Issue |
A&A
Volume 534, October 2011
|
|
---|---|---|
Article Number | A46 | |
Number of page(s) | 15 | |
Section | The Sun | |
DOI | https://doi.org/10.1051/0004-6361/201016108 | |
Published online | 30 September 2011 |
Alpha effect due to buoyancy instability of a magnetic layer
1
NORDITA, AlbaNova University Center,
Roslagstullsbacken 23,
10691
Stockholm,
Sweden
e-mail: piyalic@nordita.org
2
Department of Astronomy, AlbaNova University Center, Stockholm
University, 10691
Stockholm,
Sweden
Received:
5
November
2010
Accepted:
13
May
2011
Context. A strong toroidal field can exist in form of a magnetic layer in the overshoot region below the solar convection zone. This motivates a more detailed study of the magnetic buoyancy instability with rotation.
Aims. We calculate the α effect due to helical motions caused by an unstable magnetic layer in a rotating density-stratified system with angular velocity Ω making an angle θ with the vertical. We also study the dependence of the α effect on θ and the strength of the initial magnetic field.
Methods. We carry out three-dimensional hydromagnetic simulations in Cartesian geometry. A turbulent electromotive force (EMF) due to the correlations of the small scale velocity and magnetic field is generated. We use the test-field method to calculate the transport coefficients of the inhomogeneous turbulence produced by the layer.
Results. We show that the growth rate of the instability and the twist of the magnetic field vary monotonically with the ratio of thermal conductivity to magnetic diffusivity. The resulting α effect is non-uniform and increases with the strength of the initial magnetic field. It is thus an example of an “anti-quenched” α effect. The α effect is also nonlocal, i.e. scale dependent, requiring around 8–16 Fourier modes to reconstruct the actual mean EMF based on the actual mean field.
Key words: magnetohydrodynamics (MHD) / magnetic fields / turbulence / dynamo / instabilities
© ESO, 2011
1. Introduction
The magnetic fields in many astrophysical bodies have their origin in some kind of turbulent dynamo (Priest 1982). This means that a part of the kinetic energy of the turbulent motions is diverted to enhancing and maintaining a magnetic field. This magnetic field is generally also random, but under certain conditions a large-scale magnetic field can also emerge (Parker 1979). Here by large-scale we mean length scales larger than the energy containing scale of the fluid. In particular, large-scale fields may show up when the turbulence is helical, owing, e.g., to the simultaneous presence of rotation and stratification (Moffatt 1978).
The evolution of large-scale magnetic fields can be described by the mean-field equation that emerges upon averaging of the induction equation. In the process of averaging, new terms emerge (e.g., the α effect and turbulent diffusion) that result from correlations between small-scale velocity and magnetic fields (Krause & Rädler 1980). Here one usually considers the case where the magnetic fluctuations are caused by the fluctuating velocity acting on the mean field. However, under certain conditions it might well be the other way around. Imagine, for example, the case where initially no turbulent velocity is present, but there is instead a strong large-scale magnetic field the presence of which makes the initial state unstable. The magnetic field would then be responsible for driving velocity and magnetic fluctuations at the same time. This type of scenario was first simulated in the context of accretion discs where the magneto-rotational instability drives the turbulence (Brandenburg et al. 1995), and later in the context of the magnetic buoyancy instability with shear (Cline et al. 2003). The latter is relevant to the overshoot layer of the Sun. It had already been proposed by Moffatt (1978) that, once the dynamo-generated magnetic field in this layer reaches appreciable strengths, the magnetic buoyancy instability can set in and govern the dynamics thereafter. In this paper we consider a similar case in which an otherwise stable layer of rotating fluid is made unstable by the presence of a strong magnetic field. This is known as the magnetic buoyancy instability.
If the length scale of the initial magnetic field is larger than or of the same order as
the height of the fluid layer, the magnetic Boussinesq approximation is often used. Linear
stability analyses for such cases (Acheson 1979; Hughes 1985a,b)
have shown that the initial state is unstable to both two-dimensional perturbations in the
plane perpendicular to the initial magnetic field and three-dimensional perturbations. The
two-dimensional perturbations do not bend the the initial magnetic field lines and are known
as interchange modes. The three-dimensional perturbations are often called undular modes.
Here we concentrate on the other limit where the length scale of initial magnetic field is
significantly smaller than the height of the fluid layer. Such a magnetic field is also
unstable to both the interchange and undular modes. The linear phase of this instability
with stratification and rotation has been studied in detail by Schmitt (1984, 1985) and Ferriz-Mas et al. (1994) for ideal MHD. A necessary but
not sufficient condition for instability is (Eq. (3.3) of Acheson 1979) (1)which essentially means
that the magnetic field modulus B decreases faster with height
z than the density ρ. Fan (2001) has studied the linear stability of the same problem in absence of
rotation, yet in a parameter range where only the undular modes are unstable. Gilman (1970) considered the influence of large-scale
shear and the case where the thermal diffusivity is large compared to viscosity and magnetic
diffusivity. Silvers et al. (2009) have studied the
same but in the absence of rotation. While the focus of the Gilman (1970) study has been on the formation of flux tubes from a pre-existing
toroidal magnetic layer in a stably stratified atmosphere, in Silvers et al. (2009) a magnetic layer was generated from an initially
vertical magnetic field in presence of strong shear. It was further shown that, when the
ratio of magnetic to thermal diffusivities is sufficiently low, magnetic buoyancy can still
operate.
In this paper we are not only interested in the linear growth of this instability but also in the turbulent transport coefficients characterizing the turbulence driven by such an instability. There has been several earlier attempts to calculate the coefficient α for the magnetic buoyancy instability. Brandenburg & Schmitt (1998) performed numerical calculations in presence of rotation and determined α for the resulting turbulence. Thelen (2000) has used the linearly unstable eigenmodes to calculate α, but this calculation is not applicable to the fully-developed turbulent regime which is our main interest here. Wissink et al. (2000) calculated αyy in the nonlinear stage of the instability, but both works did not take into account the turbulent diffusivity. More recently Davies & Hughes (2011) have used the same technique as Thelen (2000) to compute the turbulent electromotive force (EMF). They, however, claim that the complexity of their results renders a discussion of the EMF in terms of α effect and turbulent diffusivity impossible. All these works rely upon fully compressible numerical calculations and hence are closest to the present one. In the concluding section of this paper we shall compare their results with ours.
The focus of this work is twofold. Firstly, we study the nature of the magnetic buoyancy instability in its initial linear stage. In particular, we study its dependence on various parameters such as magnetic and thermal Prandtl numbers, angular velocity, strength of the initial field, etc., and compare against the linear theory and previous numerical work. Secondly, to gain some insights on whether the instability constitutes a viable dynamo process, we measure the mean-field transport coefficients, namely the tensors α and η using the quasi-kinematic test-field (QKTF) method (Schrinner et al. 2005, 2007). Note that, with one exception (Vermersch & Brandenburg 2009), the QKTF method has not been applied previously to the calculation of transport coefficients for the case of an inhomogeneous turbulence induced by the mean magnetic field. Therefore we validate the QKTF method for this problem by reconstruction of the turbulent EMF from the turbulent transport coefficients. For a review on the transport coefficients and their determination using test fields; see Brandenburg et al. (2010).
![]() |
Fig. 1 The Cartesian simulation domain with respect to spherical coordinates. |
2. The model
We consider a setup similar to that described in Brandenburg
& Schmitt (1998). The computational domain is a cuboid with constant
gravity, gz, pointing in the negative
z direction, and rotation Ω making an angle
θ with the vertical. The box may be thought to be placed at a colatitude
θ on the surface of a sphere with its unit vectors
pointing along the local θ, φ, r
directions, respectively, as shown in Fig. 1.
We solve the following set of MHD equations. The continuity equation is given by
(2)where
D/Dt ≡ ∂/∂t + U·∇
denotes the Lagrangian derivative with respect to the local velocity of the gas
U. Assuming an ideal gas, we express the pressure in terms
of density, specific entropy s, and sound speed
cs, which, in turn, is a function of ρ
and s. Thus the momentum equation in a frame of reference rotating with
angular velocity Ω reads
(3)where
J is the current density, B
is the magnetic field, ν is the constant kinematic viscosity, and
S is the traceless rate-of-strain tensor. The sound speed is
related to temperature by
with
cp and cv the specific heat at
constant pressure and constant volume, respectively, and
γ = cp/cv
is here fixed to 5/3. The induction equation is solved in terms of the
magnetic vector potential A, such that
∇ × A = B,
hence
(4)where
η denotes constant molecular magnetic diffusivity.
Finally, we have for the entropy equation with temperature T and constant
radiative (thermal) conductivity K(5)where the temperature
is related to the specific entropy by
(6)We use the fully
compressible Pencil Code1 for all our
calculations.
For all quantities, periodic boundary conditions in the x and y directions are adopted. In the z direction we use the no-slip boundary condition for the velocity and the vertical field condition Bx = By = 0, that is, Jz = 0, for the magnetic field as a proxy for vacuum boundaries. We keep the temperature at the top and the (radiative) heat flux at the bottom fixed. Their values were chosen to conform with the initial temperature profile of the (not magnetically modified) polytrope described below.
2.1. Initial state
The base state is a polytrope that is,
p = CρΓ, with index
m = 1/(Γ − 1) = 3. The initial z
profiles of density, pressure, temperature and entropy are given by
(7)where
Φ is a non-dimensional gravitational potential given by
with the reference
point z0 chosen to be at the bottom of the domain and the
values at this point given by ρ0,
,
and s0. Here,
cs0 is the reference sound speed to which we also refer to
when calculating Mach numbers.
As the adiabatic index here is mad = 1/(γ − 1) = 3/2, the subadiabaticity in the domain is very large, namely ∂lnT/∂lnP − (∂lnT/∂lnP)ad = −0.15. Thus, the initial stratification is highly stable to convection in the absence of any magnetic field, guaranteeing that turbulence is generated solely by the buoyancy instability.
The initial magnetic field is a horizontal layer of thickness
2HB, where
By has the profile
(8)and the reference Alfvén
speed is defined by
with μ0 being the vacuum permeability. If not indicated
otherwise, the initial magnetic field strength is fixed to
vA0/cs0 = 0.5.
In order to satisfy the condition (1)initially, we have to ensure
2HB < Hρ(zB),
where
Hρ(z) = |∇lnρ(z)|-1
is the local density scale height. When choosing
zB − z0 = 0.3Lz
this is satisfied for
2HB < 0.1Lz + 4T0(cp − cv)/3|gz|
which is surely true for the choice
HB = 0.05Lz.
Upon addition of a magnetic field, we modify the base state such that the density profile
remains unchanged. In order to obey magnetostatic equilibrium, pressure and temperature
are adjusted in the following way: (9)The entropy is
then re-calculated from Eq. (6). This leads
to a local minimum in the temperature profile which would smooth out over a diffusion time
scale
, being in any
case much larger than the e-folding time of the instability. Another
common choice is to keep the temperature profile unchanged, but this implies a decrease of
the density in the magnetic layer, making it immediately buoyant in presence of small
perturbations.
The initial velocity components Ux and
Uy are specified such that they contain
about 20 localized eddies in the plane
z = zB with Mach numbers
of about 10-5. Also the initial vertical velocity,
Uz is Gaussian random noise with the same
Mach number and the rms value of the initial kinetic helicity, scaled with the product of
initial rms velocity and vorticity, , is
4 × 10-6. Here and in the following angular brackets mean volume averaging.
Non-dimensional control parameters characterizing the buoyancy instability.
2.2. Control parameters, non-dimensional quantities, and computational grid
The problem posed by (2)through (5)is governed by five independent dimensionless
parameters; (i) the Prandtl number
Pr = ν/χ0, with the
temperature conductivity
χ0 = K/ρ0cp;
(ii) the magnetic Prandtl number
PrM = ν/η; (iii) the
“magnetic Taylor number” ; (iv) the
rotational inclination (colatitude), θ, and (v) the normalized
gravitational acceleration
. In addition there are
two independent parameters of the initial equilibrium (vi) the normalized pressure scale
height at the bottom,
and (vii) the
initial Lundquist number,
Lu0 = vA0HB/η,
based upon the thickness of the magnetic layer. In addition to this, we also have the
non-dimensional sound speed,
cs0Ly/η.
In this paper we shall keep the normalized pressure scale height and the sound speed
fixed, while varying both Prandtl numbers, TaM, θ and
Lu0. The definitions as well as the values or ranges of the control
parameters are summarized in Table 1. We have also
included in the same table two dependent parameters namely the Roberts number
Rb = PrM/Pr = χ0/η
and the modified initial plasma beta
in the midplane of the
magnetic layer. It is defined as the ratio of the total pressure
ptot = p + pM
to the magnetic pressure
, because
then adopts a simple
1/B2 dependence on the magnetic field;
cf. Eq. (9).
The computational domain is defined by
|x| ≤ Lx/2,
|y| ≤ Ly/2,
−Lz/4 ≤ z ≤ 3Lz/4,
Lx = Lz = Ly/3,
thus its aspect ratio is 1:3:1. The results will be presented in non-dimensional form,
velocity in units of the reference Alfvén speed, vA0, time in
units of the corresponding Alfvén travel time in the y direction,
tA0 = Ly/vA0,
and magnetic field in units of B0 or the rms value
.
It is instructive to look upon the relevant definitions of the fluid Reynolds number, Re,
and the magnetic Reynolds number, Rm, for this problem where
the turbulence is driven solely by the instability of the magnetic layer. From first
principles, Re characterizes the ratio of the advective term
⟨ (U·∇U)2 ⟩ 1/2
and the approximate viscous term
⟨ (ν∇2U)2 ⟩ 1/2
in the Navier-Stokes equation, while Rm characterizes the
ratio of
⟨ (∇ × (U × B))2 ⟩ 1/2
and
⟨ (η∇2B)2 ⟩ 1/2
in the induction equation with the angular brackets representing volume averaging. Let us
denote these ab initio definitions as “term based” and refer to them by Re ∗
and . Note that
with the term-based definitions Rm/Re may
well deviate from PrM. Alternatively, we can define a length scale
LU = Urms/2πWrms
from the rms values of velocity and vorticity and define the more conventional
length-based Reynolds numbers
Re = UrmsLU/ν
and
Rm = UrmsLU/η.
The calculations were carried out on equidistant grids with resolutions of either 643 or 1283. For numerical testing we have also performed a few runs with 2563 or 1282 × 256 resolutions.
List of runs with high resolution.
2.3. The test-field method
Here, for the sake of completeness, we give a concise description of the test-field method we use to calculate the turbulent transport coefficients. For a more detailed discussion we refer the reader to Brandenburg et al. (2008c) and Rheinhardt & Brandenburg (2010).
Let us define mean magnetic and velocity fields,
and
,
where overbars denote Reynolds averaging. Fluctuations are defined correspondingly as
and
.
Following these conventions, the induction equation may be correspondingly averaged,
resulting in
(10)where
is the mean EMF resulting from the fluctuating fields u and
b. The essence of mean-field magneto-hydrodynamics is to
provide an expression for
as a
functional of the large scale magnetic field and its derivatives. A simple ansatz reads
(11)where
α and η are called
transport coefficients. Note that a much more general representation of
is
given by the convolution integral
(12)with an
appropriate tensorial kernel G or Green’s function. When
considering Fourier transforms with respect to x′
(and now dropping the time dependence) we have
(13)with
being the Fourier transform of
G(x,x′)
with respect to x′. As shown in Brandenburg et al. (2008c) a decomposition in the form
(14)with
real
and
is always achievable. Further, the traditional (local) coefficients are given by
and
.
The aim of the test-field method is to provide an expression for
G as a functional of properties of the mean and
fluctuating fluid velocity, and
u by utilizing direct numerical simulations. Subtracting
the averaged induction Eq. (10)from the
original one (4), written in terms of
B rather than A, we obtain
the following equation for the fluctuating magnetic field b.
(15)with
. The
superscripts pq now indicate that this equation is solved for several
suitably chosen test fields
with
p,q = 1,2. This is sufficient as we define the mean as
horizontal (xy) average. Thus
, and
only the components αij and
ηij3,
i,j = 1,2, are of interest. In the following we will
hence refer to the rank-2 tensor
ηij = −ηik3ϵjk3
only which allows to rewrite the second term in (11)into
.
Equation (15) is the central relation
invoked by the test-field method. The test-field suite of the Pencil Code
has the provision for using either harmonic test fields, i.e.,
(16)or linear test fields i.e.,
(17)Harmonic test fields with varying wavenumber k allow us to
calculate for each k the 2 × 2 tensors
and
by solving an algebraic system of 8 equations given by
(18)As shown above,
expressing the mean EMF by
and
avoids the shortcomings of a truncation like in and provides the desired non-local
relationship between
,
and
.
In the present context, the buoyancy-driven turbulence and hence the transport
coefficients necessarily show an intrinsic inhomogeneity both due to stratification and
the mean magnetic field itself. No specific complication is induced by it as
and
emerge straightforwardly from the testfield method as functions of z and
k.
In the nonlinear situation, the Green’s function approach remains valid if
is
considered as a functional of U and
which
is then linear and homogeneous in the latter. However, we have to label
G by the
actually acting upon U, that is,
, and can thus only make
statements about the transport tensors for just the particular
at
hand. Hence, the tensors have to be labelled likewise:
,
.
3. Results
3.1. Nature of the instability
To start with we have performed a number of runs with different values of Pr and
PrM, but all other dimensionless parameters held fixed, see Table 2. In particular, we have used a value of
TaM = 3.24 × 1010 for the magnetic Taylor number and
Lu0 = 500 for the initial Lundquist number. Table 2 shows the Reynolds numbers according to the two alternative
definitions provided in Sect. 2.2. Note that with the
exception of the run B128f, Re from the length-based and the term-based definitions are in
agreement. Also the ratio from the
term-based definitions approaches PrM reasonably.
We first show the temporal evolution of the magnetic field for a few representative cases in Fig. 2. In all of them, we can clearly distinguish a first stage of exponential growth from a subsequent saturation phase. The x and z components of the magnetic field are generated at the expense of its y component. Although there exists a persistent energy source in the form of a constant heat flux into the domain, the final saturated stage always undergoes a slow decay. This decay is most clearly visible in By. Thus the instability is not able to maintain a dynamo on its own.
![]() |
Fig. 2 Time evolution of the runs in Table 2.
Upper panel: mean squared values of velocity and generated
magnetic field components Bx,
Bz of run B128c
(Pr = PrM = 1) scaled by |
We suppose that the magnetic layer formed by Bx, though having a vertical scale suited to maintain the instability, is eventually not strong enough to take over the role of the initial magnetic layer. Let us first discuss the initial linear stage of the instability.
![]() |
Fig. 3 Top: velocity components Uy (in color), Ux and Uz (vectors) in the plane y = 0. Bottom: Ux (in color), Uy and Uz (vectors) in the plane x = 0. Both during the linear evolution phase of the run B128a. The initial condition for U was set to be harmonic in x and y with kx = 8 × 2π/Lx and ky = 2π/Ly. |
3.2. Linear stage
At first we verify that the instability is indeed driven by magnetic buoyancy. As the
coefficients in Eqs. (2)–(5)are constant, the initial state (7)depends only on z, and the
boundary conditions in the x and y directions are
periodic, all eigensolutions
ψ = (ρ,u,b,s)
of the linearized problem must have the form (19)where m
and n are integers and
ω = ωR + iωI.
For n ≠ 0 the undular modes are obtained. Corresponding dispersion
relations ω(m,n) have been found by Fan (2001), using variational minimization of energy,
and by Schmitt (1985), using eigenmode
decomposition of linearized MHD equations. The former considers the non-rotating case, and
the latter the rotating one. Both works use the anelastic approximation and their results
are strictly applicable only for ideal MHD. Fan
(2001) reports only non-oscillatory modes. Schmitt (1985), who further uses the magnetostrophic approximation, finds that
the growing modes also have an oscillatory part, with the ratio
ωR/ωI
decreasing with latitude. Note however that the results of Schmitt (1985) do not directly apply to the present case because our boundary
conditions are different2.
For the runs in Table 2 we find that the early exponential growth phase is dominated by a mixture of several eigenmodes having almost the same growth rates, but different x wavenumbers, 5 ≤ m ≤ 8, whereas for all of them n = 1. This is further substantiated by the occasional occurrence of a beating in the x dependence with a modulation period given by m = 1 which indicates the presence of two almost equally strong eigenmodes with m and m ± 1. By setting the initial conditions to be harmonic in x and y it is also possible to select single eigenmodes cleanly. Figure 3 shows the velocity pattern of the fastest growing mode with m = 8, n = 1. According to the terminology of Hughes (1985a) we may qualify our eigenmodes as undular as they change periodically in the direction of the initial magnetic field. Our results are consistent with the findings of Thelen (2000) where for moderate rotation the fastest growing mode had always the smallest possible (non-vanishing) wavenumber in the direction of the field while the wavenumber perpendicular to the field was high.
The growth rates ωI = ℑ(ω) presented in Table 2, have to be considered as average values for a (not precisely known) number of eigenmodes. While the ωI could be easily identified from the averaged quantities shown in Fig. 2, it is difficult to access the oscillation frequencies ℜ(ω). This is because they are small compared to the growth rates and saturation sets in too early to allow for the observation of a complete oscillation period. Nevertheless, some indications for temporal variations in the eigenmode geometries have been found. Generally, we observe an increase of the growth rate with increasing magnetic Prandtl number, but a decrease with increasing Prandtl number. We find that the growth rate increases with the Roberts number as shown in Fig. 4. This means that increasing the efficiency of heat conduction in comparison to magnetic diffusion destabilizes the sub-adiabatic stratification in the system in agreement with the destabilizing effect of thermal diffusion studied by Acheson (1979).
![]() |
Fig. 4 Dependence of growth rate ωI on the inverse Roberts number derived from the runs in Table 2. Solid line: best linear fit. Size of circles codes for the value of Rm (length based, see Table 2). |
3.3. Dependence on initial magnetic field and rotation
Another piece of evidence for the magnetic character of the instability is its dependence
on the initial magnetic field strength. From Fig. 5
we see a clear increase of the growth rate and saturation level with decreasing
, that is, increasing
Lu0, while keeping the rotation rate fixed at
TaM = 3.24 × 1010 (see also the last three runs in Table 3 below). Schmitt
(2003) predicted in the magnetostrophic approximation a growth rate ∝
inversely proportional to
.
Next we keep constant at 1.51 and
decrease ζ gradually from 4.09 × 105 to 0. Inspecting
Fig. 5, we find that growth rate and saturation
level of
increase monotonically and
reach their maxima at ζ = 0 (Ω = 0) while the saturation time is
decreasing. The impeding effect of rotation onto the instability at large Ω is plausible
in view of the Taylor-Proudman theorem because the unstable eigenmodes do show pronounced
z gradients in U, see Fig. 3.
![]() |
Fig. 5 Dependence of the instability on initial magnetic field strength, expressed by
|
3.4. Saturated stage
At later time the instability reaches saturation, characterized by turbulent magnetic, velocity, density and temperature fields, that decay slowly thereafter. However, in most of the analysis below, this decay will be ignored and the turbulence considered as approximately statistically stationary. The turbulence is necessarily both inhomogeneous and anisotropic and we shall further show that it is also helical. Under such conditions we expect the emergence of a mean EMF. Indeed magnetic fields perpendicular to the initial magnetic layer are produced having non-vanishing horizontal averages.
In order to give a better idea of the 3D geometry of the magnetic field we provide in Fig. 6 a volume rendering of By at a time after tsat for the runs B128e and B128f (see Table 2) which differ only in their magnetic Prandtl numbers. Notice how the magnetic layer breaks into flux tubes – similar to what is seen in Fig. 3 of Fan (2001) and also in Matthews et al. (1995). More relevantly, it is also similiar to the findings of Wissink et al. (2000) who include rotation. The difference between the two runs is most striking in the nature of corrugation in the surface shown. We attribute it to larger twist in the rising tubular structures in the run B128e compared to B128f which becomes clearly visible in the field line pictures also depicted in Fig. 6 (right).
Figure 7 demonstrates the breakup of the magnetic layer into tubular structures of concentrated magnetic field which are also regions of low density, hence rising. Notice also the high density regions just above and below these tubular structures. They show a significantly lower temperature than their surroundings (Fig. 7, bottom).
Studies of rising flux tubes in simulations of the solar convection zone have shown that
their ability to rise depends crucially on how twisted they are (see, e.g., Emonet & Moreno-Insertis 1998). Hence it
behooves us to measure the twist of the large-scale magnetic field in our simulations. For
a quantitative measurement we utilize the dimensionless parameter
ϵJ = ⟨ J·B ⟩ /JrmsBrms,
the relative current helicity, essentially measuring the overall degree
of alignment between B and J.
Here, angular brackets denote volume averages. The corresponding horizontally averaged
quantity is . Figure 8 shows the normalized current helicity density
J·B/JrmsBrms
(filled contours) as well as By in the plane
y = 0 for run B128e. Notice that the contours are bend leftward because
of the Coriolis force with
Ω·ẑ > 0.
The contour plots of By in this figure also
show the formation of rising tubular structures from the magnetic layer.
In Fig. 9 we show the dependence of ϵJ on Rb and the profile of εJ(z) for some selected runs. The latter quantity can be as strong as 30% near to the initial location of the magnetic sheet, but the total helicity reaches only values of a few per cent. The clear dependence of ϵJ on Rb is in contrast to the only weak dependences on PrM and Pr individually. This is an important result from this section. Our conjecture is that, at large Rb, the magnetic buoyancy instability may play an important role in the formation of twisted flux tubes in the Sun, where Rb ≫ 1 is expected.
![]() |
Fig. 6 Left: volume rendering of the By = 0.1B0 isosurface. Right: field lines, colored according to the value of vorticity Wy for runs B128f (top) and B128e (bottom) at t = 2tA0 (saturated stage). The contours of By are shown on side walls. |
To demonstrate the emergence of a mean magnetic field we present in Fig. 10 time-height plots of
and
for the run B128g
(note that
).
There, t ≈ 1.6 tA0 marks the end of the
exponential growth phase after which a strong growth of
, obviously at the
expense of
, sets in.
reaches its
maximum around t ≈ 3tA0 and is then subject
to the overall decay. Note the strong vertical concentration of
, approximately
antisymmetric about the midplane of the magnetic sheet.
3.5. Calculation of turbulent transport coefficients
The turbulence resulting from the buoyancy instability generates a mean magnetic field
component from an initial
which is also
modified compared to its initial shape (see Fig. 10). We employ the quasi-kinematic test-field method to calculate
transport coefficients like the α and
η tensors which describe this process. So far, the
method has mostly been applied in situations where a hydrodynamic background was already
present in absence of the mean magnetic field (see, e.g. Brandenburg et al. 2008a,b,c). Here, in contrast, the (magnetohydrodynamic)
turbulence results entirely from the instability of a pre-existing mean magnetic field,
By0(z). In other words,
our simulations do not posses a kinematic stage in which the influence of
would
be negligible. The applicability of QKTF to such a situation has been questioned by Courvoisier et al. (2010). We emphasize, however, that
Eq. (15)continues to be valid also in
that case and hence all conclusions drawn from it, because the decisive criterion for its
validity is that there exists no magnetic turbulence in the absence of the mean
magnetic field, that is,
b → 0 if
(see Rheinhardt & Brandenburg 2010, for a more
thorough explanation). As this is the case here the QKTF method is applicable.
Furthermore, we explicitly check the applicability of QKTF to this problem by
reconstructing the turbulent EMF from the turbulent transport coefficients we calculate.
The only peculiarity occurring is the fact that all components of
α and η vanish for
, because
fluctuating velocity and magnetic fields develop only after the instability has set in.
Another aspect not considered in most previous test-field studies is the strong intrinsic
inhomogeneity of the turbulence not only as a consequence of the strong z
dependence of
, but
also due to the stratified density background. Thus the transport coefficients need to be
determined as z dependent quantities.
In the next section we demonstrate that the test-field method works reasonably well for our problem. Note that to calculate the transport coefficients in addition to the usual MHD equations four additional evolution equations of the form (15)for four independent test-fields have to be solved. Hence the test-field runs are computationally almost thrice as expensive. We have thus reduced resolution to 643 grid points for all these runs.
![]() |
Fig. 7 Relative density perturbation, δρ/ρi (top) and relative temperature perturbation δT/Ti (bottom), with δρ = ρ − ρi, δT = T − Ti and ρi(z), Ti(z) taken from Eq. (7), in the plane y = 0 at t = tsat ~ 2tA0 for the run B128h. Both plots overlaid with contours of By (solid lines). |
![]() |
Fig. 8 Top: relative current helicity J·B/JrmsBrms for run B128d at t = tsat ~ 2tA0; arrows: vx, vz. Bottom: By/B0 for the same run; arrows: Bx, Bz. Both panels show the plane y = 0. |
![]() |
Fig. 9 Top: dependence of the total relative current helicity
ϵJ on inverse Roberts number. Size of
circles codes for value of Rm (length based).
Bottom: dependence of |
![]() |
Fig. 10 Time-height diagram for |
3.5.1. Reconstruction of the mean EMF
To validate the test-field method we first confirm that the quantity
,
taken directly from the direct numerical simulation (DNS), can be reproduced by
employing the relation (11)between
and
with the tensors
and
determined using the quasi-kinematic test-field method. In mathematical terms,
(20)with
(21)where
the superscript R indicates reconstruction. Here, the boundary condition for
B gives rise to the selection of discrete cosine and
sine modes with wavenumbers kc and
ks, respectively. The additional argument
is
to indicate that the quantities
,
related to
in Eq. (13), as well as the tensors
and
,
are valid just for that mean field
which is present in the main run. As a consequence, the reconstruction of the mean EMF
can be successful only when employing exactly this
in
(21). That is, the mean field
representation of the turbulence in terms of
and
has, at this level, merely descriptive rather than predictive potential. In order to
possess the latter, α and η
have to be formulated as functionals of
and
z. Consequently, the mean EMF will appear as a nonlinear functional
of
.
Since it is additionally also non-local, we are confronted with a task
of considerable mathematical complication which is not addressed here.
Let us denote as the
reconstructed EMF according to Eq. (20)
truncated by k′ ≤ K′, with
kc,s = 2πk′/Lz.
Here k′ can now take both integer and half-integer values
where the integer (half-integer) values of k′ correspond to
the family of sine (cosine) modes in Eq. (21). An initial estimate of K′ required for a
reasonable reconstruction of
was
obtained from the power spectra of both
and
. It turned out
that
has significant
spectral power up until k′ = 16, whereas for
the power
spectrum has levelled off already at k′ = 8.
The components of the tensors
and
also show rather different spectral behavior, both in the midplane of the magnetic layer
and near the midplane of the box as seen in Fig. 11. Given that the spectrum of the mean field clearly converges to zero, this
guarantees that the mean EMF will also spectrally converge. The tensor components either
appear to converge to a constant value or to zero. For coefficients like
and
,
converging to a constant, this asymptotic, say
,
can in principle be separated and would appear as a part
of the convolution kernel
of (12), representing a local
part of the relation between
and
.
This is, however, not done here. Instead we simply employ the k
dependent coefficients up to the maximum required by the spectra of
,
that is, k′ = 16. Note that the values for
k′ = 0 are not relevant here as, due to the boundary
conditions,
does not possess a k′ = 0 contribution.
The α tensor can be decomposed in symmetric and antisymmetric parts,
the latter being representable by a turbulent pumping velocity
γ, which gives rise to the term
in the mean EMF. Since
, the
only relevant component of γ is here
.
Analytic results indicate that in a wide range of situations, the turbulent pumping is
directed away from the region of strong turbulence (“turbulent diamagnetism”, see Krause & Rädler 1980).
is presented in Fig. 11 together with
in one panel. As
decreases in modulus beyond k′ = 1 and even changes sign at
k′ = 8 while
increases in modulus, but preserves its sign, we observe a sign change in
at k′ = 4 and k′ = 3 in the
midplane of the magnetic layer and above it, respectively (see also Fig. 18 below). A similar dependence of turbulent pumping
on wavenumber has been found by Käpylä et al.
(2009) in DNS of convection. Consequently, magnetic fields with scales up to
≈ (5...6)HB, that is,
k′ > 4 are pumped into the layer,
but those with larger scales are pumped away from it the former being contrary to the
standard concept of “turbulent diamagnetism”. It is thus difficult to comment on the
transport of the total
by
.
Only if the pumping were oriented away from the magnetic layer for all the wavenumbers
of the dominating constituents in
it
would lead to a broadening of the initial layer i.e., a reduction of
and would hence inhibit the
instability. With respect to its saturation, however, the strong turbulent magnetic
diffusion given by
is likely to be more important, as for low k′ it reaches
≈ 20 times the molecular value, cf. Fig. 11.
The result of the assembly of from (20)with (21), is presented in Fig. 12,
middle column. From simple visual inspection we find it to be a faithful reproduction of
from the DNS shown in the left column. Clearly, a naive application of the test-field
procedure with harmonic test fields with only the lowest
k′ = 0.5 results in an inadequate description as shown in
the right column.
![]() |
Fig. 11 Dependence of |
We define two measures for the quality of the mean EMF reconstruction at a given
K′ namely and the
correlation coefficient rK′
defined as
(22)where the subscript
“z,t” denotes that the averaging has been carried out over the
vertical coordinate z as well as over the temporal range
1.2tA0 ≤ t ≤ 3.4tA0.
The relative error of the reconstruction,
, and the
correlation coefficient, rK′,
are plotted in Fig. 13 as a function of the
truncation wavenumber K′. The
reach a
minimum value of about 35% and 30% for
and
, respectively,
and level off around K′ = 8. This implies that including
higher harmonic test fields beyond k′ = 7 does not improve
the reconstructed EMF. Similarly,
rK′ for
converges to a value of 0.98 (0.93) at K′ = 4 (8). It is
important to note that even though the tensor components
and
do not converge with increasing k′, the reconstructed EMFs
do as a consequence of the sufficiently fast convergence of
with k′. Also, including transport coefficients for
k′ ≥ 8 does not improve the reconstruction any further.
Clearly, one reason behind the discrepancies is that we have neglected memory effects in
the turbulent transport coefficients (see Hubbard
& Brandenburg 2009). It can be particularly important in the present
situation as we are not in a statistically stationary regime. Apart from this,
enhancement of resolution might further improve the results.
![]() |
Fig. 12 Reconstruction of the mean EMF for the run TF30+ using
|
![]() |
Fig. 13 Quality of the mean EMF reconstruction as a function of the truncation wavenumber
K′: |
3.5.2. Dependence of the transport tensors on inclination
In view of the solar dynamo problem it is important to look
at
and
as functions of the rotational inclination θ or latitude
λ = 90° − θ. We expect the growth rate of
the instability to increase from the equator to the pole (Schmitt 2003). This can be explained by the buoyant nature of the
turbulence, for which vertical motions are essential: At the poles, the effect of the
Coriolis force on them is weakest whereas they are strongly deflected at the equator.
This is indeed confirmed by Fig. 14, where the
growth rate is seen to decrease continuously when changing θ from
0° (pole) towards 90° (equator). The corresponding runs are
summarized in Table 3.
Let us now consider the symmetry properties of our solution with respect to
λ = 0, which is the solar equator. Moving from the northern
hemisphere at λ to the southern at − λ, that is
changing θ to 180° − θ, but keeping all
other problem parameters invariant, is equivalent to inverting the sign
of Ωz. The same can be accomplished by reflecting the
corresponding rigid rotation about the plane x = 0. Hence we can
construct the solution
(ρ,U,B,s)
of (2)–(5)for − λ simply by reflecting the solution for
λ properly about the same plane. Under this reflection, polar vectors
like velocity transform as, (23)and axial vectors like
the magnetic field as
(24)(Note that the
gravitational acceleration is invariant under this reflection.) Hence, for the initial
magnetic field, By0(z),
the transition to − λ requires only a sign inversion. However, since
the induction equation is linear in B, and Lorentz force
as well as Ohmic dissipation are quadratic, inverting the sign of
By0(z) would just
transform the solution
{ ρ,U,B,s }
to
{ ρ,U, − B,s } ,
that is, would leave the turbulence essentially unchanged and can be omitted. Moreover,
as the transport coefficients which express the correlation properties of the turbulent
velocity u are functions of z only, the
reflection operation cannot change their magnitudes. With respect to possible sign
inversions in the coefficients upon reflection, we note that
and
,
being polar vectors, invert the sign of their x components under
reflection, but keep their y components unchanged. The axial vector
behaves just the opposite way. Thus, we have
for i = 1,2 (no summation) and
for i ≠ j, whereas
for i = 1,2 and
for i ≠ j when moving from λ to
− λ. Consequently, it appears that the results for the southern
hemisphere can be derived from those for the northern by simple operations. Strictly
speaking however, this is only true when the initial condition for
U is also reflected upon the transition from
λ to − λ. From a naive point of view we might
suppose that omitting this reflection can hardly be of any importance, because the
initial conditions are anyway random. But we have found this not to be true.
In Fig. 15 we show
z–λ plots of and
,
constructed from the runs with Ω ≠ 0. We expect
and
to be
antisymmetric about the equator. This is indeed true for higher latitudes but not in the
neighbourhood of the equator where they are rather symmetric. Violation of the expected
symmetry properties can also be inferred from the extrema of
given in
Table 3.
Symmetry can formally be restored as follows. From Fig. 15 we construct another set of plots by “symmetrizing”
and
, but
“anti-symmetrizing”
and
, that is,
setting
to
and
to
, analogously for
.
The result is shown in Fig. 16. In the same way,
we have obtained the transport coefficients shown in Figs. 17 and 18 for the test-field
wavenumber k′ = 0.5, that is, for the largest possible
wavelength (cosine mode). We find that the moduli of all turbulent transport
coefficients increase monotonically for all z as we go from the equator
to the poles, except
and
which exhibit local extrema at |λ| ≈ 30°.
List of runs with test-field calculations.
![]() |
Fig. 14 Dependence of the instability on rotational inclination θ in
terms of rms value of generated field components
|
A careful look at the “symmetrized” and “anti-symmetrized” plots reveals a
discontinuity at the equator best seen in the plots of
(Figs. 17). This phenomenon is counterintuitive
at first: if we regard the computational box as a local area within a spherical body,
changing λ from positive to negative values implies travelling along a
meridian across the equator. If quantities like
and
change their
sign there, they should vanish for continuity reasons and the same should hold true for
and
.
Instead of looking at the “symmetrized” quantities we can reflect the initial velocity
according to
(23)upon the transition
λ → − λ. This restores exact
symmetry/antisymmetry about the equator and we again observe a discontinuity at the
equator.
In conclusion, near the equator the initial conditions play a crucial role in
determining the saturated values of and
and
consequently the values of the turbulent transport coefficients. As the coefficients in
are typically connected with the kinetic helicity density
hK, we have calculated it for the saturated states. Indeed,
at the equator the solutions resulting from the initial condition
and its
reflected counterpart show opposite signs of hK. This is an
example of spontaneous symmetry breaking with bifurcation into helical solutions of
opposite handedness which enables finite values of
and
as well as of
at the equator. (Note, that
are not related to the non-vanishing helicity at the equator and consequently show no
discontinuity in their anti-symmetrized profiles in Fig. 17.) As a prerequisite for such a bifurcation, the unstable eigenmodes of the
buoyancy instability must be degenerate. That is, two linearly independent modes with
opposite signs of helicity, but identical growth rate must exist. This has been noticed
for Ω ≠ 0 at the equator (Thelen 2000) and also
holds true for Ω = 03. A detailed discussion of
this phenomenon for Ω = 0 is the subject of a separate paper Chatterjee et al. (2011).
![]() |
Fig. 15
|
![]() |
Fig. 17
|
![]() |
Fig. 18 Pumping velocity |
3.5.3. Dependence on initial magnetic field strength
In standard mean-field theory for a prescribed hydrodynamic background the turbulent
transport coefficients usually decrease as the mean magnetic field increases
(“quenching”). The present problem is different, however, because the instability and
hence the turbulence is just caused by the initial (mean) magnetic field. In Fig. 19 we show the z averages of
and
at tsat as functions of the initial magnetic field
.
These components have been selected because they essentially do not invert sign in the
domain during the entire evolution. Because of its importance for the generation of
from
we have also
given
in this figure. However, as this coefficient shows a significant sign change with
respect to z, its maximum value with respect to z and
the time span
0 < t < 2tsat
was plotted.
Clearly, ,
,
and
increase with the initial magnetic field strength supporting earlier ideas of a possible
“anti-quenching” in the case of the buoyancy instability (see, e.g., Brandenburg et al. 1998). Note, however, that the
dependence on the initial field strength in Fig. 19 might differ from the dependence on the local
field strength to which the term “quenching” usually refers.
![]() |
Fig. 19 Vertical averages of |
4. Conclusions
We have studied in detail the generation of the α effect due to the buoyancy instability of a horizontal magnetic layer in a stratified atmosphere by using direct numerical simulations. We find that both the magnetic energy and the current helicity (twist) in the system increase monotonically with the ratio of thermal conductivity to magnetic diffusivity, i.e. the Roberts number Rb (Fig. 4). This agrees with earlier analytical work of Gilman (1970) and Acheson (1979), as well as numerical work of Silvers et al. (2009), who find that efficient thermal diffusion or heat exchange can destabilize a stable stratification. The increase of the resulting twist with Rb is an important result since the buoyancy instability would thus produce twisted flux tubes from a magnetic layer, if it existed in the overshoot layer of the Sun where Rb ≫ 1. Vasil & Brummel (2008) also report the formation of twisted flux tubes from a horizontal magnetic layer, but in their case it is due to the action of shear on a weak vertical magnetic field. We further find that the growth rate of the buoyancy instability is reduced in presence of rotation compared to the case with Ω = 0.
Most of the earlier work on the magnetic buoyancy instability resorted to either the magnetic Boussinesq approximation or the anelastic approximation. Here we have performed fully compressible numerical calculations (for differences between anelastic and compressible calculations see also Berkoff et al. 2010). Let us now relate our results to corresponding earlier calculations. The work of Wissink et al. (2000) comes closest to our own as they not only included rotation, but also employed a thin magnetic layer. However, both thermal and magnetic boundary conditions were different from ours and the initial magnetostatic equilibrium was established by lowering the density in the layer, thus making it immediately Rayleigh-Taylor unstable. Moreover, their simulations were done for different parameters (our values in brackets): Rb = 17 [0.25...1] (bottom), density contrast 5.8 [223] , TaM ≈ 107 [1010] and Lu0 ≈ 50 [500...600]. Hence, a quantitative comparison is not possible, although the occurrence of similar tubular structures in the saturated stage (see our Fig. 6 and their Fig. 4) indicates that those might be a robust feature of magnetic-buoyancy driven turbulence.
Thelen (2000) has also performed fully compressible
calculations and provides the α coefficient, but only as a result from
individual unstable eigenmodes in the linear stage. In contrast, our calculations of
α and magnetic diffusivity were done in the saturated stage and are hence
not comparable to the results of Thelen (2000).
Furthermore, in his analysis, the relation
αyy ⟨ B0y ⟩ = ⟨ u × b ⟩ y
was employed with ⟨ · ⟩ being either a volume or horizontal average. In the first case it
yields the correct αyy, but this is valid only
for uniform mean fields. In the second case the result is not a measure of the actual
αyy, because the contribution of
to
is ignored. Hence, in both cases no description of the mean EMF that would be useful for a
mean-field dynamo model is provided. Further, he admits only a local relationship between
and
.
Kersalé et al. (2007) considered a non-rotating box with a linear vertical profile of the horizontal magnetic field, which is supported by the boundary condition. Perturbations to this state were subjected to perfect conductor boundary conditions, and the background state was a polytrope with index 1.6 along with Pr = 1/Rb = 0.02. Hence, both their model and the employed parameters are too different from ours to allow a meaningful comparison of the results. Nevertheless, their magnetic field structure for random initial conditions (their Fig. 4) resembles our finding in Fig. 6.
We have run our simulations only until the time taken by the initial magnetic layer to break up due to the back-reaction of the unstable modes and ohmic diffusion. In the absence of any other forcing such as a strong shear, the buoyancy instability is found to be incapable of sustaining itself past the break-up phase since the scale height of the magnetic layer becomes comparable to that of the density. We may hence say that strong shear is not imperative to the production of tubular structures from the magnetic layer, but is likely to play a key role in keeping the layer from breaking up. It may also be possible that turbulent downward pumping (Nordlund et al. 1992) arrests the decay of a magnetic layer in the overshoot region. However, it is not yet clear if such a layer exists at all in the real Sun and, moreover, if it is subject to the buoyancy instability there.
For mean fields defined by horizontal averaging we have “measured” the turbulent transport
coefficients using the technique of the quasi-kinematic test-field method. In order to prove
that the
and
tensors obtained from this method are reasonably accurate, we have verified the agreement
between
from the DNS and the representation of
by the convolution
, with
corresponding tensors
and
obtained for harmonic test
fields with wavenumbers
0 ≤ (Lz/2π)k ≤ 16.
A technique for the quantitative assessment of this agreement is presented. We find that,
even in presence of magnetically driven turbulence, the obtained
and
provide a reasonably accurate description of the turbulent EMF; see Figs. 12 and 13 (although
the relative errors are still about 30 ... 35%). This is an
important outcome of our study.
We find that , determined using
a harmonic test field with the lowest wavenumber that fits in the vertical extent of the
box, already comprises a considerable part of the total EMF. Hence it can be enlightening to
look at the turbulent transport coefficients obtained by the QKTF method for this wavenumber
only. At finite Ω their dependence on (solar or stellar) latitude λ is of
particular interest. The component
contributes to the generation of
from the strong
initial field
in the layer. The
off-diagonal components of
contribute to a vertical turbulent pumping velocity directed away from the region of
turbulence surrounding the magnetic layer. The effect of this pumping systematically
broadens magnetic structures with increasing latitude. We find that all transport
coefficients except
and
increase with latitude and are significantly reduced near the equator due to the suppressing
effect of the Coriolis force on the instability. For the first time the turbulent magnetic
diffusivity given by the diagonal components of
has been computed for the magnetic buoyancy instability; see Fig. 17. In particular, near the magnetic layer, the diagonal component
is 25 times larger than the molecular value η.
The buoyancy instability has the property that the EMF, as indicated by the extrema of
, increases
progressively with the magnitude of the magnetic field in the horizontal layer (compare
solid and dashed lines in Fig. 5). Indeed, as a
remarkable result, we demonstrate that the most significant coefficients
,
,
and
increase with the initial field strength; see Fig. 19.
This feature makes the buoyancy instability an attractive candidate for generating an
α-effect inside the Sun. Unlike turbulent convection which yields an
α that is quenched for strong mean magnetic fields, this one increases
with increasing field strength. Our findings support suggestions by Brandenburg et al. (1998) that, if turbulent transport coefficients are
caused by flows that are magnetically driven like in the present case or, e.g., in the
magneto-rotational instability, then both α and η may
increase with magnetic field strength. This trend is sometimes referred to as
“anti-quenching” and could be useful to explain the steepness of the observational relation
between the ratio of dynamo cycle to rotation frequencies,
ωcyc/Ω and inverse Rossby number for
stellar data (Brandenburg et al. 1998; Saar & Brandenburg 1999).
Although highly desirable, modelling of α and η as functionals of position and the mean magnetic field for use in a predictive mean field dynamo model is a difficult proposition that needs to be postponed to future work.
Note also that Hughes (1985a,b) uses the magnetic Boussinesq approximation to obtain the linear stability diagram for the case where the characteristic scale of the magnetic field is larger than the depth of the fluid layer. In this paper we consider the opposite limit, i.e., the limit of a thin magnetic layer.
Another example of spontaneous symmetry breaking in magnetohydrodynamics has been recently observed in the case of the Tayler instability of a purely toroidal magnetic field (Gellert et al. 2011).
Acknowledgments
We thank Alexander Hubbard for reading the manuscript carefully. The computations have been carried out on the National Supercomputer Centre in Linköping and the Center for Parallel Computers at the Royal Institute of Technology in Sweden. This work was supported in part by the European Research Council under the AstroDyn Research Project No. 227952 and the Swedish Research Council Grant No. 621-2007-4064.
References
- Acheson, D. J. 1979, Sol. Phys. 62,23 [Google Scholar]
- Berkoff, N. A., Kersalé, E., & Tobias, S. M. 2010, GAFD, 104, 545 [Google Scholar]
- Brandenburg, A. 1998, in Theory of Black Hole Accretion Discs, ed. M. A. Abramowicz, G. Björnsson, & J. E. Pringle (Cambridge University Press), 61 [Google Scholar]
- Brandenburg, A., & Schmitt, D. 1998, A&A, 338, L55 [NASA ADS] [Google Scholar]
- Brandenburg, A., Nordlund, Å., Stein, R. F., & Torkelsson, U. 1995, ApJ, 446, 741 [NASA ADS] [CrossRef] [Google Scholar]
- Brandenburg, A., Saar, S. H., & Turpin, C. R. 1998, ApJ, 498, L51 [NASA ADS] [CrossRef] [Google Scholar]
- Brandenburg, A., Rädler, K.-H., & Schrinner, M. 2008a, A&A, 482, 739 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Brandenburg, A., Rädler, K.-H., Rheinhardt, M., & Käpylä, P. J. 2008b, ApJ, 676, 740 [NASA ADS] [CrossRef] [Google Scholar]
- Brandenburg, A., Rädler, K.-H., Rheinhardt, M., & Subramanian, K. 2008c, ApJ, 687, L49 [NASA ADS] [CrossRef] [Google Scholar]
- Brandenburg, A., Chatterjee, P., Del Sordo, F., et al. 2010, Phys. Scr., T142, 014028 [Google Scholar]
- Chatterjee, P., Mitra, D., Brandenburg, A., & Rheinhardt, M. 2011, Phys. Rev. E, 84, 025403(R), DOI: 10.1103/PhysRevE.84.025403 [NASA ADS] [CrossRef] [Google Scholar]
- Cline, K. S., Brummell, N. H., & Cattaneo, F. 2003, ApJ, 599, 1449 [NASA ADS] [CrossRef] [Google Scholar]
- Courvoisier, A., Hughes, D. W., & Proctor, M. R. E., 2010, Proc. Roy. Soc. Lond., 466, 583 [Google Scholar]
- Davies, C. R., & Hughes, D. W. 2011, ApJ, 727, 112 [NASA ADS] [CrossRef] [Google Scholar]
- Emonet, T., & Moreno-Insertis, F. 1998, ApJ, 492, 804 [NASA ADS] [CrossRef] [Google Scholar]
- Fan, Y. 2001, ApJ, 546, 509 [NASA ADS] [CrossRef] [Google Scholar]
- Ferriz-Mas, A., Schmitt, D., & Schüssler, M., 1994, A&A, 289, 949 [NASA ADS] [Google Scholar]
- Gellert, M., Rüdiger, G., & Hollerbach, R. 2011, MNRAS, 414, 2696 [NASA ADS] [CrossRef] [Google Scholar]
- Gilman, P. A., 1970, ApJ, 162, 1019 [NASA ADS] [CrossRef] [Google Scholar]
- Hubbard, A., & Brandenburg, A. 2009, ApJ, 706, 712 [NASA ADS] [CrossRef] [Google Scholar]
- Hughes, D. W. 1985a, GAFD, 32, 273 [CrossRef] [Google Scholar]
- Hughes, D. W. 1985b, GAFD, 34, 99 [CrossRef] [Google Scholar]
- Käpylä, P. J., Korpi, M., & Brandenburg, A. 2009, A&A, 500, 633 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Kersalé, E., Hughes, D. W., & Tobias, S. M. 2007, ApJ, 663, L113 [NASA ADS] [CrossRef] [Google Scholar]
- Krause F., & Rädler K.-H. 1980, Mean-Field Magnetohydrodynamics and Dynamo Theory (Oxford: Pergamon Press) [Google Scholar]
- Matthews, P. C., Hughes, D. W., & Proctor, M. R. E. 1995, ApJ, 448, 938 [NASA ADS] [CrossRef] [Google Scholar]
- Moffatt, H. K. 1972, J. Fluid Mech., 53, 385 [Google Scholar]
- Nordlund, Å., Brandenburg, A., Jennings, R. L., et al. 1992, ApJ, 392, 647 [NASA ADS] [CrossRef] [Google Scholar]
- Parker, E. N. 1979, Cosmical magnetic fields (Oxford: Clarendon Press) [Google Scholar]
- Priest, E. R. 1982, Solar Magnetohydrodynamics (D. Reidel, Dordrecht) [Google Scholar]
- Rheinhardt, M., & Brandenburg, A. 2010, A&A, 520, A28 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Saar, S. H., & Brandenburg, A. 1999, ApJ, 524, 295 [NASA ADS] [CrossRef] [Google Scholar]
- Schmitt, D. 1984, in ESA, ed. The Hydromagnetics of the Sun (N85-25091 14-92), 223 [Google Scholar]
- Schmitt, D. 1985, Dynamowirkung magnetostrophischer Wellen (Ph.D. Thesis, University of Göttingen) [Google Scholar]
- Schmitt, D. 2003, in Advances in nonlinear dynamos, ed. A. Ferris-Mas, M. Núñez Jinénez (Taylor & Francis), 83 [Google Scholar]
- Schrinner, M., Rädler, K.-H., Schmitt, D., Rheinhardt, M., & Christensen, U. 2005, Astron. Nachr., 326, 245 [NASA ADS] [CrossRef] [Google Scholar]
- Schrinner, M., Rädler, K.-H., Schmitt, D., Rheinhardt, M., & Christensen, U. 2007, GAFD, 101, 81 [Google Scholar]
- Silvers, L. J., Vasil, G. M., Brummel, N. H., & Proctor, M. R. E. 2009, ApJ, 702, L14 [NASA ADS] [CrossRef] [Google Scholar]
- Thelen, J. C. 2000, MNRAS, 315, 155 [NASA ADS] [CrossRef] [Google Scholar]
- Vasil, G. M., & Brummel, N. H., 2008, ApJ, 686, 709 [NASA ADS] [CrossRef] [Google Scholar]
- Vermersch, V., & Brandenburg, A. 2009, Astron. Nachr., 330, 797 [NASA ADS] [CrossRef] [Google Scholar]
- Wissink, J. G., Hughes, D. W., Matthews, P. C., & Proctor, M. R. E. 2000, MNRAS, 318, 501 [NASA ADS] [CrossRef] [Google Scholar]
All Tables
All Figures
![]() |
Fig. 1 The Cartesian simulation domain with respect to spherical coordinates. |
In the text |
![]() |
Fig. 2 Time evolution of the runs in Table 2.
Upper panel: mean squared values of velocity and generated
magnetic field components Bx,
Bz of run B128c
(Pr = PrM = 1) scaled by |
In the text |
![]() |
Fig. 3 Top: velocity components Uy (in color), Ux and Uz (vectors) in the plane y = 0. Bottom: Ux (in color), Uy and Uz (vectors) in the plane x = 0. Both during the linear evolution phase of the run B128a. The initial condition for U was set to be harmonic in x and y with kx = 8 × 2π/Lx and ky = 2π/Ly. |
In the text |
![]() |
Fig. 4 Dependence of growth rate ωI on the inverse Roberts number derived from the runs in Table 2. Solid line: best linear fit. Size of circles codes for the value of Rm (length based, see Table 2). |
In the text |
![]() |
Fig. 5 Dependence of the instability on initial magnetic field strength, expressed by
|
In the text |
![]() |
Fig. 6 Left: volume rendering of the By = 0.1B0 isosurface. Right: field lines, colored according to the value of vorticity Wy for runs B128f (top) and B128e (bottom) at t = 2tA0 (saturated stage). The contours of By are shown on side walls. |
In the text |
![]() |
Fig. 7 Relative density perturbation, δρ/ρi (top) and relative temperature perturbation δT/Ti (bottom), with δρ = ρ − ρi, δT = T − Ti and ρi(z), Ti(z) taken from Eq. (7), in the plane y = 0 at t = tsat ~ 2tA0 for the run B128h. Both plots overlaid with contours of By (solid lines). |
In the text |
![]() |
Fig. 8 Top: relative current helicity J·B/JrmsBrms for run B128d at t = tsat ~ 2tA0; arrows: vx, vz. Bottom: By/B0 for the same run; arrows: Bx, Bz. Both panels show the plane y = 0. |
In the text |
![]() |
Fig. 9 Top: dependence of the total relative current helicity
ϵJ on inverse Roberts number. Size of
circles codes for value of Rm (length based).
Bottom: dependence of |
In the text |
![]() |
Fig. 10 Time-height diagram for |
In the text |
![]() |
Fig. 11 Dependence of |
In the text |
![]() |
Fig. 12 Reconstruction of the mean EMF for the run TF30+ using
|
In the text |
![]() |
Fig. 13 Quality of the mean EMF reconstruction as a function of the truncation wavenumber
K′: |
In the text |
![]() |
Fig. 14 Dependence of the instability on rotational inclination θ in
terms of rms value of generated field components
|
In the text |
![]() |
Fig. 15
|
In the text |
![]() |
Fig. 16 Same as Fig. 15, but with
symmetrized |
In the text |
![]() |
Fig. 17
|
In the text |
![]() |
Fig. 18 Pumping velocity |
In the text |
![]() |
Fig. 19 Vertical averages of |
In the text |
Current usage metrics show cumulative count of Article Views (full-text article views including HTML views, PDF and ePub downloads, according to the available data) and Abstracts Views on Vision4Press platform.
Data correspond to usage on the plateform after 2015. The current usage metrics is available 48-96 hours after online publication and is updated daily on week days.
Initial download of the metrics may take a while.