<<

A&A 478, 1–8 (2008) DOI: 10.1051/0004-6361:20077172 & c ESO 2008 Astrophysics

Stability of toroidal magnetic fields in rotating stellar zones

L. L. Kitchatinov1,2 and G. Rüdiger1

1 Astrophysikalisches Institut Potsdam, An der Sternwarte 16, 14482, Potsdam, Germany e-mail: [lkitchatinov;gruediger]@aip.de 2 Institute for Solar-Terrestrial Physics, PO Box 291, Irkutsk 664033, Russia e-mail: [email protected]

Received 26 January 2007 / Accepted 14 October 2007

ABSTRACT

Aims. Two questions are addressed as to how strong magnetic fields can be stored in rotating stellar radiation zones without being subjected to pinch-type instabilities and how much radial mixing is produced if the fields are unstable. Methods. Linear equations are derived for weak disturbances of magnetic and velocity fields, which are global in horizontal dimen- sions but short–scaled in radius. The linear formulation includes the 2D theory of stability to strictly horizontal disturbances as a special limit. The eigenvalue problem for the derived equations is solved numerically to evaluate the stability of toroidal field patterns with one or two latitudinal belts under the influence of rigid rotation. Results. Radial displacements are essential for magnetic instability. It does not exist in the 2D case of strictly horizontal disturbances. Only stable (magnetically modified) r-modes are found in this case. The instability recovers in 3D. The minimum field strength Bmin for onset of the instability and radial scales of the most rapidly growing modes are strongly influenced by finite diffusion, the scales are indefinitely short if diffusion is neglected. The most rapidly growing modes for the have radial scales of about 1 Mm. In the upper part of the solar radiation zone, Bmin 600 G. The toroidal field can exceed this value only marginally, for otherwise the radial mixing produced by the instability would be too strong to be compatible with the observed lithium abundance. Key words. instabilities – magnetohydrodynamics (MHD) – : interiors – stars: magnetic fields – Sun: magnetic fields

1. Introduction produces is not well-known so far. As the radial mixing is rel- evant to the transport of light elements, a theory of the mixing The old problem of the hydromagnetic stability of stellar ra- compared with the observed abundances can help to restrict the diation zones attains renewed interest in relation to Ap amplitudes of the internal magnetic fields (Barnes et al. 1999). magnetism (Braithwaite & Spruit 2004), the solar In the present paper the vertical mixing produced by the Tayler (Gilman 2005), and the transport processes in stably stratified instability is estimated and then used to evaluate the upper limit stellar interiors (Barnes et al. 1999). on the magnetic field amplitude. Which fields the stellar radiative cores can possess is rather In this paper, the equations governing the linear stability of uncertain. The resistive decay in radiation zones is so slow that toroidal magnetic fields in a differentially rotating radiation zone primordial fields can be stored there (Cowling 1945). Whether are derived. They are solved for two latitudinal profiles of the 5 the fields of 10 G that can influence g-modes of solar oscilla- toroidal field with one and with two belts in latitude but only tions (Rashba et al. 2006) or whether still stronger fields that for rigid rotation. The unstable modes are expected to have the can cause neutrino oscillations (Burgess et al. 2003) can indeed longest possible horizontal scales (Spruit 1999). Accordingly, survive inside the Sun mainly depends on their stability. our equations are global in horizontal dimensions. They are, Among several instabilities to which the fields can be sub- however, local in radius, i.e. the radial scales are assumed to be jected (Acheson 1978), the current-driven (pinch-type) instabil- short, kr  1(k is radial wave number). The computations con- ity of toroidal fields (Tayler 1973) is probably the most relevant firm that the most rapidly growing modes have kr ∼ 103,but one because it proceeds via almost horizontal displacements. they are global in latitude. The derived equations reproduce the The radial motions in radiative cores are strongly suppressed 2D approximation as a special limit. An exactly solvable case by . Watson (1981) estimated the ratio of radial (ur) shows that the Tayler instability is missing in the 2D approxi- to horizontal (uθ) velocities of slow (subsonic) motions in ro- mation. It recovers, however, in the 3D case. Finite diffusion is 2 2 tating stars as ur/uθ ∼ Ω /N , where Ω is the basic angular found important for the instability. It prevents the preference for velocity and N the buoyancy (Brunt-Väisälä) frequency. This the indefinitely short radial scales found for ideal MHD. The dis- ratio is small in radiative cores (Fig. 1). If the radial velocities turbances that first become unstable as the field strength grows are completely neglected, the stability analysis can be done in a have small but finite radial scales. For those scales, the minimum 2D approximation with purely horizontal displacements (Watson field producing the instability is strongly reduced by allowing for 1981; Gilman & Fox 1997). Some radial motion still is, however, finite thermal conductivity. This field amplitude is about 600 G excited. How much radial mixing the Tayler (1973) instability for the upper part of the solar radiation zone. Considering the

Article published by EDP Sciences and available at http://www.aanda.org or http://dx.doi.org/10.1051/0004-6361:20077172 2 L. L. Kitchatinov and G. Rüdiger : Magnetic instabilities in stellar radiation zones

velocity, u, in spherical geometry in terms of the two scalar po- tentials for the poloidal, Pv, and toroidal, Tv,flows   e e 1 ∂T ∂2P u = r LˆP − θ v + v r2 v r sin θ ∂φ ∂r∂θ   2 eφ ∂T 1 ∂ P + v − v , r ∂θ sin θ ∂r∂φ 1 ∂ ∂ 1 ∂2 Lˆ = sin θ + (3) sin θ ∂θ ∂θ sin2 θ ∂φ2 (cf. Chandrasekhar 1961). Here and in the following, distur- bances are marked by dashes. For zero , the whole class of disturbances corresponding to the poloidal flow vanishes. Then the remaining toroidal flow can only produce toroidal magnetic disturbances so that, in the expression for the  Fig. 1. The buoyancy frequency (2) in the upper part of the radiative magnetic field perturbations (B ), core of the Sun after the solar structure model by Stix & Skaley (1990).   The zone of the model includes the overshoot layer so that e e 1 ∂T ∂2P B = r LˆP − θ m + m N/Ω is large immediately beneath the . r2 m r sin θ ∂φ ∂r∂θ   2 eφ ∂T 1 ∂ P + m − m , (4) r ∂θ sin θ ∂r∂φ transport of chemical species by the Tayler instability, we find = that the field strength can only be slightly above this marginal the poloidal field potential becomes zero, Pm 0. value. Otherwise, the intensity of radial mixing would not be It can be seen from Eq. (3) that the horizontal part of the compatible with the observed lithium abundance. poloidal flow can remain unchanged when the radial velocity is reduced if the radial scale of the flow is reduced proportionally. The disturbances can thus avoid the stabilizing effect of the strat- 2. The model ification by decreasing their radial scale instead of losing their poloidal parts. 2.1. Background state and basic assumptions Our stability analysis is local in the radial dimension; i.e. we use Fourier modes exp(ikr) with kr  1. It will be confirmed The stability of the rotating radiation zone of a star containing a posteriori that the most unstable modes do indeed prefer short magnetic field is considered. The field is assumed axisymmetric radial scales. The analysis remains, however, global in horizon- and purely toroidal, so it can be written as tal dimensions. Our formulation is very close to the thin layer √ approximation by Cally (2003) and Miesch & Gilman (2004), B = eφr sin θ µ0ρ ΩA (r,θ) (1) though we do not impose sharp boundaries and use other de- pendent variables. We also include finite diffusion and focus on in terms of the Alfvén angular frequency ΩA. In this equation, ρ marginal stability instead of growth rates for highly supercritical is the density, r,θ,φ are the usual spherical coordinates, and eφ fields. is the azimuthal unit vector. Equation (1) automatically ensures The instabilities of toroidal fields or differential rotation that the toroidal field component vanishes – as it must – at the ro- proceed via non compressive disturbances. The characteristic tation axis. Centrifugal and magnetic forces are assumed weak growth rates of the instabilities are low compared to p-modes compared to , g/r  Ω2 and g/r  Ω2 . Deviation of frequencies. In the short-wave approximation, the velocity field A  the fluid stratification from spherical symmetry can thus be ne- can be assumed divergence-free, div u = 0. Note that even a glected. slow motion in a stratified fluid can be divergent if its spatial The stabilizing effect of a subadiabatic stratification of the scale in the radial direction is not small compared to scale height. radiative core is characterized by the buoyancy frequency, In the short-wave approximation (in radius), the divergency can, however, be neglected. g ∂s N2 = , (2) Cp ∂r 2.2. Equations We start from the linearized equations for small velocity pertur- where s = C ln (P/ργ) is the specific entropy of . The v bations, i.e. ratio N/Ω is high in the radiative cores of not too rapidly rotating stars like the Sun (see Fig. 1).       ∂u + ·∇  +  ·∇ + 1 ∇ ·  The larger N, the more the radial displacements are opposed (u ) u u u B B ∂t µ0ρ by the buoyancy force. Radial velocities should therefore be low.      1 They are often neglected in stability analysis, which might be − (B ·∇) B − B ·∇ B = − ∇P + ν∆u (5) dangerous as certain instabilities may even be suppressed by the ρ neglect (we shall see later how this indeed happens for the Tayler magnetic field, instability). The consequences of neglecting the radial velocity perturba- ∂B    = ∇× u × B + u × B − η∇×B , (6) tions, ur, can be seen from the expression for (divergence-free) ∂t L. L. Kitchatinov and G. Rüdiger : Magnetic instabilities in stellar radiation zones 3 and entropy, The equation for the poloidal flow then reads        2 ν ∂s   Cpχ  ωˆ LˆV = −λˆ LˆS − i LˆV + u ·∇s + u ·∇s = ∆T . (7) ˆ 2 ∂t T λ       ∂ µΩˆ ∂W ∂Ωˆ The background flow is rotation with angular velocity Ω(r,θ) −2µΩˆ LˆW − 2 1 − µ2 − 2m2 W and the mean magnetic field is the toroidal one of the form (1). ∂µ ∂µ ∂µ Perturbations of velocity and magnetic field are expressed in     ∂ µΩˆ A Ωˆ 2 ∂B 2 ∂ A terms of scalar potentials according to Eqs. (3) and (4). The +2µΩˆ A LˆB + 2 1 − µ + 2m B identities   ∂µ ∂µ ∂µ       ∂ µΩˆ A   Ωˆ r r ·∇×B = LˆT , r r · B = LˆP , − Ωˆ Lˆ − − − 2 ∂ A ∂A m m   m A A 2m A 2m 1 µ     2  ∂µ ∂µ ∂µ  ˆ 3  ˆ 2 ∂ ˆ r r ·∇×u = LTv, r r ·∇×∇×u = − L + r LPv   Ωˆ   ∂r2 ∂ µ ∂Ωˆ ∂V +mΩˆ LˆV + 2m V + 2m 1 − µ2 , (9) ∂µ ∂µ ∂µ are used to reformulate the equations in terms of the poten- tials (using the potentials helps to reduce the equation system). where µ = cos θ,and The radial component of Eq. (6) then gives the equation for N the poloidal magnetic field and the radial components of curled ˆ = λ Ω (10) Eqs. (5) and (6) give the equations for toroidal flow and toroidal 0kr magnetic field, respectively1. can be understood as special normalization for radial wave- The perturbations are considered as Fourier modes in time, length. The first term on the righthand side describes the stabiliz- azimuth, and radius in the form of exp (i(−ωt + mφ + kr)).Foran ing effect of the stratification. It vanishes for small λˆ. Apart from instability, the eigenvalue ω should possess a positive imaginary this stabilizing buoyancy term, the wavelength is only present in part. Only the highest order terms in kr for the same variable diffusive terms. The second term of the RHS includes the action were kept in the same equation. of finite viscosity, When deriving the poloidal flow equation, the term νN2 was transformed as = . ν 3 (11)     Ω r2 1 1 0 r ·∇×∇× ∇P = −r ·∇× (∇ρ) × (∇P) = ρ ρ2 Similarly, we use 2 2   ηN χN r 1 r   g = , = (12) ·∇× ∇ × ∇ = − ·∇× ×∇  = Lˆ  η Ω3 2 χ Ω3 2 ( s) ( P) g s s . 0r 0r Cp ρ Cp rCp for the diffusive parameters η and χ below (note that the buoy- This transformation neglects the pressure disturbances in the ancy frequency N drops off the diffusive terms of the result- baroclinic forcing. More precisely, local thermal disturbances ing equations). The second and the following lines of Eq. (9)   are assumed to occur at constant pressure so that ρ /ρ = −T /T describe the influences of the basic rotation and the toroidal  = −  or s Cpρ /ρ. This assumption is again justified by the in- field. Only latitudinal derivatives of Ω and ΩA appear. All ra- compressible nature of the perturbations. Another interpretation dial derivatives are absorbed by disturbances that vary on much of this assumption is given by Acheson (1978), who assumed shorter radial scales than Ω or ΩA. The complete system of five zero disturbances of total (including magnetic) pressure to in- equations also includes the equation for toroidal flow volve magnetic buoyancy instability. In our derivations, assump-         ν tions of constant total or only gas pressure are identical because ωˆ LˆW = −i LˆW + mΩˆ LˆW − mΩˆ A LˆB ff kr −1 λˆ 2 the e ect of magnetic buoyancy appears in higher order in ( )       compared to the terms kept in the equations to follow. Note that ∂2 ∂2 −mW 1 − µ2 Ωˆ + mB 1 − µ2 Ωˆ the pressure disturbances are neglected in the barocline term ∂µ2 ∂µ2 A alone and the geostrophic balance is still possible.           ∂ 2 ∂ 2 In order to use normalized variables, the time is measured in + LˆV 1 − µ Ωˆ − LˆA 1 − µ Ωˆ A − ∂µ ∂µ units of Ω 1 (Ω is a characteristic angular velocity), the veloci-     0 0     ties are scaled in units of rΩ , the normalized eigenvalue (ω ˆ )is ∂ 2 ∂Ωˆ ∂V 0 + 1 − µ2 − 2 1 − µ2 Ωˆ measured in units of Ω0, and the other normalized variables are ∂µ ∂µ ∂µ       Ωˆ   k 1 k ∂ 2 2 ∂ A 2 ∂A = √ = √ = − 1 − µ − 2 1 − µ Ωˆ A , (13) A 2 Pm, B 2 Tm, V 2 Pv, Ω0r µ0ρ Ω0r µ0ρ Ω0r ∂µ ∂µ ∂µ 1 ikrg Ω Ω W = T , S = s, Ω=ˆ , Ωˆ = A . (8) the equation for toroidal magnetic field Ω r2 v C rN2 Ω A Ω         0 p 0 0 η ωˆ LˆB = −i LˆB + mLˆ Ωˆ B − mLˆ Ωˆ AW Introducing the factors kr in the normalizations of poloidal po- λˆ 2    tentials makes them equal by order of to the toroidal ∂Ωˆ ∂ 2 ∂Ωˆ ∂A −m2 A − 1 − µ2 potentials, while in Eqs. (3) and (4) it was Pv rTv, Pm rTm. ∂µ ∂µ ∂µ ∂µ    1 The radial component of Eq. (5) curled twice provides the equation ∂Ωˆ A ∂ 2 ∂Ωˆ A ∂V +m2 V + 1 − µ2 , (14) for the poloidal flow. ∂µ ∂µ ∂µ ∂µ 4 L. L. Kitchatinov and G. Rüdiger : Magnetic instabilities in stellar radiation zones the poloidal field equation present in all instabilities allowed by our equations, the symme-         try notations are related to the potential W of that flow. Standard η ωˆ LˆA = −i LˆA + mΩˆ LˆA − mΩˆ A LˆV (15) notations, Sm and Am, are used for the modes with symmet- λˆ 2 ric and antisymmetric W about the equator, respectively, and and the entropy equation, m is the azimuthal wave number. The symmetry of magnetic disturbances with the same symmetry notation differ, however,

= − χ + Ωˆ + Lˆ between the cases of one and two belts of background toroidal ωˆ S i S m S V. (16) Ω = λˆ 2 field. For the case of one belt, A const, it is

In the simplest case of Ω=ΩA = 0 and vanishing diffusivities, Sm modes : W, B symmetric, V, A, S antisymmetric, the only nontrivial solution of the above equations are short (in radius) gravity waves with Am modes : W, B antisymmetric, V, A, S symmetric. N  For two belts of toroidal fields of Eq. (19) the symmetry nota- ω = l(l + 1), l = 1, 2, ... (17) kr tions imply Instabilities are thus only possible because of magnetic field or Sm modes : W, A symmetric, V, B, S antisymmetric, nonuniform rotation. It should be kept in mind that the above equations are only valid for kr  1. We shall see that the maxi- Am modes : W, A antisymmetric, V, B, S symmetric. mum growth rates do indeed belong to the short radial scales. In this latter case, Sm modes have vector field B which is   mirror-symmetric about equatorial plane (symmetric Bφ , Br   2.3. 2D approximation (λˆ  1) and antisymmetric Bθ ) and mirror-antisymmetric flow u (sym- metric u  and antisymmetric u , u ) and the other way round N/Ω λˆ θ φ r The ratio of 0 in radiation zones is so high (Fig. 1) that for Am modes. The symmetry definition for two belts of toroidal kr  (10) can also be high in spite of 1. Equation (9) in the field are the same as in Gilman et al. (2007). leading order in λˆ then gives S = 0. Next, Eqs. (16) and (15) successively yield V = 0andA = 0. Diffusive terms can also be neglected. The equation system reduces to two coupled equa- 3. Results and discussion tions for toroidal magnetic field and toroidal flow (Gilman & Fox 1997), 3.1. Nonexistence of 2D magnetic instabilities       Ω Ω Lˆ = Ωˆ Lˆ − Ωˆ Lˆ The particular case where and A are both constant strongly ωˆ W m W m A B simplifies the 2D Eqs. (18). The equations, then, attain con- 2    2    stant coefficients and can be solved analytically. The solution − ∂ − 2 Ωˆ + ∂ − 2 Ωˆ mW 1 µ mB 1 µ A , in terms of Legendre polynomials W, B ∼ Pm(µ)givesthe ∂µ2 ∂µ2 l       eigenfrequencies Lˆ = Lˆ Ωˆ − Lˆ Ωˆ ωˆ B m B m AW . (18)    ωˆ 1 2 1 Here the wave number k drops out and the equations describe 2D = 1 − ± Ωˆ 2 1 − + (20) + A + 2 + 2 disturbances within decoupled spherical shells. The condition m l(l 1) l(l 1) l (l 1) Ω  for validity of the 2D approximation is, therefore, N/( kr) 1. of magnetically modified r-modes (Longuet-Higgins 1964; Papaloizou & Pringle 1978) describing stable horizontal patterns 2.4. Equatorial symmetry of fluctuating and background drifting in longitude. We shall see in Sect. 3.3.1 that the case of fields constant Ω and ΩA shows Tayler instability in the 3D case. The 2D approximation, however, misses the instability. In the present paper the interaction of toroidal magnetic fields Though the result is obtained for the particular case of con- and the basic rotation is formulated only for rigid rotation. This stant ΩA, it is most probably valid in general. Indeed, the is to single out the instabilities that are magnetic by origin (uni- toroidal field can be understood as consisting of closed flux form rotation has a minimum energy for given angular momen- tubes. Noncompressive disturbances conserve the volume and tum so that rotational energy cannot feed any instability). The magnetic flux of the tubes so that the magnetic energy of a tube results for the joint stability of differential rotation and magnetic is proportional to the square of its length (Zeldovich et al. 1983). field will be the subject of a separate publication. The 2D disturbances also conserve the area of a segment on a For the toroidal field, two simple geometries are considered. spherical surface encircled by a magnetic field line. The circu- Ω First, the quantity A is taken as a constant so that the toroidal lar lines of background field have minimum length for any given field has only one belt symmetric with respect to the equator. encircled area. Any 2D disturbance increases the length, thereby Second, two magnetic belts are considered with equatorial anti- increasing magnetic energy. Therefore, there is no possibility of symmetry, i.e. with a node of Bφ at the equator, feeding a 2D instability by magnetic energy release. A more for- Ω = Ω mal proof of this statement can use the expression for magnetic A b cos θ, (19) disturbances =  where b is the normalized field strength, b 1 means equiparti- B = ∇ × (ξ × B) tion of rotational and magnetic energies. Two types of equatorial symmetry are allowed by the equa- in terms of the displacement vector ξ. It is easy to see that the tions for the linear disturbances in Sect. 2.2. Different dependent contribution of B to the magnetic energy density integrated over variables of the same symmetry type can, however, have differ- a spherical surface of constant r is positively definite for 2D dis- ent equatorial symmetries. As the toroidal flow disturbances are placements ξ = ∇ × (rψ) (ψ is an arbitrary scalar function). L. L. Kitchatinov and G. Rüdiger : Magnetic instabilities in stellar radiation zones 5

The non-existence of Tayler instability in 2D clearly shows that radial displacements are necessary for the instability. Therefore, radial mixing is unavoidable. Transport of chemical species by the Tayler instability will be estimated in Sect. 3.3.2. The conclusion about the absence of 2D instabilities is of course restricted to the case of rigid rotation. Rotational shear can excite its own 2D instability (Watson 1981; Dziembowski & Kosovichev 1987; Garaud 2001) fed by rotational energy re- lease, and magnetic field can catalyze the process (Gilman & Fox 1997). Gilman et al. (2007) find that 3D instabilities in the solar tachocline with strong toroidal fields differ only slightly from their 2D counterparts. This is probably because of large differ- ential rotation in the tachocline. We shell see that the results for rigid rotation differ strongly between the 2D and 3D cases. The amount of differential rotation required for the difference to fade is still unknown.

3.2. Ideal fluids, χ = η = ν = 0 Fig. 2. The stability map for constant Ω and zero diffusivities. There is We proceed by discussing numerical solutions for Eqs. (9), A no instability for weak fields with ΩA < Ω. The threshold field strength (13)–(16) for 3D disturbances. Consider first the case of ideal for the instability increases with increasing vertical wavelength λˆ. fluids with vanishing diffusion. Constant ΩA give the simplest realization of the Tayler instability. The instability criterion for nonaxisymmetric disturbances (Goossens et al. 1981) reads

∂   2µ2 − m2 ln B2 <  , for µ ≥ 0. (21) ∂µ φ µ 1 − µ2

It is satisfied with constant ΩA for m = 1 and close to the pole. The axisymmetric instability with m = 0 does not exist in this case. From Sect. 3.1 we know that the instability can appear only in 3D formulation when radial displacements are allowed. Figure 2 gives the resulting stability map. Two features are re- markable: (i) rotation suppresses the instability so that subeqi- upartition fields, ΩA < Ω, are stable, and (ii) the unstable dis- turbances for the smallest field producing the instability have indefinitely short radial scales, λˆ → 0. Buoyancy frequency, N, enters the equations for linear disturbances of Sect. 2.2 in com- bination (10) with radial wave number, k. Therefore, the distur- bances can avoid the stabilizing effect of stratification by reduc- ing their radial scale. This is the strategy the Tayler instability Fig. 3. The small λˆ part of the stability map for two belts of toroidal follows. Another possibility would be zero entropy disturbances magnetic field (19). Rotation does not suppress instability in this case. Ω Ω (the λˆ-parameter (10) enters the equations as a factor of entropy Parameter b is the polar value of the ratio A/ . perturbation). Entropy perturbations can be avoided with 2D dis- turbances only. Tayler instability cannot use this opportunity be- cause it has no 2D counterpart. Eq. (19) confirm this expectation. The stability map of Fig. 3 Rotational quenching of Tayler instability is a controversial shows that the disturbances with sufficiently small radial scales issue. Spruit (1999) finds that rotation modifies the instability but remain unstable when the (normalized) toroidal field b is small2. does not switch it off. Cally (2003), however, concluded that the The high sensitivity of the Tayler instability in ideal fluids polar kink instability (m = 1 mode of Tayler instability) is totally to details of the Ω (µ) profile is mainly of academic interest. suppressed if Ω is smaller than Ω near the pole. Simulations of A A Figures 2 and 3 show that the smallest b producing the instability Braithwait (2006) also show that the instability does not develop corresponds to indefinitely small radial scales. Finite diffusion, when the toroidal field is below a critical value roughly equal to therefore, must be included. We shall see that the strong differ- the equipartition level of Ω Ω. A ence between the cases of one and two belts of toroidal field The controversy can be resolved by observing that the ro- fades when diffusion is allowed for. tational quenching is sensitive to details of the toroidal field profile. Already, Pitts & Tayler (1985) suggested that suppres- 2 There is no contradiction with the result by Cally (2003) that the sion of the instability by rotation may be exceptionally strong polar kink instability does not develop whenever the polar value of the when the ratio of Alfvén velocity to rotation velocity is uni- ratio ΩA/Ω is below one. Two lines of Fig. 3 do not converge at small form, ΩA/Ω=const., but rotation is unlikely to lead to com- b, clearly indicating that the kink instability is not of the polar type (cf. plete stability for the general configuration of the toroidal field. Fig. 6) so that the boundary layer analysis by Cally (2003) does not Computations for the case of two belts of the toroidal field of apply. 6 L. L. Kitchatinov and G. Rüdiger : Magnetic instabilities in stellar radiation zones

Fig. 4. The same as in Fig. 2 but with the finite diffusivities (22). The critical field strengths for the onset of the instability are strongly re- duced compared to ideal MHD.

Fig. 6. Toroidal field lines of the most rapidly growing eigenmodes of Fig. 5 for weak background toroidal field, ΩA = 0.1Ω (upper panel), and strong field, ΩA = 10Ω (lower panel).

3.3.1. Fields with equatorial symmetry The reduction of the critical field for onset of the instability is exceptionally strong for the case of uniform ΩA. The stability map for diffusion parameters of Eq. (22) is shown in Fig. 4. The minimum field producing the instability is reduced by more than two orders of magnitude compared to ideal fluids (Fig. 2). The characteristic minima in Fig. 4 correspond to small, λˆ ∼< 1, but finite vertical scales. For strong fields (ΩA > Ω) the basic rotation and diffusion are not important and there only characteristic frequency to scale the growth rates is ΩA. The dependence of Fig. 5 does indeed approach the σ ∼ Ω relation (σ is the growth rate) in the strong- Fig. 5. The normalized growth rateσ ˆ as function of the magnetic field A amplitude for the S1 mode and for λˆ = 0.6 (where the line of Fig. 4 has field limit. The growth rate drops by almost four orders of mag- Ωˆ a minimum). The dotted lines for strong, ΩA > Ω, and weak, ΩA < Ω, nitude when A is reduced below 1. In the weak-field regime it ˆ ∼ Ω2 Ω fields give approximations by the linear,σ ˆ ∼ ΩA, and parabolic,σ ˆ ∼ is σ A/ (Spruit 1999). The growth rates for weak fields Ωˆ 2 (ΩA < Ω), where the instability exists due to finite diffusion, A, laws, respectively. are rather small. Nevertheless, the growth rate of, e.g.,σ ˆ = 10−5 means the e-folding time of about 1000 yr for the Sun, which is very short compared to evolutionary time scales. We shall see 3.3. Finite that the instability of weak fields is actually too vigorous to be compatible with observed lithium abundance. From now on the values The structures of the unstable modes differ between strong and weak field regimes. For the weak fields, the growing dis- −10 −8 −4 ν = 2 × 10 , η = 4 × 10 , χ = 10 , (22) turbance pattern is distributed over the entire sphere. For strong fields, it is much more concentrated at the poles (Fig. 6) but still which are characteristic of the upper part of the solar radiative remains global in latitude. core, are used for the diffusion parameters (11) and (12). The relation χ  η  ν of Eq. (22) is typical of stellar radiation 3.3.2. Fields with equatorial antisymmetry zones. For ideal fluids the Tayler instability operates with extremely Now a toroidal field with two belts of opposite polarity (19) small radial scales. When there is no stabilizing stratification, is considered with the diffusivity set (22). Such a field ge- however, finite vertical scales are preferred (Arlt et al. 2007). ometry can result from the action of differential rotation on The thermal conductivity decreases the stabilizing effect of the dipolar poloidal fields. Figure 7 shows the corresponding sta- stratification and reduces the critical field strengths for the insta- bility map. It is now similar to the map of Fig. 4 for one bility if the (normalized) radial scale λˆ is not too small. field belt. Strong dependence on the ΩA(µ)-profile is no longer L. L. Kitchatinov and G. Rüdiger : Magnetic instabilities in stellar radiation zones 7

Fig. 9. The growth rate as a function of the toroidal field amplitude for λˆ = 0.1. The dotted line shows the parabolic approximationσ ˆ = 0.1b2. Fig. 7. Stability map for the field model (19) with two belts and equato- The scale on the right gives the radial diffusivity of chemical species rial antisymmetry. estimated after (25).

by the parabolic lawσ ˆ = 0.1b2, Eq. (25) can be rewritten in terms of Bφ (Eq. (23)) as   2 Bφ D 7 × 104 cm2 s−1. (26) T 1kG As is known, diffusivities in excess of 104 cm2 s−1 in the upper radiative core are not compatible with the observed solar lithium abundance. Hence, the toroidal field amplitude can only slightly exceed the marginal value of about 600 G. In our (simplified) Fig. 8. Growth rates in units of Ω for b = 0.01 (left)andb = 0.1(right). formulation, the observed solar lithium abundance seems to The highest rates exist for λˆ ∼< 0.1 independent of the magnetic field exclude3 any concept of hydromagnetic dynamos driven by amplitude, all for most rapidly growing S1 modes. Tayler instability in the upper radiation zone of the Sun. We should not forget, however, that the superrotation (∂Ω/∂r > 0) at the bottom of the convection zone in the equatorial region acts to observed. Axisymmetric instability does not exist for pro- ff file (19). Instability is again only found for the kink mode of stabilize. A mild stabilizing e ect can also be expected from the m = 1. compositional gradient due to gravitational settling of chemical In physical values, one finds for the upper part of solar radi- species (Richard et al. 1996; Vauclair 1998). The critical field ation zone (with ρ 0.2 g/cm3) amplitudes for Tayler instability may, thus, be somewhat higher than the computed 600 G. Even this value is, however, above 5 Bφ 10 b G (23) the toroidal field strengths obtained in magnetic models for the laminar solar tachocline (cf. Rüdiger & Kitchatinov 2007). The and axisymmetric tachocline models are, therefore, expected to be λ 10λˆ Mm, (24) stable to nonaxisymmetric disturbances. so that from Fig. 7 follows a critical magnetic field for the insta- bility slightly below 600 G. The mode that first becomes unstable 4. Summary if the field exceeds this critical value has a vertical wavelength The linear stability of toroidal magnetic field in rotating stel- between 1 and 2 Mm. For larger field strengths, of course, there lar radiation zones is analyzed assuming that the vertical scale is a range of unstable wavelengths. The maximum growth rates of the fluctuations is short compared to the local radius (“short- ∼< appears, however, at wavelengths 1 Mm (Fig. 8). wave approximation”). The analysis is global in horizontal di- The flow field of the instability also mixes chemical species mensions. Stability computations confirm that the most rapidly in radial direction. Such an instability can thus be relevant to the growing perturbations have short radial scales: kr ∼ 103. radial transport of the light elements (Barnes et al. 1999). The ff ff   We have shown that the pinch-type instability of toroidal e ective di usivity, DT u  (u and  are rms velocity and fields require nonvanishing radial displacements. The instability correlation length in radial direction) can be roughly estimated does not exist in a 2D approximation with zero radial velocities. from our linear computations assuming that the instability sat- The maximum amplitude of stable toroidal magnetic fields for urates when mixing frequency u/ approaches the growth rate  the Sun we found is about 600 G. This value results only for σ u / and  λ/2. With Eq. (24), this yields rigid rotation. It will most probably increase if the stabilizing 9 2 −1 Ω DT 7 × 10 σˆ cm s , (25) influence of the positive radial gradient of in the equatorial region of the tachocline is included in the model. whereσ ˆ is the normalized growth rate given in Fig. 9. For the range of 0.01 < b < 0.2 where the plot is closely approximated 3 Or rise new problems. 8 L. L. Kitchatinov and G. Rüdiger : Magnetic instabilities in stellar radiation zones

The field strength in the upper part of the solar radiative Braithwaite, J., & Spruit, H. C. 2004, Nature, 431, 819 interior can only marginally exceed the resulting critical val- Burgess, C., Dzhalilov, N. S., Maltoni, M., et al. 2003, ApJ, 588, 65 ues. Otherwise the instability would produce too strong a ra- Cally, P.S. 2003, MNRAS, 339, 957 Chandrasekhar, S. 1961, Hydrodynamic and hydromagnetic stability (Oxford: dial mixing of light elements. After our results, all the axisym- Clarendon Press), 622 metric hydromagnetic models of the solar tachocline (Rüdiger Cowling, T. G. 1945, MNRAS, 105, 166 & Kitchatinov 1997; MacGregor & Charbonneau 1999; Garaud Dziembowski, W., & Kosovichev, A. G. 1987, Acta Astron., 37, 341 2002; Sule et al. 2005; Kitchatinov & Rüdiger 2006) have sta- Garaud, P. 2001, MNRAS, 324, 68 ∼ 5 Garaud, P. 2002, MNRAS, 329, 1 ble toroidal fields. On the contrary, the strong fields 10 G, Gilman, P. A. 2005, Astron. Nachr., 326, 208 which are able to modify g-modes, or even stronger fields that Gilman, P. A., & Fox, P.A. 1997, ApJ, 484, 439 may be relevant to neutrino oscillations (Burgess et al. 2003) are Gilman, P. A., Dikpati, M., & Miesch, M. S. 2007, ApJS, 170, 203 strongly unstable with e-folding times shorter than one rotation Goossens, M., Biront, D., & Tayler, R. J. 1981, Ap&SS, 75, 521 period. Kitchatinov, L. L., & Rüdiger, G. 2006, A&A, 453, 329 Longuet-Higgins, M. S. 1964, Proc. Roy. Soc. Lond. A, 279, 446 The 3D computations of joint instabilities of toroidal fields MacGregor, K. B., & Charbonneau, P. 1999, ApJ, 519, 911 and differential rotation (Gilman & Fox 1997; Cally 2003; Miesch, M. S., & Gilman, P. A. 2004, Sol. Phys., 220, 287 Rüdiger et al. 2007) will be discussed in a separate paper. Papaloizou, J., & Pringle, J. E. 1978, MNRAS, 182, 423 Another tempting extension is to include of the poloidal field. Pitts, E., & Tayler, R. J. 1985, MNRAS, 216, 139 Rashba, T. I., Semikoz, V.B., & Valle, J. W. F. 2006, MNRAS, 370, 845 The field can be important in view of the very short vertical Richard, O., Vauclair, S., Charbonnel, C., & Dziembowski, W. A. 1996, A&A, scales of the unstable modes. 312, 1000 Rüdiger, G., & Kitchatinov, L. L. 1997, Astron. Nachr., 318, 273 Acknowledgements. This work was supported by the Deutsche Forschungs- Rüdiger, G., & Kitchatinov, L. L. 2007, New J. Phys., 9, 302 gemeinschaft and by the Russian Foundation for Basic Research (project Rüdiger, G., Hollerbach, R., Schultz, M., & Elstner, D. 2007, MNRAS, 377, 1481 05-02-04015). Spruit, H. C. 1999, A&A, 349, 189 Stix, M., & Skaley, D. 1990, A&A, 232, 234 Sule, A., Rüdiger, G., & Arlt, R. 2005, A&A, 437, 1061 References Tayler, R. J. 1973, MNRAS, 161, 365 Arlt, R., Sule, A., & Rüdiger, G. 2007, A&A, 461, 295 Vauclair, S. 1998, SSRv, 85, 71 Acheson, D. J. 1978, Phil. Trans. Roy. Soc. Lond. A, 289, 459 Watson, M. 1981, Geophys. Astrophys. Fluid Dyn., 16, 285 Barnes, G., Charbonneau, P., & MacGregor, K. B. 1999, ApJ, 511, 466 Zeldovich, Ya. B., Ruzmaikin, A. A., & Sokolov, D. D. 1983, Magnetic fields in Braithwaite, J. 2006, A&A, 453, 687 astrophysics (New York: Gordon and Breach Science Publishers)