechanism and torsional oscillations. Astrophys. J. 647, 662–675. (doi:10.1086/505170) 36 Brown, B. P., Browning, M. K., Brun, A. S., Miesch, M. S. & Toomre, J. 2010 Persistent magnetic wreathes in a rapidly rotating sun. Astrophys. J. 711, 424–438. (doi:10.1088/ 0004-637X/711/1/424) 37 Racine, É., Charbonneau, P., Ghizaru, M., Bouchat, A. & Smolarkiewicz, P. K. 2011 On the mode of dynamo action in a global large-eddy simulation of solar convection. Astrophys. J. 735, 46. (doi:10.1088/0004-637X/735/1/46) 38 Rempel, M. 2007 Origin of solar torsional oscillations. Astrophys. J. 655, 651–659. (doi:10. 1086/509866) 39 Strassmeier, K. G. 2009. Starspots. Astron. Astrophys. Rev. 17, 251–308. (doi:10.1007/ s00159-009-0020-6) 40 Schrijver, C. J., DeRosa, M. L. & Title, A. M. 2003 Asterospheric magnetic fields and winds of cool stars. Astrophys. J. 590, 493–501. (doi:10.1086/374982) 41 Pizzolato, N., Maggio, A., Micela, G., Sciortino, S. & Ventura, P. 2003 The stellar activity-rotation relationship revisited: dependence of saturated and non-saturated X-ray emission regimes on stellar mass for late-type dwarfs. Astron. Astrophys. 397, 147–157. (doi:10.1051/0004-6361:20021560) 42 Saar, S. H. 2009 The activity cycles and surface differential rotation of single dwarfs. ASP Conf. Ser. 416, 375–384. (doi:10.1086/177517) 43 Donahue, R. A., Saar, S. H. & Baliunas, S. L. 1996 A relationship between mean rotation period in lower main-sequence stars and its observed range. Astrophys. J. 466, 384–391. (doi:10.1086/177517) 44 Reiners, A. & Schmitt, J. H. M. M. 2003 Rotation and differential rotation in field F- and G-type stars. Astron. Astrophys. 398, 647–661. (doi:10.1051/0004-6361:20021642) 45 Barnes, J. R., Collier Cameron, A., Donati, J.-F., James, D. J., Marsden, S. C. & Petit, P. 2005 The dependence of differential rotation on temperature and rotation. Mon. Not. R. Astron. Soc. 357, L1–L5. (doi:10.1111/j.1745-3933.2005.08587.x) 46 Christensen, U. R., Holzwarth, V. & Reiners, A. 2009 Energy flux determines magnetic field strength of planets and stars. Nature 457, 167–169. (doi:10.1038/nature07626) 47 Rempel, M. 2008 Solar and stellar activity cycles. J. Phys. Conf. Ser. 118, 012032. (doi:10.1088/1742-6596/118/1/012032) 48 Lanza, A. F. 2010 Stellar magnetic cycles. In Solar and stellar variability: impact on earth and planets, Proc. IAU Symp. 264 (eds A. Andrei, A. Kosovichev & J.-P. Rozelot), pp. 120–129. Cambridge, UK: Cambridge University Press. Phil. Trans. R. Soc. A (2012) Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 3068 M. S. Miesch 49 Schekochihin, A. A., Cowley, S. C., Taylor, S. F., Maron, J. L. & McWilliams, J. C. 2004 Simulations of the small-scale turbulent dynamo. Astrophys. J. 612, 276–307. (doi:10. 1086/422547) 50 Tobias, S. M., Cattaneo, F. & Boldyrev, S. 2011 MHD turbulence and dynamos. In The nature of turbulence (eds P. A. Davidson, Y. Kaneda & K. R. Sreenivasan). Cambridge, UK: Cambridge University Press. 51 Ponty, Y., Mininni, P. D., Montgomery, D. C., Pinton, J.-F., Politano, H. & Pouquet, A. 2005 Numerical study of dynamo action at low magnetic Prandtl numbers. Phys. Rev. Lett. 94, 164502. (doi:10.1103/PhysRevLett.94.164502) 52 Iskakov, A. B., Schekochihin, A. A., Cowley, S. C., McWilliams, J. C. & Proctor, M. R. E. 2007 A numerical demonstration of fluctuation dynamo at low magnetic Prandtl numbers. Phys. Rev. Lett. 98, 208501. (doi:10.1103/PhysRevLett.98.208501) 53 Cattaneo, F. 1999 On the origin of magnetic fields in the quiet photosphere. Astrophys. J. 515, L39–L42. (doi:10.1086/311962) 54 Vögler, A. & Schüssler, M. 2007 A solar surface dynamo. Astron. Astrophys. 465, L43–L46. 55 Berger, M. A. & Field, G. B. 1984 The topological properties of magnetic helicity. J. Fluid Mech. 147, 133–148. (doi:10.1017/S0022112084002019) 56 Seehafer, N. 1996 Nature of the a effect in magnetohydrodynamics. Phys. Rev. E 53, 1283–1286. (doi:10.1103/PhysRevE.53.1283) 57 Brandenburg, A. & Subramanian, K. 2005 Astrophysical magnetic fields and nonlinear dynamo theory. Phys. Rep. 417, 1–209. (doi:10.1016/j.physrep.2005.06.005) 58 Gruzinov, A. V. & Diamond, P. H. 1994 Self-consistent theory of mean-field electrodynamics. Phys. Rev. Lett. 72, 1651–1653. (doi:10.1103/PhysRevLett.72.1651) 59 Pouquet, A., Frisch, U. & Léorat, J. 1976 Strong MHD helical turbulence and the nonlinear dynamo effect. J. Fluid Mech. 77, 321–354. (doi:10.1017/S0022112076002140) 60 Brandenburg, A. 2001 The inverse cascade and nonlinear alpha-effect in simulations of isotropic helical hydromagnetic turbulence. Astrophys. J. 550, 824–840. (doi:10.1086/319783) 61 Alexakis, A., Mininni, P. D. & Pouquet, A. 2006 On the inverse cascade of magnetic helicity. Astrophys. J. 640, 335–343. (doi:10.1086/500082) 62 Tobias, S. M., Cattaneo, F. & Brummell, N. H. 2008 Convective dynamos with penetration, rotation, and shear. Astrophys. J. 685, 596–605. (doi:10.1086/590422) 63 Brandenburg, A. 2009 Advances in theory and simulations of large-scale dynamos. Space Sci. Rev. 144, 87–104. (doi:10.1007/s11214-009-9490-0) 64 Vishniac, E. T. & Cho, J. 2001 Magnetic helicity conservation and astrophysical dynamos. Astrophys. J. 550, 752–760. (doi:10.1086/319817) 65 Brown, B. P., Miesch, M. S., Browning, M. K., Brun, A. S. & Toomre, J. 2011 Magnetic cycles in a convective dynamo simulation of a young solar-type star. Astrophys. J. 731, 69. (doi:10.1088/0004-637X/731/1/69) 66 Browning, M. K., Miesch, M. S., Brun, A. S. & Toomre, J. 2006 Dynamo action in the solar convection zone and tachocline: pumping and organization of toroidal fields. Astrophys. J. 648, L157–L160. (doi:10.1086/507869) 67 Miesch, M. S., Brown, B. P., Browning, M. K., Brun, A. S. & Toomre, J. 2010 Magnetic cycles and meridional circulation in global models of solar convection. In Astrophysical dynamics, from stars to galaxies, Proc. IAU Symp. 271 (eds N. Brummell, A. S. Brun, M. S. Miesch & Y. Ponty), pp. 261–269. Cambridge, UK: Cambridge University Press. 68 Ghizaru, M., Charbonneau, P. & Smolarkiewicz, P. K. 2010 Magnetic cycles in global large-eddy simulations of solar convection. Astrophys. J. 715, L133–L137. (doi:10.1088/20418205/715/2/L133) 69 Dikpati, M. & Charbonneau, P. 1999 A Babcock–Leighton flux transport dynamo with solar-like differential rotation. Astrophys. J. 518, 508–520. (doi:10.1086/307269) 70 Munoz-Jaramillo, A., Nandy, D., Martens, P. C. H. & Yeates, A. R. 2010 A double-ring algorithm for modeling solar active regions: unifying kinematic dynamo models and surface flux-transport simulations. Astrophys. J. 720, L20–L25. (doi:10.1088/2041-8205/720/1/L20) 71 Ossendrijver, M., Stix, M., Brandenburg, A. & Rüdiger, G. 2002, Magnetoconvection and dynamo coefficients. II. Field-direction dependent pumping of magnetic field. Astron. Astrophys. 394, 735–745. (doi:10.1051/0004-6361:20021224) Phil. Trans. R. Soc. A (2012) Downloaded from http://rsta.royalsocietypublishing.org/ on June 17, 2017 Review. The solar dynamo 3069 72 Guerrero, G., Dikpati, M. & de Gouveia Dal Pino, E. M. 2009 The role of diffusivity quenching in flux-transport dynamo models. Astrophys. J. 701, 725–736. (doi:10.1088/ 0004-637X/701/1/725) 73 Munoz-Jaramillo, A., Nandy, D. & Martens, P. C. H. 2011 Magnetic quenching of turbulent diffusivity: reconciling mixing-length theory estimates with kinematic dynamo models of the solar cycle. Astrophys. J. 727, L23–L28. (doi:10.1088/2041-8205/727/1/L23) 74 Brandenburg, A. 2005 The case for a distributed solar dynamo shaped by near-surface shear. Astrophys. J. 625, 539–547. (doi:10.1086/429584) 75 Dikpati, M. & Gilman, P. A. 2006 Simulating and predicting solar cycles using a flux-transport dynamo. Astrophys. J. 649, 498–514. (doi:10.1086/506314) 76 Munoz-Jaramillo, A., Nandy, D. & Martens, P. C. H. 2009 Helioseismic data inclusion in solar dynamo models. Astrophys. J. 698, 461–478. (doi:10.1088/0004-637X/698/1/461) 77 Schrijver, C. J. & Harvey, K. L. 1994 The photospheric magnetic flux budget. Solar Phys. 150, 1–18. (doi:10.1007/BF00712873) 78 Beer, J., Tobias, S. M. & Weiss, N. O. 1998 An active sun throughout the maunder minimum. Solar Phys. 181, 237–148. (doi:10.1023/A:1005026001784) 79 Dikpati, M. & Gilman, P. A. 2001 Flux-transport dynamos with a-effect from global instability of tachocline differential rotation: a solution for magnetic parity selection in the sun. Astrophys. J. 559, 428–442. (doi:10.1086/322410) 80 Parker, E. N. 1993 A solar dynamo surface wave at the interface between convection and nonuniform rotation. Astrophys. J. 408, 707–719. (doi:10.1086/172631) 81 Charbonneau, P. & MacGregor, K. B. 1997 Solar interface dynamos. II. Linear, kinematic models in spherical geometry. Astrophys. J. 486, 502–520. (doi:10.1086/304485) 82 Dikpati, M., Gilman, P. A., de Toma, G. & Ulrich, R. 2010 Impact of changes in the sun’s conveyor-belt on recent solar cycles. Geophys. Res. Lett. 37, L14107. (doi:10.1029/ 2010GL044143) 83 Miesch, M. S., Elliott, J. R., Toomre, J., Clune, T. C., Glatzmaier, G. A. & Gilman, P. A. 2000 Three-dimensional spherical simulations of solar convection. I. Differential rotation and pattern evolution achieved with laminar and turbulent states. Astrophys. J. 532, 593–615. (doi:10.1086/ 308555) 84 Miesch, M. S., Brun, A. S., DeRosa, M. L. & Toomre, J. 2008 Structure and evolution of giant cells in global models of solar convection. Astrophys. J. 673, 557–575. (doi:10.1086/523838) 85 Jiang, J., Chatterjee, P. & Choudhuri, A. R. 2007 Solar activity forecast with a dynamo model. Mon. Not. R. Astron. Soc. 381, 1527–1542. (doi:10.1111/j.1365-2966.2007.12267.x) 86 Yeats, A. R., Nandy, D. & Mackay, D. 2008, Exploring the physical basis of solar cycle predictions: flux transport dynamics and persistence of memory in advection- versus diffusiondominated solar convection zones. Astrophys. J. 673, 544–556. (doi:10.1086/524352) 87 Guerrero, G. & de Gouveia Dal Pino, E. M. 2008 Turbulent magnetic pumping in a Babcock–Leighton solar dynamo model. Astron. Astrophys. 485, 267–273. (doi:10.1051/ 0004-6361:200809351) Phil. Trans. R. Soc. A (2012)
© Copyright 2026 Paperzz