Articles

LOCAL STUDY OF ACCRETION DISKS WITH A STRONG VERTICAL MAGNETIC FIELD: MAGNETOROTATIONAL INSTABILITY AND DISK OUTFLOW

and

Published 2013 March 21 © 2013. The American Astronomical Society. All rights reserved.
, , Citation Xue-Ning Bai and James M. Stone 2013 ApJ 767 30 DOI 10.1088/0004-637X/767/1/30

0004-637X/767/1/30

ABSTRACT

We perform three-dimensional, vertically-stratified, local shearing-box ideal MHD simulations of the magnetorotational instability (MRI) that include a net vertical magnetic flux, which is characterized by midplane plasma β0 (ratio of gas to magnetic pressure). We have considered β0 = 102, 103, and 104, and in the first two cases the most unstable linear MRI modes are well resolved in the simulations. We find that the behavior of the MRI turbulence strongly depends on β0: the radial transport of angular momentum increases with net vertical flux, achieving α ∼ 0.08 for β = 104 and α ≳ 1.0 for β0 = 100, where α is the height-integrated and mass-weighted Shakura–Sunyaev parameter. A critical value lies at β0 ∼ 103: for β0 ≳ 103, the disk consists of a gas pressure dominated midplane and a magnetically dominated corona. The turbulent strength increases with net flux, and angular momentum transport is dominated by turbulent fluctuations. The magnetic dynamo that leads to cyclic flips of large-scale fields still exists, but becomes more sporadic as net flux increases. For β0 ≲ 103, the entire disk becomes magnetically dominated. The turbulent strength saturates, and the magnetic dynamo is fully quenched. Stronger large-scale fields are generated with increasing net flux, which dominates angular momentum transport. A strong outflow is launched from the disk by the magnetocentrifugal mechanism, and the mass flux increases linearly with net vertical flux and shows sign of saturation at β0 ≲ 102. However, the outflow is unlikely to be directly connected to a global wind: for β0 ≳ 103, the large-scale field has no permanent bending direction due to dynamo activities, while for β0 ≲ 103, the outflows from the top and bottom sides of the disk bend towards opposite directions, inconsistent with a physical disk wind geometry. Global simulations are needed to address the fate of the outflow.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

The magnetorotational instability (MRI; Balbus & Hawley 1991) is believed to be the primary mechanism for driving accretion in a wide range of astrophysical disks by generating turbulence to provide outward angular momentum transport. The robustness of the MRI and its physical properties have been widely studied by means of both local shearing box simulations (e.g., Hawley et al. 1995; Brandenburg et al. 1995) and global simulations (e.g., Hawley 2001; Fromang & Nelson 2006). Local shearing-box simulations have the advantages of being able to achieve high resolution at modest computational cost, and are the primary method for exploring the detailed physics of the MRI.

Four types of shearing-box MRI simulations exist, depending on whether vertical gravity from the central object (which gives density stratification) is included and whether the simulations include net vertical magnetic flux. While unstratified MRI simulations provide important benchmarks on the physical properties of the MRI turbulence (e.g., Hawley et al. 1996; Sano et al. 2004; Fromang & Papaloizou 2007; Fromang et al. 2007; Lesur & Longaretti 2007), real disks involve vertical density stratification. So far, most vertically stratified shearing-box simulations focus on a magnetic field configuration with zero net vertical magnetic flux (e.g., Stone et al. 1996; Miller & Stone 2000; Ziegler & Rüdiger 2000; Hirose et al. 2006; Davis et al. 2010; Shi et al. 2010; Flaig et al. 2010; Simon et al. 2012). However, the inclusion of a net vertical magnetic field in stratified shearing-box simulations, which is probably closer to reality, has only rarely been explored, most likely due to numerical difficulties (Stone et al. 1996; Miller & Stone 2000): either the numerical code crashes or the disk is violently disrupted in a few orbits.

This least-explored regime of stratified shearing-box with net vertical flux has drawn more and more attention in recent years. This is partly due to improvement of the numerical schemes, but more importantly, the physical significance of the MRI in this regime and its intimate connections to astrophysical applications make it worth more detailed study. In particular, the presence of jets and outflows in a wide range of accretion systems such as protostars (e.g., Reipurth & Bally 2001; Cabrit 2007), X-ray binaries (e.g., Mirabel & Rodríguez 1999; Fender 2006), and active galactic nuclei (e.g., Begelman et al. 1984; Harris & Krawczynski 2006) is indicative of the presence of a large-scale (poloidal) magnetic field threading through the disk, which is essential for jet/outflow acceleration and collimation via the magnetocentrifugal mechanism (Blandford & Payne 1982; Pudritz & Norman 1983).

Recently, Suzuki & Inutsuka (2009) and Suzuki et al. (2010) successfully conducted stratified shearing-box simulations that include a net vertical magnetic field with an outflow boundary condition using characteristic decomposition. Their net vertical field, characterized by β0 (ratio of gas to magnetic pressure at midplane), is rather weak (β0 ⩾ 104). They reported the launching of disk outflow and found that the averaged mass outflow rate increases linearly with the magnetic pressure of the net vertical field. It was speculated that the outflow serves as the base of a Blandford–Payne type magnetocentrifugal wind (MCW).

In this paper, we conduct shearing-box simulations of the MRI that include a relatively strong net vertical magnetic flux with 104 ⩾ β0 ⩾ 100, which has largely been unexplored in the literature. The main motivation of our work is two-fold as we elaborate below.

First, we aim at studying the properties of the MRI turbulence in more realistic magnetic field geometry and their dependence on the net poloidal magnetic flux threading through accretion disks. In particular, how the Shakura–Sunyaev α parameter, which is of the most common interest, depends on the net magnetic flux. Due to the vast dominance of zero net vertical flux stratified shearing-box simulations, the α parameter has been taken for granted to be of the order of 0.01, while the α value inferred from observations of fully ionized disks is at least an order of magnitude higher (e.g., see discussions by King et al. 2007 and references therein). Unstratified shearing-box simulations (e.g., Hawley et al. 1995; Bai & Stone 2011) unambiguously found that as long as the most unstable linear MRI mode fits into the simulation domain, the turbulent stress increases roughly linearly with 1/β0. It is natural to expect the same trend in stratified shearing-box simulations, and hence the net vertical magnetic flux is likely to be a crucial ingredient in real accretion disks. Moreover, as in the zero net-flux cases, we expect buoyancy and open vertical boundaries in our stratified simulations to give rise to novel features on the properties of the MRI turbulence, particularly the MRI dynamo and the generation of large-scale fields (Lesur & Ogilvie 2008; Gressel 2010; Guan & Gammie 2011).

Second, our study will address the potential connection between the MRI and the MCW (Blandford & Payne 1982). The MCW extracts both mass and angular momentum from accretion disks, and is a competing mechanism for driving disk accretion and evolution. While the wind acceleration and collimation of the MCW is largely known, the wind launching process that loads mass from the disk onto the wind is still not well understood. Most semi-analytical works and numerical simulations either assume a razor-thin disk treated as a boundary condition with artificial mass injection (e.g., Li 1995; Ouyed & Pudritz 1997; Krasnopolsky et al. 1999), or proceed with an unresolved disk by adopting some forms of artificial diffusion, (e.g., Kato et al. 2002; Casse & Keppens 2002; Zanni et al. 2007). Such diffusion, which presumably arises from the MRI, is necessary to allow the gas to slide through the disk that avoids rapid accumulation of magnetic flux to maintain steady state accretion. However, the microphysics (e.g., the MRI) within the accretion disk, which is crucial for wind launching, is not modeled directly. In fact, it is numerically very challenging to properly resolve the MRI turbulence in global wind simulations, which requires an extended three-dimensional (3D) computational domain with large numerical resolution. On the other hand, local shearing-box simulations provide superb resolution with a realistic computational cost, and hence are a powerful tool for studying the wind launching process from first principle.

Recently, Fromang et al. (2012) conducted a series of stratified shearing-box MRI simulations with β0 = 104. After carefully examining various numerical issues, they reported simultaneous MRI turbulence and launching of an MCW, whereas angular momentum transport is still dominated by the MRI rather than the MCW. In parallel, Lesur et al. (2013) conducted a series of stratified shearing-box simulations in both one-dimension and three-dimensions with β0 ∼ 10, where the most unstable linear MRI modes just marginally fit into the disk. They found that such MRI modes directly produce magnetically driven outflows, which are time varying due to secondary instabilities. Meanwhile, Moll (2012) studied the wind launching process in two-dimensional (2D) shearing-box simulations with β0 ∼ 1–10 (where MRI is suppressed) and found long-wavelength "clump" instability for β0 ≲ 1 and speculated its connection to the mass-flux instability found in earlier analytical works (Lubow et al. 1994b; Cao & Spruit 2002). Our simulations fill the gap in parameter space explored by these authors and reveal the interesting transition from the MRI-dominated transport to potentially wind-dominated transport. Together with the earlier semi-analytical study of wind launching by Ogilvie (2012), which applies to the regime of β0 ≲ 1, as well as the simulations by Suzuki & Inutsuka (2009), which applies to β0 ≳ 104, the series of studies cover the complete parameter range relevant to MRI turbulence and wind launching in realistic accretion disks under the local shearing-box framework.

This paper is organized as follows. In Section 2, we describe the methodology and setup of our simulations. We present our simulation results in Section 3, focusing on the properties of the MRI and angular momentum transport, and in Section 4, we focus on the dynamics of the outflow and its connection to global disk winds. Each result section is broken into three subsections, each devoted to one aspect of the results. The implications of our results for global disk evolution are discussed in Section 5, before the conclusion.

2. SIMULATION SETUP

We use the Athena magnetohydrodynamic (MHD) code (Stone et al. 2008) and perform 3D local MHD simulations of gas dynamics using the shearing-box approach (Goldreich & Lynden-Bell 1965). Ideal MHD equations are written in a Cartesian coordinate system in the corotating frame at a fiducial radius R with Keplerian frequency Ω, with ${\boldsymbol{e}}_x,{\boldsymbol{e}}_y,{\boldsymbol{e}}_z$ denoting unit vectors pointing to the radial, azimuthal, and vertical directions, respectively, and ${\boldsymbol{\Omega }}$ is along the ${\boldsymbol{e}}_z$ direction. The MHD equations read

Equation (1)

Equation (2)

Equation (3)

where ${\sf T}$ is the total stress tensor

Equation (4)

ρ is the gas density, ${\boldsymbol{u}}$ is the gas velocity. We use an isothermal equation of state $P=\rho c_s^2$. Note that the unit for the magnetic field is such that the magnetic pressure equals B2/2, which avoids the extra factor of $\sqrt{4\pi }$. The vertical gravity from the central object (−Ω2z) is included in the momentum equation, and hence our simulations have vertical density stratification.

The shearing-box source terms (Coriolis force and tidal gravity) have been readily implemented in Athena (Stone & Gardiner 2010). An orbital advection scheme is adopted where the system is split into an advective part that corresponds to the background shear flow $-3\Omega x/2{\boldsymbol{e}}_y$, and another part that only evolves the velocity fluctuations ${\boldsymbol{v}}$:

Equation (5)

Most of our analysis deals with ${\boldsymbol{v}}$, except in the study of conservation laws along field lines, where ${\boldsymbol{u}}$ will also be used (see Section 4.1.2).

We use standard shearing-box boundary conditions in the radial direction (Hawley et al. 1995). In the vertical direction, we adopt a simple zero-gradient outflow boundary condition for velocities and magnetic fields, with density extrapolated assuming vertical hydrostatic equilibrium (Simon et al. 2011). In addition, vertical velocity in the ghost cells is set to zero in the case of inflow. It was pointed out that a strict zero-gradient boundary condition (including zero vertical density gradient) would prevent MHD-driven outflow (Lesur et al. 2013), while we find that the outflow can be launched once a vertical density gradient is present at the interface of the vertical boundary.

It is well known that in the presence of a pure vertical background magnetic field, the linear MRI modes consist of counter-motions in the radial direction that alternate along the vertical direction (i.e., the channel flow). The channel flow is an exact solution of the MHD equations, even in the nonlinear regime, provided that the magnetic field remains sub-thermal (Goodman & Xu 1994). It can achieve a very large amplitude before being broken up by parasitic instabilities (Latter et al. 2009; Pessah & Goodman 2009). Recently, Latter et al. (2010) studied the MRI modes in the presence of vertical stratification and their parasitic instabilities. They found that the channel modes strongly amplify the magnetic fields to thermal strength before they can be destroyed. On the simulation side, the channel flows in the initial stage of the simulations not only place enormous demands on numerical codes, but also give rise to eruptive behavior that depletes the disk in just a few orbits (Miller & Stone 2000). In view of the transient nature of the channel flow, as well as its large numerical demand, we initialize our simulations in a way that strong channel flow can be avoided, as described below.

Fiducially, we use a box size of (4H, 8H, 12H) in (x, y, z) resolved by (96, 96, 288) grid points, where Hcs/Ω is the thermal scale height. The relatively large horizontal box size is necessary to properly capture the mesoscale structures of the MRI turbulence such as the zonal flow (Johansen et al. 2009; Simon et al. 2012). The initial density profile is taken to be Gaussian ρ = ρ0exp (− z2/2H2). We use the natural unit in our simulations, with cs = 1, Ω = 1, H = 1, and ρ0 = 1. We initialize our simulations with zero net vertical magnetic flux configuration, i.e., a weak vertical magnetic field that varies sinusoidally in the radial (x) direction. We run the zero net-flux simulation to t = 120Ω−1 when turbulence is fully developed, then we start to add net vertical magnetic flux, which is applied uniformly in the grid (hence it does not introduce any divergence error) and the amount added is proportional to the time step dt. This process continues to t = 240Ω−1, when the net vertical magnetic field strength reaches the desired level B0, characterized by midplane plasma $\beta _0=2\rho _0c_s^2/B_0^2$. The simulations are run for about another 150 orbits to t = 1200Ω−1.

For our simulations, we consider β0 = 102, 103, and 104, corresponding to runs B2, B3, and B4. Ignoring vertical stratification, the most unstable MRI wavelength at the disk midplane is about $9.18\beta _0^{-1/2}H$ (Hawley et al. 1995). Including vertical stratification, calculations by Latter et al. (2010) suggest a resolution of about 25 cells per H for β0 = 100, and 50 cells per H for β = 103, to properly resolve the MRI channel modes. Our fiducial resolution meets this criterion for β0 = 100. For β0 = 103, we conduct an additional simulation (run B3-hr) with twice the resolution, which then properly resolves the most unstable linear modes. For this run, we initialize the simulation with a full net vertical magnetic flux, together with a sinusoidally varying vertical field component, so as to avoid an excessively strong initial channel flow. The list of simulation runs is provided in Table 1.

Table 1. Summary of Simulation Runs

Run β0 Domain Size Resolution Durationa
B2 102 4H × 8H × 12H 96 × 96 × 288 960Ω−1
B3 103 4H × 8H × 12H 96 × 96 × 288 960Ω−1
B4 104 4H × 8H × 12H 96 × 96 × 288 960Ω−1
B3-hr 103 3H × 6H × 12H 144 × 144 × 576 720Ω−1

Note. aDuration of the run with net flux fully added.

Download table as:  ASCIITypeset image

A density floor of ρFloor = 10−4, 3 × 10−5, and 10−5 are applied for simulations with β0 = 102, 103, and 104, respectively, to avoid numerical difficulties at magnetically dominated (low plasma β) regions. We will see in Figure 3 that horizontally averaged densities in the saturated states of all our simulations are at least two orders of magnitude larger than the density floor. In addition, we apply another density floor that is ten times smaller than ρfloor to all the left and right states inside the MHD integrator (we use the corner transport upwind (CTU) integrator with third order spatial reconstruction; see Stone et al. 2008 for details). This procedure is essential and greatly improves the robustness of the Athena MHD code.

Our simulations with β0 = 102 and 103 occasionally produce an extremely strong magnetic field in very localized regions (a few grid cells in the entire computational domain), which severely limits the simulation time step. In order to proceed, we have included some hyper-resistivity η∝J2 in regions with excessively large J, where J is the current density. More specifically, we define $j_d\equiv |\nabla \times {\boldsymbol{B}}|/\sqrt{\rho }$, and set Ohmic resistivity $\eta =0.1(c_s^2/\Omega)\cdot (j_d/10^3\Omega)^2$ when jd > 102Ω, and η = 0 otherwise. We have observed that for the β0 = 102 and 103 cases, hyper-resistivity is applied to at most 1% of all grid cells (most of the time less than 0.1%) and does not affect the general properties of the MRI turbulence.

Our simulations always launch a strong outflow (see Figure 9) that would drain the gas in the simulation box in tens to hundreds of orbits. In order for the system to reach a steady state that allows us to characterize the dynamical structure of the turbulent disk, we enforce the total mass contained in the simulation domain to be constant by multiplying the density of all grid cells by a common factor after each time step. The momentum remains fixed in each grid cell. This approach is similar to that of Ogilvie (2012), who considered much stronger magnetic fields that would launch even stronger disk outflow. As will be discussed briefly in Section 3.2.3, such mass addition does not significantly affect the energetics of the system for all our simulation runs.

In Figure 1, we show a snapshot of our simulation run B3 in the saturated state of the MRI turbulence. We clearly see the large density fluctuations as a result of the vigorous MRI turbulence, and a substantial fraction of the gas is in supersonic motion. Meanwhile, there is a systematic vertical velocity at the vertical boundaries, indicating a strong outflow. In the following two sections, we discuss in detail the properties of the MRI turbulence and the outflow, respectively.

Figure 1.

Figure 1. One snapshot from our simulation run B3. The surface plot shows the isosurface of gas density, increasing logarithmically from blue with ρ = 0.1 (cutoff) to red with ρ = 2.0. The streamlines indicate gas velocity, increasing linearly from white with v = 0 to dark red with v = 5cs.

Standard image High-resolution image

3. SIMULATION RESULTS: MRI TURBULENCE

One of the main diagnostics for the MRI turbulence is the rϕ component of the stress tensor, consisting of the sum of Reynolds and Maxwell stresses,

Equation (6)

and they characterize the rate of radial angular momentum transport. We are interested in the horizontally averaged vertical profiles (notated with an over bar, here $\overline{T}_{r\phi }$) of the stresses, and they are usually normalized by $\rho _0c_s^2$. Another useful measure is the mass weighted average of the stresses, normalized to the gas pressure (Shakura & Sunyaev 1973):

Equation (7)

where α denotes either Reynolds or Maxwell stresses (with subscript), or their sum (without subscript), which characterizes the total rate of radial angular momentum transport in the disk.

We show the evolution of αRey and αMax in Figure 2. The total stress in the initial zero net-flux simulation saturates at 〈αRey〉 ≈ 5.0 × 10−3 and 〈αMax〉 ≈ 2.0 × 10−2, respectively, consistent with previous works (e.g., Simon et al. 2012). The stresses increase steadily as we gradually add net vertical magnetic flux from t = 120Ω−1, and saturate as the full net flux is in place at t = 240Ω−1. We wait for about another 20 orbits and take the data from t = 360Ω−1 to the end of the simulations at t = 1200Ω−1 for all the time averages in this section, which amounts to about 130 orbits. For our high-resolution run B3-hr, we consider data from t = 180Ω−1 to the end of the simulation at t = 720Ω−1, which amounts to about 85 orbits. Throughout this paper, we use the angle bracket 〈 · 〉 to denote the time average.

Figure 2.

Figure 2. Time history of mass weighted Maxwell (solid) and Reynolds (dashed) stresses from simulations B2, B3, and B4. They are restarted from the end of the zero net-flux simulation at t = 120Ω−1. Net vertical magnetic field is added gradually to its full strength at t = 240Ω−1.

Standard image High-resolution image

In the saturated state, a common indicator for proper resolution of the MRI turbulence and convergence is the quality factor, defined as

Equation (8)

which is the ratio of a characteristic MRI wavelength to the grid scale. Here, j denotes either the $\hat{\boldsymbol{y}}$ or the $\hat{\boldsymbol{z}}$ dimension. Sano et al. (2004) found that Qz ≳ 6 is required for the MRI turbulence to be well resolved. Hawley et al. (2011) further suggested that Qz ≳ 10 and Qy ≳ 20 suffice to properly resolve the MRI turbulence. Although these results are mainly based on either unstratified simulations or stratified simulations with zero net vertical magnetic flux, we take them as a useful reference. In our simulations, the quality factors minimize at the disk midplane, where we find 〈Qz〉 to be about 87, 50, and 12 for runs B2, B3, and B4, respectively, and 〈Qy〉 is about 185, 110, and 25 for the three runs. For the high-resolution run B3-hr, we find 〈Qz〉 ∼ 110 and 〈Qy〉 ∼ 285. In all cases, the quality factors are well above the suggested value.

3.1. DISK STRUCTURE

3.1.1. Vertical Profiles

In the presence of a strong net vertical magnetic flux, the systems all saturate into very powerful MRI turbulence that substantially changes the structure of the disk. In Figure 3, we show the vertical profiles of various horizontally averaged quantities. The upper two panels are for gas density $\langle \bar{\rho }\rangle$ and plasma $\langle \bar{\beta }\rangle$ (ratio of gas to magnetic pressure). We see that the density profile deviates substantially from Gaussian due to strong magnetic support. Only for β0 = 104 is the Gaussian kernel still present; meanwhile, for β0 ⩽ 103, the entire disk becomes magnetically dominated, with plasma β ≲ 1 everywhere. In all cases, the averaged plasma β saturates at about 0.1 towards the disk surface, while in the extreme situation of β0 = 102, plasma β is about 0.1 everywhere as MRI saturates. The magnetic pressure profile shows a large vertical gradient when β ≲ 1 (as can be seen by dividing the first two panels of Figure 3). The magnetic pressure is dominated by contributions from the toroidal field (as can be compared with the third panel of the Figure), especially when β0 ≲ 103. Correspondingly, the disk is substantially puffed up.3 We also note a potential caveat: the fact that the entire disk becomes magnetically dominated implies that in real disks, there can be strong magnetic pressure gradient support in the radial direction. This may lead to substantial sub-Keplerian rotation, and in turn modify the behavior of the MRI, a situation that requires global simulations to be properly addressed (see further discussions in Section 5.2).

Figure 3.

Figure 3. Vertical profiles of (from top to bottom) gas density $\langle \bar{\rho }\rangle$, plasma $\langle \bar{\beta }\rangle$, magnetic pressure of the toroidal field $\langle \bar{P}_{B, y}\rangle$, turbulent magnetic energy $\langle \bar{E}_{\rm B,turb}\rangle,$ and turbulent kinetic energy $\langle \bar{E}_{\rm K,turb}\rangle$. In each panel, we show the profiles for runs B2 (β0 = 102, solid), B3, (β0 = 103 dashed), and B4 (β0 = 104, dash-dotted), respectively, and the thin dashed lines correspond to run B3-hr (β0 = 103 at high resolution).

Standard image High-resolution image

In the lower two panels of Figure 3, we show the profiles of horizontally averaged turbulent magnetic energy $\langle \bar{E}_{\rm B, turb}\rangle$ and kinetic energy $\langle \bar{E}_{\rm K, turb}\rangle$. Here, we have subtracted contributions from horizontally averaged mean fields and mean velocities at each vertical layer in the calculations: $\bar{E}_{\rm K, turb}\equiv \bar{E}_K-\bar{\rho }(\bar{v})^2/2$, $\bar{E}_{\rm B, turb}\equiv \bar{E}_B-(\bar{B})^2/2$. From the figure, we see that for both magnetic and kinetic components, turbulent energy increases substantially by about one order of magnitude as β0 changes from 104 to 103. As the net vertical field increases further, to β0 = 102, however, turbulent energy roughly stays at the same level as the β0 = 103 case. This indicates that the strength of the MRI turbulence saturates as β0 ≲ 103. On the other hand, the amplitude of the mean flow and mean magnetic field continue to increase with the net vertical magnetic field, and their contributions to the total energy gradually dominate that from turbulent energy. We will discuss this fact further in Section 3.2. Here, we note another caveat that in the presence of a strong mean toroidal field at the level of equipartition or higher, the linear dispersion relation of the MRI is strongly modified by field curvature (Pessah & Psaltis 2005), which is ignored in the shearing-box approximation, and again requires global simulations to be addressed properly.

For β0 ≲ 103, turbulence in all regions of the disk becomes trans-sonic or super-sonic, as one compares the profiles $\langle \bar{E}_{\rm K, turb}\rangle$ and $\langle \bar{\rho }\rangle$. This result implies strong compression and rarefaction, and hence large density variations. Looking at the density field, we find that the density fluctuations in our simulations generally show elongated structure that is slightly tilted anti-clockwise from the azimuthal direction, indicating spiral density waves excited by the MRI turbulence (Heinemann & Papaloizou 2009), although these structures are short lived due to the strong turbulence. To address such density variations, we measure the standard deviations of density fluctuations δρ in each horizontal layer, and normalize them to the horizontally averaged densities. The obtained profile of 〈δρ21/2/ρ for all simulations is shown in Figure 4. We see that density variations are significant in all cases. For runs B3 and B4, the midplane density fluctuation increases with the net magnetic flux, and reaches about 0.5 for β0 ≲ 103. The fluctuation level increases from the disk midplane, and reaches of the order of unity at disk surface as the disk become more and more magnetically dominated. The case with β0 = 102 shows features distinctive from higher β0 cases: except for a small bump around the disk midplane, the density fluctuation level stays roughly constant at about 0.5. We note that the turbulent kinetic energy profile in run B2 is similar to that in run B3, while the density near the disk surface in run B2 is a factor of several higher than run B3, indicating smaller turbulent velocity, and hence the profile of 〈δρ21/2/ρ for run B2 drops below that for run B3, which is essentially a reflection of weaker turbulence for run B2.

Figure 4.

Figure 4. Vertical profiles of density fluctuations 〈δρ21/2(z) normalized to horizontally averaged density $\langle \bar{\rho }\rangle$ for all simulation runs.

Standard image High-resolution image

3.1.2. Autocorrelation Function

Another useful diagnostic of the MRI turbulence is the autocorrelation function (ACF) of kinetic and magnetic energy fluctuations, which is defined as (Guan et al. 2009; Simon et al. 2012)

Equation (9)

where f denotes ${\boldsymbol{B}}$ for magnetic energy and $\sqrt{\rho }{\boldsymbol{v}}$ for kinetic energy, and the angle bracket denotes time average. Here, we consider a 2D ACF in the horizontal plane at some fixed vertical height z, and we subtract the horizontally averaged quantities in the evaluation of $\delta {\boldsymbol{f}}$: $\delta {\boldsymbol{f}}\equiv {\boldsymbol{f}}(x,y,z)-\bar{\boldsymbol{f}}(z)$. This is necessary, as the mean flow and mean field can become particularly strong in the presence of a net vertical magnetic field.

In Figure 5, we show the ACFs for magnetic energy fluctuations in all our simulations B2, B3, B3-hr, and B4, for both near the midplane region (left, for z ⩽ 2.5H) and the disk surface (right, 2.5Hz ⩽ 5H). The ACFs for the kinetic energy exhibit similar patterns, and hence we do not involve a separate figure to show them. The ACF for run B4 at the midplane where the net vertical magnetic flux is relatively small (β0 = 104) is very similar to that found in zero net-flux simulations (Guan et al. 2009; Simon et al. 2012), containing a narrow and elongated centroid that is tilted with respect to the azimuthal axis by about ζB ∼ 15°. Such a tilt angle is related to the empirical correlation between the plasma β and the stress parameter α ≈ 1/2β (Guan et al. 2009; Bai & Stone 2011), which also applies for the midplane region of our run B4 as we can compare Figures 3 and 6. Moving up to the disk surface, we find that the centroid becomes broader in the upper layer, and the tilt angle also becomes smaller. These features indicate a longer correlation length in the radial direction, as well as more isotropy in the magnetically dominated disk corona. The ACFs in run B3 at the midplane have a similar tilt angle as in run B4, typical for MRI turbulence, while at both midplane and corona, the peaks in the ACFs are as broad as the corona region of run B4, which is in line with the fact that the entire disk has become magnetically dominated.

Figure 5.

Figure 5. Two-dimensional autocorrelation function (ACF) in the horizontal plane for magnetic energy fluctuations for our simulations B4 (β0 = 104), B3, B3-hr (β0 = 103), and B2 (β0 = 102), from top to bottom. Left panels correspond to the ACFs for |z| < 2.5H, while right panels are for |z| > 2.5H. Note that the domain size for run B3-hr is slightly smaller.

Standard image High-resolution image
Figure 6.

Figure 6. Vertical profiles of Maxwell stress (solid) and Reynolds stress (dashed) for runs B2, B3-hr, and B4 (from top to bottom panels). In each panel, we show the total stress (bold) as well as contribution from turbulent fluctuations (thin). The stresses are normalized by $\rho _0c_s^2$.

Standard image High-resolution image

Our run B2 with β0 = 100 exhibits an extreme situation: fluctuations are highly elongated in the azimuthal direction, with the measured correlation length comparable to the azimuthal size of our simulation box, which might suggest that our azimuthal box size is not sufficient to properly resolve the MRI turbulence. Nevertheless, the shape of the ACF indicates that the fluctuations are quasi-axisymmetric (which is also evident as one views the raw simulation data), and the radial fluctuations are well fitted into our simulation box.

3.1.3. Numerical Convergence

Due to the large computational cost, we have not carried out a full resolution study in this paper. However, we do expect our simulation results to converge in runs B2 and B3-hr, where we have adequate resolution to properly resolve the linear MRI modes as suggested by Latter et al. (2010). Based on the quality factor criterion (Hawley et al. 2011), we expect all our simulations to resolve the MRI turbulence well. Below, we compare the results between our simulation runs B3 and B3-hr and briefly discuss numerical convergence.

We see from Figure 5 that the shape of the ACFs in our runs B3-hr and B3 are similar, but the ACF in run B3-hr is more centrally-peaked than its low-resolution counterpart. Since the ACF is simply the Fourier transform of the power spectrum, this indicates that the high-resolution run gives a flatter power spectrum with more power on the small scales. In addition, comparing various profiles in Figure 3 between runs B3 and B3-hr, we see again that the profiles are very similar, with higher resolution giving slightly higher turbulent energy. This is also reflected in Table 2, where the Shakura–Sunyaev α from run B3-hr is about 25% higher. Similarly, in Figure 4, we see that run B3-hr leads to slightly larger density fluctuations. Later, in Figure 9, slightly higher mass outflow rate is achieved in run B3-hr.

Table 2. Energetics and Angular Momentum Transport

Run 〈αMax 〈αRey $\langle T_{z\phi }^{\rm Max}\rangle$ $\langle T_{z\phi }^{\rm Rey}\rangle$ Psh $\dot{E}_K$ $\dot{E}_P$ $\dot{m}_w$
B2 0.92 0.12 0.061 0.048 3.90 0.89 1.26 0.0944
B3 0.37 0.077 0.0085 0.0041 1.69 0.079 0.15 0.0138
B3-hr 0.47 0.086 0.010 0.0046 2.08 0.086 0.21 0.0160
B4 0.061 0.014 0.00072 0.00036 0.28 0.0065 0.010 0.0013

Note. All in natural units, with ρ0 = cs = Ω = 1.

Download table as:  ASCIITypeset image

The systematically higher saturation amplitude of the MRI in the high-resolution run discussed above might suggest that our simulations for β0 = 103 have not fully reached numerical convergence. However, such a higher saturation amplitude may not be due to under-resolved MRI turbulence, and several other factors are likely to play a role. While a slightly smaller horizontal box size in run B3-hr can be a potential cause, the fact that the peaks of the ACF are well contained in the box makes it somewhat unlikely (Simon et al. 2012). Another possibility is that in stratified MRI simulations, the property of the MRI turbulence is largely affected by the dynamo process (see detailed study in Section 3.3), which generates a strong mean toroidal field that changes sign over time. It appears that the instantaneous large-scale toroidal field generated in run B3-hr is on average stronger than that in run B3, as can be judged from Figure 7, and both are above equipartition field strength. Such a larger mean toroidal field leads to stronger magnetic support to the disk, resulting in slightly different vertical disk structures between the two runs B3 and B3-hr. In this respect, one should be careful when discussing the issue of convergence, since comparing turbulent properties between simulations with different disk structures can be misleading. Since the mechanism of the dynamo is far from being understood, and the dynamo in the presence of strong net vertical flux behaves very differently from the case with zero net vertical flux, we defer this convergence issue for future study.

Figure 7.

Figure 7. The time evolution of the horizontally averaged profiles of toroidal magnetic field in all simulation runs, from top to bottom: B2, B3, B3-hr, and B4. Here, we mark t' = 0 as the time when full net vertical flux is added (t = 240Ω−1 for B2, B3, and B4, t = 0 for B3-hr). The white contours mark plasma β = 1.

Standard image High-resolution image

In summary, although quantitative differences exist between runs B3 and B3-hr, the physical properties of the MRI turbulent disk from the two runs all agree with each other. We therefore consider both runs as valid representations of the physical system, with run B3-hr as a more accurate model.

3.2. Energetics and Angular Momentum Transport

There are two sources of angular momentum transport in accretion disks, namely, the radial transport (or re-distribution) via turbulence and the vertical transport via disk wind. We study them separately in this subsection and then briefly discuss the energetics in our simulations.

3.2.1. Radial Transport of Angular Momentum

The radial transport of angular momentum is characterized by the α parameters defined in Equation (7). In Figure 6, we show the vertical profiles of Maxwell and Reynolds stresses for all our simulations. Both the Maxwell and Reynolds stresses increases monotonically with net vertical magnetic flux, and approach a few times $0.1\rho _0c_s^2$ for β0 = 102. In the mean time, the profile gradually changes from centrally peaked for large β0, to a more or less flat profile at β0 = 102. In particular, the slope of the wing beyond ±2H becomes more level as β0 decreases. We can further compare our run B4 with zero net vertical flux stratified MRI simulations (Stone et al. 1996; Miller & Stone 2000; Guan & Gammie 2011; Simon et al. 2012), where the slope of the wing is even steeper. These results suggest a general trend that the surface region plays a more and more important role in the radial angular momentum transport as the net vertical magnetic field increases. Also, we see that the Maxwell stress is always larger than the Reynolds stress, but the Reynolds stress becomes more and more important toward the disk surface.

To further decode the mechanism for the radial transport of angular momentum, we separate the contributions from the mean field/mean flow ($\langle \bar{\rho }\bar{v}_x\bar{v}^{\prime }_y\rangle$ or Reynolds stress and $\langle -\bar{B}_x\bar{B}_y\rangle$ for Maxwell stress) and the turbulent stress (the rest). We remind the reader that we compute the mean stress from horizontally averaged quantities in each snapshot and then take the time average. As the net vertical field increases, the turbulent component increases steadily with net field and saturates for β ≲ 103 in a way similar to the turbulent energy density profile in Figure 3. There are two intriguing features in Figure 6. First, in the disk surface region (near our vertical boundary), the stress is dominated by the mean field/mean flow components (i.e., magnetic breaking) rather than turbulence, and the contribution from the mean field/mean flow components rapidly increases as the net field increases. In run B4, the mean field component dominates the wings of the Maxwell stress, while in run B3-hr (and B3), the mean field component dominates the entire disk. Second, for a sufficiently strong background magnetic field β0 ≲ 102, the turbulent stress become completely unimportant at all locations, and the turbulent Maxwell stress can even become negative. Interestingly, the central region where the turbulent Maxwell stress is positive coincides with the bump in the density fluctuation (Figure 4).

The decline of the turbulent component and rise of the mean field/mean flow component of the stress as one increases net vertical flux strongly contrasts with conventional unstratified shearing-box simulations. In those simulations, the initial net toroidal magnetic field is mostly set to zero. Although the net toroidal magnetic flux is not a conserved quantity, it generally stays very close to zero for the duration of the simulations. With stratification, the large-scale toroidal field evolves substantially, and for β0 ≲ 103, it well exceeds equipartition strength (see Figures 7 and 12). Although the linear growth of the MRI is not directly coupled to the toroidal field, the nonlinear saturation certainly depends on it (Hawley et al. 1995). Therefore, it is not appropriate to compare unstratified MRI simulations with stratified simulations with the same β0 unless the unstratified simulation contains a substantial net toroidal magnetic field.

Integrating the stresses, one obtains the α parameter defined in Equation (7), and the results are listed in Table 2. The α parameter increases monotonically with net magnetic flux, from a total of about 〈α〉 ≈ 0.075 for β0 = 104 to 〈α〉 ≈ 1.0 for β0 = 100. The ratio of 〈αMax〉 to 〈αRey〉 increases from about 4 for β0 = 104, which is similar to that in zero net vertical flux simulations (e.g., Davis et al. 2010), to about 7 for β0 = 102, which is mainly due to an increasing contribution from the mean (rather than turbulent) magnetic field.

We note that in run B2, the Maxwell stress only decreases very slowly with height, while the Reynolds stress increases with height, making the volume integrated stress limited by the vertical extent of our simulation box. However, we argue that increasing the box size does not necessarily improve the situation. In this case, the radial angular momentum transport is dominated by the mean field and mean flow. As we shall discuss in Section 4, the strength of the large-scale mean field at the disk surface is intimately connected to the global condition of the disk and can not be determined in the local shearing-box approach. Therefore, uncertainty still remains as one uses a taller box. The measured 〈α〉 value for run B2 should be only taken as a reference.

3.2.2. Vertical Transport of Angular Momentum

Angular momentum can also be extracted from the disk directly in the vertical direction through outflow and magnetic fields, which is determined by the zϕ components of the Reynolds and Maxwell stresses exerted at vertical boundaries:

Equation (10)

where, in practice, subscripts bot and top mean that we measure these quantities at the last layer of the top and bottom of the computational domain. The torque exerted by the wind stresses can be obtained by simply multiplying $\langle T_{z\phi }^{\rm Rey}\rangle$ and $\langle T_{z\phi }^{\rm Max}\rangle$ by the radius R. Including the contributions from radial and vertical transports, the accretion rate can be approximately written as4

Equation (11)

where $\dot{M}_r$ and $\dot{M}_z$ represent contributions to the accretion rate from radial and vertical transport, respectively. Qualitatively, one may replace the vertical integral ∫dz by the disk scale height H. We see that given the same level of stress Trϕ and Tzϕ, vertical transport is more efficient than radial transport by a factor of about R/H.

Being a local shearing-box simulation, however, R is unspecified, and moreover, the location of the central object is unspecified (can either at inner or outer $\hat{\boldsymbol{x}}$ direction), making the sign of R ambiguous. A physical disk wind requires that the sign of Tzϕ on the two vertical boundaries be the opposite and are steady, and is closely related to the symmetry of the wind solution as we will discuss in detail in Section 4.3. In this subsection, we only consider the absolute value of Tzϕ and average it over the two vertical boundaries (i.e., assuming the flow structure and magnetic configuration has the physical geometry),5 hence the relative importance between radial and vertical transport can be rewritten as

Equation (12)

We list the values of $T^{\rm Rey}_{z\phi }$ and $T^{\rm Max}_{z\phi }$ from each simulation run in Table 2, with both normalized by $\rho _0c^2_s$. The wind stresses appear to depend more sensitively on β0, and increase by about two orders of magnitude from β = 104 to β = 100. This may suggest that the importance of wind transport relative to turbulent transport increases with net vertical field. Taking H/R = 0.1, we find that wind transport, if present, should be small (less than 25%) compared with radial transport for β0 = 104, which is consistent with the results by Fromang et al. (2012). It would become more or less comparable to radial transport for β0 = 103, while it would dominate radial transport for β0 = 102. In addition, the contributions from Maxwell $T_{z\phi }^{\rm Max}$ and Reynolds $T^{\rm Rey}_{z\phi }$ components are roughly 2: 1 in runs B3 and B4, while the Reynolds component becomes more important for run B2. However, again, as we shall discuss in Section 4, the wind stress can only be determined from the global approach, while our measured 〈Tzϕ〉 values should also be taken as a reference.

3.2.3. Energetics

Using an isothermal equation of state, total energy is not conserved in the simulation. However, it is important to examine the energy budget for consistency, such that the work done by the Keplerian shear (i.e., from radial shearing-box boundaries) exceeds the energy loss from the vertical boundaries, with the rest of the energy presumably escaping from the disk in the form of radiation.

In ideal MHD, the energy flux is given by

Equation (13)

where epsilon is the internal energy. Consider the net energy flux entering the two sides of the radial boundaries, where we compute ${\boldsymbol{F}}_E\cdot {\boldsymbol{e}}_x$ and integrate over height and azimuth. Note that ${\boldsymbol{u}}$ contains background shear, and one should shift uy by 3ΩLx/2 between the two boundaries. Correspondingly, only terms containing ${\boldsymbol{u}}$ contribute, while all other terms cancel due to the shear-periodic boundary condition. The non-vanishing terms turn out to be Maxwell and Reynolds stress, and the work done by the radial boundaries per unit horizontal area per unit time read

Equation (14)

where $\Sigma =\int \langle \bar{\rho }(z)\rangle dz$ is the column density.

The energy loss from vertical boundaries is given by

Equation (15)

where ${\boldsymbol{n}}$ is the unit vector pointing away from the disk in the vertical direction, and we sum over contributions from the top and bottom vertical boundaries. The first terms in the bracket represent the kinetic energy loss and the PdV work done by the mass outflow, and the second term represents energy loss from the Poynting flux.6 We have ignored the energy loss term associated with internal energy due to the mass outflow, since we keep feeding mass to the system which balances this term exactly. This term is comparable to the PdV work term which, as we shall see, is always small compared with the total $\dot{E}$ (hence our mass addition do not substantially alter the energy balance).

In Table 2, we also list the values of Psh, $\dot{E}_K$, and $\dot{E}_P$, normalized in natural units. We see that Psh is always larger than the sum of $\dot{E}_K$ and $\dot{E}_P$, and hence energy conservation is not violated. Also, we see that $\dot{E}_P$ is always larger than $\dot{E}_K$ by a factor of about 1.5 to 2, and hence the Poynting flux dominates energy loss. In addition, under the assumption of the isothermal equation of state adopted in our simulations, we find that as net vertical magnetic field increases, $\dot{E}_K+\dot{E}_P$ roughly increases as 1/β0, which is much faster than the rate Psh increases. This may suggest that radiative efficiency decreases as a larger fraction of work done by the shear is converted into the wind energy loss.

3.3. Dynamo Activities

It is well known in stratified shearing-box simulations with zero net vertical magnetic flux that the mean toroidal field component experiences periodic flips on a time scale of 10 orbits (e.g., Brandenburg et al. 1995; Shi et al. 2010; Davis et al. 2010; Simon et al. 2012). Since a zero background vertical field does not give any instability on its own in the linear theory, the MRI activity in these simulations, together with the periodic flips, strongly suggest that a magnetic dynamo is in operation. It is generally believed that shear, turbulence, and large scale azimuthal field are responsible for sustaining the dynamo cycles (Vishniac & Brandenburg 1997; Yousef et al. 2008; Lesur & Ogilvie 2008; Gressel 2010), yet the detailed mechanism remains unsettled.

In the presence of a net vertical magnetic field, the MRI can be sustained from the linear modes, hence a dynamo is not necessary to explain the turbulence. With very weak vertical net magnetic flux, dynamo behavior was also observed in Suzuki & Inutsuka (2009), Suzuki et al. (2010) and Fromang et al. (2012), whose simulations correspond to β0 = 104–106, although it was not studied in detail. Our simulations allow us to systematically study the evolution of the MRI dynamo with net vertical magnetic flux. In Figure 7, we show the space-time plot of the horizontally averaged azimuthal magnetic field for all our simulation runs. It is obvious that for β0 = 104, the mean azimuthal field still undergoes cyclic oscillations between positive and negative signs, while for β0 = 100, the dynamo behavior disappears, where the mean azimuthal field in the entire simulation box is completely dominated by one sign at all times. The case with β0 = 103 is somewhat marginal, where the dynamo-like behavior is present for part of the time, while for the rest of the time the mean azimuthal field is predominantly one sign.

In zero net vertical flux stratified shearing-box simulations (see Figure 15 of Simon et al. 2012 for a most clear rendering), the dynamo pattern is highly repeatable in the space-time plot, also known as the "butterfly diagram," with a well-defined period of about 10 orbits. The dynamo cycles in our run B4 with β0 = 104, however, are highly irregular. Such behavior is also observed in Fromang et al. (2012). Also including the marginal case of run B3 with β0 = 103, we see that there is a systematic trend that as net vertical field increases, the dynamo cycle weakens by becoming more irregular with less periodicity. The flipping of the mean azimuthal field becomes more sporadic, and the mean time interval for field flipping also becomes longer. As β0 reaches below 103, the flipping time interval virtually becomes infinity, and the dynamo is completely quenched.

We note that the most unstable linear MRI mode is properly resolved in run B2, which does not show dynamo behavior, but may not be well resolved in runs B3 and B4, which show dynamo activities. Hence, questions arises over whether the dynamo activity depends on the proper resolution of the linear MRI modes. Our run B3-hr, which is the same as run B3 but properly resolves the most unstable MRI mode, is designed to clarify this potential ambiguity. We see that the space-time pattern for the mean azimuthal field is very similar in the two runs: dynamo activity appears for only part of the time. This comparison further justifies that the dynamo behaviors observed in our simulations are real.

It is natural to ask about what physical effects control the dynamo activities. Phenomenologically, we see that in the presence of the dynamo, the disk is separated into two distinct regions, namely, the region near the midplane with a relatively weak magnetic field where the dynamo appears to be developed, and a highly magnetized region outside. In Figure 7, we have overplotted white contours that mark the total plasma β = 1 (i.e., total magnetic pressure equals the gas pressure). Remarkably, these white contours perfectly separate the two regions for both runs B3 and B4, while for run B2, the entire disk is magnetically dominated, and hence there is no contour at all times. It is intriguing to notice that the dynamo is present whenever the magnetic field strength in the disk midplane region is below equipartition strength.

Meanwhile, the criterion of β = 1 also applies to magnetic buoyancy (i.e., interchange and Parker-type instabilities, Newcomb 1961; Parker 1966), as discussed in a number of zero net-flux simulations (Blaes et al. 2007; Shi et al. 2010; Guan & Gammie 2011; Simon et al. 2011). With an isothermal equation of state, and assuming a zero mean vertical field, the fluid is buoyantly unstable if the magnetic energy density decreases with height (Guan & Gammie 2011). It turns out that the magnetic field strength tends to be constant in the gas pressure dominated disk midplane regions, presumably due to efficient turbulent mixing, while it falls off with height in the magnetically dominated corona. Correspondingly, for regions with β < 1, the fluid is unstable due to magnetic buoyancy, while for regions with β > 1, the fluid is buoyantly (marginally) stable and dynamo activities produce cyclic alternations of the mean magnetic field. Although the analysis of the Parker instability in previous works assumes zero mean vertical field, it continues to appear to be valid in our simulations with a net vertical field, which is likely because the net vertical field in our simulations is much weaker (by at least a factor of 10) than the mean azimuthal field.

The strong azimuthal magnetic field in our simulations with β0 ≲ 103 is to some extent similar to the simulations performed by Johansen & Levin (2008). They initiated their simulations by a pure azimuthal field with equipartition strength (with zero net vertical magnetic flux), and found that the interplay between the Parker instability at the disk surface layer and the MRI near the midplane effectively produces a magnetic dynamo, and the azimuthal flux is well confined to the disk. Simulations of Gaburov et al. (2012) for the tidal disruption of a molecular cloud approximately realize the situation. The key difference in our case is that a strong outflow is produced in the presence of net vertical magnetic field (see Section 4), hence azimuthal flux continuously escape the simulation box and has to be continuously generated within the disk. Therefore, although the Parker instability is likely to be present in the entire disk when β0 ≲ 103, the dynamo mechanism of Johansen & Levin no longer operates in the presence of strong net vertical magnetic field. On the other hand, magnetic dissipation due to the Parker instability is likely to play an important role in the thermal structure of the disk (Uzdensky 2012), which deserves future exploration.

4. SIMULATION RESULTS: OUTFLOW

It has been found in Suzuki & Inutsuka (2009) and Suzuki et al. (2010) that the inclusion of a net vertical magnetic field leads to a strong mass outflow in shearing-box MRI simulations, and the rate of the mass outflow is approximately proportional to 1/β0. The mass outflow was interpreted as the launching of a wind from the disk and would potentially connect to a Blandford & Payne (1982)-type global disk wind. The connection between an MRI-driven outflow and the global disk wind has been further investigated by Fromang et al. (2012), who focused on the case with β0 = 104. They argue that the MRI is able to launch the MCW that is strongly time-dependent. A similar conclusion has been made by Lesur et al. (2013) for the case with β0 ∼ 10, where there is only one single MRI mode fitted into the disk with a physical symmetry for the wind, and secondary instabilities are likely to make it time-dependent. Very strong outflows are also observed in our simulations. However, we argue that the outflow seen in shearing-box MRI simulations is unlikely to be directly connected to a global disk wind, as we elaborate below.

4.1. Structure of the Outflow

4.1.1. Critical Points

The global disk wind model by Blandford & Payne (1982) states that when the magnetic field at the surface of a thin disk is inclined relative to the disk normal for more than 30°, outflowing gas can be accelerated centrifugally along field lines, extracting angular momentum from the disk, and is gradually collimated by the magnetic hoop stress. A crucial ingredient of this picture is the wind launching/mass loading from the disk, which is intrinsically connected to the gas dynamics in the disks. The wind launching process has been studied analytically under the local shearing-box approximation (Wardle & Koenigl 1993; Ogilvie & Livio 2001; Ogilvie 2012), where all the radial gradients except the shear are omitted to reduce the problem to be fully one-dimensional (1D). The 1D solutions are then matched to global wind solutions, where the mass loading rate is determined as an eigen-value problem.

A physical wind solution should pass three critical points, namely, the slow and fast magnetosonic points, and the Alfvén point. In the 1D (laminar) model, they are given by

Equation (16)

for fast and slow magnetosonic points, and

Equation (17)

for the Alfvén point, where $v_{Az}=\sqrt{B_z^2/\rho }$ is the vertical component of the Alfvén velocity, and vA is the full Alfvén velocity. The requirement that the flow passes smoothly through these critical points poses three eigen-value problems that would determine three physical quantities, such as the mass loading rate and the field inclination at disk surface. In the local shearing-box approach, the fast magnetosonic point, and sometimes the Alfvén point, are sufficiently high above the disk where the local approximation no longer applies, and they need to be captured in the global models. The lack of one or two critical points implies that the solution obtained within the computational domain has one or two degrees of freedom.

In Figure 8, we show the time and horizontally averaged profiles of the gas vertical velocity as a function of height. In all cases, the vertical velocity rapidly increases towards the disk surface, and becomes transsonic or supersonic as the flow leaves the simulation box. We also calculate the location of the critical points based on Equations (16) and (17), where the Alfvén velocity is based on the time average of the absolute value of the horizontally averaged magnetic field. We find that for all cases, the fast magnetosonic points are beyond our simulation box (the fast magnetosonic speed is about 3cs or larger at vertical boundaries), while the Alfvén (marked in the Figure) and slow magnetosonic points are well contained in the box. This fact means that the flow structure is not fully determined and has one degree of freedom.

Figure 8.

Figure 8. Horizontally-averaged vertical velocity profiles for fiducial resolution simulations B2, B3, and B4 with β0 = 102 (solid), 103 (dashed), 104 (dash-dotted) respectively. Also marked are the locations of Alfvén points (solid dots) and slow magnetosonic points (circles). The fast magnetosonic points are beyond the simulation domain.

Standard image High-resolution image

4.1.2. Conservation Laws

It is well known that for a laminar magnetized wind, the gas follows the magnetic field lines, and the flow is characterized by a number of conserved quantities along the streamlines (Blandford & Payne 1982; Pelletier & Pudritz 1992). These conservation laws provide very useful diagnostics that help us understand the mechanism for the launching and acceleration of the outflow, as well as the angular momentum transfer processes. The forms of these conservation laws in the shearing-box framework are derived in Lesur et al. (2013). In the shearing-box, it was shown that poloidal gas streamlines do not necessarily follow exactly the poloidal field lines, which brings new terms to the conservation law. Nevertheless, the difference is generally small. Since strong turbulence are present in all our simulations while these conservation laws are derived for a laminar flow, they only hold approximately and serve for diagnostic purpose.

The starting point is mass conservation, where

Equation (18)

where an overline indicates horizontal average, and κ is constant. We show the profile of $\overline{\rho v_z}/\overline{B}_z$ (or κ) for all our simulation runs in the top panel of Figure 9. Because we constantly add mass to the simulation box to maintain a steady state, the vertical profile of $\overline{\rho v_z}$ does not become flat until beyond z ∼ ±3H. There is some asymmetry in the high-resolution run B3-hr, which is most likely due to its shorter run time and the lack of statistics. The asymptotic values of $\overline{\rho v_z}$ are used to calculate the mass loss rate $\dot{m}_w$ from the simulations

Equation (19)

and we have listed the value of $\dot{m}_w$ in Table 2. Further details about the mass outflow rate will be discussed in Section 4.2.

Figure 9.

Figure 9. Top: time averaged vertical profile of mass outflow rate normalized to the mean vertical magnetic field (or κ, see Equation (18)) for all our simulation runs. Bottom: time averaged mass loss rate from the simulation domain as a function of net vertical flux. The thick/thin symbols correspond to fiducial/high resolution simulations (B3-hr/B3).

Standard image High-resolution image

As long as κ is constant, as is valid beyond z ± 3H in our simulations, specific angular momentum and energy are conserved along streamlines. Following Lesur et al. (2013), the specific angular momentum reads

Equation (20)

where ${\mathcal {L}}\equiv v_y+\Omega x/2$ is the fluid part of the specific angular momentum. The partition between ${\mathcal {L}}$ and $\bar{B}_y/\kappa$ then describes the angular momentum exchange between the gas and the magnetic field.

Energy conservation is given by the Bernoulli invariant, which reads (Lesur et al. 2013, without correction terms)

Equation (21)

where EBer represents the specific energy along a streamline, with the four terms denoting kinetic energy, enthalpy, potential energy, and the work done by the magnetic torque, respectively, and

Equation (22)

Note that the full velocity ${\boldsymbol{u}}$ (rather than ${\boldsymbol{v}}$) enters EK, and the potential energy for the shearing-box is

Equation (23)

We extract the mean flow of the gas and integrate to obtain the streamline. Because the disk midplane regions are generally the most turbulent, with κ varying with height, we do not expect conservation laws to hold in these regions and only trace streamlines from z = 3H to z = 6H, setting x = 0 at z = 3H (note that by construction, the conservation laws do not depend on the choice of the zero point of x). This treatment covers the most interesting surface regions of the disk where the launching and acceleration of the outflow take place. For illustration purpose, we only consider runs B2 and B4. For run B2, the large-scale flow and magnetic structures are more or less steady (as seen from Figure 7), and hence we take the long-term average to obtain the desired physical quantities along streamlines. For run B4, due to the dynamo activities, we only pick up a short period between t' = 560Ω−1 and t' = 600Ω−1, where the magnetic structure in the upper side of the disk is quasi-steady, as judged from Figures 7 and 11.

In Figure 10, we show the vertical profiles of individual terms in the specific energy and angular momentum for streamlines obtained from runs B2 and B4. We see from the black dashed lines that energy and angular momentum are approximately conserved beyond about 4H for both cases. The rise of the blue lines in the energy plots indicate flow acceleration. This is mostly compensated for by the reduction of potential energy (red) and the work done by the magnetic torque (green). Similarly, in the bottom panels of Figure 10, we see that the increase of fluid angular momentum is compensated for by the reduction of magnetic torque. Therefore, it becomes clear that the acceleration is magnetocentrifugal in nature. Note that the role played by centrifugal potential ϕ and the magnetic torque $-\bar{B}_yv_y^*/\kappa$ can be exchanged by shifting the zero point of x, hence they represent the same effect. For example, if we were to trace the streamlines from the disk midplane for the case of run B2, then we would find that the acceleration is almost completely due to the centrifugal potential, since for Run B2 the poloidal field lines are almost always inclined for more than 45° relative to disk normal, as can be seen in Figure 12.

Figure 10.

Figure 10. Energy (top) and angular momentum (bottom) along a streamline in the disk surface from z = 3H to 6H for run B2 (left, with long time average) and B4 (right, averaged between t' = 560Ω−1 and t' = 600Ω−1). For the energy plot, the various terms in Equation (21) are represented by: EBer (black dashed), EK (blue solid), ET (magenta solid), EB (green solid), and ϕ (red solid). For the angular momentum plot, the various terms in Equation (20) are represented by: f (black dashed), ${\mathcal {L}}$ (blue solid), $-\bar{B}_y/\kappa$ (red solid).

Standard image High-resolution image

4.2. Mass Loss Rate

In the bottom panel of Figure 9, we show the total mass outflow rate $\dot{M}_w$ measured from the two vertical boundaries for all our simulations, with the numbers given in Table 2. We see that $\dot{m}_w$ scales roughly linearly with 1/β0, which is consistent with Suzuki & Inutsuka (2009) and Suzuki et al. (2010), while extending the range of β0 from 104 in their case down to 102. From the definition of the Alfvén point, the mass loss rate can be expressed as

Equation (24)

where |A denotes the value at the Alfvén point. A linear scaling of $\dot{m}_w$ with 1/β0 indicates that $\rho \propto B_z^2\propto 1/\beta _0$ at the Alfvén point, which is in line with the fact that the location of the Alfvén points decreases in height as the net vertical field increases (the change in density profile due to magnetic pressure support is not as significant). Moreover, based on the same scaling, vz at the Alfvén point is expected to be constant, which is roughly the case as seen from Figure 8.

4.2.1. Determining $\dot{m}_w$

In Figure 9, we also indicate the scaling relation of $\dot{m}_w$ with β0 from Suzuki & Inutsuka (2009). We find that while our measured $\dot{m}_w$ follows the same trend, our proportional coefficient is about a factor of five smaller than theirs. Besides the fact that our vertical box size is slightly larger (12H versus $8\sqrt{2}\approx 11.3H$), this difference is mainly due to the different outflow boundary conditions implemented in our simulations from theirs. In fact, the rate of mass outflow $\dot{m}_w$ is not a well determined quantity in local shearing-box type simulations because of the additional degree of freedom. Moreover, in both Fromang et al. (2012) and Lesur et al. (2013), it was found that $\dot{m}_w$ decreases as the vertical size of the simulation box increases, which again suggests that $\dot{m}_w$ can not be reliably obtained from shearing-box simulations.

Since the vertical gravity increases linearly with z in shearing-box approximation, the gas has to overcome a stronger and stronger potential well in order to flow out as the box height increases. In reality, the vertical increases only as z/(R2 + z2)3/2, where R is the distance to the central object, and hence shearing-box approximation fails at zR. Fromang et al. (2012) considered a higher order expansion of the real potential which would also require the radial potential to be modified to the same order. They noticed that this procedure would introduce curvature and no longer fit into the shearing-box framework.

The mass loss rate in real systems should be a well defined quantity, and we speculate that it should be comparable to the mass loss rate in shearing-box simulations performed with box size Lz ≈ 2R. Suppose the mass outflow rate scales as

Equation (25)

where A is a dimensionless constant, the index ν describes the scaling of $\dot{m}_w$ on the box height, and we have assumed that $\dot{m}_w\propto 1/\beta _0$, as suggested by the simulations discussed in the previous subsection, with B0 denoting the background vertical field strength. With a little algebra, we find

Equation (26)

In particular, if ν ≈ 1 (S. Fromang 2011, private communication), then our hypothesis (Lz = 2R) yields $\dot{m}_w\approx AB_0^2/2R\Omega$. This means that the mass loss rate can be found once the coefficient A is known. Moreover, $\dot{m}_w$ is solely determined by the background vertical field strength and does not depend on the disk thickness or the disk surface density. On the other hand, if ν deviates from 1, then one would expect $\dot{m}_w\propto (H/R)^{\nu -1}$.

In summary, although $\dot{m}_w$ is not well determined from shearing-box simulations, we argue that by carefully studying the dependence of $\dot{m}_w$ on the height of the simulation box, it is possible to obtain a physical estimate of the mass loss rate.

4.2.2. Evolutionary Scenarios

While we have demonstrated the correlation $\dot{m}_w\propto 1/\beta _0$, we have not discussed the range for which it is applicable. Reading from Table 2, we see that $\dot{m}_w$ for β0 = 100 is somewhat smaller than the expected linear correlation, which is suggestive of saturation. Indeed, the rate of mass outflow measured in Lesur et al. (2013) for a similar numerical setup (their run 1Dz6) with β0 ∼ 10 gives $\dot{m}_w=0.234$ (multiplied by 2 to account for two surfaces), which is only a factor of 2.5 times higher than our case with β0 = 100 ($\dot{m}_w=0.0944$). With an even stronger magnetic field that is too strong for the MRI to operate, Ogilvie (2012) showed that the rate of mass outflow should rapidly decrease once β0 drops below 1. Therefore, the scaling relation $\dot{m}_w\propto 1/\beta _0$ holds up to β0 ≳ 100, beyond which the mass loss rate slows down and will eventually fall off rapidly for a strong super-equipartition field.

All together, these results suggest an evolutionary scenario. Assuming no magnetic flux evolution, the disk would lose mass at a constant rate (assuming ν = 1) until the midplane net vertical field exceeds equipartition. Lesur et al. (2013) explored such a scenario in their 1D simulations and confirmed the steady decrease of mass loss rate with increasing magnetization in the strong field regime. A strongly magnetized thin accretion disk may gradually evolve into such a stage, where the rate of mass loss is self-regulated so that it loses mass faster as more mass is fed into the disk (increasing β0, but still β0 < 1), while it loses mass slower if the mass loss is not compensated (decreasing β0). The disk structure would be similar to a jet-emitting disk (Combet & Ferreira 2008), which is tenuous with super equipartition field. Alternatively, if the evolution starts from a strongly magnetized thick disk, then the mass loss rate can be so large that the disk will be depleted in a few orbits. For example, without feeding mass to our run B2, the mass loss time scale is only about five orbits. Such rapid mass loss may lead to catastrophic disruption of the disk, such as in core-collapse supernovae where a rapidly rotating progenitor core is threaded by extremely strong magnetic fields (Burrows et al. 2007).

4.3. Fate of the Outflow

In this subsection, we discuss whether the outflow launched from shearing-box MRI simulations can be connected to a global Blandford–Payne type disk wind. The key to this question lies in the symmetry. We note that due to the neglect of curvature, the shearing-box approximation does not specify which side of the radial domain the central object is located. For a physical disk wind, symmetry requires that the outflow from the top and bottom sides of the disk should be inclined towards the same radial and azimuthal directions (which would be considered as radially outward and trailing), and remain so at all times. Since gas flows along magnetic field lines, the mean magnetic field at the top and bottom sides of the disk should be bent to the same direction.

The first difficulty for the outflow-wind connection is the magnetic dynamo, applicable to the cases with β0 ≳ 103 as we discussed in Section 3.3. The constant flipping of the mean azimuthal (as well as radial, see Figure 11) field lines means that the direction of the outflow would constantly swap between radially inward and outward, either of the directions would be invalid for a global disk wind. Although our simulations with β0 = 104 are in many aspects very similar to simulations by Suzuki & Inutsuka (2009), Suzuki et al. (2010), and Fromang et al. (2012), they argue that such phenomenon leads to the time variability of the wind. However, we stress here that the simple fact of the cyclic sign change of the large-scale field is already inconsistent with global wind geometry.

The second difficulty is that even were there no dynamo, such as the case for β0 = 102, the natural symmetry from the simulation is unphysical for it to be connected to a global disk wind. In Figure 12, we show the vertical profiles of the magnetic field and velocity fields for run B2. Clearly, the outflows at the top and bottom sides of the disk are inclined toward opposite directions, which is inconsistent with a global disk wind. Even in the presence of the dynamo, there is a good chance that field lines at the top and bottom sides tend to incline toward different directions (as can be tracked in Figures 7 and 11).

Figure 11.

Figure 11. Same as Figure 7, but for run B4 only, and color represents the horizontally averaged radial magnetic field.

Standard image High-resolution image

Symmetry is also directly related to the vertical angular momentum transport by the outflow, as we discussed in Section 3.2. With the wrong symmetry, the wind stress (Equation 10) at the top and bottom sides of the disk have the same sign, and cancel each other, hence the net vertical angular momentum transport vanishes.

One might attribute this "wrong-symmetry" problem to the one extra degree of freedom in shearing-box simulations, which could possibly be avoided by performing global simulations. While it is certainly important to explore the same problem in global geometry, it has been suggested by Ogilvie (2012) that global effects can be effectively taken into account by applying some additional constraints on the vertical boundaries. Ogilvie constructed a 1D model for launching a disk wind/jet in the shearing-box approximation, where a much stronger net vertical field β0 ⩽ 1 was assumed so as to suppress the MRI. He assigned the two free parameters at the vertical boundaries to be the strength of the radial and azimuthal magnetic fields (he has two parameters because the Alfvén point turns out to be beyond his computational domain), and was able to successfully construct laminar wind solutions with the desired symmetry for the given constraints.

In view of this possibility, we further modify our vertical boundary conditions so that the mean radial (or azimuthal) magnetic field in the last four grid zones at the vertical boundary are set to some fixed value, which is done by first calculating the mean radial (or azimuthal) field in these zones, and then subtracting a constant radial (or azimuthal) field uniformly in all these cells to make the mean field equal to the desired value without violating the divergence-free condition. As we see in Figure 12, the mean radial field at the vertical boundaries is roughly the same as the mean vertical field in strength. Therefore, we demand that the field incline by 45° at the vertical boundaries in the $\hat{x} \hbox{--}\hat{z}$ plane, and require that the mean radial fields at the two vertical boundaries have opposite signs, which conforms to the desired symmetry. However, after performing various experiments on smaller domain test runs with parameters similar to runs B2, B3, and B4, we find that such a treatment does not have an evident physical effect on the profile of the mean flow and mean field, except for introducing a strong current sheet at the vertical boundaries. This fact has also been discussed in Lesur et al. (2013), and we conclude that the poloidal field inclination at the disk surface, particularly the symmetry of the mean horizontal fields, is insensitive to the imposed vertical boundary condition once the Alfvén point is contained within the simulation domain.

Figure 12.

Figure 12. Time-averaged vertical profiles of magnetic field (left) and velocity (right) for run B2 with β0 = 100. All in code units.

Standard image High-resolution image

In summary, we conclude that whether or not the MRI turbulence is accompanied by dynamo activities, the outflow from the shearing-box MRI simulations is unlikely to be directly connected to a global disk wind. The fate of the outflow remains to be explored using global simulations.

5. DISCUSSION AND CONCLUSIONS

5.1. Summary

In this paper, we have successfully performed local stratified shearing-box simulations of the MRI that include a strong vertical magnetic flux with the midplane plasma β of the net vertical field ranging from β0 = 104 to 102, a regime that has not been explored, though it is very likely relevant to accretion disks in many astrophysical systems. Such a magnetic configuration gives rise to very vigorous MRI turbulence, and simultaneously launches an outflow. We studied the properties of the MRI turbulence as well as the disk outflow in detail and our major findings are summarized below.

For the properties of the MRI turbulence, we find

  • 1.  
    With a relatively weak net vertical field of β0 ≳ 103, the disk consists of a gas pressure dominated disk midplane and a magnetic dominated disk corona (β ∼ 0.1–1). By contrast, the entire disk becomes magnetically dominated when β0 ≲ 103. The strong magnetic support substantially modifies the vertical structure of the disk and makes it substantially thicker.
  • 2.  
    Turbulent magnetic and kinetic energies increase monotonically with net vertical magnetic flux, and saturate when β0 ≲ 103. The MRI also generates a large scale toroidal magnetic field whose strength increases monotonically with net vertical field, and dominates the total magnetic energy over contribution from turbulence for β ≲ 103.
  • 3.  
    The Shakura–Sunyaev parameter α increases monotonically with net vertical magnetic flux, ranging from α ∼ 0.08 at β0 = 104 to α ≳ 1.0 for β = 102. Maxwell stress dominates Reynolds stress by a factor of 4–7. For a weak net vertical flux of β0 ≳ 103, turbulent fluctuations dominate the contribution to α, while for β0 ≲ 103, radial transport of angular momentum is dominated by the large scale fields.
  • 4.  
    The MRI dynamo that generates cyclic flips of the mean toroidal field persists in the presence of a weak net vertical magnetic flux, but becomes more sporadic with less periodicity as the net flux increases. The dynamo is completely suppressed when the net flux is strong, with β0 < 103, where the mean toroidal field exceeds the equipartition strength and never flips.

Our results demonstrate the crucial dependence of the behavior of the MRI turbulence on the net vertical magnetic flux. In particular, there is a critical net vertical magnetic flux of β0 ≈ 103 near which many aspects of the MRI turbulence change qualitatively. Additionally, more careful convergence study is still needed, and deeper understanding about the properties with MRI turbulence would further benefit from the study of Prantl number dependence (Fromang et al. 2007; Longaretti & Lesur 2010; Fromang et al. 2012).

For the properties of the disk outflow, we find

  • 1.  
    The slow magnetosonic point and the Alfvén point are always contained in our simulation box. The location of these critical points shifts towards the disk midplane as the net vertical magnetic flux increases. The launching and acceleration of the outflow are due to the magnetocentrifugal mechanism, where the surface magnetic field is sufficiently inclined.
  • 2.  
    There is a robust trend that the outflow mass loss rate $\dot{m}_w$ increases with increasing net vertical field, and tends to saturate at β0 ≲ 102. Although $\dot{m}_w$ can not be well determined from shearing-box simulations, its exact value in real systems is likely to be much smaller than that measured from local simulations and may depend on the aspect ratio H/R.
  • 3.  
    The outflow from the MRI simulations is unlikely to be directly connected to a global disk wind for geometric reasons. For β0 ≳ 103, the large scale radial and azimuthal field lines are constantly flipped due to the dynamo activities without a permanent bending direction. For β0 ≲ 103, the large scale magnetic fields do not change sign across the midplane, and hence the field lines at the two sides of the disk bend to opposite directions, inconsistent with a global wind geometry.
  • 4.  
    Angular momentum transport by the disk outflow is likely to play a minor role compared with that by the MRI turbulence for relatively weak net vertical magnetic flux. With stronger net vertical flux (β0 ≲ 103), the symmetry of the flow (as the dynamo is suppressed) makes net vertical angular momentum transport zero, but gives an oppositely directed inflow/outflow with substantial mass flux.

It has been reported that $\dot{m}_w$ from shearing-box MRI simulations decreases with increasing height of the simulation domain, which is related to the intrinsic limitations of the shearing-box framework. We propose that a careful study of the height dependence can help determining the true value of $\dot{m}_w$ in real systems. Our results leave an open question on the fate of the outflow, which can only be reliably explored using global simulations and will be discussed further in Section 5.3.

5.2. Implications for Global Disk Evolution

We note that shearing-box simulations with zero net vertical flux tend to give α ∼ 0.01–0.02 with confirmed numerical convergence (Miller & Stone 2000; Hirose et al. 2006; Shi et al. 2010; Davis et al. 2010; Simon et al. 2012). The vast dominance of simulations of this type has, to some extent, implicitly created a misconception that the α value resulting from the MRI turbulence if of the order of 0.01. On the observational side, the value of α can be estimated from the viscous timescale of transient systems such as dwarf novae, X-ray transient, and FU Ori bursts, where it was found to be of the order of 0.1 or larger (Smak 1999; Dubus et al. 2001; Zhu et al. 2007). Such apparent discrepancy leads King et al. (2007) to question the efficiency of the MRI in driving disk accretion and evolution. The dependence of α on the net vertical magnetic flux has already been extensively discussed in unstratified shearing-box simulations (Hawley et al. 1995; Sano et al. 2004; Pessah et al. 2007; Longaretti & Lesur 2010). Our simulations further confirm and quantify this trend using more realistic simulations with vertical stratification, and demonstrate that the above discrepancy can be naturally resolved if accretion disks are threaded by some large scale poloidal magnetic fields.

The strong dependence of α on β0 immediately indicates that the evolution of accretion disks strongly depends on the distribution of the poloidal magnetic flux through the disk. On the other hand, because of flux freezing in ideal MHD, the distribution of the magnetic flux is intimately connected to disk evolution (as well as turbulent diffusion and reconnection due to the MRI). Therefore, the mutual dependence of disk evolution and magnetic flux evolution makes the problem of disk accretion intrinsically non-local, and hence a full understanding of accretion disk structure and evolution requires global simulations.

Most global simulations to date adopt a magnetic field geometry that is analogous to zero net-flux shearing-box: the simulations are initialized either with single/multiple poloidal magnetic loops for a thick disk torus (Hawley 2000; De Villiers et al. 2003; Penna et al. 2010), or with pure toroidal magnetic field in a thin power-law disk (Fromang & Nelson 2006; Beckwith et al. 2011; Flock et al. 2011). These simulations generally show values of α that are slightly higher than 0.01 (∼0.025). In these simulations, the MRI generates a net vertical magnetic flux in the local patches of the disk connected by coronal loops and the local stress correlates with the net vertical flux (Tout & Pringle 1992; Sorathia et al. 2010). Combined with our study, the local net vertical flux in these global simulations is sufficient to account for the slightly higher total stress α than that in zero net flux shearing-box simulations. However, the distribution of the local net vertical magnetic flux falls off quickly with a stronger net flux, hence the total value of α can not be enhanced substantially.

Initial magnetic field geometry plays an important role in global simulations of the MRI, although the situation with large scale poloidal fields threading the disk have rarely been explored. Semi-analytical treatment reveals that the global distribution of large-scale magnetic flux in accretion disks depends largely on the ratio of turbulent viscosity and resistivity (Lubow et al. 1994a), as well as the vertical structure of the disk (Guilet & Ogilvie 2012). Together with these works, our simulations provide strong motivation for exploring the consequence of large scale poloidal field in global simulations. Such simulations will not only expand our knowledge the MRI, but will also address the fate of the outflow, which we discuss in the next section.

5.3. Disk Wind Launching in Global Simulations?

It has been recognized that a strong net vertical field that approaches equipartition strength at the disk midplane is necessary for launching a steady-state wind (Wardle & Koenigl 1993; Ferreira & Pelletier 1995; Ogilvie & Livio 2001). The requirement may be relaxed to lower magnetizations (Murphy et al. 2010), while for a magnetization that is too low the outflow is either suppressed or highly unsteady (Tzeferacos et al. 2009). Numerous global simulations have studied the launching of disk wind in the context of protostellar disks (e.g., Kato et al. 2002; Casse & Keppens 2002; Zanni et al. 2007)). However, as noted in Section 1, the use of artificial diffusion in these simulations does not properly reflect the disk microphysics. For a strong net vertical field that would suppress the MRI, the artificial diffusion has an unknown origin (and the non-ideal MHD effects can not be properly treated as artificial diffusion, Bai & Stone 2012). For a weaker net vertical field, the MRI is the presumed source of artificial diffusion. However, our evidence against direct wind launching all originates from the microphysical effects of the MRI (such as the MRI dynamo) that can not be captured in these simulations.

Another family of global simulations focus on non-radiative accretion disks around black holes (BH). These simulations either use psudo-Newtonian gravitational potential for the BH or adopt a general-relativistic prescription. Despite the non-Keplerian nature in the vicinity of the BHs, the physics of the MRI that is not too close to the innermost stable orbit is similar. Most of these simulations adopt similar poloidal loop field geometry and focus on jet launching, which deals with the innermost part of the accretion flow, a few of them have included large scale poloidal fields. In particular, Beckwith et al. (2009) initialized the simulations with pure vertical magnetic fields with β0 ∼ 102. However, upon saturation, the launching of a Blandford–Payne type wind was not observed, which was attributed to a global "coronal mechanism" that governs the magnetic flux evolution. Also, their simulations have a relatively small radial domain and a very thick disk. Significant outflow has been reported in 2D MHD simulations (Stone & Pringle 2001; Yuan et al. 2012), where MRI is a transient phenomenon. More recently, Tchekhovskoy et al. (2011), McKinney et al. (2012) and Narayan et al. (2012) studied the accretion flow with a large scale initial poloidal field in the inner disk with β0 ∼ 102, which leads to a "magnetically arrested" or "magnetically chocked" accretion flow due to the accumulation of magnetic flux in the vicinity of the BH. The MRI in the main disk was properly resolved in these simulations, however, they do not report any significant disk outflow in the inner region of the disks, and found that the jet power is directly correlated with BH spin.

Together, it appears that a Blandford–Payne type disk wind has not been unambiguously identified in global simulations with properly resolved MRI turbulence. Future studies should focus on the largely unexplored area of 3D global simulations for thin accretion disks in the presence of large-scale poloidal magnetic field. A large simulation domain in both radius and polar angle, fine resolution in the disk, and careful treatment of boundary conditions are required (T. Suzuki 2012, private communication), and such simulations are essential to ascertain the nature of the outflow, and to reveal the underlying physics governing the evolution of accretion disks.

We thank Sebastian Fromang and Geoffroy Lesur for sharing with us their submitted manuscripts, and the referee, Geoffroy Lesur for helpful comments that improved the clarity of this paper. We also acknowledge Adam Burrows, Sebastian Fromang, Shu-ichiro Inutsuka, Anders Johansen, Geoffroy Lesur, Ramesh Narayan, Gordon Ogilvie, Dimitrios Psaltis, Aleksander Sadowski, Takeru Suzuki, Taku Takeuchi, Alexander Tchekhovskoy, and Feng Yuan for useful discussions. X.N.B. acknowledges support for program number HST-HF-51301.01-A provided by NASA through a Hubble Fellowship grant from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Incorporated, under NASA contract NAS5-26555. J.M.S. acknowledges support from the National Science Foundation through grant AST-0908269. This work used computational facilities provided by PICSciE at Princeton University.

Footnotes

  • We note that Lesur et al. (2013) and Moll (2012) observed strong compression of the disk in their simulations where a stronger vertical field (β0 ⩽ 10) is used and reflection symmetry about the midplane is assumed. The compression is due to the large poloidal field curvature around the disk midplane. We do not find any disk compression in the case of β0 ⩾ 102, mainly because no reflection symmetry is enforced to build up the poloidal field curvature. In fact, our simulation results show opposite symmetry across the midplane and the poloidal field curvature near the midplane is about zero. For further discussions about symmetry, see Section 4.3.

  • With the assumption that the vertical integral of Trϕ is independent of radius. This corresponds to Equation (18) of Fromang et al. (2012) with p + q + 2 = 1/2.

  • We will discuss in Section 4.3 that the outflow is unlikely to be directly connected to an ordered disk wind, mainly due to geometric reasons. Let us set aside the issue with outflow geometry here for the purpose of discussion.

  • In principle, one should use the total velocity ${\boldsymbol{u}}$ rather than ${\boldsymbol{v}}$ in Equation (15). However, linear terms that contain background shear should vanish over space and time average. The quadratic term that contains background shear corresponds to the kinetic energy of the Keplerian profile, which is compensated exactly by the addition of mass to the simulation domain.

Please wait… references are loading.
10.1088/0004-637X/767/1/30