19
Mon. Not. R. Astron. Soc. 315, 165–183 (2000) Non-linear av -dynamos driven by magnetic buoyancy J.-C. Thelen w ² Department of Applied Mathematics, University of Leeds, Leeds LS2 9JT Accepted 2000 January 17. Received 1999 October 4; in original form 1999 April 8 ABSTRACT It is now widely accepted that the large-scale solar magnetic field is generated by some kind of non-linear av -dynamo. However, these dynamos have the disadvantage that the non- linearities, and in particular the a -effect, are chosen in an ad hoc fashion and are only vaguely related to the underlying physical processes. Here, on the other hand, an av -dynamo with an a -effect based on magnetic buoyancy instabilities is described. This a -effect has one main advantage over previous descriptions, namely that it is derived from a numerical model of the physical processes that are thought to occur at the base of the convection zone. In this paper we investigate a one-dimensional and a two-dimensional av -dynamo in a cylindrical and a spherical shell, respectively. In both models the v -effect is described by a simple shear flow, while a is proportional to the radial gradient of the magnetic field because, simply speaking, this gradient determines when magnetic buoyancy instabilities occur and hence when the a -effect sets in. Key words: instabilities – MHD – Sun: magnetic fields – sunspots. 1 INTRODUCTION Lower main sequence stars display a cyclic magnetic activity, which is characteristic of stars with a deep convection zone. The degree of magnetic activity depends on both the angular velocity and the spectral type of the star and it is considered to be a function of the inverse Rossby number s Vt c ; where V is the angular velocity and t c is a suitable convective time-scale (Noyes, Weiss & Vaughan 1984). The most active stars are young stars with very short rotation rates of the order of a day or less. As the star becomes older its rotation rate decreases owing to magnetic braking, i.e. owing to the torque exerted by the stellar wind, and the star becomes less active. The Sun, which concerns us here, at 4:5 10 9 yr; is a weakly active middle aged star with a magnetic cycle which oscillates irregularly with a mean period of 22 yr, taking into account the reversal of the polarity of the magnetic field (Weiss 1993). It is widely believed that the magnetic cycle is generated by some type of hydrodynamic dynamo. The most common approach to study such a dynamo is to make use of mean-field dynamo theory. Here, the magnetic field is split into its mean and fluctuating parts and an equation for the mean part is derived by parametrizing the interaction between the small-scale magnetic field and the turbulent velocity field. The most important parameter is the a -effect which relates the mean electric current to the mean magnetic field. In most mean-field dynamos the a -effect is thought to be the result of helical motions resulting from the interactions of the convective turbulence and rotation (Parker 1955). The first kinematic mean-field dynamo was developed by Steenbeck & Krause (1969) and their simple model already agreed with some of the solar observations such as the butterfly diagram and Hale’s polarity laws. Similar models were investigated by Roberts (1972), Jepps (1975) and more recent kinematic models include those by Schmitt (1987), Prautzsch (1993) and Parker (1993). However, all the above mentioned kinematic models have the disadvantage that the growth of the magnetic field is unlimited. In order to inhibit this unlimited growth, it is necessary to include non-linear effects into the basic dynamo equations. Several field limiting mechanisms have been proposed so far. First, a -quenching, which is in essence an arbitrary non-linearity, based on the plausible physical argument that a strong magnetic field exhibits a strong resistance to deformations by small-scale motions. The arbitrari- ness results from the fact that the exact physics of the solar interior, such as the flux tubes and the form of convection, are not known and that therefore it is necessary to rely on a parametriza- tion. Dynamo models incorporating this type of non-linearity have been investigated by Brandenburg et al. (1989a), Brandenburg, Tuominene & Moss (1989b), Covas et al. (1997a,b), Jennings (1991) and Jennings & Weiss (1991). Secondly, the effectiveness of the v -effect can be reduced, either by v -quenching or by the ‘Malkus–Proctor effect’. v -quenching, on the one hand, is an arbitrary parametrization of the effect of the mean field on the q 2000 RAS w Current address: Department of Astronomy and Astrophysics, University of Chicago, 5640 S. Ellis, Chicago, IL, 60637, USA. ² E-mail: [email protected]

Non-linear αω-dynamos driven by magnetic buoyancy

Embed Size (px)

Citation preview

Mon. Not. R. Astron. Soc. 315, 165±183 (2000)

Non-linear av-dynamos driven by magnetic buoyancy

J.-C. Thelenw²Department of Applied Mathematics, University of Leeds, Leeds LS2 9JT

Accepted 2000 January 17. Received 1999 October 4; in original form 1999 April 8

A B S T R A C T

It is now widely accepted that the large-scale solar magnetic field is generated by some kind

of non-linear av-dynamo. However, these dynamos have the disadvantage that the non-

linearities, and in particular the a -effect, are chosen in an ad hoc fashion and are only

vaguely related to the underlying physical processes.

Here, on the other hand, an av -dynamo with an a -effect based on magnetic buoyancy

instabilities is described. This a -effect has one main advantage over previous descriptions,

namely that it is derived from a numerical model of the physical processes that are thought

to occur at the base of the convection zone.

In this paper we investigate a one-dimensional and a two-dimensional av -dynamo in a

cylindrical and a spherical shell, respectively. In both models the v -effect is described by a

simple shear flow, while a is proportional to the radial gradient of the magnetic field

because, simply speaking, this gradient determines when magnetic buoyancy instabilities

occur and hence when the a -effect sets in.

Key words: instabilities ± MHD ± Sun: magnetic fields ± sunspots.

1 I N T R O D U C T I O N

Lower main sequence stars display a cyclic magnetic activity,

which is characteristic of stars with a deep convection zone. The

degree of magnetic activity depends on both the angular velocity

and the spectral type of the star and it is considered to be a

function of the inverse Rossby number s � Vtc; where V is the

angular velocity and tc is a suitable convective time-scale (Noyes,

Weiss & Vaughan 1984). The most active stars are young stars

with very short rotation rates of the order of a day or less. As the

star becomes older its rotation rate decreases owing to magnetic

braking, i.e. owing to the torque exerted by the stellar wind, and

the star becomes less active. The Sun, which concerns us here, at

4:5 � 109 yr; is a weakly active middle aged star with a magnetic

cycle which oscillates irregularly with a mean period of 22 yr,

taking into account the reversal of the polarity of the magnetic

field (Weiss 1993).

It is widely believed that the magnetic cycle is generated by

some type of hydrodynamic dynamo. The most common approach

to study such a dynamo is to make use of mean-field dynamo

theory. Here, the magnetic field is split into its mean and

fluctuating parts and an equation for the mean part is derived by

parametrizing the interaction between the small-scale magnetic

field and the turbulent velocity field. The most important

parameter is the a -effect which relates the mean electric

current to the mean magnetic field. In most mean-field dynamos

the a -effect is thought to be the result of helical motions resulting

from the interactions of the convective turbulence and rotation

(Parker 1955).

The first kinematic mean-field dynamo was developed by

Steenbeck & Krause (1969) and their simple model already agreed

with some of the solar observations such as the butterfly diagram

and Hale's polarity laws. Similar models were investigated by

Roberts (1972), Jepps (1975) and more recent kinematic models

include those by Schmitt (1987), Prautzsch (1993) and Parker

(1993).

However, all the above mentioned kinematic models have the

disadvantage that the growth of the magnetic field is unlimited. In

order to inhibit this unlimited growth, it is necessary to include

non-linear effects into the basic dynamo equations. Several field

limiting mechanisms have been proposed so far. First, a-quenching,

which is in essence an arbitrary non-linearity, based on the plausible

physical argument that a strong magnetic field exhibits a strong

resistance to deformations by small-scale motions. The arbitrari-

ness results from the fact that the exact physics of the solar

interior, such as the flux tubes and the form of convection, are not

known and that therefore it is necessary to rely on a parametriza-

tion. Dynamo models incorporating this type of non-linearity have

been investigated by Brandenburg et al. (1989a), Brandenburg,

Tuominene & Moss (1989b), Covas et al. (1997a,b), Jennings

(1991) and Jennings & Weiss (1991). Secondly, the effectiveness

of the v-effect can be reduced, either by v-quenching or by the

`Malkus±Proctor effect'. v-quenching, on the one hand, is an

arbitrary parametrization of the effect of the mean field on the

q 2000 RAS

w Current address: Department of Astronomy and Astrophysics, University

of Chicago, 5640 S. Ellis, Chicago, IL, 60637, USA.

² E-mail: [email protected]

differential rotation via either small scales or large scales

(Brandenburg et al. 1991; Jennings 1991; Jennings & Weiss

1991). The `Malkus±Proctor effect', on the other hand, does not

involve any parametrization and takes into account the effect of

the large-scale magnetic field on the driving velocity via the

Lorentz force (Malkus & Proctor 1975; Brandenburg et al. 1989a;

Schmitt & SchuÈssler 1989; Belvedere, Pitadella & Proctor 1990;

Tobias 1996, 1997). The final field limiting mechanism that has

been proposed is flux escape by magnetic buoyancy. This process

is based on the fact that the magnetic flux will eventually leave the

regions of field amplification owing to buoyancy and related

instabilities and hence will no longer contribute to the generation

of the magnetic field. (Schmitt & SchuÈssler 1989; Moss et al.

1990; Jennings 1991). An interesting model has been constructed

by Ferriz-Mas, Schmitt & SchuÈssler (1994) in which the dynamo

is driven by a non-axisymmetric instability of toroidal magnetic

flux tubes. The non-linear effect which finally limits the amplitude

of the dynamo results in a natural way from the mechanism which

drives the dynamo, namely the eruption of magnetic flux tubes

from the dynamo region.

An important question concerning these dynamo models is that

of their location. The assumption that the a -effect results from the

interaction of turbulent convection and rotation implies that the

dynamo is situated in the solar convection zone. However, there

are several problems associated with this idea. The most important

one concerns the difficulty of storing the magnetic flux in the

convection zone for periods comparable to the solar cycle.

Moreover, from self-consistent convection zone dynamos it emerges

that the toroidal magnetic field belts move toward the poles,

instead of towards the equator, for a surface differential rotation

which is consistent with that of the Sun (Gilman & Miller 1981;

Gilman 1983; Glatzmaier 1984, 1985). Therefore, for a variety of

reasons, based on results from magneto-convection (Spiegel &

Weiss 1980; Galloway & Weiss 1981), self-consistent dynamo

calculations of Gilman and Glatzmaier and helioseismology

(Brown et al. 1989), it is nowadays assumed that the dynamo is

located in the overshoot zone, a thin layer, situated between the

convection zone and the radiative zone. Locating the dynamo and

hence the magnetic field into the overshoot zone, nevertheless,

gives rise to another problem. As was mentioned before, dynamo

action in the convection zone relies on cyclonic motions, which

are caused by the interaction of convection and rotation. This

implies that in the overshoot zone the dynamo would be very

weak, or would even fail to operate, because of its subadiabatic

stratification. There are, however, several ways around this

problem. First, it is possible to split the dynamo into two parts,

with the v-effect and the a -effect confined to half space z , 0

and z . 0; representing the overshoot zone and the convection

zone, respectively. Such a dynamo has been investigated by Parker

(1993) in the kinematic regime and Tobias (1996, 1997) in the

non-linear regime. Secondly, convective motions are not neces-

sarily needed in order to obtain an a -effect. It has been shown by

Moffatt (1978), Schmitt (1985) and Thelen (1998, 2000) that a

mean-electromotive force, and hence an a -effect, can instead be

generated by helical waves resulting from magnetic buoyancy

instabilities. Schmitt (1987) and Prautzsch (1993) investigated

dynamo models in which the generation of the toroidal magnetic

field results from the induction effect of unstable magnetostrophic

waves. Their a-effect has the following latitudinal dependence: it

is negative near the equator and changes sign at mid-latitude to

become positive near the pole (Schmitt 1985). However, it was

shown by Thelen (1998, 2000) that this latitudinal dependence of

the a -effect is only valid in the case where the viscosity and the

magnetic diffusivity are neglected. If, on the other hand, viscous

effects are taken into account the latitudinal dependence can no

longer be determined, but some important insights into the radial

dependence of a can still be gained.

In this paper we therefore propose an av-dynamo that operates

in the overshoot zone with an a -effect based on magnetic

buoyancy instability. In particular, a is assumed to be proportional

to the radial gradient of the toroidal magnetic field since the

simplest criterion for magnetic buoyancy instabilities is that the

toroidal field strength decreases with height. In order to avoid any

unnecessary parametrizations the field growth is limited by the

`Malkus±Proctor' effect. Such a dynamo can operate in the over-

shoot zone without any difficulties since it does not depend on

turbulent convection. Moreover, this a has an advantage over

previous prescriptions in the sense that it is derived from a

numerical model of the physical processes which are thought to

occur in the overshoot zone. It should also be noted that the

cylindrical dynamo model has been chosen for simplicity since its

aim is not to obtain an accurate model but to show that magnetic

buoyancy instabilities are capable of a dynamo and to gain a basic

understanding of the underlying physical processes.

This paper is organized as follows. In Section 2 we describe

both the cylindrical and the spherical model, present the governing

equations and briefly explain the form of the a -effect. The results

for the cylindrical and the spherical dynamo models will be

presented in Sections 3 and 4, respectively. Finally in Section 5 a

conclusion is given.

2 M AT H E M AT I C A L F O R M U L AT I O N

2.1 The cylindrical shell model

Here, the overshoot zone is modelled by a cylindrical shell,

containing an electrically conducting and incompressible plasma,

with inner radius r0 and outer radius r1. Using mean-field dynamo

theory, the velocity field and the magnetic field are assumed to

have a small-scale structure. All quantities are assumed to be

axisymmetric, i.e. independent of f . Furthermore, the velocity is

assumed to be purely toroidal so that there is no meridional flow.

Initially, there is no toroidal magnetic field and the cylindrical

shell contains only a weak poloidal seed field. This poloidal seed

field is stretched into a field with a toroidal component by the

imposed differential rotation. This constitutes the v-effect which

is modelled by a simple, prescribed, radial shear flow, given by:

vpres�r� � Vpres�r�rf : �1�The a -effect which regenerates the poloidal field component from

a toroidal field component results from magnetic buoyancy

instabilities and is made to depend on the radial gradient of the

toroidal magnetic field strength. This is a reasonable assumption

since it has been shown that the interaction of magnetic buoyancy

instabilities and rotation gives rise to an a -effect (Moffatt 1978;

Schmitt 1985; Thelen 1998, 2000). Moreover, a simple criterion

for magnetic buoyancy instability is given by

2­BT

­r. 0 �2�

since, simply speaking, all that is required for magnetic buoyancy

instability to set in is a decrease with height of the magnetic field.

Furthermore, it was also shown that a is largest at the upper

interface between the magnetic layer and the non-magnetic layer,

166 J.-C. Thelen

q 2000 RAS, MNRAS 315, 165±183

i.e. in the region where the magnetic field falls off rapidly with

height. Hence, here a is assumed to be proportional to the

gradient of the toroidal magnetic field. Thus we choose to model

a (r, t) by

a�r� � 2a0

­jBTj­r

; if­jBTj

­r, 0;

0; otherwise:

8<: �3�

From (3) it follows that initially a is zero since there is no toroidal

field and, assuming that the toroidal field BT is a smooth function

of r, a weak toroidal field corresponds to a weak a . This is

different from most av-dynamos where a (r) is always non-zero

and a is largest when the magnetic field is smallest.

The dynamical effect of the magnetic field is chosen to come

into play through the `Malkus±Proctor' effect, i.e. through its

effect on the f -component of the velocity (Malkus & Proctor

1975). The Lorentz force, which gives the back-reaction of the field

on the flow, induces a large-scale toroidal velocity vind subject to

viscous damping, in the opposite direction of the prescribed

azimuthal velocity. The total azimuthal velocity is then given by:

V tot � vpres � vind: �4�The Lorentz force and viscous damping are then the only terms to

appear in the f-component of the momentum equation. The

magnetic pressure and hence the gas pressure drop out owing to

axisymmetry and gravity only acts in the radial direction.

Similarly the inertial term vanishes because of axisymmetry and

because the flow is purely azimuthal.

This system is governed by the mean-field dynamo equations

­BT

­t� �72 2 r22�BT � r�BP ´ 7�V; �5�

­A

­t� �72 2 r22�A� DaBT �6�

and the momentum equation

­vind

­t� PM�72 2 r22�vind � �BP ´ 7�BT; �7�

where D � a0V00d3=h2 is the dynamo number and PM � nT=h is

the magnetic Prandtl number. Here BT and BP denote the toroidal

magnetic field and poloidal magnetic field, respectively. Equa-

tions (5)±(7) have been non-dimensionalized by introducing the

following variables:

r � r*d; t � t*d2

h; Bi � Bi*

d

h

����������m0r0

p; �8�

where d is the layer depth and d2/h is the magnetic diffusion time.

In order to reduce the system (5)±(7) to one dimension only one

Fourier mode is retained and the solutions A and BT are assumed

to be of the form:

A�r; z; t� � R�A�r; t� exp�ikz��; �9�BT�r; z; t� � R�B�r; t� exp�ikz��; �10�where Aà and Bà are complex. Note that for such a magnetic field

the Lorentz force, and therefore the induced velocity will not be

z-independent. In order to eliminate the z-dependence two

different approaches have been used. First, the induced velocity

is assumed to be independent of z, i.e. the Lorentz force has

been truncated at zeroth order in the z-direction. This is the

equivalent of only retaining the z-averaged component of the

Lorentz force. Thus after substituting (9) and (10) into (5)±(7),

and dropping `hats' for convenience, the following system of

partial differential equations is obtained:

­B

­t� �72 2 r22 2 k2�B 2 irk

­V

­rA; �11�

­A

­t� �72 2 r22 2 k2�A� DaB; �12�

­vind

­t� PM�72 2 r22�vind � ik

4

­

­r�A*B 2 AB*�

� �� ik

2r�A*B 2 AB*�: �13�

Secondly, a higher order system is obtained by retaining the second-

order terms of the Lorentz force and by truncating the other non-

linear terms at the third order in z. This gives rise to an additional

equation for the induced velocity. Proceeding as above we obtain:

­B

­t� �72 2 r22 2 k2�B 2 irk

­V

­rA� irkv

­A*

­r

� ikvA*� irk

2

­v

­rA*; �14�

­A

­t� �72 2 r22 2 k2�A� DaB; �15�

­vind

­t� PM�72 2 r22�vind � ik

4

­

­r�A*B 2 AB*�

� �� ik

2r�A*B 2 AB*� �16�

and

­wind

­t� PM�72 2 r22 2 4k2�wind � ik

2

­A

­rB 2

­B

­rA

� �: �17�

Here `*' denotes the complex conjugate, vind is real and wind is

complex. The two systems (11)±(13) and (14)±(17) will

subsequently be studied. Here V is again given by:

V � Vpres � vind=r; �18�while v is given by

v � wind=r; �19�where vind is the induced velocity resulting from the zeroth-order

terms in the Lorentz force and wind is the induced velocity

resulting from to the second-order terms. These two systems are

then solved subject to the following boundary conditions. First, it

is assumed that the base of the overshoot zone is a perfect

conductor. Hence the radial component of the magnetic field and

the tangential component of the electric field vanish there, i.e.

A � 0 and­Br

­r� 0 at r � r0: �20�

Secondly, the top boundary of the overshoot zone is assumed to be

perfectly insulating. Hence there are no currents in the region above

our layer which implies that the toroidal magnetic field has to

vanish and that the field matches a potential field at the surface, i.e.

B � 0 and­A

­r� cA � 0 at r � r1; �21�

where r1 � r0 � 1 since the layer depth has been chosen as

the unit of length. The boundary condition for the velocity has

been chosen to be stress-free at the top and slip-free at the

Non-linear av-dynamos driven by magnetic buoyancy 167

q 2000 RAS, MNRAS 315, 165±183

bottom, i.e.

V tot � 0 at r � r1 and­V tot=r

­r� 0 at r � r0: �22�

It is important to notice that equations (11)±(13) and (14)±(17)

have been obtained by averaging over latitude and retaining the

radial dependence. This is contrary to most one-dimensional

av-dynamo models which normally only keep the latitudinal

dependence (see for example Weiss, Catteneo & Jones 1984;

Jennings 1991). This approach results directly from our prescrip-

tion of the a -effect since a is driven by the radial gradient of the

toroidal magnetic field and therefore it has to be a function of

radius.

2.2 The spherical shell model

Here, the one-dimensional dynamo, described in the previous

section, is extended to a two-dimensional spherical shell model.

The two dynamo models are basically identical except for the

description of the a -effect. In the one-dimensional model it was

assumed that magnetic buoyancy, and hence the a -effect sets in as

soon as the magnetic field strength falls off with height (see

equation 3). This is strictly speaking not true since in general the

magnetic buoyancy instability must be vigorous enough to

overcome the stabilizing influence of the stratification of the

atmosphere and various other effects. Thus the radial gradient of

the magnetic field strength must exceed a certain threshold value

in order for magnetic buoyancy to occur. Therefore, in this model,

the a -effect will only set in if the following condition is satisfied:

­jBTj­r

, 2C; �23�

where BT is the toroidal field strength and C is assumed to be a

measure of the stabilizing effect of the atmospheric stratification.

This is based on the instability criterion for axisymmetric modes:

2g

gc2s

d

dzln

BT

r

� �.

N2

V2�24�

or for non-axisymmetric modes:

2g

gc2s

d

dz�ln�BT�� . k2

y 1� k2z

k2x

� �� N2

V2; �25�

where N is the Brunt±Vaiasala frequency (Acheson 1979).

Moreover, since the latitudinal dependence of a is not known,

we choose, for the sake of simplicity, a to be positive in the

northern hemisphere and negative in the southern hemisphere. It

should, however, be noted that results from helioseismology

suggest that in the neighbourhood of the equator a needs to be

negative at the base of the convection zone. Thus we choose to

model a (r, u , t) by:

a�r; u� � a�r� sin�2u�; 2 p=2 # u # p=2; �26�where

a�r� � 2a0BT

­jBTj­r

2 C; if BT

­jBTj­r

, 2C;

0; otherwise:

8<: �27�

Here, C, is added to the stability criterion in order to avoid a

spatial discontinuity which would otherwise occur. For physical

reasons, only positive values of C will be considered since these

correspond to a stably stratified atmosphere. Finally, C . 0

ensures that the a -effect vanishes for zero toroidal magnetic field.

In order to get an equatorward migration of the toroidal field belt

it is required that a (­v /­r) is negative in the northern hemisphere

and positive in the southern hemisphere. Moreover, the angular

velocity has to be symmetric about the equator and has to vanish

at the poles.

The governing equations of the two-dimensional spherical shell

dynamo are given by:

­A

­t� �72 2 r22 sin22 u�A� DBTa�r; u�; �28�

­BT

­t� �72 2 r22 sin22 u�BT � r sin uBP7

VT�r; u�fr sin u

� �; �29�

­vind

­t� PM�72 2 r22 sin22 u�vind � ��7 ^ B� ^ B�f: �30�

where D is the dynamo number and PM is the magnetic Prandtl

number. As in the previous section, equations (28)±(30) have been

non-dimensionalized using the layer depth d and the diffusion

time d2/h as units of length and time.

The above equations are then solved subject to the boundary

conditions (20), (21) and (22). Furthermore, the toroidal field B,

the vector potential A, the prescribed velocity vpres and the

induced velocity vind have to vanish at the poles. Thus:

A � BT � upres � vind � 0 at u � ^p=2: �31�

3 R E S U LT S F R O M C Y L I N D R I C A L S H E L L

DY N A M O

3.1 Zeroth-order truncation of the Lorentz force in Z

For the one-dimensional av-dynamo calculation performed here

a grid size of 51 grid points, uniformly spaced over r0 # r # r1

and a time-step of Dt � 1024 have been used. The inner radius has

been fixed at r0 � 2 while the magnetic Prandtl number was

chosen to be PM � 1: Changes in r0 and PM have no influence on

the qualitative behaviour of the system. Here it should also be

noted that equations (5)±(7) do not lend themselves to a linear

stability analysis since, as will be shown below, the initial

bifurcation point from the trivial steady state solution occurs at

infinity. This makes it impossible to determine the mode of maxi-

mum growth rate kmax and for this reason k has been arbitrarily

fixed at k � 1: This is one of the main disadvantages of this

model, although, changing the value of k, will most probably only

affect the amplitudes and the frequencies of the vector potential A

and the toroidal field, while their qualitative behaviour remains

unchanged.

The prescribed angular velocity is given by:

Vpres � 1:8 cos2�p�r 2 r0�=2�: �32�Taking the initial toroidal field to be B0 � 0 and the vector

potential A0 to be some seed field, finite amplitude solutions are

obtained for Ds $ 80: For these particular initial conditions lower

values of D do not yield dynamo action. However, note that the

dynamo number Ds at which non-trivial solutions occur depends

on the strength of the initial field. This is illustrated in Table 1,

which was obtained by keeping B0 � 0 fixed and varying the

strength of the initial poloidal seed field. Here we used

EA � 1

2

�1

0

jA0�r�j2 dr �33�

168 J.-C. Thelen

q 2000 RAS, MNRAS 315, 165±183

as a measure of the energy contained in the poloidal field. Thus as

the strength of the initial field is reduced Ds moves towards larger

values of D, while if the strength of the initial field is increased Ds

moves towards smaller values of D. The data in Table 1 suggests

that, as EA ! 0; the dynamo number Ds / E20:5A : Thus as EA

tends to zero Ds tends to infinity and the bifurcation from the

trivial solution to the finite amplitude solution is subcritical with

the bifurcation point at infinity. Furthermore, there seems to exist

a `critical' dynamo number Dc < 20 below which no dynamo

action occurs even for very strong magnetic fields. Here, it should

be noted that `critical' does not refer to the linear instability

threshold. Instead, by `critical' we mean that it is the lowest value

of the dynamo number D for which a non-trivial, stable solution

has been obtained. The simplest possible scenario that may give

rise to an oscillatory finite amplitude solution is that of a

subcritical Hopf bifurcation occurring at infinity with an unstable

branch extending back to Dc < 20; where a saddle-node bifur-

cation occurs as indicated in Fig. 1. This saddle-node bifurcation

is inherent to all three models discussed in this paper. Its existence

is a result of the fact that a finite value of D is required for dynamo

action. For D � 0 or for very small values of D the dynamo fails

to set in, independently of the initial field strength, because the

a -effect or the v-effect are not vigorous enough to maintain

dynamo action. The bifurcation at infinity, on the other hand,

results from the disappearanace of the a -effect in the absence of a

toroidal magnetic field. If the initial poloidal seed field is weak,

only a weak toroidal field will be generated by the differential

rotation, which translates into a small a . If D is large, this will be

enough to regenerate the poloidal field at a faster rate than the rate

at which it is destroyed by diffusion. If D is too weak the poloidal

field will decay and hence the dynamo will fail to set in. If,

however, the initial seed field is strong, then a strong toroidal field

and hence a large a is obtained and only a small D is required to

obtain dynamo action.

All the solutions obtained for the system of equations (11)±(13)

are periodic with a single frequency. Apart form the Hopf bifur-

cation at infinity and the saddle-node bifurcation at Dc, no other

bifurcations have been found. Furthermore, as D! 1 the induced

velocity tends to become uniform in space and time. Table 2

shows the amplitude of A, B and the frequency of the oscillations

for different values of the dynamo number D for PM � 1: As

D! 1 the following result is obtained:

jAj ! 1;jBj ! 0;

f ! const:

A similar result was obtained by Weiss et al. (1984) who

investigated a simple non-linear model of an oscillatory stellar

dynamo. They reduced the partial differential equations describing

the dynamo to a system of ordinary differential equations and

found that their simplest model, a fifth-order system, displayed a

similar behaviour to the one described above, while their higher

order systems were much more complicated.

Assuming that the induced velocity vind ! 2vpres; i.e. that the

shear becomes very small as D! 1; it is possible to show that:

A / D1=3; �34�B / D21=3: �35�Under the above assumption it is possible to eliminate the second

term on the right-hand side of equation (11) by putting:

­V

­r� ­Vpres

­r� ­�vind=r�

­r� 0 �36�

and hence:

vind � 2Vpresr � const: as D! 1: �37�Putting A � ~ADa and B � ~BDb and balancing the terms in

equations (11)±(13) it can easily be shown that (34) and (35) hold.

The constant of proportionality depends on the magnetic

Prandtl number and the layer depth. An increase in PM

corresponds to a decrease in the magnetic diffusivity and the

magnetic field behaves more and more as if it moved with the

plasma. Thus both the a and the v-effect become more effective

in generating the magnetic field, which lead to an increase in the

amplitude. This is shown in Fig. 2. Similarly the amplitudes of A

and B increased with decreasing layer depth. However, the exact

effect of varying the layer thickness could not be determined

because the value of the preferred latitudinal mode k, which most

likely changes with r0, is not known. In this case the value of k

becomes important since the chosen value of k may be more

favourable to dynamo action for thin layers than for thick layers.

Table 1. The dynamo number Ds at which non-trivial solutions occur as a function of the initialfield strength, for PM � 1 and r0 � 2:

EA 1/4 1/8 1/16 1/32 1/64 1/256

Ds 72 102 140 196 280 558

Figure 1. A possible bifurcation sequence that gives rise to periodic finite

amplitude solutions. A subcritical Hopf bifurcation occurs at D � 1: The

unstable branch (dashed line) extends back to Dc where a saddle-node

bifurcation occurs giving rise to a stable branch (continuous line) of finite

amplitude solutions.

Table 2. Amplitude of A, B and their frequencyfor PM � 1; r0 � 2 and for different values ofD.

D Amplitude Amplitude FrequencyA B

80 ^18.00 ^1.68 0.8100 ^21.36 ^1.64 0.7200 ^36.39 ^1.50 0.6400 ^56.88 ^1.33 0.6600 ^70.27 ^1.22 0.6800 ^79.97 ^1.13 0.6

1000 ^87.48 ^1.05 0.62000 ^110.6 ^0.84 0.64000 ^135.0 ^0.65 0.6

Non-linear av-dynamos driven by magnetic buoyancy 169

q 2000 RAS, MNRAS 315, 165±183

Thus the system (11)±(13) seems to be very robust. Changes in

the initial conditions, of the parameters D, PM and the layer depth

do not affect the overall behaviour of the system.

3.2 Second-order truncation of the Lorentz force in z

Here the influence of the dynamo number D on the behaviour of

the higher order system (14)±(17) will be studied. As it turns out

the behaviour of the solutions is much more complicated than for

the previous system, where only a simple periodic solution was

obtained. Here solutions with higher periods and chaotic solutions

occur and a much more elaborate bifurcation structure, including

features like hysteresis, has been obtained.

Fig. 3(a) shows the bifurcation diagram for 0 # D # 400; while

for larger values of D the behaviour is represented schematically

by Fig. 3(b). Fig. 3(a) is obtained by plotting EA, defined by

EA � 1

2�t1 2 t0��t1

t0

�1

0

jA�r; t�j2 dr dt �38�

against the dynamo number D. As in the previous section EA may

be considered to be a measure of the poloidal magnetic field

strength. [t0, t1] is the time interval over which the vector potential

has been averaged. Taking as initial condition B0 � 0 and A0 to be

some initial poloidal seed field, we obtain a periodic finite

amplitude solution for D $ 80: Note, however, that D � 80 is not

the critical dynamo number, because it was shown in the previous

section that the dynamo number at which the dynamo sets in

depends on the strength of the initial condition. The bifurcation

from the trivial state to the periodic solution is thus subcritical

with the bifurcation point at infinity. By taking the solution of

D � 80 as the initial condition and following the solution

backwards it turns out that it is possible to obtain dynamo action

for values of D as low as D < 20: Thus Dc < 20 is the `critical'

dynamo number since, as explained in Section 3.1, for D , Dc the

only stable solution which has been found is the trivial solution. If,

on the other hand, the dynamo is increased from D � 80 onwards,

the periodic solution persists up to D � 210: As the dynamo

number D is increased through D � 210 this lower branch

becomes unstable and an `upper' branch of periodic finite

amplitude solutions appears. This second solution distinguishes

itself from the previous one by having a larger frequency, a

smaller amplitude of the vector potential A and a larger amplitude

of the toroidal field B. This second solution is stable in the interval

165 # D # 260: As D is further increased beyond D � 260; this

solution becomes unstable and a period-five solution is obtained,

which persists in the interval 260 # D # 296: In order to gain a

better understanding of the behaviour of the system near D � 200

and D � 260 consider Fig. 4. This is a schematic representation of

the bifurcation diagram 3(a) in the interval 150 # D # 280: The

continuous lines represent the stable solutions obtained from our

numerical code, while the dashed lines show the simplest way in

which these stable branches might be connected by unstable

solutions. From this pictures it emerges that hysteresis is likely to

occur in the interval 165 # D # 210 and in the immediate vicinity

of D � 260:Finally, before chaotic solutions occur for D $ 302; a

Figure 2. (a) The amplitude of the vector potential as a function of D for

different values of r0. (b) The amplitude of the vector potential as a

function of D for different values of PM.

Figure 3. (a) Schematic representation of the bifurcation diagram up to

D � 310: Here stars represent periodic solutions and crosses chaotic ones.

(b) Occurrence of solutions at higher values of D. CH� chaotic, 1-P�period-one, 2-P� period-two.

Figure 4. Schematic representation of the bifurcation diagram 3(a). The

continuous lines indicate stable solutions found by our numerical code,

while the dashed lines represent possible unstable solutions.

170 J.-C. Thelen

q 2000 RAS, MNRAS 315, 165±183

period-seven solution is obtained in the short interval 299 #D # 302: The period-five and the period-seven solutions are

shown in Figs 5(a) and 6(a) for D � 270 and D � 300

respectively. In order to verify that the last two types of solutions

are indeed periodic and not quasiperiodic, their Poincare sections

and power spectra have been computed. The Poincare section is

obtained by projecting the trajectories of the solution in phase

space onto a suitable plane, in this case onto the (B, Vtot)-plane.

From the power spectrum it can clearly be seen that the

subharmonics are located at odd multiples of f1/5 and f1/7,

where f1 is the predominant frequency. This is shown in Figs 5(c)

and 6(c). Figs 5(b) and 6(b) show the corresponding PoincareÂ

sections. From Fig. 5(b) it is also apparent that the solution in

phase space intersects the (B, Vtot)-plane at five distinct points and

does not describe a closed curve which would be indicative of a

quasiperiodic solution. However, from the Poincare section for the

seven-periodic solution (see Fig. 6b) it is not that obvious anymore

that the phase-space solution intersects the (B, Vtot)-plane at seven

points. The spread of points in the Poincare section could be from

either numerical noise or the presence of a very long orbit, i.e.

very small frequency. Furthermore, since the value of D � 300 is

very close to the point where the seven periodic solution becomes

Figure 5. (a) B, A, Vtot of the period-five solution as a function of time at r � 2:4 for D � 270; PM � 1 and r0 � 2: (b) Poincare section of the period-five

solution. (c) Power spectrum of A.

Non-linear av-dynamos driven by magnetic buoyancy 171

q 2000 RAS, MNRAS 315, 165±183

unstable, the spread of points could also have been caused by a

very long initial transient. Chaotic solutions are obtained in the

range 302 # D # 4000: However, inside this interval 302 # D #4000 there exist several windows of periodicity. They are

enumerated below.

(i) Period-two solution in 610 # D # 780:(ii) Single periodic solution in 1160 # D # 1350:(iii) Period-two solution in 1680 # D # 2100:(iv) Single periodic solution in 2750 # D # 2800:

Thus, in this case, the route to chaos is the following: the trivial

solution gives rise to a periodic solution via a subcritical

bifurcation with the bifurcation point at infinity. A further

subcritical bifurcation gives rise to a second periodic solution

and chaos is then reached via a branch of period-five and a branch

of period-seven solutions. It should be noted no quasiperiodic

solutions have been obtained here.

A similar type of behaviour, namely the occurrence of jumps in

the energy levels, indicating the possibility of hysteresis (see Figs

3 and 4) and the existence of periodic solutions in the chaotic

interval, has also been observed by Covas et al. (1998). However,

they only obtained windows of periodicity in the case of thick

Figure 6. (a) B, A, Vtot of the period-seven solution as a function of time at r � 2:4 for D � 300; PM� 1 and r0 � 2: (b) Poincare section of the period-seven

solution. (c) Power spectrum of A.

172 J.-C. Thelen

q 2000 RAS, MNRAS 315, 165±183

shells, i.e. with an inner radius of r0 � 0:25; and that these

vanished for relatively thin shells with an inner radius of r0 � 2;such as have been considered in this section. Moreover, the

coexistence of two or three different stable solutions for the same

dynamo number depending on the initial conditions has also been

found by Belvedere et al. (1990) and Schmitt & SchuÈssler (1989).

This suggests that stars of a similar age and structure might

display different types of behaviour at any given time.

The system (11)±(13) and the system (14)±(17) have one char-

acteristic in common, namely that as D! 1; the amplitude of the

vector potential A goes to infinity, the amplitude of the toroidal

magnetic field B tends to zero and the amplitude of the induced

velocity vind, owing to the zeroth-order terms in the Lorentz force,

tends to some constant, i.e. it becomes steady. From this we may

conclude that the complicated behaviour of the system (14)±(17)

is because of wind, which is induced by the higher order terms

included in the Lorentz force.

3.3 The influence of the magnetic Prandtl number

Finally the consequences of changing the magnetic Prandtl

number on the different types of solutions will be considered.

Recent results from two-dimensional mean-field dynamos,

including the Malkus±Proctor effect, suggest that the dynamics

of the system depend on the magnetic Prandtl number. The period

of modulation of solutions, for example, seems to increase with

decreasing magnetic Prandtl number and grand minima only seem

to occur for PM ! 1: Physically this is because in order to obtain

grand minima the energy needs to be transferred between the

magnetic and the velocity field and there has to be a delay in the

quenching mechanism. In the case of the Malkus±Proctor effect

this delay corresponds to the time-scale on which the energy is

returned to the magnetic field, which is controlled by PM (Tobias

1996; Knobloch, Tobias & Weiss 1997). This might also explain

the absence of this type of modulated solution in our case, where

the magnetic Prandtl number PM is of order unity since the

exchange of energy between the magnetic field and the velocity

field occurs on too short a time-scale for the magnetic field to

decay by a considerable amount and thus the amplitude of the

oscillations remains constant. Furthermore, it was shown by Weiss

et al. (1984) and Jones, Weiss & Cattaneo (1985) that for low-

order models the bifurcation structure depended on the choice of

PM, with the periodic solutions remaining stable for PM . 1:In this investigation the following values of the magnetic

Prandtl number have been considered: PM � 2; PM � 0:5; PM �0:25: First, note that increasing PM leads to an increase in the

magnetic energy. This is, as expected, because for large PM the

magnetic field lines are stretched more rigorously and thus

stronger fields are obtained. However, the magnetic Prandtl

number seems to play an important part in the dynamics of our

model. For PM � 2 only periodic solutions have been obtained

which seems to confirm the above result that quasiperiodic

solutions only appear for PM , 1: Moreover, except for the short

interval of the period-three solution, the oscillatory solutions of

period one persist over long intervals of the dynamo number D.

This solution is still stable at D � 2000; at which stage the

computations have been stopped. At PM � 1 the solutions are still

periodic, but, as has been shown before, the overall behaviour is

less uniform and the bifurcation structure is more complicated.

The first quasiperiodic solution appears for PM � 0:5 in the short

interval 400 # D # 450 although the periodic solutions are still

predominant. However, PM � 0:5 is apparently still not small

enough to obtain `grand minima' type modulation. Thus in order

to check whether our model gives rise to such a behaviour some

calculations have been performed at lower values of PM. At PM �0:25 quasiperiodic solutions appear over a larger range of the

control parameter D. Furthermore the decrease of the magnetic

Prandtl number leads to the appearance of oscillatory solutions

with very large periods, which were absent in the case of PM � 2:For example, in the case of PM � 0:25 a period-nine solution was

obtained. No solutions have been obtained for lower values of the

magnetic Prandtl number, mainly because in order to ensure

convergence, a very large number of grid points is required, which

makes the investigation of the parameter space impractical. It

seems, however that `grand minima' type solutions, if they occur

in our model at all, will appear only for PM ! 1; which is in

agreement with Tobias (1996).

4 R E S U LT S F R O M S P H E R I C A L S H E L L

DY N A M O

4.1 PM . 1

The system of equations (28)±(30) is solved pseudospectrally by

expanding the vector potential A, the toroidal field B and the

induced velocity vind in Chebychev polynomials and associated

Legendre polynomials of order 1. By the term pseudospectral we

mean that the non-linear terms are evaluated in the physical space.

The time-stepping is performed by using a semi-implicit time-

integration scheme; the diffusion terms are calculated by a Crank±

Nicolson scheme in spectral space, while the non-linear terms are

calculated in physical space by using a two-step Adams±

Bashforth method. Both time-integration schemes are of second

order. The calculations have been performed by using either 32 �32 or 48 � 48 grid points, which is high enough a resolution in

order to ensure an exponential decay of high frequency modes. For

most of the calculations the inner radius is fixed at r0 � 2 and the

magnetic Prandtl number is given by PM � 3: For the same reason

as in Section 3, no linear stability analysis has been performed.

Taking B0 � 0 and A0 to be some seed field an oscillatory

dipole solution occurs at DD � 50; while an oscillatory quadru-

pole solution occurs at DQ � 70: It should, however, be noted that

the dynamo at which dynamo action occurs depends on the strength

of the initial field, i.e. as the initial field strength decreases DD and

DQ increase. Therefore, the bifurcation point from the trivial

steady state to a finite-amplitude solution is subcritical with the

bifurcation point at infinity. For a more detailed explanation see

Section 3. There also exists a critical dynamo number Dc below

which the dynamo fails to set in independently of the initial field

strength. In our calculation it actually turns out that DD (DQ) is

also the critical dynamo number for the oscillatory dipole

(quadrupole solution). This results from the relatively strong

initial poloidal field strength which was chosen here.

Next, the nature of the solution is investigated as the dynamo

number D is increased. First, stability tests have been performed in

order to see whether the pure dipole and the pure quadrupole

solutions are stable with respect to perturbations of the other

parity. In order to perform these calculations, EA, which is

considered to be a measure of the poloidal magnetic field strength,

is calculated according to:

EA � 1

2�t1 2 t0��t1

t0

�S

jA�r; u; t�j2 dS dt; �39�

where S is a meridional section of the spherical shell. Fig. 7 shows

Non-linear av-dynamos driven by magnetic buoyancy 173

q 2000 RAS, MNRAS 315, 165±183

the plot of EA versus the dynamo number D. Note that the energy

EA of both the dipole mode and the quadrupole mode seems to

level off as the dynamo number is increased. This is not too

surprising because it was shown previously that the amplitude of A

is proportional to D1/3 for large values of D.

However, the number of active regions in each hemisphere is

not exactly the same at any given time and hence it is reasonable

to assume that the solar magnetic field, which is responsible for

the appearance of sunspots, is slightly asymmetric about the

equator. Moreover, just as the Sun entered and emerged from the

Maunder minimum, sunspots were distributed asymmetrically

around the equator (Ribes & Nesme-Ribes 1993). Thus, since the

solar magnetic field seems to contain contributions from both

parities, the question arises as to what happens if the initial seed

field is of mixed parity. In order to study this case, the parity P of

the solution is considered. The parity is defined by:

P � ESA 2 EA

A

ESA � EA

A

; �40�

where ESA and EA

A denote the `energy' of the symmetric

(quadrupole) and antisymmetric (dipole) part of the vector poten-

tial in the conducting spherical shell (cf. Brandenburg et al.

1989a). ESA �EA

A � can easily be obtained from (39) by only

considering the coefficients with odd (even) n. Note that P � 1

corresponds to a pure quadrupole solution and that P � 21 to a

pure dipole solution. The stability of these mixed-mode solutions

is then investigated by adding an amount e of a quadrupole

component to a dipole and vice versa. Thus initially we have:

A0 � AA0 � eAS

0 �41�or

A0 � AS0 � eAA

0 : �42�Fig. 8 gives the evolution of the parity as a function of time for

D � 100; PM � 3; r0 � 2; C � 0 and different values of e . From

this it can clearly be seen that the final solution always has parity

P � 21; i.e. the final solution is always antisymmetric, except in

the case where the initial condition is a pure quadrupole. This also

holds at larger values of D. As the dynamo D is increased from Dc,

the dipole solution, represented by the continuous line in Fig. 7, is

found to be stable to quadrupole perturbations, while the quadru-

pole solution, represented by the dashed line, is unstable to dipole

perturbations. This, and the fact that they appear at slightly lower

values of D, suggests that, here, dipole solutions are preferred.

Furthermore, the evolution from the mixed-parity initial condition

to the pure dipole happens on a fairly short time-scale. The time

the solution takes to reach the dipole state is of order O(1),

independently of the initial value of the parity P, i.e. indepen-

dently of the relative strengths of the initial quadrupole and dipole

Figure 7. Plot of EA versus D for the dipole solution (A) and the

quadrupole solution (S) for PM � 3; r0 � 2 and C � 0: The continuous

line indicates that the dipole solution is stable to quadrupole perturbations,

while the dashed line indicates that the quadrupole solution is unstable to

dipole perturbations.

Figure 8. Evolution of the parity for D � 100; PM � 3; r0 � 2; C � 0 and different initial values of P.

174 J.-C. Thelen

q 2000 RAS, MNRAS 315, 165±183

components. A similar behaviour is observed for all values of

PM $ 1: In fact, changing PM does not alter the qualitative

behaviour of the system provided that PM $ 1 is satisfied, except

for an increase in the magnetic energy contained in the magnetic

field. This is because by increasing PM the magnetic diffusivity is

effectively reduced and hence the a -effect and the v-effect

become more effective in generating the magnetic field.

It should also be noted that all the solutions found in the large

magnetic Prandtl number regime are periodic in time, except for

two steady quadrupole solutions, which we will come to shortly.

The absence of quasiperiodic and chaotic solutions corresponds to

that found by Jones et al. (1985), who investigated the bifurcation

structure of a sixth order dynamo model and found that

quasiperiodic types of solutions only occur for PM , 1:Next the influence of the layer depth on the dynamo is

investigated. The following values for the inner radius r0 are

considered: r0 � 1 and r0 � 3: Higher values of r0 (thinner layers)

require a very high resolution and hence a large number of time-

steps in order to ensure stability and accuracy and are, therefore,

impractical to study. Increasing the inner radius r0 from r0 � 2 to

r0 � 3; i.e. decreasing the layer depth, leads to an increase in both

EA and EB and to a decrease in the critical dynamo number Dc.

Thus the dynamo becomes easier to excite, independently of the

initial conditions. In thick shells �r0 � 1� this situation is still valid

for the dipole solution, i.e. the energy contained in the magnetic

field is less than for either r0 � 2 or r0 � 3; and a larger D is

required for dynamo action. However, if the initial condition is

chosen to be a pure quadrupole then the situation changes

completely. First, instead of an oscillatory solution, a steady

solution is obtained and, secondly, this steady solution sets in at a

very low dynamo number compared to the corresponding dipole

solution. Unlike in all the previous cases, where the oscillatory

dipole was preferred over the oscillatory quadrupole, here the

steady quadrupole seems to be preferred over the oscillatory

dipole.

In order to investigate this behaviour the following initial

conditions are considered: B0 � 0 and A0 is chosen such that

EAA � ES

A �P � 0�; i.e. the initial energy contained in the dipole

component is the same as the initial energy contained in the

quadrupole component. Starting with this initial condition, a

steady quadrupole solution is obtained at D � 25; which is also

the critical dynamo number for the quadrupole solution. However,

for the same initial condition, a periodic dipole appears at D �350 and, following this solution backwards, the critical dynamo

number for the dipole solution is found to be D � 150: The graph

of EA versus the dynamo number D for both the dipole and the

quadrupole solution is shown in Fig. 9. For D , 350; the energy

contained in the quadrupole is larger than the energy contained in

the dipole, while for D $ 350 the reverse is true. Furthermore,

both the dipole and the quadrupole solutions are stable to

quadrupole and dipole perturbations, respectively. Thus, both

dipolar and quadrupolar solutions are possible stable solutions. A

similar behaviour was also observed by Brandenburg et al.

(1989a) in an investigation of an av-dynamo including the

Malkus±Proctor effect, except that in this case antisymmetric

solutions appeared at lower values of D than did symmetric

solutions. The final nature of the solution depends on the parity of

the initial condition, as can be seen from Fig. 10. From this graph

it emerges that at D � 225 the quadrupole solution is obtained for

20:2 # P # 1 and that the dipole solution is obtained for 21 #P 2 0:2: As D is increased further, dipole solutions are obtained

for values of P . 20:2: At D � 325 oscillatory dipole solutions

are obtained for P belonging to the interval 21 # P # 0:25 and

for D � 500 oscillatory dipole solutions occur for 21 # P #0:75:

Finally, returning to r0 � 2; the influence of the subadiabatic

stratification of the atmosphere, measured by the parameter C, is

considered. Here four different initial conditions are investigated,

namely

(i) a pure dipole poloidal seed field,

(ii) a mainly dipole poloidal seed field with a small quadrupole

component,

(iii) a pure quadrupole poloidal seed field,

(iv) a mainly quadrupole poloidal seed field with a small dipole

component.

First, for the pure dipole initial condition, increasing C,

(increasing the stabilizing effect of the atmospheric stratification)

does not result in a qualitative change in the nature of the periodic

solution. An increase in C just leads to a decrease in the magnetic

energy and moves the dynamo number at which dynamo action

occurs towards larger values of D. For large values of C, in this

case C . 1; it is difficult to obtain dynamo action even at large

values of D since the subadiabatic stratification suppresses the

a -effect very effectively. This is especially true for the initial

conditions with B0 � 0 since it then becomes nearly impossible to

obtain a steep enough gradient for dynamo action. A similar result

is obtained for a mainly dipolar initial condition. However, a pure

quadrupolar poloidal seed field results in a different behaviour.

For small values of C �C , 0:3� the solution is a periodic

quadrupole which behaves exactly as the dipole solution above.

As C is increased above C � 0:3; a steady quadrupole is obtained,

until at C � 1:5 the dynamo fails to set in. Finally, the behaviour

of the system for a mainly quadrupole poloidal seed field is shown

in Fig. 11. Here, the final solution, after the transients have died

away, is a pure dipole for 0 # C , 0:60: For larger values, how-

ever, the solution switches parity and becomes a steady quadru-

pole until for C . 1:5 the magnetic field starts to decay.

The investigation of the large Prandtl number regime gives rise

to two questions in particular. What are the physical reasons for

the appearance of the steady quadrupole mode and why is it, in

Figure 9. Plot of EA versus D for the dipole solution (A) and the

quadrupole solution (S) for PM � 3; r0 � 1 and C � 0: As opposed to

Fig. 7, both the dipole and quadrupole are stable to quadrupolar and

dipolar perturbations.

Non-linear av-dynamos driven by magnetic buoyancy 175

q 2000 RAS, MNRAS 315, 165±183

certain cases, preferred over the oscillatory dipole mode? Steady

quadrupole solutions have also been found by Jennings & Weiss

(1991) and Schmalz & Stix (1991) in one-dimensional Cartesian

dynamo models and by Schmitt & SchuÈssler (1989) in a one-

dimensional spherical shell dynamo. These models all have one

thing in common, namely, that they neglect curvature. Indeed, it

was shown by Jennings (1991), who carried out an intensive study

of the effects of geometry on dynamo models, that the steady

mode found by Schmitt and SchuÈssler vanished if the radial

structure was included in the model. Thus, one possible

explanation is that the appearance of the steady quadrupole is

the result of the geometry of the model under consideration. This

is supported by the fact that steady modes disappear in a fully

three-dimensional model. Knobloch & Landsberg (1996) showed

that for a three-dimensional magnetic field the primary bifurcation

has to be a Hopf bifurcation and that the resulting modes are either

odd or even with respect to the equator. However, in the special

case of axisymmetry, the initial instability does not have to be an

oscillatory mode, i.e. in axisymmetric dynamos steady solutions

are allowed. For the axsimmetric dynamo considered here, it

seems that the steady quadrupole mode is easier to excite than the

oscillatory dipole, i.e. less energy needs to be liberated in order to

excite the steady quadrupole. First, the steady quadrupole solution

appears at large values of C, i.e. in a strongly subadiabatically

stratified atmosphere. Thus, a convectively stable atmosphere

seems to exert a stronger stabilizing influence on oscillatory

modes than on steady modes. Secondly, in the case of r0 � 1 the

quadrupole appears at much lower values of D than the oscillatory

dipole, which indicates that the field-inducing effects need to be

stronger in order to obtain oscillatory modes. Nevertheless, if

there is enough energy in the system to destabilize the oscillatory

dipole, this becomes the preferred mode.

4.2 PM , 1

In this section, only mixed parity initial conditions have been

considered. First, mixed mode initial conditions are physically

more relevant since the Sun's magnetic field is neither a pure

dipole nor a pure quadrupole. Secondly, they give rise to a more

interesting behaviour, since pure parity initial conditions only

seem to result in pure parity solutions. Moreover, the lowest value

of PM at which it is practical to investigate the parameter space is

PM � 0:5: For smaller values it is difficult to ensure stability and

accuracy. For most of the calculations, r0 and C are chosen to be

r0 � 2 and C � 0:As previously, the trivial steady state solution A � B � vind � 0

becomes unstable to an oscillatory dipole at the critical dynamo

number Dc < 100: For D , Dc; no dynamo action has been

obtained. Since the onset of the dynamo depends on the strength

Figure 11. Evolution of the parity for D � 200; PM � 3; r0 � 1; P � 0:75

and different initial values of C.

Figure 10. Evolution of the parity for D � 225; PM � 3; r0 � 1; C � 0 and different initial values of the parity P, namely P < 0:75; 0.25, 0.25, 0, 20.25,

20.50 and 20.75.

176 J.-C. Thelen

q 2000 RAS, MNRAS 315, 165±183

of the initial condition, the bifurcation from the trivial steady state

to the dipole is subcritical with the bifurcation point at infinity.

This stable oscillatory dipole persists in the interval 100 # D #350: The time evolution of EA, EB, the kinetic energy EK and the

parity are shown in Fig. 12. The parity remaining constant at

P � 21 indicates that the solution is dipolar. The butterfly

diagram corresponding to the solution is shown in Fig. 13, and it

can clearly be seen that, first, the dynamo wave appears at high

latitudes and migrates toward the equator and secondly, that the

toroidal field is antisymmetric about the equator. As the dynamo

number is increased past D � 350; this oscillating dipole solution

loses its stability in a Hopf bifurcation to a mixed parity solution.

Just beyond the bifurcation point, the field is mainly dipolar with a

weak quadrupole component. As D is increased, the quadrupole

component grows and the magnetic energies and the parity

become modulated. As D is further increased, this modulation

becomes more and more apparent. As can be seen from the time

series of this mixed-mode solution, shown in Fig. 14, the parity

varies both on the time-scale of the magnetic cycle and on the

longer modulation time-scale. This type of modulation, called

Figure 12. Time evolution of EB, EA, EK and the parity P for D � 200; PM � 0:5; r0 � 2 and C � 0:5:

Figure 13. Butterfly diagram for D � 200; PM � 0:5; r0 � 2 and C � 0:

Non-linear av-dynamos driven by magnetic buoyancy 177

q 2000 RAS, MNRAS 315, 165±183

type I modulation, has been observed before by Moss et al. (1990),

Tobias (1997) and Covas (1998). At D < 625; this quasiperiodic

mixed mode solution loses its stability to a periodic mixed mode

solution. This type of solution remains stable in the interval 625 #D # 750: As can be seen from Fig. 16, which shows the time

evolution of the solution at D � 700; there is only one time-scale

present. As opposed to the previous quasiperiodic solution, the

parity is now solely oscillating with the frequency of the magnetic

cycle. However, in both these cases, the contribution of the

quadrupole component is much smaller than that of the dipole

component and hence the composition of the magnetic field is

mainly dipolar, except in the vicinity of the equator, where the

symmetry is broken, as can be seen from the butterfly diagram in Fig.

17. As a result of this there exists a small asymmetry between the

northern and the southern hemisphere. A similar mode was obtained

by Jennings (1991) and Jennings & Weiss (1991). They also obtained

a second type of periodic mixed mode solution, in which the

magnetic field is restricted to one hemisphere, while the other

hemisphere remains nearly free of magnetic activity. Such a mode

was not obtained here. As D is increased past D � 750; the periodic

Figure 14. Time evolution of EB, EA, EK and the parity P for D � 575; PM � 0:5; r0 � 2 and C � 0:

Figure 15. Butterfly diagram for D � 575; PM � 0:5; r0 � 2 and C � 0:

178 J.-C. Thelen

q 2000 RAS, MNRAS 315, 165±183

mixed mode solution loses stability, probably in a Hopf bifurcation,

and a second modulated mixed mode solution is obtained. From the

butterfly diagram in Fig. 19 it is clear that the asymmetry between

the northern and the southern hemisphere has become much more

pronounced. This modulated solution persists up to D � 1050;where, for numerical reasons, the calculations were stopped.

It should be noted that in this model no type II modulation has

been observed. Type II modulation is characterized by a change in

the amplitude which is not accompanied by a change in the parity.

The absence of this type of solution might result from the fact that

the calculations were stopped at D � 1050; but more likely the

reason is that the magnetic Prandtl number is not small enough. It

has been suggested by Tobias (1996, 1997) that the magnetic

Prandtl number controls the modulation of the frequency and that

the time-scale of the modulation varies as P0:5M : This implies that in

order for the ratio of the magnetic frequency to the modulation

frequency to be similar to that found on the Sun, i.e. so that the

minima occur roughly every 20 activity cycles, a value of PM �0:025 is required which is well beyond the limits of our numerical

scheme.

Figure 16. Time evolution of EB, EA, EK and the parity P for D � 700; PM � 0:5; r0 � 2 and C � 0:

Figure 17. Butterfly diagram for D � 700; PM � 0:5; r0 � 2 and C � 0:

Non-linear av-dynamos driven by magnetic buoyancy 179

q 2000 RAS, MNRAS 315, 165±183

In the low magnetic Prandtl number regime considered here, the

thickness of the spherical shell seems to play an important role in

the behaviour of the dynamo. In the case of a thick layer, i.e.

r0 � 1; none of the interesting solutions observed for r0 � 2

occurs, but the behaviour is identical to that described in the

previous section for r0 � 1 (see Fig. 9), except that EA and EB are

much smaller. For thin shells �r0 � 3�; on the other hand, the

behaviour of the system is similar to that described above. The

trivial steady state solution gives way to a periodic dipole solution

at the critical dynamo number Dc � 50; which persists in the

interval 50 # D # 100: As the dynamo number is increased past

D � 100; a modulated mixed mode solution is obtained, which

remains stable for values of D up to D � 300: Unfortunately, for

much larger values of D difficulties are encountered with respect

to numerical stability and accuracy and, therefore, the calculations

were stopped.

Finally, the consequence that changing C has on the nature of

the solution is investigated. Increasing C (making the atmosphere

more convectively stable) does not alter the bifurcation structure,

which has just been described, i.e. the periodic dipole solutions,

Figure 18. Time evolution of EB, EA, EK and the parity P for D � 1000; PM � 0:5; r0 � 2; and C � 0:

Figure 19. Butterfly diagram for D � 1000; PM � 0:5; r0 � 2; and C � 0:

180 J.-C. Thelen

q 2000 RAS, MNRAS 315, 165±183

the quasiperiodic mixed mode solutions and the periodic mixed

mode solution still appear in the same order. Even the bifurcation

points remain more or less at the same values of the dynamo

number D. The only difference which occurs is that the energy

contained in the magnetic field is reduced as C is increased.

However, it should be noted that only a small increase in C is

necessary in order to prohibit dynamo action. In this case, the

dynamo fails to set in for values of C < 0:1; compared to C < 1

in the large magnetic Prandtl number regime. This is probably

owing to the fact that for small PM the field inducing effects are

less effective and hence that the a -effect is more easily suppressed

by the subadiabatic stratification of the atmosphere.

Next we compare these solutions to the behaviour of the

magnetic field observed on the Sun. First, note that in this model

the dynamo waves appear at a relatively high latitude (<508)when compared to the butterfly diagrams obtained from observa-

tions, which indicate that sunspots appear between 30 and 40

degrees of latitude. Secondly, consider the butterfly diagram of the

periodic mixed mode solution, shown in Fig. 17. This solution is

nearly dipolar but there exists a slight north±south asymmetry

which is most obvious in the neighbourhood of the equator. As

similar asymmetry is actually observed on the Sun since the onset

of the solar cycles as well as the time of maximal sunspot intensity

are different in the southern and northern hemisphere. During the

normal solar cycle this asymmetry is very weak and therefore

Sokoloff & Nesme-Ribes (1994) suggested that there exists a

connection between this type of mixed parity solution and the

asymmetry in the solar cycle. Finally, let us turn to the two

quasiperiodic mixed mode solutions whose butterfly diagrams are

shown in Figs 15 and 19 and whose time evolution is shown in

Figs 14 and 18. They do not seem to correspond to either the

behaviour of solar magnetic fields during grand minima or that

displayed during normal solar activity. During the Maunder

minimum sunspots were mostly located in the southern hemi-

sphere between 0 and 220 degrees of latitude while very few

sunspots were observed in the northern hemisphere (Ribes &

Nesme-Ribes 1993; Sokoloff & Nesme-Ribes 1994). On the other

hand, during normal cycle activity the number of sunspots and

hence the magnetic field, varies nearly periodically with an

average period of 11 yr and does not seem to be modulated on a

longer time-scale. However, aysmmetric fields, which are distri-

buted over both hemispheres have also been observed. Observa-

tional data from the solar sunspot cycle in the Maunder minimum

(1645±1715) shows that when the magnetic field was weak, i.e.

just before and just after the Maunder minimum, both quadrupole

and dipole components were present and sunspots were distributed

asymmetrically (Ribes & Nesme-Ribes 1993). Numerical results,

mimicking this behaviour have recently been obtained by Beer,

Tobias & Weiss (1998).

5 C O N C L U S I O N

In the previous sections, we described an av-dynamo whose

a -effect is essentially the result of the interaction of magnetic

buoyancy instabilities and rotation. Our dynamo model contains

the three following non-linearities, namely the a -effect, the

v -effect and the Lorentz force. The a -effect is considered to be

proportional to the radial gradient of the magnetic field if the field

falls off with height and is chosen to be zero otherwise. The

Lorentz force allows us to take into account the back-reaction of

the magnetic field on the flow, which implies that the growth of

the magnetic field is limited by v-quenching.

The main consequence of the above a -effect is that the

bifurcation from the trivial state solution to a finite-amplitude

solution is subcritical and occurs at infinity. This result holds for

both the one-dimensional and the two-dimensional models

described here. The occurrence of the bifurcation point at infinity

is a consequence of the fact that for weak toroidal magnetic fields

the radial gradient of the field and hence a is small and that,

therefore, a large dynamo number D is required in order to obtain

dynamo action. Thus as the magnetic field strength tends to zero

the dynamo number D tends to infinity. This behaviour is in

contrast to that of many mean-field dynamo models investigated

so far, where the bifurcation from the trivial solution to the finite

amplitude solution occurs at a finite value of D. In these models ais either independent of the magnetic field strength or a is a

decreasing function of the magnetic field (Jepps 1975; Schmitt &

SchuÈssler 1989; Brandenburg 1989a; Covas 1998), as opposed to

the a chosen here which is an increasing function of the magnetic

field strength.

The fact that the initial bifurcation occurs at infinity is not very

likely in an astrophysical context. It would be more realistic to

have a subcritical (or supercritical) bifurcation at a finite value of

D. This could have been achieved by introducing an extra term

into the a -effect. The reason why it was decided not to do this was

to avoid introducing any unnecessary parameters into the

equations, whose physical meaning was not well understood.

The main results obtained from the cylindrical shell model are

the following. The first system, which is obtained by truncating

the Lorentz force at zeroth order is very robust with respect to

changes in the dynamo number D, the magnetic Prandtl number

PM and the layer depth. Only a single periodic finite amplitude

solution has been obtained, which remains stable for all values of

D and PM. Moreover, as the dynamo number tends to infinity, the

toroidal magnetic field strength goes to zero, the vector potential

and hence the poloidal field tends to infinity and the frequency

becomes constant. Moreover, the total velocity becomes uniform

in space and time as D is increased. This implies that the Lorentz

force drives a differential rotation which cancels out the

prescribed shear flow as D! 1: A similar result was obtained

by Weiss et al. (1984). The dynamics of the second system, on the

other hand, are far more complicated. First, a more complex

bifurcation structure is obtained. As opposed to the previous case,

where only one simple periodic solution was found, here solutions

with higher periods, such as period-five and period-seven

solutions, as well as chaotic solutions appear. Furthermore,

hysteresis occurred, i.e. in certain regions of parameter space

the solutions depend on the choice of the initial conditions. This is

interesting since it indicates that two stars of similar age and

structure might display two different types of behaviour. Secondly,

the behaviour of this system is no longer as robust as in the

previous case, in the sense that the bifurcation structure depends

now on the boundary conditions, the layer thickness and the

magnetic Prandtl number. The magnetic Prandtl number also plays

a very important role in the dynamics of the spherical shell

dynamo. For PM . 1 the solutions are in general periodic dipoles

or quadrupoles, while for PM , 1 quasiperiodic solutions appear.

This result corroborates that of Jones et al. (1985), namely that

periodic finite amplitude solutions only become unstable for

PM , 1: However, physically it is not clear why the behaviour of

the dynamo changes so abruptly at PM � 1: This question is

probably linked to the fact that in the large magnetic Prandtl

number regime only pure parity solutions seem to exist because

either the dipole or the quadrupole mode is damped if mixed

Non-linear av-dynamos driven by magnetic buoyancy 181

q 2000 RAS, MNRAS 315, 165±183

parity initial conditions are used. Thus type-I modulation,

characterized by a modulated parity, does not occur since it relies

on an exchange of energy between the dipole mode and the

quadrupole mode. Furthermore, type-II modulation, characterized

by either a constant quadrupole or dipole parity and a modulated

magnetic field, only arises at very low magnetic Prandtl numbers

and if the dynamo number D is sufficiently subcritical. This type

of modulation is based on an exchange of energy between the

magnetic field and the velocity field and can only occur if there is

a delay in the quenching mechanism. In our model this delay is

controlled by the magnetic Prandtl number and corresponds to the

time-scale on which the energy is returned to the field (Knobloch

et al. 1997). For small magnetic Prandtl numbers the behaviour of

the dynamo changes completely with the appearance of modulated

mixed mode solutions. However, type-II modulation, has not been

obtained, because the magnetic Prandtl number is not small

enough and the dynamo number is not sufficiently supercritical

enough. This is perhaps not too surprising since Tobias (1997)

showed that large values of PM shift type-II modulation to larger

values of D. Thus in order to obtain this type of modulation it is

necessary to decrease PM to values lower than those investigated

here. Decreasing PM, however, leads to difficulties since the

resolution and the number of time steps have to be increased in

order to ensure numerical stability. This problem is further

compounded by the fact that a decrease in PM reduces the

effectiveness of the field-inducing effects and that, therefore,

larger values of D are required in order to obtain dynamo action.

The results obtained here clearly indicate that it is possible to

drive an av -dynamo by an a-effect, which is based on magnetic

buoyancy instabilities. It should, however, be noted that the

behaviour of the solution is not in agreement with that observed in

solar type stars. By measuring the Ca+ emissions from active stars

it is possible to determine their rotation rates. It has been shown

that the most active stars are stars with short rotation periods and

moreover that their cycle frequency increases with the rotation

rate (see Weiss et al. 1984; Weiss 1993, and references therein).

Thus we would expect the toroidal magnetic field strength, which

can be considered to be a measure of the magnetic activity, to

increase with increasing D. However, in the system considered

here the toroidal magnetic field strength decreases. Furthermore,

modulation associated with changes in the parity have not been

observed in the Sun so far (Knobloch et al. 1997). Nevertheless,

some features of our results are consistent with solar observations.

First, the frequency of the oscillations increases with the dynamo

number. Secondly, in some regions of parameter space, our results

agree with observations of sunspots, which indicate that the solar

magnetic field is mainly dipolar and that it exhibits a small north±

south symmetry. Even though our results do not totally agree with

observations it is possible to conclude that such a dynamo may

well operate in the convectively stable overshoot zone or at the

base of the convection zone since the two key ingredients

necessary for dynamo action, namely the v-effect, resulting from

the strong radial gradient of the angular velocity, and the a -effect,

resulting from the interaction of magnetic buoyancy instabilities

and rotation, are present.

The main advantage of the dynamo model described in this

paper is that the a -effect is derived from a numerical model of the

physical processes operating in the overshoot zone. Therefore, this

a represents a physically more accurate model than the prescrip-

tions of the a -effect used in many models to date, which rely on

an ad hoc parametrization of the effects which are thought to

occur in the solar interior. Moreover, the growth of the magnetic

field is limited by taking account of the back-reaction of the field

on the flow through the Lorentz force and therefore it emerges

self-consistently from our model. Hence no extra parametrization

is needed. The main problem, on the other hand, is that, owing to

the prescription of the a -effect, it becomes impossible to

determine the latitudinal modes of maximum growth rate.

Furthermore, the latitudinal dependence of the a -effect is not

known and therefore it has been prescribed in such a way as to

give equatorward migration of the magnetic field. There are,

however, no physical reasons for this prescription. A further

disadvantage of these models is their inability to cope with very

low magnetic Prandtl numbers. This problem seems to be, at least

partially, inherent to the model under consideration. As PM tends

to zero the diffusive term, which balances the Lorentz force in

the momentum equation becomes negligible. Consequently, the

induced velocity grows rapidly until the direction of the flow is

reversed again by the Lorentz force. This growth of the induced

velocity results in a strong radial gradient of the azimuthal

velocity which is thought to lead to problems if the resolution is

too small.

AC K N OW L E D G M E N T S

I am grateful to David Hughes for all the helpful discussions. I am

indebted to Rainer Hollerbach for valuable help on spectral

methods and to an anonymous referee for some very helpful

comments on this paper. I would also like to thank PPARC for

paying my tuition fees.

R E F E R E N C E S

Acheson D. J., 1979, Phil. Trans. R. Soc. Lond., 289, 459

Beer J., Tobias S., Weiss N. O., 1998, Sol. Phys., 181, 237

Belvedere G., Pitadella R. M., Proctor M. R. E., 1990, GAFD, 51, 263

Brandenburg A., Krause F., Meinel R., Moss D., Tuominen I., 1989a,

A&A, 213, 411

Brandenburg A., Tuominene I., Moss D., 1989b, GAFD, 49, 129

Brandenburg A., Moss D., RuÈdiger G., Tuominen I., 1991, GAFD, 61, 179

Brown T. M., Christensen-Dalsgaard J., Dziembowski W. A., Goode P.,

Gough D. O., Morrow C. A., 1989, ApJ, 343, 526

Covas E., Tworkowski A. S., Brandenburg A., Tavakol R., 1997a, A&A,

317, 610

Covas E., Tworkowski A. S., Tavakol R., Brandenburg A., 1997b, Sol.

Phys., 172, 3

Covas E., Tavakol R., Tworkowski A. S., Brandenburg A., 1998, A&A,

329, 350

Ferriz-Mas A., Schmitt D., SchuÈssler M., 1994, A&A, 289, 949

Galloway D. J., Weiss N. O., 1981, ApJ, 243, 945

Gilman P. A., 1983, ApJS, 53, 243

Gilman P. A., Miller J., 1981, ApJS, 46, 211

Glatzmaier G. A., 1984, J. Comput. Phys., 55, 461

Glatzmaier G. A., 1985, ApJ, 291, 300

Jennings R. L., 1991, GAFD, 57, 147

Jennings R. L., Weiss N. O., 1991, MNRAS, 252, 249

Jepps S. A., 1975, JFM, 67, 625

Jones C. A., Weiss N. O., Cattaneo F., 1985, Phys. D, 14, 161

Knobloch E., Landsberg A. S., 1996, MNRAS, 278, 294

Knobloch E., Tobias S. M., Weiss N. O., 1997, MNRAS, 307, 1

Malkus W. V. R., Proctor M. R. E., 1975, JFM, 67, 417

Moffatt H. K., 1978, Magnetic Field Generation in Electrically Conducting

Fluids. Cambridge Univ. Press, Cambridge

Moss D., Tuominen I., Brandenburg A., 1990, A&A, 228, 284

Noyes R. W., Weiss N. O., Vaughan A. H., 1984, ApJ, 287, 769

Parker E. N., 1955, ApJ, 121, 293

182 J.-C. Thelen

q 2000 RAS, MNRAS 315, 165±183

Parker E. N., 1993, ApJ, 408, 707

Prautzsch T., 1993, in Proctor M. R. E., Matthews P. C., Rucklidge A. M.,

eds, Solar And Planetary Dynamos. Cambridge Univ. Press, Cambridge

Ribes J. C., Nesme-Ribes E., 1993, A&A, 276, 549

Roberts P. H., 1972, Phil. Trans. R. Soc. Lond. A, 272, 663

Schmalz S., Stix M., 1991, A&A, 245, 645

Schmitt D., 1985, PhD thesis, Univ. GoÈttingen

Schmitt D., 1987, A&A, 174, 281

Schmitt D., SchuÈssler M., 1989, A&A, 223, 343

Sokoloff D., Nesme-Ribes E., 1994, A&A, 228, 293

Spiegel E. A., Weiss N. O., 1981, Nat, 287, 616

Steenbeck M., Krause F., 1969, Astron. Nachr., 21, 1285

Thelen J.-C., 1998, PhD thesis, Univ. Leeds

Thelen J.-C., 2000, MNRAS, 315, 155 (this issue)

Tobias S. M., 1996, A&A, 207, L21

Tobias S. M., 1997, A&A, 322, 1007

Weiss N. O., 1993, in Proctor M. R. E., Gilbert A. D., eds, Lectures On

Solar and Planetary Dynamos. Cambridge Univ. Press, Cambridge

Weiss N. O., Cattaneo F., Jones C. A., 1984, GAFD, 30, 305

This paper has been typeset from a TEX/LATEX file prepared by the author.

Non-linear av-dynamos driven by magnetic buoyancy 183

q 2000 RAS, MNRAS 315, 165±183