Issue |
A&A
Volume 508, Number 3, December IV 2009
|
|
---|---|---|
Page(s) | 1117 - 1133 | |
Section | Astrophysical processes | |
DOI | https://doi.org/10.1051/0004-6361/200912879 | |
Published online | 04 November 2009 |
A&A 508, 1117-1133 (2009)
MHD simulations of accretion onto a dipolar magnetosphere
I. Accretion curtains and the disk-locking paradigm
C. Zanni1,2 - J. Ferreira2
1 - INAF - Osservatorio Astronomico di Torino, Strada Osservatorio 20, 10025, Pino Torinese, Italy
2 -
Laboratoire d'Astrophysique de Grenoble, 414 rue de la Piscine,
BP 53, 38041 Grenoble, France
Received 13 July 2009 / Accepted 30 September 2009
Abstract
Aims. We investigate the accretion process from an accretion
disk onto a magnetized rotating star with a purely dipolar magnetic
field. Our main aim is to study the mechanisms that regulate the
stellar angular momentum. In this work, we consider two effects that
can contrast with the spin-up torque normally associated with
accretion: (1) the spin-down torque exerted by an extended
magnetosphere connected to the disk beyond the corotation radius;
(2) the spin-down torque determined by a stellar wind flowing
along the opened magnetospheric field lines.
Methods. Our study is based on time-dependent axisymmetric
magnetohydrodynamic numerical simulations of the interaction between a
viscous and resistive accretion disk with the dipolar magnetosphere of
a rotating star. We present the first example of a numerical experiment
able to model at the same time the formation of accretion curtains, the
effects of an extended stellar magnetosphere and the launching of a
stellar wind.
Results. In the examples presented, the spin-down torque related to the star-disk interaction can extract only
of the accretion torque, due to the weakness of the extended
connection. Not even the spin-down torque exerted by a stellar wind is
strong enough (
): despite a huge lever arm (
), the mass-loss rate (
)
is too low to provide an efficient torque.
Conclusions. We argue that, at least in the case of typical classical T Tauri stars (
,
)
rotating at
of
their break-up speed, the disk spin-down is unlikely to balance the
accretion torque (``disk locked'' equilibrium). A massive stellar
wind (
)
could in principle succeed, but its mass and energy fluxes are quite
demanding, both from a theoretical and an observational point of view.
Key words: stars: rotation - stars: magnetic fields - accretion, accretion disks - ISM: jets and outflows - methods: numerical - magnetohydrodynamics (MHD)
1 Introduction
Different classes of magnetized stellar-type objects (T Tauri protostars, white dwarfs, neutron stars) are often surrounded by accretion disks and it is likely that the stellar magnetosphere is able to disrupt the inner part of the disk and control the accretion flow down to its surface. According to this picture, the accretion disk is truncated at a few stellar radii by the interaction with the stellar magnetic field, the accretion flow is channeled into funnel flows which follow the magnetospheric field lines and terminates with a shock on the stellar surface. This general scenario has been applied to a variety of different astrophysical systems: X-ray pulsars (Bildsten et al. 1997), magnetic cataclysmic variables (Warner 2003) and classical T Tauri stars (CTTS, Bouvier & Appenzeller 2007). In the particular case of CTTS, which will be the main subject of our investigation, this scenario is supported by much observational evidence: CTTS are magnetically active protostars, characterized by surface magnetic fields of a few kG (Johns-Krull 2007); both CO line profiles (Najita et al. 2003) and infrared colors (Kenyon et al. 1996) are indicative of emission from a Keplerian accretion disk truncated at a few stellar radii; inverse P-Cygni profiles with strong redshifted absorption are interpreted as due to free-falling gas flowing along magnetospheric field lines anchored at the inner disk edge (Edwards et al. 1994); both the optical excess and UV continuum can be fitted by accretion-shock models (Gullbring et al. 2000).
The study of different aspects of this scenario has been based on analytical models (Ghosh & Lamb 1979a,b; Collier Cameron & Campbell 1993; Koldoba et al. 2002; Li & Wilson 1999; Kluzniak & Rappaport 2007; Wang 1995) while, more recently, numerical simulations have aimed at giving a global view of the magnetic star-disk interaction (Bessolaz et al. 2008; Goodson et al. 1999; Miller & Stone 1997; Romanova et al. 2002; Küker et al. 2003; Long et al. 2005). All models agree on the fact that, for accretion to occur, the truncation radius
of a Keplerian disk is likely to be located inside the corotation radius
,
since accretion would be inhibited by a centrifugal barrier if the
inner disk edge rotates slower than the stellar angular speed
(``propeller'' regime, e.g. Illarionov & Sunyaev 1975).
Assuming that later on the disk angular momentum is transported from
the truncation region to the star along the magnetospheric field lines,
the star experiences a spin-up torque of the order of
.
On the other hand, many of these accreting systems show a peculiar
angular momentum evolution. Many X-ray pulsars are characterized by
several episodes of spin-up/spin-down torque reversal (Nelson et al. 1997). CTTS are characterized by slow rotation periods (3-10 days, Bouvier et al. 1993), corresponding to
of their break-up speed: this means that a large fraction of the
stellar angular momentum already has been extracted during the embedded
phase. Moreover, the rotation period seems to stay constant during the
T Tauri phase (Irwin et al. 2007),
despite the fact that the protostar is still actively accreting and
contracting. In all these systems an efficient mechanism of angular
momentum removal is therefore required.
Different processes have been proposed to balance the spin-up torques due to accretion. In the classical Ghosh & Lamb picture, a spin-down torque is exerted along the magnetospheric field lines connecting the star to the disk beyond the corotation radius, where the disk rotates slower than the star: if the spin-down balances the spin-up torque, the stellar rotation is ``disk-locked''. This scenario often has been advocated to explain the spin-down of accreting stars, both in the case of X-ray pulsars (Ghosh & Lamb 1979b; Campbell 1987) and young protostars (Armitage & Clarke 1996; Yi 1995; Collier Cameron & Campbell 1993; Königl 1991). On the other hand the limits of the ``extended magnetosphere'' scenario have been discussed by several authors: in a general way, the magnetic field twisting due to the star-disk differential rotation leads to an inflation of the magnetospheric structure, in order to relax the build-up of toroidal field pressure. Uzdensky et al. (2002) have shown that, if the poloidal field distribution does not change in time at the midplane of the disk, the inflation occurs at mid-latitudes and, after a maximum twist has been reached, the magnetic structure opens up and at least part of the magnetic connection with the star is lost. On the other hand, if the poloidal field is able to diffuse quickly enough inside the disk, Bardou & Heyvaerts (1996) have shown that field inflation happens mostly along the midplane of the disk: in this case, the star-disk connection can in principle be kept, but the radial diffusion strongly reduces the poloidal field intensity. Both mechanisms lead to a reduction of the spin-down torque associated with the extended magnetosphere: the first one by lowering the maximum attainable value of the field twist and the size of the connected region, as discussed by Matt & Pudritz (2005a), the second one by decreasing the intensity of the poloidal field, as investigated by Agapitou & Papaloizou (2000). The relative importance of the two phenomena depends on the ratio between the inflation and the diffusion timescales.
Other proposed mechanisms are based on the presence of outflows, extracting angular momentum from the central parts of the star-disk system. Stellar winds, similarly to relativistic pulsar winds, can remove angular momentum directly from the star along the opened field lines of the magnetosphere. On the other hand, since the rotational energy of the slowly rotating star is not enough to power these outflows, an additional energy input is required (Ferreira et al. 2006), possibly related to the energy deposited by accretion onto the stellar surface (``accretion powered'' stellar winds, Matt & Pudritz 2005b). The efficiency of the wind braking torque is therefore deeply affected by the energetics of the outflow and the opened magnetic configuration (i.e. the magnetic lever arm).
Different classes of outflows loaded with disk material can
play an important role too. The original idea of the ``X-Wind''
scenario involves an outflow launched along the opened magnetic field
lines which thread the disk due to the inflation and opening of the
closed magnetosphere (Cai et al. 2008; Shu et al. 1994):
according to this picture, such an outflow could be able to extract a
large fraction of the disk angular momentum before it falls onto the
star so that accretion would exert a null torque. This scenario also
envisages that the launching region of the ``X-Wind'' and the
truncation radius of the disk are located in a tiny region (of the
order
)
around the corotation radius, which represents a saddle
(``X'' point) of the gravito-centrifugal potential.
The ``reconnection X-winds'' proposed by Ferreira et al. (2000)
are fed with disk material but powered by the stellar rotational
energy, thus exerting an effective spin-down torque without being
affected by the energetic problems of purely stellar winds: this model
requires the presence of a large scale magnetic field threading the
disk which reconnects with the stellar magnetic field corresponding to
an X-point located at the disk midplane. Romanova et al. (2009)
have proposed a somewhat different class of disk-loaded outflows
(``conical'' winds), but their impact on the angular momentum evolution
of the central star is still unclear.
This paper is the first of a series of works dedicated to the numerical modeling of different spin-down mechanisms of magnetized rotating stars. Our axisymmetric models will be based on time-dependent magneto-hydrodynamic (MHD) simulations of the interaction between a viscous and resistive accretion disk with a dipolar magnetosphere aligned with the rotation axis of the central star. At least in the case of CTTS this is a particular case, since spectropolarimetric measurements indicate that the topology of the magnetosphere is often complex and multipolar, where the dipolar component is not the dominant one (Donati et al. 2007,2008). Also, the polar axes are misaligned with respect to the rotation axis of the star. On the other hand, the dipolar component is the one that can affect the disk dynamics to a larger extent: the dipolar field strengths and mass accretion rates that we will employ are consistent with observations of strongly magnetized CTTS like BP Tau (Donati et al. 2008). Moreover, our arguments will be presented in a very general way so that they can be applied, at least qualitatively, to any class of magnetized accreting stars.
In this paper, we present the first numerical time-dependent global simulations of the ``extended magnetosphere'' scenario originally depicted by Ghosh & Lamb: this is the first example of a dynamical simulation of the interaction between an accretion disk and a stellar magnetosphere connected to a large extent of the disk, below and beyond the corotation radius. This allows us to treat both the spin-up and the spin-down torques associated with the star-disk magnetic interaction and discuss their efficiency. At the same time, we are able to include in the models the effects of a stellar wind and discuss its properties and characteristics. In Sect. 2 we present the numerical methods that we employ, the initial setup and the boundary conditions of our numerical experiments. In Sect. 3 we describe the angular momentum balance of the star-disk system: we consider, in succession, the torques acting on the accretion disk (Sect. 3.1); the formation of accretion funnels, the associated angular momentum transfer (Sect. 3.2) and the effective torque acting on the star, taking into account both the interaction with the accretion disk and the stellar wind (Sect. 3.3). The analysis of the electric currents, presented in Sect. 4, allows us to describe the magnetic coupling between the different parts of the system. In Sect. 5 we discuss the efficiency of the different spin-down mechanisms while in Sect. 6 we summarize the results and draw our conclusions.
2 Numerical simulations
2.1 MHD equations
We numerically solve the MHD system of equations, including resistive and viscous terms:
This includes the equations of mass, momentum and energy conservation and the induction equation (Faraday's law). The equations have been solved in nondimensional form, thus without






where








![]() |
(2) |
where


A cooling term
is included to remove the viscous and Ohmic heating. For simplicity,
radiative transfer is not taken into account and the cooling is treated
as optically thin radiative losses. This term has been mainly
introduced to avoid the irreversible heating and the consequent thermal
thickening of the accretion disk and is not meant to be representative
of the actual radiative mechanism acting inside the disk. As a
consequence of its presence, the dynamics are adiabatic. This is a
caveat that is generally common to current MHD simulations of this
kind. Including radiative transfer and exploring thermal effects within
the disk is a very demanding task that remains to be done.
The cooling function
becomes effective whenever the entropy of the disk plasma becomes
greater
than its initial value. This value was found to be a good compromise
between avoiding runaway irreversible heating of the disk and
preventing the thermal heightscale of the disk from becoming too small
and numerically under-resolved.
The system of Eqs. (1) is solved numerically exploiting the MHD module provided with the PLUTO code (Mignone et al. 2007).
PLUTO is a modular Godunov-type code aimed at solving the equations of
hydrodynamics and magneto-hydrodynamics in both classical and special
relativistic regimes. The classical MHD module has been configured
to perform second-order piecewise linear reconstruction of primitive
variables, with a Van Leer limiter for the density and
magnetic field components and a minmod limiter for the thermal pressure
and velocity components. To compute the intercell fluxes,
a Roe-type Riemann solver has been employed, while second order in
time has been achieved using a Runge-Kutta scheme. The constrained
transport (CT) approach of Balsara & Spicer (1999) has been used to control the solenoidality of the magnetic field (
).
Since the stellar magnetosphere is modeled initially as a force-free
potential magnetic field, we adopted a ``field-splitting'' technique (Tanaka 1994; Powell et al. 1999)
to compute only the deviation from the initial magnetic field: this
approach is crucial to correctly treat current-free strong magnetic
fields numerically. The viscous and resistive terms have been treated
explicitly, using a second-order finite difference approximation for
the dissipative fluxes and checking the diffusive timestep. The
simulations have been carried out in 2.5 dimensions, that is, in
spherical coordinates (
)
assuming axisymmetry around the rotation axis of the disk and the star.
In the following we will indicate the spherical radius with the capital
letter R, while the cylindrical radius
will be marked by the lower case.
2.2 Initial conditions
The initial conditions are made up of three parts: the viscous
accretion disk, a surrounding rarefied corona and the stellar magnetic
field. The accretion disk is set following the three-dimensional models
of Keplerian accretion disks originally developed in
Kluzniak & Kita (2000) and further discussed by Regev & Gitelman (2002) and Umurhan et al. (2006).
These are polytropic (
)
solutions obtained through an expansion up to terms of order
of the equations of viscous hydrodynamics, where
is the disk aspect ratio, given by the ratio between the isothermal sound speed
and the Keplerian speed
evaluated on the midplane of the disk. With these assumptions and taking
,
the density and thermal pressure are determined by the vertical equilibrium of the disk
![]() |
(3) |
![]() |
(4) |
where







where



The magnetic Prandtl number

![]() |
(7) |
The rotation speed is determined by the radial equilibrium of the disk:
![]() |
(8) |
with

The rotation of the disk is therefore slightly sub-Keplerian: the small


Finally the meridional flow is purely radial (spherical) with
![]() |
(9) |
As pointed out by Kluzniak & Kita (2000), for


The corona is modeled as a polytropic (
)
hydrostatic atmosphere. Its density and pressure are therefore given by the expressions:

where


Finally the stellar magnetosphere is modeled initially as a purely
dipolar field aligned with the rotation axis of the star-disk system.
The magnetic field is defined by the flux function
:
![]() |
(10) |
where



The disk surface is defined by the pressure equilibrium


To reduce the effects of the strong transients due to the initial differential rotation between the disk and the corona, the magnetic surfaces anchored in the disk are set to rotate at the Keplerian angular speed calculated at the radius at which they enter the disk. The magnetic surfaces anchored on the star but not entering the disk rotate at the same angular speed as the star. Obviously this velocity correction puts the atmosphere slightly out of equilibrium: we therefore add an additional small correction to the coronal density to nullify the centrifugal effects in the direction perpendicular to the magnetic surfaces.
![]() |
Figure 1:
Appearance of the initial conditions of the simulations. Colors are representative of the logarithmic density in units of
|
Open with DEXTER |
2.3 Computational domain and boundary conditions
We restrict our study to axisymmetric models assuming also a planar
symmetry with respect to the disk midplane. The two-dimensional
computational domain therefore encompasses an angular sector going from
the rotation axis (
)
to the disk midplane (
).
Suitable boundary conditions are set to impose the required symmetries.
The sector is delimited by an inner and an outer radius, the inner one
corresponding to the stellar radius
.
The angular coordinate
has been discretized with
points.
In order to have grid cells with approximately the same volume, the
radial coordinate has been discretized on a stretched grid with a
spacing
.
Following this constraint, we used a radial grid made up of NR = 214 points, equivalent to a grid outer radius
.
Particular attention has been devoted to impose the boundary conditions
on the stellar surface, which we want to model as a perfect conductor
rotating at an angular speed
.
This means that in the reference frame co-rotating with the star, the
electric field is zero or, in other words, that the flow speed is
parallel to the magnetic field:
The poloidal magnetic field is forced to maintain its initial (dipolar) value: this is a reasonable assumption as long as the matter flow at the stellar surface is sub-Alfvénic, thus slightly affecting the magnetic structure. We used a linear extrapolation of density and thermal pressure along the field lines to impose the boundary values of these variables.
Following Eq. (11), the condition
imposes that the poloidal velocity must be parallel to the poloidal magnetic field (
).
On the outflowing material we imposed a continuity of the speed value
along the field lines, while for the infalling material we imposed the
continuity of the axisymmetric MHD invariant
.
This condition ensures that the infalling plasma is perfectly absorbed by the ``star''. The condition
is also used as a boundary condition for the CT method.
The condition on the poloidal electric field given by Eq. (11) imposes that the matter must rotate with a toroidal speed
Notice that the matter accreting along the funnels (







where




Notice that, since the magnetic surfaces anchored inside the disk are
set initially in corotation with it, the initial rotation rate of
the field lines is different from the stellar angular speed
and therefore introduces a differential rotation at the stellar surface.
To minimize the effects due to this differential rotation, the stellar rotation speed in Eq. (13) is initially set to the value
corresponding to the initial rotation of the magnetic surfaces. It is then gradually brought to the value
on a time scale corresponding to two stellar rotations, i.e.
.
This gradual acceleration of the stellar rotation initially triggers a
weak torsional Alfvén wave propagating from the star to the disk
surface. Finally the toroidal speed in the ghost cells is set to
.
At the outer radius delimiting the computational domain we imposed a
power-law extrapolation for density and pressure and a linear
extrapolation for all the other variables. Moreover, to impose the
boundary condition on the toroidal field along the open field lines
anchored on the surface of the star, we used an approach similar to the
one adopted on the stellar boundary. We only replaced the ``coupling''
time
of Eq. (13)
with a few times the Alfvén crossing time of the entire computational
domain. This condition ensures that the outer boundary would not exert
any artificial torque on the central star, even when the plasma
crossing this boundary is not outflowing at a super-Alfvénic speed.
2.4 Units and normalization
As customary, we normalized the MHD system of Eqs. (1) and the initial conditions presented in Sect. 2.2 to perform the calculations. We employed the stellar radius
as the unit length while, given the stellar mass
,
the velocities
have been normalized on the Keplerian speed at the stellar radius
.
Finally, by assigning the reference density
,
the unit for the magnetic field is given by
.
Consequently, the unit time is
,
mass accretion (or ejection) rates will be expressed in units of
,
powers in units of
and torques in units of
.
To make a direct application to young stars, we report here a standard normalization for a star of mass
,
radius
and taking
10-11 g cm-3.
![]() |
|||
![]() |
|||
![]() |
|||
![]() |
|||
![]() |
(14) |
2.5 The simulation parameters
Once they are normalized, the initial and boundary conditions presented in Sect. 2.2 depend on six dimensionless parameters: the disk thermal aspect-ratio ,
the equatorial stellar field strength
,
the stellar rotation rate
,
the density contrast between the disk and the corona
,
the viscous and resistive coefficients
and
.
In the simulations presented in this paper we assume a disk aspect-ratio
and a stellar magnetic field
.
In the standard ``YSO'' units given in Sect. 2.4 this corresponds to a field intensity
G.
Since in the initial conditions we truncate the disk at the radius
where the magnetic field is in equipartion with the thermal pressure of
the disk, these two parameters also determine the initial ``fiducial''
truncation radius
:

The star is taken to be rotating at one tenth of its break-up speed (





The aim of this paper is to numerically test the feasibility and the
efficiency of an ``extended magnetosphere'' scenario, where the stellar
magnetosphere connects to the disk below and beyond the corotation
radius. As discussed in Sect. 1,
the build-up of the toroidal field pressure due to the differential
star-disk rotation can lead to the inflation and opening of the
magnetosphere (Uzdensky et al. 2002), thus
reducing the size of the connected region. A high disk resistivity
can prevent, at least in part, this phenomenon so that the
magnetic connection can be maintained even beyond
.
In fact, the field line opening typically happens when a maximum
critical field twisting is attained: the disk resistivity allows some
azimuthal slippage of the field lines
relative to the disk material, thus reducing the growth of the twist
and, at least in some part of the disk, preventing it from reaching its
critical value. Moreover, the disk resistivity allows some radial
diffusion of the poloidal field through the disk, so that the magnetic
structure can also expand radially while keeping its connection. In
order to obtain the largest magnetospheric connection possible, we
therefore set the parameter
.
The viscosity parameter
is also set to unity (
). As discussed in Sect. 2.2, a value
allows the disk to accrete across its entire height. The viscous accretion rate
of the Kluzniak & Kita (2000) solution scales approximately with
and
as:
![]() |
Figure 2:
Time evolution of logarithmic density maps in units of
|
Open with DEXTER |
We will present results obtained with four different simulations characterized by the same parameters and different ``stellar'' boundary conditions: the reference simulation, whose boundary conditions have been described in Sect. 2.3, will be discussed in Sects. 3 and 4, while the simulations done to test different boundary conditions will be presented in Appendix A.
![]() |
Figure 3:
Left panel: radial behavior of the rotation speed of the accretion disk
|
Open with DEXTER |
3 An extended magnetosphere model
The time evolution of the reference simulation is shown in Fig. 2.
Several features can be noticed: (1) the simulation reaches a
quasi-stationary state (approximately after 15 stellar rotation
periods); (2) the formation of accretion funnel flows emerging
from the accretion disk truncated at
;
(3) an extended magnetosphere connecting the star and the disk up to
,
well beyond the corotation radius located at
;
(4) a stellar wind flowing along the opened magnetic field lines anchored on the surface of the star.
The main aim of this simulation is to study the angular momentum balance in a magnetized disk-star system characterized by a magnetosphere extending beyond the corotation radius. In Sect. 3.1 we will analyze the torques acting on the disk and discuss the disk truncation and the formation of the funnel flows. The angular momentum exchange between the star and the disk along the funnel flows will be presented in Sect. 3.2. The angular momentum balance of the star determined by the star-disk and the star-wind coupling will be discussed in Sect. 3.3.
3.1 Torques acting on the disk
The angular momentum balance can be studied by inspecting the angular momentum conservation equation in the system (1). By integrating this equation over the full disk thickness
and assuming stationarity (
)
we can write the following expression:
![]() |
(16) |
which expresses the angular momentum conservation inside an annulus of the disk of radial width

The differential torque:
![]() |
(17) |
quantifies the angular momentum advected through a disk ring

![]() |
(18) |
while

![]() |
(19) |
describes the outward radial transport of the angular momentum inside the disk itself. The specific magnetic torque is defined as:
where



conveys the angular momentum advected by the material extracted from an annulus




![[*]](/icons/foot_motif.png)
In the right panel of Fig. 3 we plot as a function of the radius the differential torques acting on the disk at a time equivalent to orbits of the central star.
It is possible to see that at large radii (
)
the accretion is controlled by the viscous torque
where it equals the accretion torque
.
We can also see in the left panel of Fig. 3 that in this region the rotation of the disk stays perfectly Keplerian. Since the accretion is controlled by an
viscous torque, the typical sonic Mach number of the accretion flow is
,
therefore strongly subsonic.
We can see that beyond the corotation radius
up to the last connected magnetic surface
the magnetic torque
is negative:
the star is transferring angular momentum to the disk due to the
fact that in this region the disk is rotating slower than the star.
Moreover, notice that this torque is effective only over a small radial
extension beyond the corotation radius. Nevertheless, the strength of
this part of magnetospheric torque is lower than the viscous one and
therefore the disk is able to cross the centrifugal barrier determined
by the stellar rotation. The magnetic torque changes sign at
and becomes the dominant braking torque at
,
where the viscous and the magnetic torques are equal. Notice that at this location the condition
still holds. This radius is sometimes taken as an estimate of the truncation radius (Armitage & Clarke 1996; Collier Cameron & Campbell 1993; Matt & Pudritz 2005a): our simulation shows that this can give an upper limit to the actual truncation radius with an error up to
.
On the other hand it is true that, if no outflows are present, the star
will basically accrete all the angular momentum owned by the disk at
this radius (Matt & Pudritz 2005a). In the region in which the magnetic torque is dominant,
takes the typical shape of a ``magnetically torqued'' disk (Kluzniak & Rappaport 2007): the stellar rotation forces the disk material to corotate at
and therefore the disk rotation becomes strongly sub-Keplerian. All the
missing angular momentum is obviously transferred to the star. When the
magnetic torque dominates,
increases and becomes almost unity (Bessolaz et al. 2008):
subsequently, when the magnetic field pressure becomes comparable to
the thermal plus ram pressure of the accreting material (
)
the accretion flow is strongly slowed down. This sudden decrease of the
Mach number compresses and adiabatically heats the disk:
the thermal pressure is now sufficient to lift up the material
from the disk and mass-load the accretion columns. This is clearly
shown in the right panel of Fig. 3: close to the truncation radius the dominant torque acting on the disk is the kinetic one (dot-dashed line):
the braking here is given by the loading of spinning mass onto the base
of the accretion columns.
Notice that in the region magnetically connected to the star the
magnetic and the kinetic torque do not match the accretion torque: this
means that the accretion flow is not perfectly stationary, as will
be shown in Sect. 3.3.
We briefly recall here some arguments about the estimate of the position of the truncation radius
as they are presented in Bessolaz et al. (2008).
A general expression for the truncation radius of an accretion
disk, assuming a dipolar structure for the stellar magnetic field, is
given by:
where













It must be also pointed out that, although the truncation radius is located at
,
the magnetic surfaces which are mass-loaded to form the funnels were initially located around corotation (
)
and subsequently they have been advected and compressed at the truncation radius by the accretion flow (Fig. 4):
the location of the accretion spots is therefore located at high
latitudes, corresponding to the footpoints of the field lines that were
anchored slightly below the corotation radius in the initial dipolar
configuration. The advection of the inner field lines yields a
deviation from the initial potential configuration and the formation of
``screening'' currents at the surface of the disk: the thermal
push triggered by the compression at the truncation radius is necessary
to cross the gravito-centrifugal barrier along ``screened'' field lines
(Li & Wilson 1999). Thermal effects can be
neglected, for example, in the case of potential dipolar fields for
(Koldoba et al. 2002) or if the magnetospheric field lines emerging from the corotation radius
are conveniently inclined towards the star (Ostriker & Shu 1995; Mohanty & Shu 2008):
in these cases the gravitational pull is sufficient to cross the
centrifugal barrier, but we showed that the interaction with an
accretion disk tends to prevent these potential configurations. The
redistribution of the initial dipolar field shown in Fig. 4 also has important consequences for the
efficiency of the star-disk magnetic coupling, as will be discussed in Sect. 3.3.1.
![]() |
Figure 4:
Initial magnetic configuration (dashed lines) and after |
Open with DEXTER |
![]() |
Figure 5:
Left panel: magnetic (solid line) and kinetic (dashed line)
angular momentum fluxes normalized over the magnetic flux (see the text
for the definitions) calculated along a field line connecting the star
with the accretion disk through the accretion column. The sum of the
magnetic and the kinetic fluxes is plotted with a dotted line.
A positive flux goes from the disk towards the star. The arc
length D goes from the star (D=0) to the midplane of the disk. Right panel:
forces projected along the same field line passing through the
accretion column: thermal pressure gradient (solid line), Lorentz force
(dashed line), gravity (dotted line) and centrifugal acceleration
(dot-dashed line). A positive force pushes along the field line
from the star towards the disk. Both snapshots are taken at
approximately |
Open with DEXTER |
3.2 Dynamics of the funnel flow
We discuss here how the funnel flows transfer angular momentum between the star and the disk. We define:
that is the ratio between the mass and the magnetic flux, and
i.e. the specific angular momentum transported by the matter and the magnetic field respectively. The quantities







The projection of the forces along a magnetic field line which follows
the accretion column is plotted in the right panel of Fig. 5:
at the base of the column the thermal pressure gradient uplifts and
pushes the matter against the action of the gravitational pinch and the
centrifugal barrier, accelerating the plasma up to the
slow-magnetosonic surface. Just after this critical point the gravity
pulls the gas towards the stellar surface while, as already
anticipated, the centrifugal acceleration reduces its strength
approaching the star. Under the action of gravity the infalling
material becomes strongly supersonic (
)
and reaches the stellar surface almost at free-fall speed (
,
see Fig. 6).
The convergence of the magnetospheric field lines towards the stellar
surface determines an adiabatic compression of the plasma:
the slow-magnetosonic Mach number
slightly
decreases and the thermal pressure gradient mildly resists the
gravitational pull. A weak Lorentz force pushes against the
matter's fall too. The Lorentz force along a magnetic surface
is related to the toroidal Lorentz force
by the simple relation:

Since the toroidal Lorentz force is slowing down the rotation of the infalling material (










![]() |
Figure 6:
Profiles along the accretion funnel of matter angular speed (
|
Open with DEXTER |
What determines the shape of the accretion spirals (leading or trailing), namely the sign of
along the accretion columns? This has an important consequence: in a leading spiral (
)
the Poynting flux transfers angular momentum from the disk to the star, while a trailing spiral (
)
transfers momentum from the star to the disk. By combining Eq. (24) with the relation between matter and field rotation
,
we obtain:
![]() |
(25) |
where




3.3 The stellar angular momentum
We finally study the angular momentum evolution of the star. The
angular momentum conservation equation integrated over the stellar
surface gives the expression:
![]() |
(26) |
where we separated the torque into magnetic:
![]() |
(27) |
and kinetic:
![]() |
(28) |
In Fig. 7 we plot the temporal evolution of the torques acting on the surface of the star normalized to the total angular momentum of the star (


![]() |
Figure 7:
Left panel: time evolution of torques acting on the surface of
the star, normalized over the stellar angular momentum. This defines
the inverse of the braking time. We plot the kinetic torque (dotted
line), the magnetic torque acting along the opened field lines
(dot-dashed line), along the magnetosphere connected to the disk below
(solid line) and beyond (dashed line) the corotation radius.
A positive torque spins up the stellar rotation. The inverse of
the braking time is given in units of t0-1: in the standard ``YSO'' normalization (see Sect. 2.4) a value 2 |
Open with DEXTER |
3.3.1 The star-disk magnetic coupling
The plot in the left panel of Fig. 7
shows that the magnetic torque calculated along the magnetosphere
connected inside the corotation radius is always spinning up the star
while the transport of the angular momentum along the lines connected
beyond the corotation radius is not sufficient to balance the accretion
torque: the braking torque represents less than
of the accretion torque. It was already possible to perceive this effect in Fig. 3,
where it was clear that the magnetic torque spinning up the disk beyond
corotation is much smaller than the one braking the disk rotation. The
spin-up torque shows a peculiar quasi-periodic oscillation: this
corresponds to a variation of the mass accretion rate
onto the surface of the star (see right panel in Fig. 7), with an average period of
.
This peculiar oscillation is a direct consequence of an unbalance
between the viscous torque, controlling the disk accretion rate on the
large scale down to
,
and the magnetic torque associated with the stellar rotation, regulating accretion below
:
the accretion rate measured at the stellar surface almost regularly
oscillates around the average disk accretion rate as determined by the
viscous torque. The truncation radius also oscillates, in good
agreement with Eq. (22).
It is important to point out that the characteristic timescale
associated with the magnetic torque is only a few times the Keplerian
timescale at corotation and it is therefore able to produce such a
rapid oscillation. The viscous timescale can determine only much slower
variations of the accretion rate, such as the long-term average
increase seen in the right panel of Fig. 7): this is caused by the small amount of heating that has been left active inside the accretion disk (see Sect. 2.1) which increases by
the thermal heightscale of the disk and the viscous accretion rate consistently with Eq. (15). There is a priori no reason for these two torques to provide the same
,
but other combinations of the parameters of the problem,
,
,
,
can result in a better matching. It is also possible to conjecture
that a solution characterized by a super-Alfvénic funnel flow would
favor the equality of the two accretion rates by imposing an additional
constraint on the accretion flow.
What is limiting the efficiency of the braking torque
associated with the star-disk interaction? The ``extended
magnetosphere'' torque can be obtained by radially integrating
Eq. (20):

and it depends on three factors: the field ``twist'' at the disk surface, given by the ratio







![]() |
Figure 8:
Upper panel: poloidal magnetic field along the midplane of the
disk (solid line). The behavior of the initial potential dipolar
field (R-3) is plotted for comparison (dashed line).The power law R-4.5 is plotted as a reference (dotted line). Lower panel: radial behavior of the magnetic field twist
|
Open with DEXTER |
![]() |
Figure 9: Time evolution of the mass flux of the stellar wind measured on the opened filed lines emerging from the stellar surface (solid line, left scale). The average magnetic lever arm of the wind is also plotted (dot-dashed line, right scale). |
Open with DEXTER |
3.3.2 The stellar wind
A more efficient braking of the star occurs along the opened field
lines anchored on its surface, due to a stellar wind which was already
clearly visible in Fig. 2: the wind
torque corresponds to approximately
of the accretion torque. The characteristics of this wind are summarized in Fig. 9, where we plot the temporal evolution of the mass-outflow rate of the wind and its average lever arm
,
given by ratio between the mean (cylindrical) radial position of the Alfvén surface and the radius of the star.
The torque exerted by the wind is given by the expression:
Equation (29) is the actual definition of




The main problem of a stellar wind launched from the surface of a star
rotating well below its break-up speed is that neither centrifugal
effects, nor a magnetic acceleration are able to give the first thrust
to the outflowing gas. In our case, the base of the wind is accelerated
by a thermal pressure gradient: on the other hand such a thermally
driven wind requires an enthalpy comparable to the gravitational
potential energy at the surface of the star. In the case of
T Tauri stars this corresponds to have temperatures of the order
K, which in turn poses severe cooling and radiative problems (Matt & Putritz 2007; Ferreira et al. 2006). Some other source of ``pressure'', possibly due to turbulent Alfvén waves (DeCampli 1981), must be introduced. Following the suggestion of Matt & Pudritz (2005b), a fraction of the accretion power could be used to drive the stellar wind:
in this particular simulation, the thermal power which drives the outflow corresponds to around
of the available accretion power. Here we considered that only the
kinetic energy of the infalling material is available to power the
wind: this corresponds to the gravitational energy liberated by the
infalling material minus the work done by the accretion torque on the
stellar rotation. Obviously a higher mass outflow rate of the stellar
wind would require a larger energy conversion efficiency.
The forces projected along a field line of the stellar wind as
a function of the cylindrical radius are plotted in the upper panel of
Fig. 10:
as already mentioned the thermal pressure gradient controls the
dynamics of the wind up to the slow-magnetosonic critical surface; then
the centrifugal acceleration pushes the outflow up to
,
while the Lorentz
force associated with a gradient of
accelerates it up to the Alfvén surface. The flow along this
particular field line has attained at the outer boundary a poloidal
speed
.
This means that not all the power available at the stellar surface has
been converted into kinetic energy. This can be understood by
inspecting the Bernoulli energy invariant E along this field line:
given by the sum of specific kinetic energy, enthalpy


where

![]() |
(31) |
It is possible to see that at the base of the wind the enthalpy is almost equal to the gravitational energy, the kinetic energy is negligible and the energy budget is dominated by the magnetic energy, which is around 4 times the thermal contribution. Consistently, taking







![]() |
Figure 10:
Upper panel: forces acting on the stellar wind calculated along
one of the opened field lines anchored on the stellar surface. We plot
the thermal pressure gradient (solid line), the Lorentz force (dashed
line), the gravity (dotted line) and the centrifugal acceleration
(dot-dashed line). Lower panel: Bernoulli (energy) invariant
(solid line) along the same field line, given by the sum of
Poynting-to-mass flux ratio (triple-dot-dashed line), kinetic energy
(dashed line), potential gravitational energy (dotted line, in
magnitude) and enthalpy (dot-dashed line). See the
text and Eq. (30) for the definition of the different terms. The snapshots are taken at |
Open with DEXTER |
![]() |
Figure 11:
Poloidal electric circuits flowing in the star-disk-wind system. Dark
(red) circuits are circulating clockwise along isosurfaces of
|
Open with DEXTER |
4 A global view: the electric circuits
The most synthetic way to have a global view of the magnetic
coupling between the different parts of the system, namely the
accretion disk, the star and the stellar wind, is to analyze the
poloidal electrical currents circulating in the star-disk-wind system.
The poloidal current, which is flowing along the isosurfaces = const., gives an indication of both the magnetic torques, proportional to the toroidal component of the vector product
,
and the poloidal Lorentz force associated with the toroidal field, which is perpendicular to the isosurfaces. In Fig. 11,
the dark (red) circuits (marked as A and C) are circulating
clockwise, while the light (yellow) one (marked as B) is
circulating anticlockwise. The dark (light) contours correspond to
positive (negative) values of
:
the dark circuits are responsible for extracting angular momentum from
the disk, while the light one extracts angular momentum from
the star.
The electromotive force responsible for the innermost dark circuit (A) results from the differential rotation between the star and the sub-corotation region of the disk. The current flows out of the inner boundary (the star), enters the disk around corotation and flows radially inwards inside the disk: this clearly gives a braking force -JrBz. It subsequently flows back to the star along the accretion columns almost parallel to the field lines: this is consistent with the fact that the field inside the accretion columns is almost force-free (see the weak Lorentz force plotted in Fig. 5). This circuit must finally close itself inside the star, where the disk angular momentum is deposited.
The electromotive force which fuels the light circuit (B) is given by the differential rotation between the star and the part of the disk connected beyond the corotation radius. The current flows out of the star and flows radially outwards inside the disk, giving a toroidal accelerating force -JrBz: this force corresponds to the small negative magnetic torque shown in Fig. 3. The current flows out from the disk surface along the current sheet separating the opened field lines anchored on the star and the opened field lines anchored inside the disk: this is the site where the magnetosphere gets inflated and opens up due to an overwhelming star-disk differential rotation. This circuit closes back on the star along the current-carrying stellar wind. This is an important point, since it shows clearly that the dynamics of the stellar wind along the opened field lines are linked to what happens in the closed magnetosphere: the current available inside the wind is the same which is flowing inside the connected magnetosphere. This current is responsible for the magneto-centrifugal acceleration and the collimation of the outflow: in that sense the dynamics of the closed magnetosphere affects the transverse equilibrium and structure of the stellar wind given by the solution of the Grad-Shafranov equation. Obviously this coupling does not affect the launching mechanism at the base of the flow which, as we saw in Sect. 7, must be of different origin. Unfortunately we cannot fully follow this magnetic coupling: the current outflowing on the current sheet turns back inside the wind well beyond the Alfvén surface, which in our simulation is very close to the outer boundary of the computational domain (see Fig. 2). In Fig. 11 we can also see that in the stellar wind the current crosses the field lines at mid-latitudes, giving a net toroidal and poloidal accelerating force, while it flows almost parallel to the field close to the poles (force-free configuration).
A second dark circuit (C) braking the disk rotation is visible at outer radii (
): this is reminiscent of the ``foot'' of the butterfly-like pattern characterizing the current circuit of disk-winds (Ferreira 1997).
In our simulation this corresponds to a magneto-centrifugal ``breeze''
whose mass flux and torque are negligible and do not affect the
underlying disk dynamics. The reason for this lies in the fact that the
local disk magnetization is too small to drive a powerful disk-wind.
5 Discussion
We have presented the outcome of an axisymmetric MHD simulation
of the interaction between a dipolar stellar magnetosphere with a
surrounding viscous and resistive accretion disk. First, we described
how the disk is truncated and the accretion flow is deviated into the
funnels. The angular momentum extracted from the disk in the
magnetically controlled region is transferred to the protostar, either
by a magnetic torque, or along the accretion curtains: the disk exerts
a spin-up torque on the star of the order
,
where
is the radius where the magnetic and the viscous torques are equal. The
study of possible spin-down mechanisms is therefore fundamental to
explain the angular momentum evolution of different classes of
magnetized accreting stars whose rotation frequency is not steadily
increasing.
The setup and the parameters of the numerical experiment were carefully
chosen to model two possible mechanisms which can brake the stellar
rotation and balance the accretion torque: (1) an ``extended
magnetosphere'' which keeps the connection with the disk beyond the
corotation radius; (2) a stellar wind launched along the opened
surfaces of the magnetosphere. In this particular simulation the
``extended magnetosphere'' torque is able to extract less than 10
of the accretion torque, while the stellar wind extracts more than 20
(see Sect. 3.3). Notice that, in the case of young forming stars, the spin-down mechanism must also balance the spin-up due to contraction.
We here discuss the possibility of increasing the efficiency of these two spin-down effects.
The feasibility of an effective ``extended magnetosphere'' torque
results from two contradictory requirements, namely the presence of an
extended magnetosphere characterized by a high value of both the field
and its twist: while an extended connectivity requires a high (and
maybe unphysical) disk resistivity (
in
our models), the same dissipative effects reduce the values of both the
field twisting and the intensity of the poloidal field, thus limiting
the efficiency of the spin-down torque. A stronger magnetic
coupling (a smaller disk resistivity) would induce a higher toroidal
twist and therefore a stronger field inflation: by decreasing the
parameter to
,
we verified that the spin-down torque is
reduced, until the magnetic connection beyond the corotation radius is lost. Besides, recent numerical simulations (Lesur & Longaretti 2009; Guan & Gammie 2009) suggest that the turbulent magnetic Prandtl number (
)
of accretion disks is likely to be of order unity and therefore the
turbulent resistivity and viscosity must vary accordingly. Moreover we
showed that an extended magnetic coupling can lead to oscillations of
the accretion rate and luminosity, due to a mismatch between the
torques which control accretion, viscous and magnetic: such a regular
oscillation, which has a period of a few stellar rotations, has never
been observed in CTTS. On the other hand it is likely that these
periodic oscillations can be avoided by tuning the parameters of the
model, so as to obtain a better matching between the viscous and
magnetic torques.
The relative efficiency of the spin-down torque can be improved
reducing the disk accretion rate (and therefore the accretion
torque) and/or increasing the strength of the stellar field so as to
enhance the magnetic coupling in the
region. However, when expressed in physical units, the accretion rate employed in our models is typical of CTTS (
), while the magnetic field intensity (
G at the stellar equator,
at the pole) is already compatible with the strongest dipolar components measured in CTTS, like BP Tau (Donati et al. 2008). Notice that both effects determine a larger truncation radius: compatibly with Eq. (22),
must be kept smaller than
in order to form the accretion columns. Moreover, an efficient disk
spin-down torque, able to balance at least the accretion torque (``disk
locked'' condition), would transfer in the region just beyond the
corotation radius a specific angular momentum comparable to the local
Keplerian one: this requires a strongly enhanced efficiency of the
``internal'' turbulent torque localized in the connected region to get
rid of this excess angular momentum (see, for example, Matt & Pudritz 2005a; Rappaport et al. 2004, Fig. 6). If the internal torque is not efficient enough, the disk stops accreting and enters a ``propeller'' regime (Ustyugova et al. 2006).
The torque of the stellar wind depends on three factors, see Eq. (29): the outflow rate
,
the average magnetic lever arm
and
the stellar rotation rate
.
Since classical T Tauri stars rotate well below their break-up speed
the outflow acceleration requires an extra energy input at the stellar
surface: in our adiabatic simulation this extra pressure is provided by
a large enthalpy, which can cause serious radiative cooling problems
and is incompatible with observations (Matt & Putritz 2007; Ferreira et al. 2006). Regardless of the actual nature of this energy input, Matt & Pudritz (2005b)
have proposed that a fraction of the energy deposited by accretion onto
the stellar surface can be used to power the stellar wind. In such a
scenario the mass-loss rate would essentially depend on the efficiency
of the energy coupling between accretion and ejection: in the
particular case of our numerical model a power coupling efficiency of
just
can drive a mass-loss rate
.
The wind torque strongly depends on the length of the lever arm (e.g.
in our simulation) which is determined both by the mass-loss rate and
the magnetic configuration. The efficiency of the stellar wind torque
therefore depends on at least two aspects of the interaction with the
accretion disk: not only the power coupling efficiency, as
discussed by Matt & Pudritz (2005b), but also the geometry of the magnetic surfaces which, as shown in Sects. 3.3.2 and 4,
is strongly affected by the interaction with the disk and the closed
magnetosphere. In a general way, for a given poloidal magnetic
field distribution on the stellar surface, the lever arm decreases when
increasing the mass outflow rate, so that the wind torque does not grow
linearly with the mass-loss rate. The Weber & Davis (1967) ``classical'' scaling provides
,
while a steeper scaling,
,
has been estimated by Matt & Pudritz (2008a). Even adopting this more favorable scaling, a high mass-loss rate
,
compatible with the entire mass flux observed in T Tauri microjets (Cabrit 2007), would be needed to balance the spin-up torque, in agreement with Matt & Pudritz (2008b). Correspondingly, this would require a high energy coupling efficiency
:
in the ``accretion powered'' stellar wind scenario this energy input
could determine a strong reduction of the fraction of accretion
luminosity which is radiated at the accretion shock.
The stellar wind could therefore represent an effective means to balance the spin-up due to accretion and contraction (not computed here) but some aspects still need some clarification: the nature of the power which drives the outflow at the stellar surface and its possible relation to the accretion energy; the topology of the magnetic surfaces as determined by a detailed model of the star-disk interaction.
6 Summary and conclusion
The axisymmetric MHD simulation presented in this paper constitutes
the first example of a numerical model of the star-disk magnetic
interaction which takes into account at the same time: (1) the
formation of accretion funnel flows and the associated spin-up torque;
(2) the spin-down torque due to a star-disk magnetic connection
which extends beyond the corotation radius; (3) the spin-down
torque exerted by a stellar wind. The parameters of the simulation have
been chosen to model a typical classical T Tauri star (
,
), accreting at
and rotating at
of its break-up speed (
days). The strength of its dipolar magnetosphere (
G) is consistent with the strongest dipolar components measured in CTTS.
We summarize here the outcome of this numerical experiment:
- 1.
- A magnetospheric star-disk connection extending beyond the
corotation radius requires a high level of turbulent disk resistivity (
in our model) in order to limit the build-up of toroidal field and the consequent inflation and opening of the magnetic structure. On the other hand, these dissipative phenomena reduce the value of the toroidal field and, in conjunction with the field advection and compression in the truncation region, decrease the intensity of the poloidal field in the
region. As a consequence, the efficiency of the ``extended magnetosphere'' torque is strongly reduced with respect to the standard Ghosh & Lamb scenario: in our numerical example the spin-down torque is equal to only
of the accretion torque. Moreover we found that an extended magnetospheric torque can lead to an oscillation of the accretion rate on a timescale of a few
which is usually not observed in CTTS. A lower accretion rate or an even stronger magnetic field could improve the relative efficiency of the spin-down torque but, in a ``disk-locked'' situation, this would require an enhanced transport inside the disk itself to get rid of the excess angular momentum extracted from the star.
- 2.
- The stellar wind modeled in our experiment is characterized by a mass loss rate
and a magnetic lever arm
. The corresponding spin-down torque is around
of the accretion torque. We pointed out that due to the slow rotation period of the star, an energy input comparable to the gravitational potential energy is necessary to drive the flow at the stellar surface. While in our model this energy input is given by enthalpy, this would determine severe cooling and radiative issues and an alternate form of driving pressure must be considered. For the modeled stellar wind the energy input corresponds to
of the available accretion power. We also emphasized that, beside this possible energetic connection between accretion and ejection, a ``topological'' link is present, due to the magnetic coupling between the opened and closed parts of the magnetosphere. By rescaling the solution that we found, we concluded that a stellar wind with
requiring an energy input equal to
of the accretion power could in principle balance the accretion torque. These properties are obviously quite demanding, from the point of view of both the observations and the theory.

The authors acknowledge support through the Marie Curie Research Training Network JETSET (Jet Simulations, Experiments and Theory) under contract MRTN-CT-2004-005592. The simulations were performed on the computing facilities of the Service Commun de Calcul Intensif de l'Observatoire de Grenoble (SCCI).
Appendix A: Boundary conditions on
A precise control of the period of rotation of the central star is
crucial in order to study the angular momentum evolution of the
star-disk system. In this Appendix we examine the effects of
different boundary conditions for the toroidal field
on the rotation rate of the inner bounary (the ``star''). We recall that our boundary condition on
was chosen to ensure an appropriate torque to yield the correct matter rotation speed, according to Eq. (12), see Sect. 2.3.
We here compare our boundary condition on the toroidal field with other
more ``standard'' ones: the usual ``outflow'' conditions, i.e.
.
The boundary conditions assumed in Romanova et al. (2002) (and in other works by the same group), i.e.
:
it is easy to see that this condition forces the poloidal electric
current to be purely radial at the stellar surface. A
condition is motivated by the fact that in this case the field and the matter rotate at the same speed.
![]() |
Figure A.1:
Effective rotation rate of the magnetic surfaces measured on the surface of the star as a function of the polar angle |
Open with DEXTER |
A first comparison is shown in Fig. A.1, where we plot as a function of the polar angle the effective rotation rate of the field lines
,
defined by inverting Eq. (12):
.
It is clear that our boundary condition forces the correct rotation rate with an error of only a few percent. The
condition is not able to force the rotation of the field lines close to the pole of the star (
)
while it produces a rotation two times higher than
at mid-latitudes. The ``outflow'' and the
conditions
are somewhat better but it is clear that all these boundary conditions
overestimate (at mid latitudes) or underestimate (close to the
pole) the correct rotation by
.
The excess is clearly due to the spin-up accretion torque, while
the lack is determined by the spin-down torque exerted by the stellar
wind. While this effect could be considered correct, it happens on
too short timescales: due to the enourmous stellar moment of inertia,
no appreciable variation of the stellar rotation speed should
happen during the duration of the simulations.
![]() |
Figure A.2: Time evolution of the
average specific angular momentum transferred from the disk to the
star. The curves correspond to different boundary conditions on the
toroidal field: the boundary condition used in this paper (solid line),
|
Open with DEXTER |
![]() |
Figure A.3:
Poloidal current circuits flowing in the inner regions of two
simulations characterized by different boundary conditions on |
Open with DEXTER |
This error on the computation of the stellar rotation period has
profound consequences on the effective magnetic coupling between the
star and the disk. In Fig. A.2 we
show the time evolution of the specific angular momentum defined by the ratio
between the torque exerted by the disk on the star along the connected
magnetosphere and the accretion rate measured on the surface of the
star. Our boundary condition shows clearly a higher accreted specific
angular momentum, suggesting that the effective stellar rotation rate
which affects the disk is higher for the
,
``outflow'' and
boundary
conditions. This can be qualitatively understood
by noticing that if the stellar rotation frequency increases and the
corotation radius approaches the truncation radius, the size of the
part of the disk which spins the star up decreases, while the region
connected beyond the corotation grows. A more quantitative
discussion can be found in Matt & Pudritz (2005a), see for example their Fig. 8. A clearer representation of this effect in shown in Fig. A.3:
we can see that the position that marks the separation between the
inner electric circuit accelerating the stellar rotation (dark red
isosurfaces) and the outer one braking the stellar rotation (light
yellow isosurfaces), which in our simulation correctly corresponds to
the corotation radius, has moved closer to the truncation radius in the
simulation.
This clearly indicates that with this boundary condition on the
toroidal field the disk ``feels'' a star rotating faster than the
assumed
,
since the corotation radius has moved closer to the star. This effect can be probably noticed in Fig. 3 in Long et al. (2005),
where the magnetic torque transfers angular momentum from the star to
the disk in the sub-corotation region, despite the disk rotation being
still faster than
(see Fig. 5
in the same paper). With a different boundary condition, the rotation
rate of the magnetic surfaces closer to the pole of the star is
systematically lower than
,
thus decreasing the magnetocentrifugal effects which can accelerate the
stellar wind beyond the slow-magnetosonic surface (see Sect. 3.3).
References
- Agapitou, V., & Papaloizou, J. C. B. 2000, MNRAS, 317, 273 [NASA ADS] [CrossRef]
- Aly, J. J. 1980, A&A, 86, 192 [NASA ADS]
- Armitage, P. J., & Clarke, C. J. 1996, MNRAS, 280, 458 [NASA ADS]
- Arons, J. 1993, ApJ, 408, 160 [NASA ADS] [CrossRef]
- Balsara, D. S., & Spicer, D. S. 1999, J. Comput. Phys., 149, 270 [NASA ADS] [CrossRef]
- Bardou, A., & Heyvaerts, J. 1996, A&A, 307, 1009 [NASA ADS]
- Bessolaz, N., Zanni, C., Ferreira, J., Keppens, R., & Bouvier, J. 2008, A&A, 478, 155 [NASA ADS] [EDP Sciences] [CrossRef]
- Bildsten, L., Chakrabarty, D., Chiu, J., et al. 1997, ApJS, 113, 367 [NASA ADS] [CrossRef]
- Bouvier, J., & Appenzeller, I. 2007, Star-Disk Interaction in Young Stars (Cambridge: Cambridge University Press), IAU Symp., 243
- Bouvier, J., Cabrit, S., Fernández, M., Martín, E. L., & Matthews, J. M. 1993, A&A, 272, 176 [NASA ADS]
- Cabrit, S. 2007, in Star-Disk Interaction in Young Stars, ed. J. Bouvier, & I. Appenzeller, IAU Symp., 243, 203
- Cai, M. J., Shang, H., Lin, H.-H., & Shu, F. H. 2008, ApJ, 672, 489 [NASA ADS] [CrossRef]
- Campbell, C. G. 1987, MNRAS, 229, 405 [NASA ADS]
- Collier Cameron, A., & Campbell, C. G. 1993, A&A, 274, 309 [NASA ADS]
- DeCampli, W. M. 1981, ApJ, 244, 124 [NASA ADS] [CrossRef]
- Donati, J.-F., Jardine, M. M., Gregory, S. G., et al. 2007, MNRAS, 380, 1297 [NASA ADS] [CrossRef]
- Donati, J.-F., Jardine, M. M., Gregory, S. G., et al. 2008, MNRAS, 386, 1234 [NASA ADS] [CrossRef]
- Edwards, S., Hartigan, P., Ghandour, L., & Andrulis, C. 1994, AJ, 108, 1056 [NASA ADS] [CrossRef]
- Elsner, R. F., & Lamb, F. K. 1977, ApJ, 215, 897 [NASA ADS] [CrossRef]
- Ferreira, J. 1997, A&A, 319, 340 [NASA ADS]
- Ferreira, J., Pelletier, G., & Appl, S. 2000, MNRAS, 312, 387 [NASA ADS] [CrossRef]
- Ferreira, J., Dougados, C., & Cabrit, S. 2006, A&A, 453, 785 [NASA ADS] [EDP Sciences] [CrossRef]
- Ghosh, P., & Lamb, F. K. 1979a, ApJ, 232, 259 [NASA ADS] [CrossRef]
- Ghosh, P., & Lamb, F. K. 1979b, ApJ, 234, 296 [NASA ADS] [CrossRef]
- Goodson, A. P., Böhm, K.-H., & Winglee, R. M. 1999, ApJ, 524, 142 [NASA ADS] [CrossRef]
- Guan, X., & Gammie, C. F. 2009, ApJ, 697, 1901 [NASA ADS] [CrossRef]
- Gullbring, E., Calvet, N., Muzerolle, J., & Hartmann, L. 2000, ApJ, 544, 927 [NASA ADS] [CrossRef]
- Illarionov, A. F., & Sunyaev, R. A. 1975, A&A, 39, 185 [NASA ADS]
- Irwin, J., Hodgkin, S., Aigrain, S., et al. 2007, MNRAS, 377, 741 [NASA ADS] [CrossRef]
- Johns-Krull, C. M. 2007, ApJ, 664, 975 [NASA ADS] [CrossRef]
- Kenyon, S. J., Yi, I., & Hartmann, L. 1996, ApJ, 462, 439 [NASA ADS] [CrossRef]
- Kley, W., & Lin, D. N. C. 1992, ApJ, 397, 600 [NASA ADS] [CrossRef]
- Kluzniak, W., & Kita, D. 2000 [arXiv:astro-ph/0006266]
- Kluzniak, W., & Rappaport, S. A. 2007, ApJ, 671, 1990 [NASA ADS] [CrossRef]
- Koldoba, A. V., Lovelace, R. V. E., Ustyugova, G. V., & Romanova, M. M. 2002, ApJ, 123, 2019 [NASA ADS]
- Königl, A. 1991, ApJ, 370, L39 [NASA ADS] [CrossRef]
- Küker, M., Henning, T., & Rüdiger, G. 2003, ApJ, 589, 397 [NASA ADS] [CrossRef]
- Kulkarni, A. K., & Romanova, M. M. 2008, MNRAS, 386, 673 [NASA ADS] [CrossRef]
- Lesur, G., & Longaretti, P.-Y. 2009, A&A, 504, 309 [EDP Sciences] [CrossRef]
- Li, J., & Wilson, G. 1999, ApJ, 527, 910 [NASA ADS] [CrossRef]
- Long, M., Romanova, M. M., & Lovelace, R. V. E. 2005, ApJ, 634, 1214 [NASA ADS] [CrossRef]
- Matt, S., & Pudritz, R. E. 2005a, MNRAS, 356, 167 [NASA ADS] [CrossRef]
- Matt, S., & Pudritz, R. E. 2005b, ApJ, 632, L135 [NASA ADS] [CrossRef]
- Matt, S., & Pudritz, R. E. 2007, in Star-Disk Interaction in Young Stars, ed. J. Bouvier, & I. Appenzeller, IAU Symp., 243, 299
- Matt, S., & Pudritz, R. E. 2008a, ApJ, 678, 1109 [NASA ADS] [CrossRef]
- Matt, S., & Pudritz, R. E. 2008b, ApJ, 681, 391 [NASA ADS] [CrossRef]
- Mignone, A., Bodo, G., Massaglia, S., et al. 2007, ApJS, 170, 228 [NASA ADS] [CrossRef]
- Miller, K. A., & Stone, J. M. 1997, ApJ, 489, 890 [NASA ADS] [CrossRef]
- Mohanty, S., & Shu, F. H. 2008, ApJ, 687, 1323 [NASA ADS] [CrossRef]
- Najita, J., Carr, J. S., & Mathieu, R. D. 2003, ApJ, 589, 931 [NASA ADS] [CrossRef]
- Nelson, R. W., Bildsten, L., Chakrabarty, D., et al. 1997, ApJ, 488, L117 [NASA ADS] [CrossRef]
- Ostriker, E. C., & Shu, F. H. 1995, ApJ, 447, 813 [NASA ADS] [CrossRef]
- Powell, K. G., Roe, P. L., Linde, T. J., Gombosi, T. I., & De Zeeuw, D. L. 1999, J. Comput. Phys, 154, 284 [NASA ADS] [CrossRef]
- Pringle, J. E., & Rees, M. J. 1972, A&A, 21, 1 [NASA ADS]
- Rappaport, S. A., Fregeau, J. M., & Spruit, H. 2004, ApJ, 606, 436 [NASA ADS] [CrossRef]
- Regev, O., & Gitelman, L. 2002, A&A, 396, 623 [NASA ADS] [EDP Sciences] [CrossRef]
- Romanova, M. M., Ustyugova, G. V., Koldoba, A. V., & Lovelace, R. V. E. 2002, ApJ, 578, 420 [NASA ADS] [CrossRef]
- Romanova, M. M., Ustyugova, G. V., Koldoba, A. V., & Lovelace, R. V. E. 2009 [arXiv:0901.4265]
- Rózyczka, M., Bodenheimer, P., & Bell, K. R. 1994, ApJ, 423, 736 [NASA ADS] [CrossRef]
- Shakura, N. I., & Sunyaev, R. A. 1973, A&A, 24, 337 [NASA ADS]
- Shu, F. H., Najita, J., Ostriker, E., et al. 1994, ApJ, 429, 781 [NASA ADS] [CrossRef]
- Tanaka, T. 1994, J. Comput. Phys., 111, 381 [NASA ADS] [CrossRef]
- Umurhan, O. M., Nemirovsky, A., Regev, O., & Shaviv, G. 2006, A&A, 446, 1 [NASA ADS] [EDP Sciences] [CrossRef]
- Urpin, V. A. 1984, Sov. Astron., 28, 50 [NASA ADS]
- Ustyugova, G. V., Koldoba, A. V., Romanova, M. M., & Lovelace, R. V. E. 2006, ApJ, 646, 304 [NASA ADS] [CrossRef]
- Uzdensky, D. A., Königl, A., & Litwin, C. 2002, ApJ, 565, 1191 [NASA ADS] [CrossRef]
- Wang, Y.-M. 1995, ApJ, 449, L153 [NASA ADS]
- Wang, Y.-M. 1996, ApJ, 465, L111 [NASA ADS] [CrossRef]
- Warner, B. 1995, Catalcysmic Variable Stars (Cambridge: Cambridge University Press)
- Weber, E. J., & Davis, L. 1967, ApJ, 148, 217 [NASA ADS] [CrossRef]
- Yi, I. 1995, ApJ, 442, 768 [NASA ADS] [CrossRef]
Footnotes
- ... code
- PLUTO is freely available at http://plutocode.to.astro.it
- ... surface
- In Eqs. (20) and (21) we indicate as poloidal the component of a vector field perpendicular to the surface of the disk defined by
. In the case of the magnetic field, we therefore have
. An analogous expression is used for the poloidal speed
in Eq. (21).
All Figures
![]() |
Figure 1:
Appearance of the initial conditions of the simulations. Colors are representative of the logarithmic density in units of
|
Open with DEXTER | |
In the text |
![]() |
Figure 2:
Time evolution of logarithmic density maps in units of
|
Open with DEXTER | |
In the text |
![]() |
Figure 3:
Left panel: radial behavior of the rotation speed of the accretion disk
|
Open with DEXTER | |
In the text |
![]() |
Figure 4:
Initial magnetic configuration (dashed lines) and after |
Open with DEXTER | |
In the text |
![]() |
Figure 5:
Left panel: magnetic (solid line) and kinetic (dashed line)
angular momentum fluxes normalized over the magnetic flux (see the text
for the definitions) calculated along a field line connecting the star
with the accretion disk through the accretion column. The sum of the
magnetic and the kinetic fluxes is plotted with a dotted line.
A positive flux goes from the disk towards the star. The arc
length D goes from the star (D=0) to the midplane of the disk. Right panel:
forces projected along the same field line passing through the
accretion column: thermal pressure gradient (solid line), Lorentz force
(dashed line), gravity (dotted line) and centrifugal acceleration
(dot-dashed line). A positive force pushes along the field line
from the star towards the disk. Both snapshots are taken at
approximately |
Open with DEXTER | |
In the text |
![]() |
Figure 6:
Profiles along the accretion funnel of matter angular speed (
|
Open with DEXTER | |
In the text |
![]() |
Figure 7:
Left panel: time evolution of torques acting on the surface of
the star, normalized over the stellar angular momentum. This defines
the inverse of the braking time. We plot the kinetic torque (dotted
line), the magnetic torque acting along the opened field lines
(dot-dashed line), along the magnetosphere connected to the disk below
(solid line) and beyond (dashed line) the corotation radius.
A positive torque spins up the stellar rotation. The inverse of
the braking time is given in units of t0-1: in the standard ``YSO'' normalization (see Sect. 2.4) a value 2 |
Open with DEXTER | |
In the text |
![]() |
Figure 8:
Upper panel: poloidal magnetic field along the midplane of the
disk (solid line). The behavior of the initial potential dipolar
field (R-3) is plotted for comparison (dashed line).The power law R-4.5 is plotted as a reference (dotted line). Lower panel: radial behavior of the magnetic field twist
|
Open with DEXTER | |
In the text |
![]() |
Figure 9: Time evolution of the mass flux of the stellar wind measured on the opened filed lines emerging from the stellar surface (solid line, left scale). The average magnetic lever arm of the wind is also plotted (dot-dashed line, right scale). |
Open with DEXTER | |
In the text |
![]() |
Figure 10:
Upper panel: forces acting on the stellar wind calculated along
one of the opened field lines anchored on the stellar surface. We plot
the thermal pressure gradient (solid line), the Lorentz force (dashed
line), the gravity (dotted line) and the centrifugal acceleration
(dot-dashed line). Lower panel: Bernoulli (energy) invariant
(solid line) along the same field line, given by the sum of
Poynting-to-mass flux ratio (triple-dot-dashed line), kinetic energy
(dashed line), potential gravitational energy (dotted line, in
magnitude) and enthalpy (dot-dashed line). See the
text and Eq. (30) for the definition of the different terms. The snapshots are taken at |
Open with DEXTER | |
In the text |
![]() |
Figure 11:
Poloidal electric circuits flowing in the star-disk-wind system. Dark
(red) circuits are circulating clockwise along isosurfaces of
|
Open with DEXTER | |
In the text |
![]() |
Figure A.1:
Effective rotation rate of the magnetic surfaces measured on the surface of the star as a function of the polar angle |
Open with DEXTER | |
In the text |
![]() |
Figure A.2: Time evolution of the
average specific angular momentum transferred from the disk to the
star. The curves correspond to different boundary conditions on the
toroidal field: the boundary condition used in this paper (solid line),
|
Open with DEXTER | |
In the text |
![]() |
Figure A.3:
Poloidal current circuits flowing in the inner regions of two
simulations characterized by different boundary conditions on |
Open with DEXTER | |
In the text |
Copyright ESO 2009
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.