Transcript
Page 1: Magnetic helicity in stellar dynamos: new numerical experiments

Astron. Nachr./AN 323 (2002) 2, 99–122

Magnetic helicity in stellar dynamos: new numerical experiments

A. BRANDENBURG � , W. DOBLER � � , and K. SUBRAMANIAN � � �

� NORDITA, Blegdamsvej 17, DK-2100 Copenhagen Ø, Denmark� Department of Mathematics, University of Newcastle upon Tyne, NE1 7RU, UK� National Centre for Radio Astrophysics - TIFR, Pune University Campus, Ganeshkhind, Pune 411 007, India

Received 2001 November 30; accepted 2002 February 5

Abstract. The theory of large scale dynamos is reviewed with particular emphasis on the magnetic helicity constraint inthe presence of closed and open boundaries. In the presence of closed or periodic boundaries, helical dynamos respond tothe helicity constraint by developing small scale separation in the kinematic regime, and by showing long time scales inthe nonlinear regime where the scale separation has grown to the maximum possible value. A resistively limited evolutiontowards saturation is also found at intermediate scales before the largest scale of the system is reached. Larger aspect ratioscan give rise to different structures of the mean field which are obtained at early times, but the final saturation field strengthis still decreasing with decreasing resistivity. In the presence of shear, cyclic magnetic fields are found whose period isincreasing with decreasing resistivity, but the saturation energy of the mean field is in strong super-equipartition with theturbulent energy. It is shown that artificially induced losses of small scale field of opposite sign of magnetic helicity as thelarge scale field can, at least in principle, accelerate the production of large scale (poloidal) field. Based on mean field modelswith an outer potential field boundary condition in spherical geometry, we verify that the sign of the magnetic helicity fluxfrom the large scale field agrees with the sign of � . For solar parameters, typical magnetic helicity fluxes lie around � �

� �

� �

per cycle.

Key words: MHD – turbulence

1. Introduction

The conversion of kinetic into magnetic energy, i.e. the dy-namo effect, plays an important role in many astrophysicalbodies (stars, planets, accretion discs, for example). The gen-eration of magnetic fields on scales similar to the scale ofthe turbulence is a rather generic phenomenon that occurs forsufficiently large magnetic Reynolds numbers unless certainantidynamo theorems apply, which exclude for example two-dimensional fields (e.g., Cowling 1934).

There are two well-known mechanisms, which allow thegeneration of magnetic fields on scales larger than the eddyscale of the turbulence: the alpha-effect (Steenbeck, Krause& Radler 1966) and the inverse cascade of magnetic helic-

Correspondence to: Axel Brandenburg, [email protected]� current address: Kiepenheuer Institute for Solar Physics,

Schoneckstr. 6, 79104 Freiburg, Germany� � current address: Inter University Centre for Astronomy and As-

trophysics, Post Bag 4, Pune University Campus, Ganeshkhind,Pune 411 007, India

ity in hydromagnetic turbulence (Frisch et al. 1975, Pouquet,Frisch, & Leorat 1976). In a way the two may be viewed asthe same mechanism in that both are driven by helicity. The� -effect is however clearly nonlocal in wavenumber space,whereas the inverse cascade of magnetic helicity is usuallyunderstood as a (spectrally) local transport of magnetic en-ergy to larger length scales. Both aspects are seen in simula-tions of isotropic, non-mirror symmetric turbulence: the non-local inverse cascade or � -effect in the early kinematic stageand the local inverse cascade at later times, when the fieldhas reached saturation at small scales (Brandenburg 2001a,hereafter referred to as B01).

It is not clear whether either of these two mechanisms isactually involved in the generation of large scale magneticfields in astrophysical bodies. Although much of the largescale magnetic field in stars and discs is the result of shearand differential rotation, a mechanism to sustain a poloidal(cross-stream) magnetic field of sufficiently large scale isstill needed. The main problem that we shall be concernedwith in this paper is that of the associated magnetic helic-

c� WILEY-VCH Verlag Berlin GmbH, 13086 Berlin, Germany 2002 0044-6337/02/0207-0099 $ 17.50+.50/0

Page 2: Magnetic helicity in stellar dynamos: new numerical experiments

100 Astron. Nachr./AN 323 (2002) 2

ity production, which may prevent the large scale dynamoprocess from operating on a dynamical time scale. However,not many alternative and working dynamo mechanisms havebeen suggested so far. In the solar context, the Babcock–Leighton mechanism is often discussed as a mechanism thatoperates preferentially in the strong field regime (e.g. Dik-pati & Charbonneau 1999). However, Stix (1974) showedthat the model of Leighton (1969) is formally equivalentto the model of Steenbeck & Krause (1969), except that inLeighton (1969) the � -effect was nonlinear and mildly sin-gular at the poles. Nevertheless, a magnetically driven � ef-fect has been suspected to operate in turbulent flows driven bythe magnetorotational instability (Brandenburg et al. 1995) orby a magnetic buoyancy instability (Ferriz-Mas, Schmitt, &Schussler 1994, Brandenburg & Schmitt 1998). In the frame-work of mean-field dynamo theory, magnetically driven dy-namo effects have also been invoked to explain the observedincrease of stellar cycle frequency with increased stellar ac-tivity (Brandenburg, Saar, & Turpin 1998). However, as longas these mechanisms lead to an � -effect, they also producemagnetic helicity and are hence subject to the same problemas before.

Although kinetic helicity is crucial in the usual expla-nation of the � effect, it is not a necessary requirement formagnetic field generation, and lack of parity invariance is al-ready sufficient (Gilbert, Frisch, & Pouquet 1988). However,as in every � effect dynamo, magnetic helicity is produced atlarge scales, which is necessarily a slow process. A mecha-nism similar to the ordinary � effect is the incoherent � ef-fect (Vishniac & Brandenburg 1997), which works in spiteof constantly changing sign of kinetic helicity (in space andtime) provided there is systematic shear and sufficient turbu-lent diffusion. This mechanism has been invoked to explainthe large scale magnetic field found in simulations of accre-tion disc turbulence (Brandenburg et al. 1995, Hawley, Gam-mie, & Balbus 1996, Stone et al. 1996). In addition, how-ever, the large scale field of Brandenburg et al. (1995) showsspatio-temporal coherence with field migration away fromthe midplane. This has so far only been possible to explainin terms of an � � dynamo with a magnetically driven � ef-fect (Brandenburg 1998). In any case, the incoherent � effect,which has so far only been verified in one-dimensional mod-els, would not lead to the production of net magnetic helicityin each hemisphere.

Yet another mechanism was suggested recently by Vish-niac & Cho (2001), which yields a mean electromotive forcethat does not lead to the production of net magnetic helicity,but only to a transport of preexisting magnetic helicity. Thisis why this mechanism can, at least in principle, work on afast time scale even when the magnetic Reynolds number islarge. Finally, we mention a totally different mechanism thatworks on the basis of negative turbulent diffusion (Zheligov-sky, Podvigina, & Frisch 2001). This is what produces turbu-lence in the Kuramoto–Sivashinsky Equation, which is thenstabilised by hyperdiffusion (i.e. a fourth derivative term).Among all these different dynamo mechanisms, the conven-tional � effect and the inverse magnetic cascade are the onlymechanisms that have been shown numerically to produce

Large scale dynamos(e.g. in the sun)

Is significant net magnetic helicityproduced in each hemisphere?

��

��

NOfast � effect:

(like in kinematiccase,

and large scaleseparation)

NOnon- � effect:(e.g., negative

magn. diffusion,incoherent � ,

Vishniac & Choeffect)

YES

standard � effect:� what happens

to magnetichelicity?

��

�� �

Alternative Afast �

effect:transfer

from smallto large

scales, nonet helicity

Alternative Bnon- �

effectdynamos:

nonumericalevidence

yet

Alternative C(i)

dissipation(ii) recon-

nection(near

surface?)

Alternative Dloss

throughboundaries:(i) equator?

(ii) outersurface

Fig. 1. Sketch illustrating the different ways astrophysical dynamosmay be able to circumvent the magnetic helicity problem.

strong large scale fields under turbulent conditions. These arehowever exactly the mechanisms that suffer from the helicityconstraint.

The purpose of the present paper is to assess the signifi-cance of the helicity constraint and to present some new nu-merical experiments that help understanding how the con-straint operates and to discuss possible ways out of thedilemma. For orientation and later reference we summarise inFig. 1 various models and the possible involvement of mag-netic helicity in them. Dynamos based on the usual � effectare listed to the right. They all produce large scale magneticfields that are helical. However, in a periodic or an infinitedomain, or in the presence of perfectly conducting bound-aries, the magnetic helicity is conserved and can only changethrough microscopic resistivity. This would therefore seem tobe too slow for explaining variations of the mean field on thetime scale of the 11 year solar cycle. Open boundary condi-tions may help, although this was not yet possible to demon-strate, as will be explained in Sect. 4 of this paper.

Given that the issue of magnetic helicity is central to ev-erything that follows, we give in Sect. 2 a brief review ofmagnetic helicity conservation, the connection between theinverse cascade and the realisability condition, and the sig-nificance of gauge-invariant forms of magnetic helicity andthe surface integrated magnetic helicity flux. Readers famil-iar with this may jump directly to Sect. 3, where we discussthe issue of magnetic helicity cancellation in kinematic and

Page 3: Magnetic helicity in stellar dynamos: new numerical experiments

A. Brandenburg et al.: Magnetic helicity in stellar dynamos 101

non-kinematic dynamos, or to Sect. 4, where we begin witha discussion of the results of B01 and Brandenburg & Dobler(2001, hereafter referred to as BD01), or to Sect. 4.3 wherenew results are presented.

2. The magnetic helicity constraint

Magnetic helicity evolution does not give any additional in-formation beyond that already contained in the inductionequation governing the evolution of the magnetic field it-self. Nevertheless, the concept of magnetic helicity provesextremely useful because magnetic helicity is a conservedquantity in ideal MHD, and almost conserved even in slightlynon-ideal conditions. The evolution equation for the mag-netic helicity can then be used to extract information thatcan be easily understood. At a first glance, however, mag-netic helicity appears counterintuitive, because it involves themagnetic vector potential, � , which is not itself a physicalquantity, as it is not invariant under the gauge transformation� � � � � � . Only the magnetic field, � � � � � , whichlacks the irrotational ‘information’ from � , is independentof � and hence physically meaningful. Therefore � � � isnot a physically meaningful quantity either, because a gaugetransformation, � � � � � � changes � � � . For periodicor perfectly conducting boundary conditions, or in infinitedomains, the integral � �

� � � � � is however gauge-invariant, because in�

� � � � � � �

� �

� � � � � �

� � � � � � (1)

the surface integral on the right hand side vanishes, and thelast term vanishes as well, because � � � � � .

The evolution of the magnetic field is governed by theinduction equation

� �

� �

� � � � � � (2)

where � is the electric field. This equation can be integratedto yield

� �

� �

� � � � � � � (3)

where � is the scalar (‘electrostatic’) potential, which is alsoreferred to as the gauge potential, because it can be chosenarbitrarily without affecting � . Common choices are �

� �

� � � � , which preserves the Coulomb gauge � � � � � , and� � � , which is convenient in MHD for numerical purposes.If � � const, another convenient gauge is � � � � � � � ,which results in

� �

� �

� � � � � � �

� ( � � const) (4)

Using Eqs (2) and (3) one can derive an equation for thegauge-dependent magnetic helicity density � � � ,

� �

� � � � � � � � � � � � � � � � � � � � � (5)

Using Ohm’s law, �

� � � � � � � , we have � � � �

� � � , where � is the microscopic magnetic diffusivity,

the magnetic permeability in vacuum, and � � � � � �

the electric current density.

2.1. Magnetic helicity conservation with periodicboundaries

Consider first the case of a periodic domain, so the divergenceterm in Eq. (5) vanishes after integration over the full volume,and thus

� �

� � � � � � � �

� � � � � � (6)

where angular brackets denote volume averages, so�

� �

� � � � � � � � � � , where � is the (constant) volume of thedomain under consideration. Note that the terms in Eq. (6)are gauge-independent [see Eq. (1)]. More important, how-ever, is the fact that the rate of change of magnetic helicity isproportional to the microscopic magnetic diffusivity � . Thisalone is not sufficient to conclude that the magnetic helic-ity will not change in the limit � � � , because the currenthelicity, � � � � � , may still become large. A similar effect isencountered in the case of ohmic dissipation of magnetic en-ergy which proceeds at the rate �

� � � � �

� �

� �

� . In thestatistically steady state (or on time averaging), the rate ofohmic dissipation must be balanced by the work done againstthe Lorentz force. Assuming that this work term is indepen-dent of the value of � (e.g. Galsgaard & Nordlund 1996) weconclude that the root-mean-square current density increaseswith decreasing � like

� � � � as � � � � (7)

whilst the rms magnetic field strength, �

, is essentially in-dependent of � . This, however, implies that the rate of mag-netic helicity dissipation decreases with � like

� �

� � � � � � �

� � � �

� � as � � � (8)

Thus, under many astrophysical conditions where the mag-netic Reynolds number is large ( � small), the magnetic helic-ity � , as governed by Eq. (6), is almost independent of time.This motivates the search for dynamo mechanisms indepen-dent of � . On the other hand, it is conceivable that magnetichelicity of one sign can change on a dynamical time scale,i.e. faster than in Eq. (8), if magnetic helicity of the other signis removed locally by advection, for example. This is essen-tially what the mechanism of Vishniac & Cho (2001) is basedupon. So far, however, numerical attempts to demonstrate theoperation of this mechanism have failed (Arlt & Branden-burg 2001). Yet another possibility is that magnetic helicityof one sign is generated at large scales, and is compensatedby magnetic helicity of the other sign at small scales, as inthe kinematic � effect. However, as we shall see in Sect. 4,this does not happen efficiently when nonlinear effects of theLorentz force come into play.

2.2. Realisability condition and connection with inversecascade

Magnetic helicity is important for large scale field genera-tion because it is a conserved quantity which cannot easilycascade forward to smaller scale. We shall demonstrate thishere for the case where the magnetic field is fully helical.The existence of an upper bound for the magnetic helicityis easily seen by decomposing the Fourier transformed mag-netic vector potential, �

, into a longitudinal component,

Page 4: Magnetic helicity in stellar dynamos: new numerical experiments

102 Astron. Nachr./AN 323 (2002) 2

� , and eigenfunctions �

� of the curl operator (also calledChandrasekhar–Kendall functions),

� �

� �

� �

� (9)

with

� � �

� � �

� � � � � � � (10)

and

� �

� �

� � � �

� �

� � � �

� �

� � � (11)

where asterisks denote the complex conjugate. The longitu-dinal part �

is parallel to � and vanishes after taking thecurl to calculate the magnetic field. In the Coulomb gauge,� � � � � , the longitudinal component vanishes altogether(apart from an uninteresting constant vector �

). Since �

and �

� are all orthogonal to each other, it is clear that, for agiven magnetic field, the Coulomb gauge minimises � �

� —or, if �

� � � � �� � , the variance � � � � � � � �

� .The (complex) coefficients �

� � � depend on � and � ,

while the eigenfunctions �

, which form an orthonormal set,

depend only on � and are given by1

� � � � � � � � � � � � � � �

� � � � � �

� �

� (12)

where � is an arbitrary unit vector not parallel to � . Withthis definition we can write the magnetic helicity and energyspectra in the form

� � � � �

� � �

� � � (13)

� � �

� � �

� � � (14)

where � is the volume of integration. (Here and elsewherethe factor

� �

is ignored in the definition of the magneticenergy.) These spectra are normalised such that�

� � � � � � � � � � � � (15)

� � � �

� � � � � (16)

where � and � are magnetic helicity and magnetic energy,respectively. From Eqs (13) and (14) one sees immediatelythat

� � �

� � �

� (17)

which is also known as the realisability condition. A fullyhelical field has therefore �

� �

.For the following it is convenient to write the magnetic

helicity spectrum in the form

� � �

� �

� � � Æ � � (18)

where the subscript � (which is here a scalar!) indicatesFourier filtering to only those wave vectors � that lie in theshell

� � Æ � � � � � � � � � Æ � � � � -shell� (19)

Note that �

and �

are in real (configuration) space, so�

� � � �

. (By contrast, �

and �

� � � � �

1 The forcing functions used in B01 were proportional to �

. Wenote that, by mistake, the square root was missing in the normalisa-tion factor � �

� in that paper.

are quantities in Fourier space, as indicated by the bold facesubscript � , which is here a vector.) In Eq. (18) the averageis taken over all points in � space. Note that in practice thehelicity is more easily calculated directly as

� -shell� �

� �

� �

� �

� �

� �

� (20)

where � �

is the solid angle element in Fourier space. Theresult is however the same. Likewise, the magnetic energyspectrum can be written as

� �

� � � Æ � � (21)

or as

� -shell�

� �

� �

(22)

We recall that for a periodic domain � is gauge invariant.Since its spectrum can be written as an integral over all space,see Eq. (18), �

is – like � – also gauge invariant.The occurrence of an inverse cascade can be understood

as the result of two waves (wavenumbers � and � ) interact-ing with each other to produce a wave of wavenumber � . Thefollowing argument is due to Frisch et al. (1975). Assum-ing that during this process magnetic energy is conserved to-gether with magnetic helicity, we have

� �

� �

� (23)

� �

� � � �

� � � �

� � (24)

where we are assuming that only helicity of one sign is in-volved. Suppose the initial field is fully helical and has thesame sign of magnetic helicity at all scales, then we have

� � � �

� and �

� � � �

� � (25)

and so Eq. (23) yields

� � �

� � � � �

� � �

� � � �

� � (26)

where the last inequality is just the realisability condition (17)applied to the target wavenumber � after the interaction. Us-ing Eq. (24) we have

� � �

� � � � �

� � � � � �

� � � �

� � (27)

In other words, the target wave vector � after the interactionof wavenumbers � and � satisfies

� �

� � �

� � � � �

� �

� � � �

(28)

The expression on the right hand side of Eq. (28) is aweighted mean of � and � and thus satisfies

� � � � � � � �

� � �

� � � � �

� �

� � � �

� � � � � � � � � � (29)

and therefore

� � � � � � � � � � (30)

In the special case where � � � , we have � � � � � , so thetarget wavenumber after interaction is always less or equal tothe initial wavenumbers. In other words, wave interactionstend to transfer magnetic energy to smaller wavenumbers,i.e. to larger scale. This corresponds to an inverse cascade.The realisability condition, �

� � �

� � �

, was perhaps the

Page 5: Magnetic helicity in stellar dynamos: new numerical experiments

A. Brandenburg et al.: Magnetic helicity in stellar dynamos 103

most important assumption in this argument. Another impor-tant assumption that we made in the beginning was that theinitial field be fully helical; see Sect. 4.5 for the case of frac-tional helicity.

We note that in hydrodynamics, without magnetic fields,the kinetic helicity spectrum, �

� � �

� �

� � � Æ � , where� � � � � is the vorticity, is also conserved by the nonlinearterms, but now the realisability condition is

� �

� �

� � �

� (31)

so it is not � , but the inverse of � , that enters the inequality.Here, �

� �

� � is related to the kinetic energy spectrum(i.e. without density), and hence one has

� �

� �

� �

� � �

� �

� �

� �

� � � �

� (32)

or

� �

� � � � � �

� �

� �

� �

� (33)

Thus � � � � � � � � � , and in the special case � � � we have� � � � � , so kinetic energy is transferred preferentiallyto larger wavenumbers, so there is no inverse cascade in thiscase.

2.3. Spectra of right and left handed components

For further reference we now define power spectra of thosecomponents of the field that are either right or left handed,i.e.

� � � �

� � �

� �

� (34)

Thus, we have �

� �

� �

and �

� �

� �

. Notethat �

and �

can be calculated without explicit decom-position into right and left handed field components using

� �

� �

� � �

� �

� �

� (35)

This method is significantly simpler than invoking explicitlythe decomposition in terms of �

.

2.4. Open boundaries and the importance of agauge-independent magnetic helicity

Let us now discuss the case where magnetic helicity is al-lowed to flow through boundaries. The first problem we en-counter in connection with open boundaries is that the mag-netic helicity is no longer gauge invariant. There are how-ever known procedures that allow � to be defined in a waythat is independent of gauge. Before we discuss this in moredetail we first want to demonstrate how the use of a gauge-dependent magnetic helicity can lead to undesired artifacts.

For illustrative purposes we consider the case of a one-dimensional mean-field �

� dynamo governed by the equation

� �

� �

� � � � �

� � (36)

where �

� �

� � is the sum of turbulent and microscopicmagnetic diffusivity, and � � � �

� �

� � � � � � � � is themagnetic field which depends only on � , and � � � � �

as well as � � � � � �

. Here we have used the gauge� � � . The overbars denote horizontal � � � � � averages and

Fig. 2. First panel: gauge-dependent magnetic helicity for an �

dynamo with vacuum boundary conditions. Second panel: gauge-independent magnetic helicities �

� �

(solid line) and �

(dashedline), compared with twice the magnetic energy, � � , for the samemodel (dotted line). Third panel: same as second panel, but with per-fect conductor boundary conditions. Here gauge-independent mag-netic helicities agree with the ordinary expression,

� � � � � .Again, the dotted line shows � � .

the corresponding mean quantities like � are therefore inde-pendent of � and � . As boundary conditions we adopt either� �

� � � � � �

� � � � � (vacuum boundary condition) or,for comparison, �

� �

� � (perfect conductor condition)on � � � � �

. We adopt � and �

quenching, i.e.

� �

� � �

� �

� �

� �

� �

� � �

� �

� �

(37)

(In the case considered here we use � � .)The result of a numerical integration of Eq. (36) is shown

in Fig. 2. Evidently, the use of a gauge-dependent magnetichelicity implies spurious contributions that prevent � � � � �

Page 6: Magnetic helicity in stellar dynamos: new numerical experiments

104 Astron. Nachr./AN 323 (2002) 2

from becoming time-independent even though the magneticenergy has long reached a steady state. The reason for thisis that for the steady state solution the right hand side ofEq. (36) is a constant which is different from zero. As anexample, consider the eigenfunction of the linearised equa-tions: � � � � � � � �

� � � � �

� � � � �

, where �

� for�

� � . In the marginally excited state, we have � � �

,so � �

� � � � � �

� � � �� � . Because of this complication, it isimportant to define magnetic helicity in a gauge-independentway. BD01 defined such a gauge-independent magnetic he-licity by linearly extending the horizontally averaged mag-netic vector potential to a periodic field which then resultsin a gauge-invariant magnetic helicity as before; see Eq. (1).This can be written in compact form as

� �

� � � � � � � �

� � � �

� �

� � � (38)

where �

and �

are the values of � on the lower and up-per boundaries, and angular brackets denote full volume av-erages, as opposed to the overbars, which denote only hori-zontal averages.2

We note that �

� �

can be written in the form proposed byFinn & Antonsen (1985),

� �

� � � � � � � � �

� � � � � �

� � � (39)

where �

is the non-helical reference field

� � � � � (40)

which is spatially constant. A corresponding �

that satisfies�

� � � �

is

� � � � �

� const� (41)

which, apart from some constant, corresponds to a linear in-terpolation of � between �

and �

. However, the constantin Eq. (41) drops out in the term �

� �

of Eq. (38). It isalso easy to see that the constant drops out in Eq. (39), be-cause � � � �

� � � . It is precisely this fact that makes (39)gauge-invariant.

In the following we compare with the magnetic helicity

in the gauge � �� � � � � � , which minimises � �

� ,

� � �

� � � � � � � � � � � � � � � � (42)

This magnetic helicity is actually also gauge-invariant, but itcannot be written in the form (39), because it is not based ona reference field. Indeed, the only value of �

that makes theexpression (39) gauge-invariant is �

� � � � , but this doesalready corresponds to �

� �

.We associate �

� � �

with the Coulomb gauge because itis based on a gauge which minimises � �

� , which is char-acteristic of the Coulomb gauge; see Sect. 2.2. However, inthe present geometry, for horizontally averaged fields, andzero net magnetic flux, a canonical gauge transform � ��

� �

� � � � � � does not change the large scale magnetic he-licity � � � � � , since �

� � . In this sense, � � � � � � in anyarbitrary gauge represents the magnetic helicity in Coulombgauge. There is, however, the extra freedom to add a constant

2 We use this opportunity to point out a sign error in Eq. (9) ofBD01, where it should read �

� �

� �

� � �

� . The contri-butions from the mean field, that are central both here and in BD01,remain however unaffected.

vector � �

to the magnetic vector potential, � �� � � �

.

It is easy to see that � �

� is minimised if �

� � � � , and thatthe definition (42) always yields the magnetic helicity whichcorresponds to this gauge.

In the second panel of Fig. 2 we have plotted both �

� �

aswell as �

� � �

. Note that � �

� � �

� � � �

� �

� and also � �

� � �

� �

� � �

� �

. If the domain was periodic, and since the compu-tational domain is � � � � � � , we would have �

� �

� .However, for vacuum boundary conditions one could still ac-commodate a wave with �

� �

� � . Then also � �

� �

� isalways less than � � �

� �

.

In the more general case of three-dimensional fields weadopt the gauge-independent magnetic helicity of Berger &Field (1984). Their definition, however, leaves the contribu-tion from the average of � undetermined and can thereforeonly be applied to the deviations from the mean field.

2.5. Resistively limited growth and open boundaries

A plausible interpretation of the resistively limited growth isthe following (Blackman & Field 2000, Ji & Prager 2001).To satisfy magnetic helicity conservation, the growth of thelarge scale field has to proceed with very little (or very slow)changes of net magnetic helicity. This is essentially the resultof B01 (see also Brandenburg & Subramanian 2000, hereafterBS00). The underlying assumption was that magnetic helic-ity of the sign opposite to that of the mean field is lost resis-tively. As a working hypothesis we assume that such lossesoccur preferentially at small scales where also Ohmic lossesare largest.

In order to speed up this process it might help to get rid ofmagnetic helicity with sign opposite to that of the large scalefield. One possibility would be via explicit losses of field withopposite magnetic helicity through boundaries (i.e. the outersurface or the equatorial plane). This possibility was first dis-cussed by Blackman & Field (2000) and more recently byKleeorin et al. (2000) and Ji & Prager (2001).

The results obtained so far seem to indicate that this doesnot happen easily (BD01). Instead, since most of the mag-netic helicity is already in large scales, most of the losses arein large scales, too. Thus, the idea of losses at small scales,where the sign of magnetic helicity is opposite to that at largescale, has obvious difficulties. In Sect. 4.1 we address thequestion of observational evidence for this. In Sect. 4 we dis-cuss quantitatively the effect of losses on large scales.

3. Magnetic helicity cancellation

3.1. Magnetic helicity in ABC-flow dynamos

Before we start discussing the evolution of magnetic helic-ity in turbulent dynamos we consider first the case of steadykinematic ABC flows, where the velocity field is given by

� �

� � � �

� � � � � � �

� � � �

� � � � � � �

� � � �

� � � � � � �

� (43)

Page 7: Magnetic helicity in stellar dynamos: new numerical experiments

A. Brandenburg et al.: Magnetic helicity in stellar dynamos 105

Fig. 3. Power spectrum, �

, of the magnetic field for an ABC flowdynamo with � � � � �

� � and � � � . Note that mostof the power is in the small scales with �

� �

� � � . The normalisedhelicity spectrum is also plotted. Note that for medium and largevalues of � the spectral magnetic helicity is very small; �

� � �

. The simulation was done with � � �

� mesh points and a magneticdiffusivity � � � � � �

� � .

Fig. 4. Cross-section of the magnetic helicity density for the ABCflow dynamo from Fig. 3. Note the intermittent nature of � � � .

with � � � � � � and �

� in the domain� � � � � � � � � � . This class of dynamo models is of in-terest because here the net magnetic helicity can be small,even though the flow itself is fully helical (Gilbert 2002, andreferences therein). The field generated by such an ABC flowdynamo does have magnetic helicity, but it is comparativelyweak in the sense that the realisability condition (17) is farfrom being saturated. This is seen in Fig. 3, where we com-pare �

with �

� � �

� . In addition, the spectral helicity is al-ternating in sign, often from one wavenumber to the next;see the second panel of Fig. 3. The magnetic helicity den-sity is also alternating in space, with a smoothly varying,mostly positive component, interspersed by negative, highlylocalised peaks as shown in Fig. 4. Moreover, as �

� � ,the net magnetic helicity goes asymptotically to zero; seeFig. 5. This result is typical of fast dynamos (Hughes, Cat-taneo, & Kim 1996, Gilbert 2002).

We note in passing that the magnetic power spectrum in-creases approximately like �

� � � (see the left panel of Fig. 3),which is similar to the results from a number of other dy-namos as long as they are in the kinematic regime; see, e.g.,Fig. 22 of B01 for a large magnetic Prandtl number calcu-lation and Fig. 17 of Brandenburg et al. (1996) for convec-

Fig. 5. Magnetic helicity, � � � � , nondimensionalised by �

� � �

,as a function of

� � � �

for a model with � � � �

� � . Here, �

� � is the smallest wavenumber in the box,which has the size � � �

� . Note that �

� � � � � �

� decreases withincreasing

like

� � � �

. Thus, in the high magnetic Reynoldsnumber limit � � � � � , the magnetic helicity of the ABC flowdynamo vanishes.

tive dynamos. Thus, most of the spectral magnetic energy isin small scales, which is quite different from the helical tur-bulence simulations of B01. In turbulent dynamos this spec-trum could be explained by the analogy between magneticfield and vorticity (Batchelor 1950). For a Kolmogorov �

� � � �

spectrum, the vorticity spectrum has a �

� � � inertial range. Inthe present case, however, this argument does not apply be-cause the flow is stationary, so there is no cascade in velocity.Thus, the origin of the �

� � � spectrum for these types of kine-matic dynamos remains unclear.

One might wonder whether nearly perfect magnetic helic-ity cancellation can also occur in the sun. Before we can ad-dress this question we need to understand the circumstancesunder which this type of magnetic helicity cancellation canoccur. It appears that nearly perfect magnetic helicity cancel-lation is an artefact of the assumption of linearity and of asmall computational domain. Increasing its size allows largerscale fields to grow even though the field has already satu-rated on the scale of the ABC flow. This is shown in Fig. 6where we plot the magnetic energy of the mean magneticfield, � , for a ‘nonlinearised’ ABC flow with � � � , and� � � � � � � � � �

� , where � is the magnetic energyand �

� . We chose a domain of size � � � � � � ,which is increased in the � direction by a factor 4. In thekinematic phase the high wavenumber modes grow fastest.Once the magnetic energy reaches saturation, the velocity isdecreased, so the high wavenumber modes stop growing, butthe remaining lower wavenumber modes are still excited. Al-though they grew slower during the earlier kinematic stage,they are now still excited and saturate at a much later time.This process continues until the mode with the geometricallylargest possible scale has reached saturation, provided thatmode remains still excited. Once a large scale field of thisform has developed, cancellation of any form has becomesmall. This is indicated by the fact, that the magnetic helic-ity is nearly maximum, i.e. � � �

� � � �

� �

with �

be-

Page 8: Magnetic helicity in stellar dynamos: new numerical experiments

106 Astron. Nachr./AN 323 (2002) 2

Fig. 6. Evolution of the magnetic energy � � �

� � �

� , for a‘nonlinearised’ ABC flow with � � � , and � � � � � � � �

� � �

scales with the magnetic energy, � . The box has a sidelength of � in the � direction. In the kinematic stage (� � � � � )modes with � � � grow fastest. After the small scale modes satu-rate, the energy of the larger scales continues to grow until the largescale reaches saturation (e.g., � � � � � saturates after � � � � � � ).The field continues to grow even further, but very slowly. Since thebox has a side length of � , the minimum wavenumber is � � � � � � .

ing the momentary wavenumber of the large scale magneticfield, which is also indicated in Fig. 6. Thus, for this type ofdynamo, small scale magnetic fields are highly intermittentand can grow fast, while large-scale fields grow once againon slow time scales only. It is this type of large scale fieldgeneration that is often thought to be responsible for generat-ing the large scale field of the sun (see Parker 1979, Krause &Radler 1980, or Zeldovich et al. 1983, for conventional mod-els of the solar dynamo), although the saturation process isstill resistively slow.

3.2. Magnetic helicity cancellation in turbulent dynamos

Magnetic helicity cancellation seems to be a generic phe-nomenon of all kinematic dynamos, including the turbulentdynamos in cartesian domains that have been studied recentlyin B01. Even though the initial field is fully helical, the signsof magnetic helicity are different at large and small scalesand can therefore cancel, at least in principle. In practicethere remains always some finite net magnetic helicity, butas the magnetic Reynolds number increases, the total mag-netic helicity decreases, exactly like for the ABC-flow dy-namos studied in Sect. 3.1. Examples of magnetic powerspectra calculated separately for the right and left handedparts of the field are shown in Fig. 7 for Runs A, B, and C ofBrandenburg & Sarson (2002, hereafter referred to as BS02).These are models with helical forcing around a wavenumber�

, like in B01. The basic parameters of the runs presentedin BS02 are summarised in Table 1. Run A has ordinarymagnetic diffusivity with � � �

� � , whilst Runs B and Chave second order hyperdiffusivity, so � �

� is replaced by� �

� in the induction equation (4), and �

� � � �

� �

and �

� � for Runs B and C, respectively. [Generally, formodels with th order hyperdiffusivity, one would have a dif-

Fig. 7. Power spectra separately for the right and left handed partsof the field, for Runs A, B, and C of BS02. Different times havebeen collapsed onto the same graph by scaling the spectra with �

� � � ,where � is the corresponding growth rate of the magnetic energy foreach run. Note that as the magnetic diffusivity is reduced, the rela-tive excess of �

over �

decreases, as expected for asymptoticmagnetic helicity conservation.

fusion term � � �

� � �

� �

� .] In all three cases the forcingof the flow is at wavenumber �

� � with the same am-plitude ( !

� � ), as in B01. At the forcing scale, Run Bis by a factor �

� � �

� � � times less resistive thanRun A. Run C is another three times less resistive. Duringthe kinematic phase shown, the spectral power increases atall wavenumbers at a rate proportional to "

� , where # � $

is the kinematic growth rate of the magnetic energy and $ thegrowth rate of the rms magnetic field.

As the magnetic Reynolds number increases, the mag-netic helicity normalised by the magnetic energy, � � � � ,decreases, as shown in Fig. 8 for Runs A, B, and C of BS02.The helicity of the forcing and the kinetic helicity of the floware positive, so the magnetic helicity at small scales is alsopositive. This results in large scale magnetic helicity that isnegative. Therefore, also the total magnetic helicity is nega-tive.

The asymptotic decrease of magnetic helicity for progres-sively smaller magnetic diffusivity (Fig. 5) can be quantita-tively described by Berger’s (1984) magnetic helicity con-straint, which we modify here for the initial kinematic evolu-tion when the magnetic energy increases exponentially at therate # � $ . We consider first the case of ordinary magneticdiffusivity and turn then to the case of hyperdiffusivity. For aclosed or periodic domain, the modulus of the time derivativeof the magnetic helicity can be bounded from above by

� �

� �

� � � � � � � � � � � � � �

� � �

� �

� � �

� �

� � �

� � �

� � � � �

� � �

� � �

(44)

where �

� � � � �

� � �

� is the rate of Joule dissipation.In the kinematic regime, �

� � � # � � � and �

� � # � �

% � �

� � � � �

� & �

� � � � �

, where % is the work done on themagnetic field by the velocity field. In the dynamo case, i.e.if # ' � , we have & ' � . Thus

#

� � � � �

� � �

� # � � & �

� � �

� � �

� � �

� � � # � & �

� � �

� � � (45)

or

� � � � �

� � � &

� � � �

� � � # �

� � �

� � (

� � �

� (46)

Page 9: Magnetic helicity in stellar dynamos: new numerical experiments

A. Brandenburg et al.: Magnetic helicity in stellar dynamos 107

Fig. 8. Magnetic helicity, normalised by the magnetic energy, forRuns A, B, and C of BS02. In Run A, � � � �

� � , whereas in Runs Band C second order hyperresistivity is used, with �

� � � � �

� �

and � �

� � , respectively. Note that the smaller the magnetic diffusiv-ity (at large and intermediate scales), the slower is the saturation ofmagnetic helicity.

where (

� � �

� � � � # �

� � � is the skin depth associated withthe growth rate # and � � &

� � � � is a parameter describing thefractional excess of magnetic energy gain over dissipation.

In the case of hyperdiffusivity, which is often used fornumerical reasons, we have

� � � � �

� �

� � �

� �

� �

� � �

� �

� � � �

� � (47)

where �

� � � �

� �

is an effective wavenumber defined suchthat the second equality in Eq. (48) below is satisfied exactlyand in Eq. (47) only approximately. The rate of change ofmagnetic helicity is then bounded by

#

� � � � �

� � �

� � � �

� �

� �

� � �

� � �

� � � �

� �

� �

� � � �

� � �

� �

� � �

� �

(48)

or#

� � � � � �

� � � �

� � � �

� � �

� � �

� � � � �

� � �

� � �

� � �

� � � �

� � � �

� � �

� # � & �

� � �

� � �

(49)

So, again,

� � � � � � � � � (

� � �

� (50)

but now (

� � �

� � �

� � � �

� � � �

� # �

� � � involves the value of�

� � � �

. Like in the case with ordinary magnetic diffusivity wehave � � &

� � � � . From the values of � � � � �

� � and (

� � �

given in Table 1 we see that this constraint is indeed well sat-isfied during the kinematic growth phase.

3.3. Magnetic helicity conversion to large scales

At least initially, i.e. when the field is still growing exponen-tially, there is a strong conversion of positive magnetic he-licity at the driving scale to negative magnetic helicity at acertain larger scale. This can be described by the magnetichelicity equation, written separately for positive and negativeparts,

� ) � � �

� (51)

Fig. 9. Relative magnetic helicity transfer from small to large scalesfor runs A (most resistive), B and C (least resistive) of BS02, asa function of time. Note that for smaller magnetic diffusivity, theratio � � � � � decreases and hence the conversion to large scalesbecomes less efficient.

where �

� � , with �

� � �

� � ,and ) describes the transfer of magnetic helicity from smallto large scales. In runs with second order hyperresistivity, theterm � �

has to be replaced by �

. The evolution of the netmagnetic helicity, �

� �

� , is of course unaffected by ) . Wenote, however, that the spectrum of the magnetic helicity doeshave non-diffusive source and sink terms such that its integralover all wavenumbers vanishes (see B01 for detailed plots).The term ) in Eq. (51) is related to the � effect (Blackman& Field 2000) or, more precisely, to the turbulent electromo-tive force. In the present case, where turbulent diffusion alsoenters, we have therefore (cf. B01)

) � � � �

� �

� � � � � (52)

Since the other two terms in Eq. (51) are known, we can di-rectly work out ) from this equation. The results for Runs A,B, and C are plotted in Fig. 9. The quantity ) � � � � is a mea-sure of the rate of helicity transfer, i.e. the difference between� -effect and turbulent diffusion. Although the ratio ) � � � �

has not yet settled to a stationary value, we compare in Ta-ble 1 the values for three different runs (at � � � � � ).

Figure 9 shows that ) � � � � approaches a finite value inthe nonlinear regime, but its value becomes smaller as themagnetic diffusivity is decreased. Thus, even though strongfinal field strengths are possible, this can still be achievedwith a residual � effect that decreases with decreasing mag-netic diffusivity. This is consistent with earlier results of Cat-taneo & Vainshtein (1991), Vainshtein & Cattaneo (1992),and Cattaneo & Hughes (1996) that � and �

are quenchedin a �

-dependent fashion. Field & Blackman (2002) havepointed out, however, that this result can also be explained bya dynamically quenched � where � and �

are not necessarilyquenched in a �

-dependent fashion.One may wonder whether the �

-dependent slow-downduring the saturation phase of the large scale field (as seen inFig. 9), applies also to intermediate scales. In Fig. 10 we showhow the magnetic power spectrum shows a bump at interme-diate wavenumbers (i.e. the wide maximum at wavenumbers

Page 10: Magnetic helicity in stellar dynamos: new numerical experiments

108 Astron. Nachr./AN 323 (2002) 2

Table 1. Summary of runs, some of which were already presented in BS02. The runs with � � � have hyperdiffusivity, so the effectivediffusivity at the forcing wavenumber, �

, and at the Nyquist wavenumber, �

� �

� � � Æ � � � � � , are different. �

� is the total number ofmesh points, and Æ � is the mesh spacing. The parameter �

� � � �

gives an approximate upper bound for � � � � � � �

� .

Run � � �

� � � �

� � � �

� �

� �

� � � �

� � � � � � �

� �

� � � �

� � � � � �

� � � � � �

A � � �

� 1 � �

� �

� �

� �

� �

� � 0.047 27 � � � � � �

� � 0.035 0.065 0.00075B � � �

� 2 � � � �

� �

� � � �

� �

� �

� � 0.070 27 � � � � � �

� � 0.018 0.025 0.00035C � � �

� 2 � �

� �

� � � �

� �

� � � �

� � 0.082 27 � � � � � �

� � 0.005 0.013 0.00020D � �

� 2 � �

� �

� � � �

� �

� � � �

� � 0.078 3 – 0.08 0.15 –E � �

� 2 � �

� �

� � � �

� �

� � � �

� � 0.12 5 – 0.040 0.065 –F � � �

� 2 � � � �

� �

� � � �

� �

� �

� � 0.12 3 – 0.0013 0.0021 –

Fig. 10. Evolution of magnetic energy spectra in the hyperdiffusiveRun B in equidistant time intervals from � � � (lowest curve), � �

� � (next one higher up), until � � � � � (peaking at the very topleft). The dotted lines give the kinetic energy.

� � � �

) travelling to the left toward smaller wavenum-bers. In order to check whether or not the speed of the spec-tral bumps in the different runs shown in Figs 7 and 10also decreases with decreasing magnetic diffusivity we plotin Fig. 11 the position of the secondary bump in the powerspectrum as a function of time. It is evident that the slope,which signifies the speed of the bump, decreases with de-creasing magnetic diffusivity. A reasonable fit to this migra-tion is given by

� �

� �

� �

� � � �

� � � �

� �

� � (53)

where the parameter �

� � � �

characterises the speed at whichthis secondary bump travels. The values obtained from fits tovarious runs are also listed in Table 1. Following argumentsgiven by Pouquet et al. (1976), �

� � � �

obtained in this way isactually a measure of the � effect. The decrease of �

� � � �

withdecreasing magnetic diffusivity adds to the evidence that the� effect is quenched in a �

dependent fashion.The �

dependence of �

� � � �

and ) � � � � , both indica-tive of inverse energy transfer, seems to be in conflict withthe usual idea of an inverse cascade where such transfers hap-pen on a dynamical time scale (e.g., Verma 2001). The onlyreasonable conclusion from this seems to be that the (nonlo-cal!) inverse transfer discussed here is in fact distinct from theusual inverse cascade. In other words, the present simulations

Fig. 11. Evolution of the position of the secondary bump in thepower spectrum.

do not support the concept of an inverse cascade in the usual(local) sense.

3.4. Magnetically driven turbulence

In the cases considered so far, the turbulence is driven by aforcing term in the momentum equation. Another way to in-ject energy into the system is by adding an irregular externalelectromotive force �

� � �

� � � � � to the system. The inductionequation (4) is then replaced by

� �

� �

� � � � � � �

� � �

� � �

� � � � � (54)

This is in some ways reminiscent of turbulence driven bymagnetic instabilities (e.g. the magnetorotational instability,

Page 11: Magnetic helicity in stellar dynamos: new numerical experiments

A. Brandenburg et al.: Magnetic helicity in stellar dynamos 109

Fig. 12. Evolution of magnetic energy in runs with magnetic driv-ing. Forcing occurs in the induction equation around wavenumber�

� � . Runs with hyperdiffusivity are denoted by �

and thosewith ordinary diffusivity by �

.

buoyancy or kink instabilities). The direct injection of energyinto the induction equation was also considered by Pouquetet al. (1976), for example, who assumed the forcing to be he-lical. The results of such a run are shown in Fig. 12. Notethat the magnetic energy evolves to larger values as the mag-netic diffusivity is decreased. This is because the injectionof magnetic energy is primarily balanced by magnetic dif-fusion, which becomes weaker as the magnetic diffusivity isdecreased. In the runs with second order hyperdiffusivity (de-noted by �

) the magnetic energy evolves to even larger val-ues. What is more important, however, is the time scale onwhich saturation is achieved. This time scale is the resistiveone, suggesting again that the transport of magnetic energyacross the spectrum from small to large scales is a resistivelylimited process.

4. Evolution of the large scale magnetic helicity

4.1. Magnetic helicity flux from global mean-field models

The only information available to date about the magnetic he-licity of the sun is from surface magnetic fields, and these dataare incomplete. Vector magnetograms of active regions shownegative (positive) current helicity on the northern (south-ern) hemisphere (Seehafer 1990, Pevtsov, Canfield, & Met-calf 1995, Bao et al. 1999, Pevtsov & Latushko 2000). Fromlocal measurements one can only obtain the current helicity,so nothing can be concluded about magnetic helicity. Typi-cally, however, current and magnetic helicities agree in sign,but the current helicity is stronger at small scales.

Berger & Ruzmaikin (2000) have estimated the flux ofmagnetic helicity from the solar surface. They discussed � -effect and differential rotation as the main agents facilitat-ing the loss of magnetic helicity. Their results indicate thatthe flux of magnetic helicity due to rotation and the actualmagnetic field obtained from solar magnetograms is nega-tive (positive) in the northern (southern) hemisphere. Thus,the sign agrees with that of the current helicity obtained us-ing vector magnetograms. Finally, Chae (2000) estimated themagnetic helicity flux based on counting the crossings ofpairs of flux tubes. Combined with the assumption that two

nearly aligned flux tubes are nearly parallel, rather than an-tiparallel, his results again suggest that the magnetic helicityis negative (positive) in the northern (southern) hemisphere.

What is curious here is that even though different scalesare involved (small and intermediate scales in active regionsand coronal mass ejections, and large scales in the differentialrotation), there is as yet no evidence that the helicity spectrumof the sun reverses sign at some scale.

We now discuss the magnetic helicity properties of mean-field � � dynamos in a sphere. There are different reasonswhy looking at mean-field dynamos can be interesting in con-nection with helical turbulence. Firstly, magnetic helicity in-volves one wavenumber factor less than the magnetic energyand tends therefore to be dominated by the large scale field.Magnetic helicity should therefore be describable by mean-field theory. Secondly, mean-field dynamos easily allow thespherical geometry to be taken into account. This is mainlybecause such models can be only two-dimensional. The re-sulting field geometry does in some ways resemble the solarfield. We recall, however, that at present there is no satisfac-tory model that explains both the equatorward migration ofmagnetic activity as well as the approximate antiphase be-tween radial and azimuthal field ( �

� � during mostof the cycle; see Stix 1976), and is still compatible with theobserved differential rotation ( � � � � * ' � at low latitudes).The sign of differential rotation does not however affect thesign of the magnetic helicity. In fact, the velocity does nevergenerate magnetic helicity, it contributes only to the flux ofmagnetic helicity; see Eqs (5) and (6).

Table 2. Properties of mean-field dynamos for different signs of �

and radial shear, �

� � � � � � . All of these models are problematicwhen applied to the sun. Case (iii), where �

� � and �

� � ,satisfies two important constraints (butterfly diagram with equator-ward migration, antiphase between radial and azimuthal field), butit is incompatible with the helioseismology data which indicate thatat least at low latitudes �

� � .

� � �

� �

� � Case (i) Case (ii)helioseismology not OK helioseismology OK

butterfly not OK butterfly OKphase rel. OK phase rel. not OK

and �

negative �

and �

negative

� � Case (iii) Case (iv)helioseismology not OK helioseismology OK

butterfly OK butterfly not OKphase rel. OK phase rel. not OK

and �

positive �

and �

positive

For the axisymmetric mean field, � , of an � � dynamowe calculate the gauge-invariant relative magnetic helicity ofBerger & Field (1984) for the northern hemisphere (indicatedby N),

� � � �

� � � � � �

� � � � (55)

Page 12: Magnetic helicity in stellar dynamos: new numerical experiments

110 Astron. Nachr./AN 323 (2002) 2

where �

� � � �

is a potential reference field inside thenorthern hemisphere, whose radial component agrees with �

(see Bellan 1999 for a review). This implies that �

� �

� �

on the boundary of the northern hemisphere (both on theouter surface and the equator).

In spherical coordinates an axisymmetric magnetic fieldcan be expressed in the form � � �

� � � �

. Thepotential reference field is purely poloidal. Therefore, the dotproduct of the toroidal components in the integral (55) is sim-ply

! � � � "

� �

� �

� �

� �

� � (56)

For the dot products of the poloidal components of the in-tegral we express the magnetic vector potential in the form� � �

� � � �

, where �

� � � � � � �

� � �

.We can then write the integrand for the poloidal componentas

� � � �

� � � � � �

� �

� �

� �

� �

� �

� �

� � � � � � � �

� �

� �

� (57)

(Note that �

� �

� � , as the potential reference field is purelypoloidal.) The integral over the divergence term vanishes be-cause �

� �

� �

on the boundary, so we have

! # � � "

� �

� �

� �

� � � � (58)

and hence

� � (59)

The time evolution of �

can be obtained from the mean-field dynamo equation (e.g. Krause & Radler 1980), whichgives

� �

� � � � ) � �

� (60)

where ) �

� � � � � was already defined in Eq. (52),� � � � � �

� is the e.m.f. of the mean field, and �

is the integrated magnetic helicity flux out of the northernhemisphere (i.e. through the outer surface and the equator),with

� �

� � � � � � � � � �

� � � � � (61)

where � � * � � + �

is the mean toroidal flow. It willbe of interest to split �

into the contributions from theouter surface integral, �

$

, and the equatorial plane, �

% �

,so �

� �

$

� �

% �

.In Table 2 we summarise the key properties of typical � �

dynamo models for different signs of � and �

� � � � � * (cf.Stix 1976). Here we have also indicated the sign of the mag-netic helicity flux that leaves from the northern hemisphere,�

. In all the models presented here the sign of �

agreeswith that of �

, which also agrees with the sign of � . This isbecause, averaged over one cycle, � � �

� � � �

& ' &

vanishes, sothe sign of � �

& ' &

must agree with that of � ) �

& ' &

which, inturn, is determined by the sign of � ; see Eq. (60).

What remains unclear, however, is whether the negativecurrent helicity observed in the northern hemisphere is in-dicative of small or large scales, and hence whether the sign

of � is positive (case iv) or negative (case ii), respectively.Only in case (iv) there would be the possibility of sheddingmagnetic helicity of the opposite sign; see Sect. 4.3 below. Incase (iii), �

is also positive, but here �

� � , which doesnot apply to the sun. In case (ii), on the other hand, �

isnegative, so shedding of negative magnetic helicity would actas a loss.

Our goal is to estimate the amount of magnetic helicitythat is to be expected for more realistic models of the solardynamo. We also need to know which fraction of the mag-netic field takes part in the 11-year cycle. In the followingwe drop the subscript N on � and � . For an oscillatory dy-namo, all three variables, � , � , and �

� � � � �

vary in an oscil-latory fashion with a cycle frequency , of magnetic energy(corresponding to 11 years for the sun – not 22 years), sowe estimate � � � � � � �

, � � � and �

� � � � �

, � , which, byarguments analogous to those used in Sect. 3.2, leads to theinequality

� � � � �

� � � (

� � �

� (62)

where (

� � �

� � � � , �

� � � is the skin depth, now associatedwith the oscillation frequency , . Thus, the maximum mag-netic helicity that can be generated and dissipated duringone cycle is characterised by the length scale � � � � �

� � ,which has to be less than the skin depth (

� � �

.

For � we have to use the Spitzer resistivity which isproportional to -

� � � � ( - is temperature), so � varies be-tween �

� �

� � at the base of the convection zone to about �

(

� �

� � near the surface layers and decreases again in thesolar atmosphere. Using , � � � � � � � � � �

� �

� �

for the relevant frequency at which � and � vary we have(

� � �

� � � � at the bottom of the convection zone and(

� � �

� � � � � � at the top. This needs to be compared withthe value � � � � �

� � obtained from dynamo models.

In order to see how the condition (62) is met by mean-field models, we present the results from a typical � � modelof the solar dynamo. Although mean-field theory has beenaround for several decades, the helicity aspect has only re-cently attracted attention. In the proceedings of a recent meet-ing devoted specifically to this topic (Brown, Canfield, &Pevtsov 1999) helicity was discussed extensively also in thecontext of mean-field theory. However, the possibility of he-licity reversals at some length scale was not addressed atthe time. We recall that in the Babcock–Leighton approachit is mainly the latitudinal differential rotation that enters.We also note that, although the latitudinal migration couldbe explained by radial differential rotation, meridional circu-lation is in principle able to drive meridional migration evenwhen the sense of radial differential rotation would otherwisebe wrong for driving meridional migration (Durney 1995,Choudhuri, Schussler, & Dikpati 1995, Kuker, Rudiger, &Schultz 2001). We therefore proceed with the simplest pos-sible model in spherical geometry. An example of the result-ing butterfly diagrams of toroidal field, magnetic helicity andmagnetic helicity flux is shown in Fig. 13.

The main parameters that can be changed in the modelare �

(we assume that � � �

� � � + , where + is colatitude),

Page 13: Magnetic helicity in stellar dynamos: new numerical experiments

A. Brandenburg et al.: Magnetic helicity in stellar dynamos 111

Table 3. Results from a mean-field dynamo model in a spherical shell with outer radius and inner radius � � � , with either purelylatitudinal shear ‘(lat)’ or purely radial shear ‘(rad)’. The values of � given here (in units of �

� ) lead to dynamos that are only slightlysupercritical. The cycle period, �

� � �

, is given in non-dimensional form and would correspond to � � � � for �

� � � � � � �

� �

� �

� � (forthe models with latitudinal shear), or � � � � � �

� �

� �

� � (for the models with radial shear). The range of �

� !

� �

" ! �

that is consistentwith observations, � � � � � � � � �

� � , is highlighted in bold face. Both �

and �

vary with the dynamo cycle period; �

� #

and �

refer to the maximum and time averaged values of �

, respectively. �

� $

� �

� � �

is the outwards magnetic helicity flux through the northernsurface, while �

% &

� �

� � �

is the magnetic helicity flux through the equator into the southern hemisphere.

'

� !

" ! �

� #

� �

� �

� $

� �

� � �

" ! �

% &

� �

� � �

" ! �

� � �

� � � �

� �

� � � �

� (lat) � � � �

� �

� � � �

� �

� � � �

� �

� � � �

� �

� � � �

� � 0.030� � � �

� �

� � � �

� (lat) � � � �

� �

� � � �

� �

� � � �

� �

� � � �

� �

� � � �

� � 0.030� � � �

� �

� � � �

� (lat) � � � �

� �

� � � �

� �

� � � �

� �

� � � �

� �

� � � �

� � 0.030� � � �

� �

� � � �

� (rad) � � � �

� �

� � � �

� �

� � � �

� �

� � � �

� �

� � � �

� � 0.047� � � �

� �

� � � �

� (rad) � � � �

� �

� � � �

� �

� � � �

� �

� � � �

� �

� � � �

� � 0.047� � � �

� �

� � � �

� (rad) � � �

� �

� � � �

� �

� � � �

� �

� � � �

� �

� � � �

� � 0.047

Fig. 13. Butterfly diagrams of the toroidal field from a mean fieldmodel, compared with butterfly diagrams of the magnetic helicitydensity near the surface and magnetic helicity flux due to differentialrotation at the surface. Note that the local magnetic helicity densityis most of the time positive (light grey) in the northern hemisphereand negative (dark grey) in the southern hemisphere. The magnetichelicity flux shows a similar variation, but there is a certain degreeof cancellation taking place at any moment in time.

� � � � � � ,

'

� � � � �

� (lat).

(=const), and the magnitude of the angular velocity, �

,in the expression

� � �

� � � � �

.

� � � � + �

� � +

� � � � �

.

� � � � + �

� � +

� (63)

which is a fit that resembles closely the surface angular ve-locity of the sun (Rudiger 1989). What matters for the ax-isymmetric dynamo is only the pole–equator difference, � � ,which is about a third of �

. Our model is made dimension-less by putting � � �

� . The governing non-dimensional

parameters are �

� �

� � �

and �

)

� � � �

� �

, where� � � �

� �

� �

� �

. This leads to the observables �

# � � �

, bywhich we mean the radial field at the pole, which is about 2gauss for the sun, and �

* � � �

, by which we mean the azimuthalfield in the main sunspot belts. The solar value of �

* � � �

is es-timated to be around � � � � , based on equipartition argu-ments and on the total emergent flux during one cycle.3 Thus,we expect �

# � � �

� �

* � � �

to be in the range � � � �

� � .

From the model results shown in Table 3 we see that,once �

# � � �

� �

* � � �

is in the range consistent with obser-vations, (highlighted in bold face), � �

� � �

� � isaround � � � � �

� � for models with latitudinal shearusing Eq. (63). Given that � � � � � ! � this means that�

� �

� � � � � � � � , which would be compatiblewith (

� � �

only near the top of the convection zone.

In units of �

* � � �

� , the magnetic helicity flux throughthe northern hemisphere is about � � � �

� � . Using typ-ical parameters for the sun, �

* � � �

� � � � and � � � � � ! � ,

we have �

* � � �

� � � �

� �

! �

� . Thus, our models pre-dict magnetic helicity fluxes of about � � � �

� (

! �

for a full 22 year magnetic cycle, or half of that for the 11year activity cycle. This is comparable to the values reportedby Berger & Ruzmaikin (2000), but less that those given byDeVore (2000). Significant magnetic helicity fluxes throughthe equator are only found if angular velocity contours crossthe equatorial plane, i.e. in the models with radial differentialrotation.

If the observed negative magnetic helicity flux on thenorthern hemisphere is associated with the large scale field, itwould suggest that �

� � . This corresponds to case (ii),where magnetic helicity flux would contribute as an addi-tional loss of large scale magnetic energy. On the other hand,if the observed negative magnetic helicity flux is associatedwith the small scale field, it could remove excess magnetichelicity of the opposite sign, which may act as a source of

3 For the sun the total emergent flux per cycle is about � �

� �

� � .Assuming that the toroidal field is distributed over a meridional sur-face of � � � � � � � � � � � � �

� � � � � � �

� �

� �

� , this amountsto an estimated toroidal field of � � � � � � .

Page 14: Magnetic helicity in stellar dynamos: new numerical experiments

112 Astron. Nachr./AN 323 (2002) 2

large scale magnetic energy; see Sect. 4.3. This correspondsto case (iv) where �

' � .

4.2. Balance equations

The models with imposed positive helical forcing at smalllength scale show clearly that there is a production of equalamounts of positive and negative magnetic helicity at smalland large scales, respectively (or vice versa if the helicity ofthe forcing is negative). As time goes on, more and morepositive and negative magnetic helicity builds up on smalland large scales. Of course, magnetic helicity always impliesmagnetic energy as well (this is related to the realisabilitycondition; see Sect. 2.2 and Moffatt 1978), and eventually thevalue of the small scale magnetic energy will have reachedthe value of the kinetic energy at that scale, so the dynamoreaches saturation. This happens first at small scales, and ifthat was where the dynamo stops, the resulting field would begoverned by small scales (Kulsrud & Anderson 1992). How-ever, this is not what happens in the majority of the numericalexperiments, and probably also not in real astrophysical bod-ies like the sun.

Before we go on, we need to address the question of op-posite helicities in the two hemispheres of astrophysical dy-namos. For reasons of symmetry, the total magnetic helic-ity is always close to zero. It is then convenient to think ofthe two hemispheres as two systems coupled with each otherthrough an open boundary (the equatorial plane). The equa-tion of magnetic helicity conservation has an additional fluxterm, � , on the right hand side, so it reads [cf. Eq. (5)]

� �

� �

� � � � � � � (64)

where � is the magnetic helicity, � the current helicity, � themicroscopic magnetic diffusivity, and � the magnetic helicityflux integrated over the surface of the given subvolume (i.e. ahemisphere in this case).4 We are interested in the evolutionof the large scale field � , and we refer to the magnetic helicityof this large scale or mean field as �

.In the absence of boundaries we have �

� � � � � ,where � � � � � and overbars denote a suitable aver-age that commutes with the curl operator and satisfies theReynolds rules (in particular, the average of fluctuations van-ishes and the average of an average gives the same; seeKrause & Radler 1980). With open boundaries, however,�

� � � � � is no longer gauge invariant. For the presentdiscussion we only need the fact that it is possible to de-fine gauge-invariant versions of magnetic helicity; see Berger& Field (1984) for the three-dimensional case and BD01 (orSect. 2.4) for the one-dimensional case that is relevant to hor-izontally averaged fields.

Splitting the field into contributions from mean and fluc-tuating components we can write

� �

� �

� �

� �

� � � �

� � �

� �

� �

(65)

This equation only becomes useful if we can make appro-priate simplifications. BD01 considered the case where the

4� is not to be confused with the rate of Joule heating, �

( � !

.

small scale field has already reached saturation and the largescale magnetic energy continues to build up further until ittoo reaches saturation.

If the large scale field is nearly fully helical we have�

� �

� , where �

is the typical wavenumber ofthe mean field. Furthermore, if the small scale field is nearlyfully helical, too, we have �

� �

� , where �

is thetypical wavenumber of the fluctuating field, or the forcingwavenumber, as in B01. Analogous relations link � to � .The relative ordering of the ratios of large and small scalemagnetic energies relative to magnetic and current helicitiesis given by

� �

� �

� �

� �

(66)

Thus, if �

� �

, as is usually the case when there is somedegree of scale separation, and �

� �

, we have � �

� �

� �

� and � �

� � � �

� , so we are left with� �

� �

� � � �

� �

� �

� (67)

or, using the above estimates between � , � and � , and as-suming, for definiteness, that �

� �

' � and �

� �

� �

(as expected in the northern hemisphere of the sun), we have� �

� �

� � � �

� �

� �

(68)

(The factor 2 in front of the � s comes from the 1/2 in thedefinition of � .) We can apply this equation to the problemwith open boundaries.

However before doing this, we recall that in the absenceof any surface fluxes of helicity ( � � � ) the assumption�

� �

does not apply near saturation. Instead, steadi-ness of � immediately requires � � �

� �

� � , so� �

� � � �

� , and hence �

� � �

� �

� �

� �

, and� �

� � � �

� for the final state. Before this final state isreached we have from Eq. (65)

� �

� �

� � � �

� � �

� (69)

which has the solution (B01)

� "

� � � �

! � � �

"

(70)

This equation is referred to as the magnetic helicity con-straint, which is probably obeyed by all helicity-driven largescale dynamos. Note, however, that no mean-field theory hasbeen invoked here. Equation (70) implies that full satura-tion is reached not earlier than after a resistive time scale,� � �

� � . We emphasise that this is the resistive time scaleconnected to the large length scale � � �

. We also empha-sise that this result is unaffected by turbulent or other (e.g.ambipolar; see BS00) magnetic diffusion.

Coming back to the problem with open boundaries wecan carry out a similar analysis by relating �

to the meanfield. Based on dimensional arguments we may assume

� �

� � �

� � �

/

� � (71)

where /

is some damping time for the large scale field. Wecan express this in terms of an effective (turbulent-like) mag-netic diffusion coefficient, �

, for the large scale field, e.g. inthe form

� � /

� � (definition of �

) � (72)

Page 15: Magnetic helicity in stellar dynamos: new numerical experiments

A. Brandenburg et al.: Magnetic helicity in stellar dynamos 113

so Eq. (68) becomes� �

� �

� � � �

� �

� �

� (73)

which has the solution

� �

� �

� "

� � �

! � � �

"

(74)

We expect �

to scale with the turbulent magnetic diffusiv-ity, �

, which describes the spreading of the mean magneticfield. This equation yields the time scale � �

� � , which isindeed shorter than the resistive time scale, � � �

� � . How-ever, if �

� � in Eq. (74), the energy of the final fieldrelative to the small scale field is decreased by the factor� � �

� � � �

� � . This was the basic result of BD01 where� �

� was found to be much smaller than � �

� . In other words,all the magnetic large scale helicity loss occurred at largescales, which is of course the scale where we wanted to gen-erate the large scale field.

The above results have shown that magnetic helicitylosses due to large scale magnetic fields are not really a sat-isfactory solution to the problem of generating strong largescale magnetic fields on fast enough time scales. Rather, onerequires that there should be a loss of magnetic helicity dueto small scale magnetic fields, i.e. a non zero �

. We mayassume that �

and �

have opposite sign, because bothcurrent and magnetic helicities also have different signs forthe contributions from small and large scales. Looking backat Eq. (74) is it clear that the �

term in the numerator canin principle enhance and even completely supersede the re-sistively limited ‘driver’, � �

, provided �

� � on thenorthern hemisphere.

Physically speaking, the role of small scale magnetic he-licity flux is to let the large scale field build up independentlyof any constraints related to the small scales. To prevent thenexcessive build-up of magnetic energy at small (and interme-diate) scales, which diffusion was able to do (albeit slowly),we need to have another more direct mechanism that is inde-pendent of microscopic magnetic diffusion. In order to in-vestigate the problem of excess magnetic helicity at smallscales we have performed a series of numerical experimentsby Fourier filtering the field and removing in regular time in-tervals the magnetic energy at the forcing and smaller scales.

4.3. Removing excess magnetic helicity

According to our working hypothesis, the reason why thelarge scale field can only develop slowly is that the build-up of the large scale field with one sign of magnetic helicityrequires the simultaneous build-up (and eventual removal) ofsmall scale field with opposite magnetic helicity. Instead ofwaiting for ohmic field destruction one might expect that alsothe explicit removal of magnetic energy at small scales wouldbe able to accomplish this goal.

A more formal argument is this. We use again Eq. (65),but without ignoring the � �

� � � term, and make the assump-tion that large and small scale fields are fully helical, andignore surface terms, because boundary conditions are as-sumed periodic. Thus, we have an additional � �

� � � termin Eq. (69), which then reads

� �

� �

� �

� �

� � �

� � �

(75)

Fig. 14. The effect of removing small scale magnetic energy inregular time intervals � � on the evolution of the large scale field(solid lines). The dashed line gives the evolution of �

� for Run 3of B01 (where no such energy removal was included) in units of�

! &

� �

� . The two solid lines show the evolution of �

after restarting the simulation from Run 3 of B01 at ! � � � � and! � � � . Time is scaled with the kinematic growth rate ! . Thecurves labelled (a) give the result for � � � � � � � !

� � and those la-belled (b) for � � � � � � !

� � . The inset shows, for a short time inter-val, the sudden drop and subsequent recovery of the total (small andlarge scale) magnetic energy in regular time intervals.

After small scale magnetic energy is removed, the right handside of Eq. (75) is out of balance. If we ignored the � �

� � �

term, the large scale magnetic energy would decay propor-tional to � � �

, because the � �

term is stillnegligible after having set �

� � . However, the �

termbegins to recover immediately which gives rise to a signif-icant contribution from the � �

� � � term. Both terms enterwith the same sign. If the magnetic Reynolds number is large,i.e. if the recovery rate of �

is faster than than the micro-scopic decay time, we have a net gain. This is indeed borneout by the simulations, where magnetic energy is removedabove the wavenumber � � � every � � � � $

� � timeunits, labelled (a) in Fig. 14, and � � � � � $

� � , labelled (b).Here, $ � � � � 0

is the kinematic growth rate of therms field strength. Shorter time intervals do not give any fur-ther enhancement; the resulting curve for � � � � � � $

� � liesalmost on top of the curve (a) that starts at $ � � � � .

We may thus conclude that the preferential removalof small scale magnetic fields does indeed enhance largescale dynamo action. Earlier experiments, which are not re-ported here, where �

was larger, and therefore the magneticReynolds number 0

� � �

smaller, did not show such anenhancement. Note however that we still do not really knowwhether the preferential loss of such small scale field doesoperate in real astrophysical bodies. Simulations have so farbeen unsuccessful in demonstrating this. Those presented inBD01 show that most of the magnetic helicity is lost throughthe large scale field. A possible reason may be that the largescale field geometry was not optimal and that in other settingslarge scale losses are reduced.

The intermediate scales are probably those where most ofthe magnetic helicity loss from the sun occurs. This argument

Page 16: Magnetic helicity in stellar dynamos: new numerical experiments

114 Astron. Nachr./AN 323 (2002) 2

is based in particular on the sign of the flux of magnetic he-licity on the solar surface: it is found to be negative (positive)on the northern (southern) hemisphere (see � 4.1 above).

According to conventional mean-field theory the sign ofthe � -effect should be positive in the northern hemisphere, sothe sign of large scale magnetic helicity and the flux thereofshould be positive. A reversal of the sign of magnetic helic-ity can then be expected on some scale smaller than the largescale of the dynamo wave ( � � –� � � ! � ). Thus, the helic-ity flux occurs on the scale of active and bipolar regions aswell as the scale of coronal mass ejections, which is about � – � ! � . These scales may therefore well correspond tothe intermediate scales referred to above. On the other hand,there is some uncertainty as to whether the dynamo alpha inthe northern hemisphere of the sun might in fact be negative.If the dynamo � is negative in the northern hemisphere, themagnetic helicity loss on the sun would be of the same signas that of the large scale dynamo-generated field, so its scalewould still belong to the large scales. However, the solutionto the helicity problem would then probably not be relatedto magnetic helicity flux, because then �

and �

, as wellas �

, would have the same sign, so �

would constitutean additional loss term, which does hence not accelerate thedynamo (BD01); see also Sect. 4.

4.4. Large aspect ratio simulations

In order to assess the role of intermediate scales, we now ex-tend the simulations of BD01 with open boundaries to do-mains that are larger in one horizontal coordinate directioncompared to before. We have already mentioned in the intro-duction that the simulations of BD01 produced most of themagnetic helicity at large scale, i.e. the same scale as that ofthe large scale field that we want to build up fast. A potentialproblem with these simulations could be a lack of sufficientscale separation. In particular, it is conceivable that interme-diate scales are not sufficiently well represented in these sim-ulations. The intermediate length scales can be expected to bewhere most of the magnetic helicity of the opposite sign re-sides (‘opposite’ is here meant to be relative to the large scalefield).

In order to study the possible role of the allowance of in-termediate length scales in the simulations we extend the sim-ulations of BD01 to the case of larger aspect ratio. The hopeis that the large scale field will now develop in the coordinatedirection that is longest, but that most of the magnetic helicityflux occurs still at a similar scale as before and that the signof magnetic helicity and its flux are now reversed relative tothe helicity at the scale of the mean field. In the simulationsof BD01 a box with unit aspect ratio was considered, andone of the coordinate directions (the � -direction) was non-periodic and magnetic helicity flux through these boundariesoccurred. We have now extended this study to the case wherethe box is four times larger in one of the other directions (the� -direction). A three-dimensional visualisation of the result-ing field is shown in Fig. 15. The � component of the mag-netic field on the � -boundaries shows a clear large scale pat-tern with a length scale larger than the extent of the domain

Fig. 16. Evolution of kinetic and magnetic energies. The magneticenergy contained in the large scale field (averaged in the � and �

directions, and denoted by "

) is shown as a dash-dotted line. Thedotted and dashed curves denote the � - and � -dependent compo-nents, respectively.

Fig. 17. Comparison of the evolution of the large scale magneticfield component that varies in the # -direction (the direction in whichthe box is longest) for runs with different aspect ratios, � , and dif-ferent values of �

� �

� � � � �

� � . Note the development of a strongerlarge scale field for � � � than for � � � .

in the � direction. The evolution of the magnetic and kineticenergies of one such run is shown in Fig. 16. We see thatkinetic and magnetic energies are approximately in equipar-tition. The large scale magnetic energy shows, as expected,variation in the � direction, i.e. the wave vector of the largescale field points in the long direction of the box.

Different runs are compared in Fig. 17. All the modelswith open boundaries have the property of developing rapidlya mean field that varies in the � -direction, which is the longdirection. When the aspect ratio � is small, however, the finalfield is not that which varies in the � -direction, but instead in

Page 17: Magnetic helicity in stellar dynamos: new numerical experiments

A. Brandenburg et al.: Magnetic helicity in stellar dynamos 115

Fig. 15. Three-dimensional visualisation of the magnetic field for a simulation where the box is four times longer in the # -direction. Vectorsin the back half of the box show the orientation of the locally averaged field. Ribbons in the front half indicate magnetic field lines; theirtwist reflects the local torsality (or helicity) of the magnetic field and has been enhanced artificially by factor of 15. The vertical magneticfield component is indicated by bright (positive) or dark (negative) colours in the top and bottom face of the box.

Table 4. Summary of the main properties of the runs with vertical field boundary condition. The parameter $

gives the fractional magneticenergy in the mean field relative to the total magnetic field.

� % � &

� �

'

� �

$

$

$

Vert 1 1 0.01 0.01 0.10 0.09 0.76 0.48 0.48Vert 2 1 0.005 0.005 0.16 0.14 0.59 0.45 0.45Vert 3 1 0.002 0.002 0.20 0.18 0.39 0.30 0.30Vert 4 1 0.002 0.001 0.19 0.17 0.24 0.20 0.20A1/10 1 0.02 0.001 0.11 0.12 0.61 0.47 0.47A1/05 1 0.02 � � � � � � 0.11 0.13 0.56 0.43 0.43A4/10 4 0.02 0.001 0.11 0.11 0.00 0.00 0.33A4/05 4 0.02 � � � � � � 0.11 0.12 0.02 0.02 0.19

the � -direction. This is seen from Table 4 where we comparethe relative strengths of the mean field,

� 1

� � � � � �

� � �

� � 2 � � � � � � (76)

where � � �

is the mean field averaged in the two directionsperpendicular to 2 (cf. B01). In all the cases with � � , �

is larger than �

and �

. This is not the case when � � � :here �

and �

are very small. Nevertheless, as the magneticReynolds number is increased, �

decreases (from 0.33 to0.19 when � is lowered from �

� � to � �

� � ). Also, com-paring �

in the � � � case with �

is the � � case, themean field is clearly less strong.

Thus, we may conclude that larger aspect ratios favourlarge scale magnetic field configurations that vary in the hori-zontal ( � ) direction instead of the vertical ( � ) direction. How-ever, at late times the saturation field strength is still decreas-ing with decreasing resistivity. It seems therefore worthwhilepursuing this work by studying still larger aspect ratios andyet different field geometries.

4.5. Fractional small scale magnetic helicity

In practice the magnetic helicity will never be near 100%;typical values are around 3-5% for convective and accretiondisc turbulence (Brandenburg et al. 1995, 1996). This hasprompted Maron & Blackman (2002) to consider the effectsof a forcing that has only fractional helicity (i.e. the realis-ability condition for the velocity is not saturated). They foundthat there is a threshold in the degree of helicity above which(large-scale) dynamo action is possible. However, these re-sults were obtained at a resolution of only " �

� meshpoints,and it is therefore possible, that this threshold effect is reallythe result of an increase of the critical magnetic Reynoldsnumber above which the large scale � -type dynamo operates,and that the model with that resolution has dropped belowthis critical value for large scale dynamo action. Below wewill estimate whether the dynamo number was in fact largeenough for this to happen.

Page 18: Magnetic helicity in stellar dynamos: new numerical experiments

116 Astron. Nachr./AN 323 (2002) 2

Fig. 18. Evolution of kinetic and magnetic energies in two runs withfractional helicity. On the left hand panel, �

� � and � � � � � �

� � ,whilst on the right hand panel, �

� � � and � � � � � �

� � .

The effect of fractional helicity of the forcing on the evo-lution of the large scale magnetic energy can be assessed interms of the magnetic helicity constraint, using a generalisedform of Eq. (69). In the case of fractional magnetic helicitythe small and large scale magnetic helicities will be less thanexpected based on the magnetic helicity constraint, so

� �

� � �

� &

(77)

and

� �

� � � �

� � �

� &

� (78)

where &

� and &

� denote the degree of the resultinghelicity on small and large scales, respectively. Then, insteadof Eq. (70), we have

&

&

� "

� � � �

! � � �

"

� (79)

which is just &

� &

times the expression on the right handside of Eq. (70). Two things can happen: if the small scalefield is only weakly helical, i.e. &

� , but still &

� , thenthe large scale magnetic energy will be decreased. On theother hand, if the large scale magnetic field is only weaklyhelical, i.e. &

� , but &

� , then the large scale mag-netic energy will actually increase relative to the predictionof Eq. (70). In practice, both effects could happen at the sametime, in which case the change of the large scale magneticenergy will be less strong.

In Fig. 18 we show the results of two runs with frac-tional helicity forcing. In both cases the magnetic energyexceeds the kinetic energy by �

� �

multiplied by an ‘ef-ficiency factor’, &

� &

. In the first case we have �

� �

and &

� &

� � � , giving super-equipartition by about 1.5.This is roughly in agreement with Fig. 18. In the secondcase we have �

� �

� � and &

� &

� � , giving super-equipartition by a factor of about 2.7. Although the run inthe second panel of Fig. 18 has not been run for a full mag-netic diffusion time (which is here ten times longer than inthe first case), it seems clear that strong super-equipartition isstill possible.

In the framework of �

� dynamo theory (Moffatt 1978,Krause & Radler 1980), the large scale field is described byEq. (36). In a periodic domain with minimum wavenumber�

the condition for large scale dynamo action is that the dy-namo number, �

� � � � �

� , exceeds unity. Standard es-timates are � � �

/ � � � � � and �

/ � �

� (e.g., Moffatt

Fig. 19. Evolution of kinetic and magnetic energies in a run with per-fectly conducting boundaries in the � -direction. Note that the mag-netic energy evolves in stages, similar to Fig. 6.

1978), where / is the correlation time. For strongly helicalturbulence with forcing at a particular wavenumber �

, wehave � � � � � � � � &

� �

� , so

� � � � �

� � �

� � � � � �

� �

� �

� &

� 3 �

� (80)

where 3 � � �

� �

is a correction factor involving the mag-netic Reynolds number, defined here as �

� �

� � . For theRuns shown in Fig. 18, and with the numbers given in theprevious paragraph, the values of �

are 1.4 and 2.7, respec-tively. Thus, the dynamo number is supercritical. This is con-sistent with the clear growth of the large scale field seen inthe simulations (Fig. 18). Maron & Blackman (2002) foundthat for �

� �

� the critical value of &

is around 0.5. Thiscorresponds to �

� � 3 � , which should still be su-percritical. However, given that only modest resolution wasused, the estimate for � may have been too optimistic.

4.6. Fractional large scale magnetic helicity

We have seen that the large scale field strength will decreaseif the small scale field is not fully helical. However, accord-ing to Eq. (79), the large scale magnetic field strength shouldincrease if the large scale field were less helical. In order todemonstrate that this is true, we show in Fig. 19 the resultsof a run where the forcing is still 100% helical, but the largescale field is not fully helical. The latter has been achieved byadopting perfectly conducting boundary conditions, in whichcase it is no longer possible to have fully helical Beltramiwaves. These calculations are otherwise similar to those pre-sented in BD01, where the effects of nonperiodic boundaryconditions was considered.

In Fig. 19 we see another remarkable feature that was al-ready seen in Fig. 6: the large scale magnetic energy growsin stages. Indeed, looking at the actual magnetic field patternone sees that the number of nodes in plots of the mean fielddecreases with time, just like in Fig. 6.

4.7. Position of the secondary peak

In the original Run 6 of B01 there was a field of intermediatescale that grew fastest at � � � . This wavenumber was found

Page 19: Magnetic helicity in stellar dynamos: new numerical experiments

A. Brandenburg et al.: Magnetic helicity in stellar dynamos 117

Fig. 20. Scale separation in runs with different resistivity (first andsecond panel). Higher resistivity results in larger scale separation(third panel).

to be in agreement with �

� �

� � � � � � �

� , which is whatone expects for an �

� dynamo. Here �

� � � �

is thesum of microscopic and turbulent magnetic diffusivity and �

represents the � effect that is caused by the helical small scalevelocity. In the simulation, the only parameter that can bechanged to make � � � � � �

� vary is the microscopic magneticReynolds number. In order to check our prediction regarding�

� �

, we reduce the value of � from � � �

� � (as in B01)to � � � �

� � . We find that now the fastest growing modeis at �

� �

� , which is, on a logarithmic scale, almostindistinguishable from the forcing wavenumber, �

� � � ; seeFig. 20. At larger scales, � � � , the exponential growth isequally fast in all modes, independent of the (microscopic)value of � .

The reason for decreasing scale separation with increas-ing value of �

is relatively easy to see. According to thedispersion relation for the �

� dynamo in an infinite domain(e.g., Moffatt 1978) the wavenumber of the fastest growingmode is

� �

� � � � � �

� � � � � 3 �

� &

� 3 (81)

So, unless �

is small, in which case 3 � � �

� �

can belarge, �

� �

will be around �

� giving a scale separation of

1:2. However, for small values of &

, the scale separation canbe much larger.

In the high �

limit, a scale separation of just a factor of2 is hardly visible during the kinematic evolution of the fieldat that scale. Once the dynamo is in the nonlinear regime, i.e.once the small scale field has saturated, � will be quenchedand �

� �

must decrease further.In order to see that the sign of the magnetic helicity is in-

deed different at small and large scales we split the field intoright and left handed parts, � � �

� �

� (see Sect. 2.2 andChristensson, Hindmarsh, & Brandenburg 2001), and showin Fig. 21 the magnetic energy spectra of �

� and �

� , �

and �

, respectively. Inverse transfer is only seen in the�

spectrum; the �

spectrum is dominated by the forc-ing scale.

5. Dynamos with shear

5.1. Evidence for magnetic Reynolds numberdependence

Finally, we turn attention to the case of dynamos with shear.The presence of shear allows for an additional induction ef-fect that is responsible for producing strong toroidal fieldfrom a poloidal (cross-stream) field component. Accordingto a result from mean-field theory such dynamos can pos-sess oscillatory solutions (e.g. Parker 1979). In a recent studyby Brandenburg, Bigazzi, & Subramanian (2001, hereafterBBS01), where a velocity shear of the form 0

� � � � � �

was applied to helically driven turbulence in a periodic box,oscillatory solutions were confirmed also for non-mean fielddynamos. The outstanding question, however, is whether theresistive time scale, which is now known to affect the satura-tion phase of the dynamo (B01, BBS01), also affects the cycleperiod. This question could not be answered conclusively inBBS01, because only one run with one value of the magneticReynolds number was considered.

In the meantime we have accumulated data from anotherrun that had twice the value of the magnetic Reynolds num-ber. The flow is driven by a forcing, which consist of twocomponents: a term varying sinusoidally in the � -directionwith wavenumber � � driving the shear and a term consist-ing of random Beltrami waves with wavenumber �

� . Asin BBS01, the ratio of the magnitudes of the two componentsof the forcing function is about 1:100, resulting in a shear ve-locity that is in the final state about 50 times larger than thepoloidal rms velocity of the turbulence. This is large enoughso that we can expect to be in the so-called ‘ � � regime’where oscillatory solutions are the rule.

Corresponding butterfly diagrams of the toroidally aver-aged toroidal field components are shown in Fig. 22. The twodiagrams are for two different � -positions, � � � � and at� � � , where the shear ) � � 0

! � "

� � � takes on its nega-tive and positive extremum, respectively. The simulation wasrestarted from the run of BBS01 at the time � � " � � � and rununtil � � � � � � with � reduced by a factor of with respectto BBS01. It turns out that the resulting magnetic field suffersa major disturbance after the magnetic diffusivity is reduced

Page 20: Magnetic helicity in stellar dynamos: new numerical experiments

118 Astron. Nachr./AN 323 (2002) 2

Fig. 21. Decomposition of magnetic field into positive and negativehelical parts. Note that the magnetic energy at the forcing scale hasincreased between � � � � � and � � � � � . The magnetic energyat large scales is dominated by helicity with negative sign (dashedcurves).

by just a factor of 2. More importantly, a clear pattern of amigratory dynamo wave is now almost absent. Instead, thetoroidal flux pattern appears to just wobble up and down inthe � direction.

One possible explanation for this result is that the dy-namo wave was just very close to the threshold betweenoscillatory and nonoscillatory behaviour. This is not verylikely however: the estimates of BBS01 indicated that thedynamo numbers based on shear, �

)

� ) � � �

� , is be-tween 40 and 80, whilst the total dynamo number ( � �

)

) is between 10 and 20 (see BBS01), and hence �

� � � �

� � � . Thus, shear dominates strongly over the� -effect ( �

)

� �

is between 150 and 300), which is typicalfor � � -type behaviour (i.e. oscillations) rather than �

� -type

Fig. 22. Butterfly diagram of �

, continuation from the run ofBBS01, but with � � � � � �

� � . Run (iii) of Table 5.

Fig. 23. Butterfly diagram of �

, continuation from the run ofBBS01, but with � � � �

� � (five times larger than in Fig. 22).Run (i) of Table 5.

behaviour which would start when �

)

� �

is below about 10(e.g. Roberts & Stix 1972).

Another possibility is that an oscillatory and migratorydynamo is still possible, but it takes longer for the systemto settle onto this solution. Unfortunately the present simula-tions have become prohibitively expensive in terms of com-puter time that this run cannot be prolonged by much fur-ther at this time. Instead, we have calculated a model withlarger magnetic diffusivity and found cycle periods that arenow somewhat shorter than before ( � � � � � � compared to � � � � � � in BBS02), but in diffusive units they are sim-ilar ( � � � � compared to � � in BBS02); see Fig. 23.This suggests that the period in this oscillatory dynamo is stillcontrolled by the microscopic magnetic diffusivity. A moredetailed comparison of various parameters with the other runpresented above and with that of BBS01 is given in Table 5.

Page 21: Magnetic helicity in stellar dynamos: new numerical experiments

A. Brandenburg et al.: Magnetic helicity in stellar dynamos 119

Table 5. Summary of the main properties of the three-dimensionalsimulations with shear. Here, � � � (

is the magnetic diffusivity inunits of the sounds speed and the wavenumber of the domain, and)

� � �

� � � � �

� � �

is the cycle frequency. In Run (iii) there is no clearcycle visible.

Run (i) (ii) (iii)� � � (

� �

� �

� � � �

� �

� � � �

� �

% � � 5 10 25 �

� � �

� � � �

30 80 200

� � �

� � � �

600 1200 2500

� � �

! &

4 6 20

� � �

! &

20 30 60

� �

� � �

� � � 0.018 0.014 0.008

� � � � � �

� 0.11 0.06 0.014)

� � �

� � � �

8. . . 9 6. . . 12 � � � �

5.2. Can magnetic helicity ride with the dynamo wave?

We now want to address the question whether magnetic he-licity can vary in space and time such that it is actually un-changed in the frame of the dynamo wave as it propagates.The background of these suggestions is as follows. Althoughthe sign of the magnetic helicity in each hemisphere remainsthe same (and is opposite in the other hemisphere), its valuestill changes somewhat as the cycle proceeds. It is conceiv-able that even a small fractional variation may prove incom-patible with the Parker-type migratory dynamo operating on adynamical time scale. One possibility worth checking is thatthe magnetic helicity pattern could propagate together withthe dynamo wave. In that case the magnetic helicity wouldbe conserved in a Lagrangian frame moving with the wave,but not in an Eulerian frame. If this transport is due to ad-vection, we would then expect systematic sign changes of 0

in a space-time diagram, which does not seem to be the casehowever; see Fig. 24.

What does seem interesting, however, is the presence ofwhat is usually (in the context of solar physics) referred toas a ‘torsional’ oscillation (Howard & LaBonte 1980). Thetoroidal flow pattern clearly traces the toroidal magnetic fieldpattern. More recently Howe et al. (2000) found anothershorter period of 1.5 years which has been seen in helioseis-mological data. If this can indeed also be interpreted as theresult of a similar magnetic field pattern (which is as yet un-detected) then this would suggest the presence of multiple pe-riods in the solar dynamo wave (cf. Covas, Tavakol, & Moss2001).

In the discussion above we had mainly bulk motions inmind that would result in a transport of magnetic helicity.This does not need to be the case because magnetic helicitycan also be transported along twisting magnetic field lines.This was used in the approach by Vishniac & Cho (2001),who proposed a dynamo effect based on a non-vanishing di-vergence of the magnetic helicity flux. If magnetic helicitywere to ride with the dynamo wave, then this could perhapscorrespond to the anticipated magnetic helicity flux. Unfor-tunately, in the parameter regime currently accessible to sim-

Fig. 24. Butterfly diagram of �

for the run of BBS01 with � � � �

� �

� � , compared with the corresponding diagrams of &

and &

forthe same time interval. Note the presence of a ‘torsional’ oscillationin &

, i.e. a superposition of the driven shear flow (second panel),but the absence of any similar pattern in &

(last panel). Run (ii) ofTable 5.

ulations the effect proposed by Vishniac & Cho has not yetbeen confirmed (Arlt & Brandenburg 2001). We may there-fore conclude that transport or advection of magnetic helicityby meridional flows remains a possibility, but has not as yetbeen verified numerically.

5.3. Helicity loss in the presence of surface shear

The calculations of BD01 had the shortcoming that largescale shear was absent. Of course, shear does not producenet magnetic helicity, but it can lead to a spatial separationwhich could be particularly useful if there are surfaces. Largescale shear in the equatorial plane would wind up a poloidalfield and hence would lead to magnetic helicity of oppositesign in the northern and southern hemispheres. In helicallyforced turbulence simulations, the various effects have onlybeen studied in isolation: shear and helicity in BBS01, sur-face shear but no helicity in Arlt & Brandenburg (2001), anddifferent helicities in the northern and southern hemispheresin Brandenburg (2001b).

Surface shear may allow for the direct production andtransport of magnetic helicity of different signs in the twohemispheres on a dynamical time scale. Consider a differ-entially rotating star permeated by an initially uniform mag-netic field parallel to the rotation axis. The differential rota-tion will wind up the magnetic field in different directions inthe two hemispheres, where the field lines describe oppositelyoriented screws. Such additional production and transport ofmagnetic helicity on a dynamical time scale may be impor-tant for the magnetic helicity problem. On the other hand, forthis idea to be relevant we have to have a pre-existing large

Page 22: Magnetic helicity in stellar dynamos: new numerical experiments

120 Astron. Nachr./AN 323 (2002) 2

scale poloidal magnetic field, so this issue does not seem toaddress the problem of resistively limited field saturation.

In order to find out whether large scale surface shear isimportant we have considered a model similar to Run C2 ofArlt & Brandenburg (2001), i.e. with finite shear in a layerof thickness � and a quiescent (non-turbulent) halo with noshear outside. As before, the shear is given by 0

� � � , whilethe boundary to the halo is at � � � . However, in con-trast to the nonhelical calculations of Arlt & Brandenburg,we have now included a helical forcing that is negative (pos-itive) above (below) the midplane. The results are in manyrespects similar to the runs of BD01 with a halo: around thetime of saturation there is a strong negative burst of smallscale magnetic helicity together with positive small scale cur-rent helicity. The resulting magnetic helicity fluxes both outof the turbulent domain as well as upwards on the two bound-aries are fluctuating about zero. Thus, the addition of shear tothe halo runs of BD01 seems to have only little effect on theevolution of helicity fluxes.

The evolution of the magnetic and kinetic energies isshown in Fig. 25. After about 0.08 resistive times the dynamohas entered the saturation phase which is then completed af-ter 0.15 resistive times altogether. This is now much fasterthan in the case of periodic boundaries. Fig. 26 shows imagesof the mean toroidal magnetic field. Especially near the lowerboundary of the disc surface (dash-dotted line) one sees theejection of magnetic structures.

As can be seen in Fig. 27, the magnetic field structure ismarkedly different at � � � compared to � � � , whereit is stronger and pushed to the disc boundaries. This fieldexpulsion to the boundaries could well be the result of usualfield migration that is expected in the presence of shear andhelicity. In terms of mean-field theory with positive � in theupper disc plane and negative below, we expect field propaga-tion away from (towards) the midplane when shear is positive(negative). This is indeed consistent with our model whereshear is positive (negative) at � � � ( � � � ).

6. Conclusions

In an attempt to clarify the problem of resistive versus dy-namical time scales in models of the solar dynamo cyclewe have tied up a number of loose ends that had been leftopen after the first exploratory simulations of B01, BD01,and BBS01. There are two distinct issues that may or maynot be controlled by the resistive time: the time it takes forthe large scale field to reach saturation and, once saturationis reached, the time scale on which the the large scale fieldcan undergo cyclic variations as seen in the sun.

Looking back at Fig. 1 in the introduction we can still ex-pect various approaches to be successful, although many oth-ers now seem to have been eliminated. One possibility wasthat the dynamo may still operate on a fast time scale if, likein the kinematic case, the wavelength of the dynamo wave isshorter than the extent of the system, and possibly only some-what larger than the scale of the forcing. The main problemwould then be to keep the process fast and to prevent the dy-namo from developing scales as large as the box size. How

Fig. 25. Semi-logarithmic and double-linear plots showing the evo-lution of kinetic and magnetic energies (dotted and solid lines),compared with the magnetic energy of the azimuthally averagedmean field. All quantities are normalised by the kinetic energy ofthe cross-stream motions. *

) � � *

� � �

refers to the total kinetic energyincluding the shear. The ratio of the kinematic growth rate to theresistive time at large scales is here ! � � �

� � � � .

Fig. 27. Butterfly diagram of the mean toroidal field of Fig. 26 at� � � � and � � � . The lines at � � � � mark the boundaries ofthe turbulent slab.

this may happen in reality is not clear. It may well be con-nected to the geometry of astrophysical dynamos which havespherical or disc geometry and are not box-like. This corre-sponds to Alternative A in Fig. 1. The argument against thispossibility is that even at intermediate scales the evolution to-

Page 23: Magnetic helicity in stellar dynamos: new numerical experiments

A. Brandenburg et al.: Magnetic helicity in stellar dynamos 121

Fig. 26. Loss of magnetic field through the upper and lower boundaries. Images of the mean toroidal field are shown at different times.Shear is positive (negative) at � � � ( � � � � ). � � � �

� � and % � � � � �

� � . The lines at � � � � mark the boundaries of the turbulentslab.

wards larger scales (i.e. the motion of the secondary bump inthe spectrum, cf. Figs 10 and 11) seems to occur on a resistivetime scale. This can be seen from the reduction of the speedof the spectral bump, �

� � � �

, as the resistivity is reduced (Ta-ble 1). So the forcing scale may have to be not much smallerthan the scale of the dynamo wave. Other plausible mecha-nisms could be based on non- � effect dynamos. Examplesare negative magnetic diffusivity effects (cf. Zheligovsky etal. 2001), the incoherent � -effect (Vishniac & Brandenburg1997), and the Vishniac–Cho effect (Vishniac & Cho 2001).These are all subsumed under Alternative B in Fig. 1.

The remaining alternatives are all based on conventional� effect dynamos, so they have finite net magnetic helicity(although of opposite sign in the two hemispheres). In or-der for the dynamo to operate fast enough, resistive magneticreconnection (with magnetic helicity destruction) has to befast enough. This could only be the case near the surfacewhere the microscopic magnetic diffusivity is sufficientlylarge. This corresponds to Alternative C in Fig. 1.

The other possibility is that the dynamo may workthrough losses of magnetic helicity of opposite sign, possi-bly at intermediate scales, for example near the surface (e.g.in coronal mass ejections and active regions) or across theequator. This is referred to as Alternative D. The simulationspresented in Sect. 4.3 have shown that this is in principle pos-sible. The problem here is that, so far, none of these mecha-nisms have been seen to occur naturally in any of the sim-ulations investigated so far. It is therefore important to movetoward more realistic simulations, perhaps in spherical geom-etry, with strong vertical stratification which could contributeto keeping the large scale field within the convection zone viaturbulent pumping, but allowing fields above a certain thresh-old to escape through the surface.

The purpose of this review was to outline the present sta-tus of our understanding of nonlinear large scale dynamos.

Things are developing rapidly with new simulations appear-ing every month. So far it has mainly been a process of elim-ination, because many different ideas have been around, andconstantly more ideas are appearing. The least explored caseis that of dynamos in realistic spherical geometries.

Although a number of direct simulations of MHD tur-bulence in spherical geometry have been analysed (Gilman1983, Glatzmaier 1985, Glatzmaier & Roberts 1995, Drecker,Rudiger & Hollerbach 2000, Ishihara & Kida 2000) the issueof fast versus resistively limited growth has not yet been in-vestigated. This is indeed not an easy task. In order to distin-guish between the two possibilities, one would need to verifyexplicitly whether a large scale field is generated on resis-tive times, or on a time scale that is independent of and muchshorter than the resistive time scale. In addition, if the modelsshow cyclic variability (which is desirable for solar dynamomodels), the cycle length needs to be (asymptotically) inde-pendent of the resistive time if the dynamo is fast. One reasonwhy such numerical experiments are difficult is that only atrelatively large magnetic Reynolds numbers (which requirehigh numerical resolution) the resistive time scale becomessufficiently long so that it can clearly be distinguished fromthe dynamical time scale. A relevant dynamical time scale isthe ratio $ � � � �

� which has to exceed a value of around 20–30 before one can see that resistively limited growth or cycleperiods have occurred (see B01). At the same time the resolu-tion must not be too high because otherwise one may not beable to run the simulations for long enough before any largescale field has occurred. The same is true of hyperresistivitywhich tends to make the large scale resistive time scale ex-tremely long so that nothing can be said about resistively lim-ited growth. This explains why the issue of resistively limitedgrowth is not yet well understood in more realistic geome-tries. Nevertheless, it would be worthwhile reanalysing datafrom recent, high-resolution dynamo simulations in the light

Page 24: Magnetic helicity in stellar dynamos: new numerical experiments

122 Astron. Nachr./AN 323 (2002) 2

of resistive limitations on the duration of the saturation phaseand, if applicable, the cycle period.

Acknowledgements. Use of the PPARC supported supercomputersin St Andrews and Leicester (UKAFF) is acknowledged. KS thanksNordita for hospitality when this work was begun. This work wassupported in part by the Leverhulme Trust (Grant F/125/AL). Weacknowledge the hospitality of the Institute for Theoretical Physicsat the University of California, Santa Barbara, where this work wascompleted. This research was supported in part by the National Sci-ence Foundation under Grant No. PHY99-07949

References

Arlt, R., Brandenburg, A.: 2001, A&A 380, 359Bao, S.D., Zhang, H.Q., Ai, G.X., Zhang, M.: 1999, A&A 139, 311Batchelor, G.K.: 1950, Proc. Roy. Soc. Lond. A201, 405Bellan, P.M.: 1999, in: M.R. Brown, R.C. Canfield, A.A. Pevtsov

(eds.), Magnetic Helicity in Space and Laboratory Plasmas,American Geophys. Union, Geophys. Monograph Vol. 111,p. 119

Berger, M.: 1984, GApFD 30, 79Berger, M., Field, G.B.: 1984, JFM 147, 133Berger, M.A., Ruzmaikin, A.: 2000, JGR 105, 10481Blackman, E.G., Field, G.F.: 2000, ApJ 534, 984Brandenburg, A.: 1998, in: M.A. Abramowicz, G. Bjornsson, J.E.

Pringle (eds.), Theory of Black Hole Accretion Discs, Cam-bridge University Press, p. 61

Brandenburg, A.: 2001a, ApJ 550, 824 (B01)Brandenburg, A.: 2001b, in: P. Chossat, D. Armbruster, O. Iu-

liana (eds.), Dynamo and dynamics, a mathematical challenge,Kluwer Acad. Publ., NATO ASI Series Vol. 26, p. 125 (astro-ph/0012112)

Brandenburg, A.: 2001c, in: P. Brekke, B. Fleck, J.B. Gurman(eds.), Recent Insights into the Physics of the Sun and Helio-sphere, IAU Symp. No. 203, ASP Conf. Ser., in press (astro-ph/0011579)

Brandenburg, A., Dobler, W.: 2001, A&A 369, 329 (BD01)Brandenburg, A., Sarson, G.R.: 2002, PRL 88, 055003 (BS02)Brandenburg, A., Schmitt, D.: 1998, A&A 338, L55Brandenburg, A., Subramanian, K.: 2000, A&A 361, L33 (BS00)Brandenburg, A., Bigazzi, A., Subramanian, K.: 2001, MNRAS

325, 685 (BBS01)Brandenburg, A., Saar, S.H., Turpin, C.R.: 1998, ApJ 498, L51Brandenburg, A., Nordlund, A., Stein, R.F., Torkelsson, U.: 1995,

ApJ 446, 741Brandenburg, A., Jennings, R.L., Nordlund, A., Rieutord, M., Stein,

R.F., Tuominen, I.: 1996, JFM 306, 325Brown, M.R., Canfield, R.C., Pevtsov, A.A. (eds.): 1999, Magnetic

Helicity in Space and Laboratory Plasmas, American Geophys.Union, Geophys. Monograph Vol. 111

Cattaneo, F., Hughes, D.W.: 1996, PhRvE 54, R4532Cattaneo, F., Vainshtein, S.I.: 1991, ApJ 376, L21Chae, J.: 2000, ApJ 540, L115Choudhuri, A.R., Schussler, M., Dikpati, M.: 1995, A&A 303, L29Christensson, M., Hindmarsh, M., Brandenburg, A.: 2001, PhRvE

64, 056405Covas, E., Tavakol, R., Moss, D.: 2001, A&A 371, 718

Cowling, T.G.: 1934, MNRAS 94, 39DeVore, C.R.: 2000, ApJ 539, 944Dikpati, M., Charbonneau, P.: 1999, ApJ 518, 508Drecker, A., Rudiger, G., Hollerbach, R.: 2000, MNRAS 317, 45Durney, B.R.: 1995, SoPh 166, 231Ferriz-Mas, A., Schmitt, D., Schussler, M.: 1994, A&A 289, 949Field, G.B., Blackman, E.G.: 2002, astro-ph/0111470Finn, J., Antonsen, T.M.: 1985, Comments Plasma Phys. Controlled

Fusion, 9, 111Frisch, U., Pouquet, A., Leorat, J., Mazure, A.: 1975, JFM 68, 769Galsgaard, K., Nordlund, A.: 1996, JGR 101, 13445Gilbert, A.D.: 2002, GApFD, in pressGilbert, A.D., Frisch, U., Pouquet, A.: 1988, GApFD 42, 151Gilman, P.A.: 1983, ApJS 53, 243Glatzmaier, G.A.: 1985, ApJ 291, 300Glatzmaier, G.A., Roberts, P.H.: 1995, Nature 377, 203Hawley, J.F., Gammie, C.F., Balbus, S.A.: 1996, ApJ 464, 690Howard, R., LaBonte, B.J.: 1980, ApJ 239, L33Howe, R., Christensen-Dalsgaard, J., Hill, F., Komm, R.W., Larsen,

R.M., Schou, J., Thompson, M.J., Toomre, J.: 2000, ApJ 533,L163

Hughes, D.W., Cattaneo, F., Kim, E.J.: 1996, PhL 223, 167Ishihara, N., Kida, S.: 2000, PSJaJ 69, 1582Ji, H., Prager, S.C.: 2001, Magnetohydrodynamics, in press (astro-

ph/0110352)Kleeorin, N.I., Ruzmaikin, A.A.: 1982, Magnetohydrodynamics 2,

17Krause F., Radler K.-H.: 1980, Mean-Field Magnetohydrodynamics

and Dynamo Theory, Pergamon Press, OxfordKuker, M., Rudiger, G., Schultz, M.: 2001, A&A 374, 301Kulsrud, R.M., Anderson, S.W.: 1992, ApJ 396, 606Leighton, R.B.: 1969, ApJ 156, 1Maron, J., Blackman, E.G.: 2002, ApJ, 566, L41Moffatt, H.K.: 1978, Magnetic Field Generation in Electrically

Conducting Fluids, Cambridge Univ. Press, CambridgeParker, E.N.: 1979, Cosmical Magnetic Fields, Clarendon Press, Ox-

fordPevtsov, A.A., Latushko, S.M.: 2000, ApJ 528, 999Pevtsov, A.A., Canfield, R.C., Metcalf, T.R.: 1995, ApJ 440, L109Pouquet, A., Frisch, U., Leorat, J.: 1976, JFM 77, 321Roberts, P.H., Stix, M.: 1972, A&A 18, 453Rudiger, G.: 1989, Differential rotation and stellar convection: Sun

and solar-type stars, Gordon Breach, New YorkSeehafer, N.: 1990, SoPh 125, 219Steenbeck, M., Krause, F.: 1969, AN 291, 49Steenbeck, M., Krause, F., Radler, K.-H.: 1966, Z. Naturforsch. 21,

369Stix, M.: 1974, A&A 37, 121Stix, M.: 1976, A&A 47, 243Stone, J.M., Hawley, J.F., Gammie, C.F., Balbus, S.A.: 1996, ApJ

463, 656Vainshtein, S.I., Cattaneo, F.: 1992, ApJ 393, 165Verma, M.K.: 2001 (nlin.CD/0103033)Vishniac, E.T., Brandenburg, A.: 1997, ApJ 475, 263Vishniac, E.T., Cho, J.: 2001, ApJ 550, 752Zeldovich, Ya.B., Ruzmaikin, A.A., Sokoloff, D.D.: 1983, Magnetic

Fields in Astrophysics, Gordon Breach, New YorkZheligovsky, V.A., Podvigina, O.M., Frisch, U.: 2001, GApFD, in

press (nlin.CD/0012005)


Recommended