A publishing partnership

LOCALITY OF MHD TURBULENCE IN ISOTHERMAL DISKS

, , , and

Published 2009 March 20 © 2009. The American Astronomical Society. All rights reserved.
, , Citation Xiaoyue Guan et al 2009 ApJ 694 1010 DOI 10.1088/0004-637X/694/2/1010

0004-637X/694/2/1010

ABSTRACT

We numerically evolve turbulence driven by the magnetorotational instability (MRI) in a three-dimensional, unstratified shearing box and study its structure using two-point correlation functions. We confirm Fromang and Papaloizou's result that shearing box models with zero net magnetic flux are not converged; the dimensionless shear stress α is proportional to the grid scale. We find that the two-point correlation of $\mbox{\boldmath $B$}$ shows that it is composed of narrow filaments that are swept back by differential rotation into a trailing spiral. The correlation lengths along each of the correlation function principal axes decrease monotonically with the grid scale. For mean azimuthal field models, which we argue are more relevant to astrophysical disks than the zero net field models, we find that: α increases weakly with increasing resolution at fixed box size; α increases slightly as the box size is increased; α increases linearly with net field strength, confirming earlier results; the two-point correlation function of the magnetic field is resolved and converged, and is composed of narrow filaments swept back by the shear; the major axis of the two-point increases slightly as the box size is increased; these results are code independent, based on a comparison of ATHENA and ZEUS runs. The velocity, density, and magnetic fields decorrelate over scales larger than ∼H, as do the dynamical terms in the magnetic energy evolution equations. We conclude that MHD turbulence in disks is localized, subject to the limitations imposed by the absence of vertical stratification, the use of an isothermal equation of state, finite box size, finite run time, and finite resolution.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Astrophysical disks appear to redistribute angular momentum rapidly, much more rapidly than one would expect based on estimates of the molecular viscosity. Classical thin accretion disk theories (Shakura & Sunyaev 1973; Lynden-Bell & Pringle 1974) solved this problem by appealing to turbulence, and modeled the effects of this turbulence as an "anomalous viscosity." The idea that turbulence plays a key role was placed on firmer foundations with the (re)discovery of the magnetorotational instability (MRI; Balbus & Hawley 1991) and subsequent numerical investigations (see Balbus & Hawley 1998 for a review). Winds or gravitational instability may drive disk evolution in certain cases, but MRI-initiated MHD turbulence appears capable of driving disk evolution in a wide variety of astrophysical disks.

We still do not know, however, whether the effects of MHD turbulence on disks are localized. It is possible that structures develop that are large compared to a scale height Hcs/Ω, and that these structures are associated with nonlocal energy and angular momentum transport. If so, disk evolution would not be well described by a theory, such as the α model, in which the shear stress depends only on the local surface density and temperature.

A related possibility, which we will not examine here, is that the time-averaged turbulent stresses $\overline{W}_{r\phi }$ might satisfy $\partial \overline{W}_{r\phi }/\partial \Sigma < 0$ (Σ≡ surface density; see Piran 1978 for a discussion). That is, the disk might be "viscously" unstable. This could cause the disk to break up into rings or even—to use a term of art—"blobs." Such an outcome would be awkward for the classic phenomenological steady disk and disk evolution models, which have had some success in modeling cataclysmic variable disks and black hole X-ray binary disks in a high, soft state (e.g., Belloni et al. 1997; Lasota 2001).

How can one probe the locality of MHD turbulence in disks? We will use the two-point correlation function of the magnetic field, velocity field, and density as determined by numerical experiments. Nonlocal transport would likely be associated with features in the two-point correlation function, as would viscous instabilities. For example, turbulence might excite waves (wakes) that carry energy and angular momentum over many H in radius before damping. These wakes would appear as extended features in the two-point correlation function.

The two-point correlation function and the power spectrum contain the same information since they are related by a Fourier transform. But they do not convey the same impression and they have different noise properties. For a one-dimensional function sampled at N points over an interval L half the sample points in the power spectrum lie between the Nyquist frequency (πN/L) and half the Nyquist frequency, while for the correlation function half the sample points lie between a separation L/4 and L/2. The two-point correlation function will therefore convey a more accurate impression of large-scale features than power spectra.

In this paper, we study models with both zero net field and net azimuthal field. We ignore mean vertical field models because we remain persuaded by the phenomenological argument of van Ballegooijen (1989) that vertical field diffuses easily out of the disk when the turbulent magnetic Prandtl number is O(1) (although there are ways of evading this argument; Spruit & Uzdensky 2005). Net azimuthal field models are, we think, most relevant to astrophysical disks. In disk galaxies—the only differentially rotating disks where we can resolve field structure—the azimuthal field dominates when averaged over areas more than a few H2 in extent (e.g., Beck 2007). Azimuthal field also dominates in global disk simulations (e.g., Hirose et al. 2004; McKinney & Narayan 2007; Beckwith et al. 2008), and in local disk simulations. In local simulations in which the mean field is allowed to evolve (e.g., Brandenburg et al. 1995; Miller & Stone 2000) an azimuthal mean field develops. Taken together these simulations and observations strongly suggest that the azimuthal field averaged over regions ∼H2 in area will never be exactly zero.

This paper is organized as follows. In Section 2, we give a simple description of the local model and summarize our numerical algorithm with orbital advection. We then study zero net flux models (as in Fromang & Papaloizou 2007; hereafter FP07); this serves as a code test and introduces the correlation function analysis. In Section 3, we explore the properties of turbulence in models with a mean azimuthal field. We report on the saturation level and correlation lengths and we discuss their dependence on the model parameters, such as resolution, box size, and initial field strength. Section 4 contains a summary and guide to our results.

2. MODEL, METHODS, AND TESTS

Our starting point is the local model for disks. It is obtained by expanding the equations of motion around a circular-orbiting coordinate origin at cylindrical coordinates (r, ϕ, z) = (ro, Ωot + ϕo, 0), assuming that the peculiar velocities are comparable to the sound speed and that the sound speed is small compared to the orbital speed. The local Cartesian coordinates are obtained from cylindrical coordinates via (x, y, z) = (rro, ro[ϕ − Ωot − ϕo], z). We assume throughout that the disk is isothermal (p = c2sρ, where cs is constant), and that the disk orbits in a Keplerian (1/r) potential.

In the local model, the momentum equation of ideal MHD becomes

Equation (1)

The final two terms in Equation (1) represent the Coriolis and tidal forces in the local frame. Note that our model is unstratified, which means that the vertical gravitational acceleration −Ω2z usually present in Keplerian disks is ignored.

Our model contains no explicit dissipation coefficients. Recent models with explicit scalar dissipation (FP07; Lesur & Longaretti 2007) have shown that the outcome (saturated field strength) depends on the viscosity ν and the resistivity η, and that ZEUS has an effective magnetic Prandtl number PrM ≡ ν/η ∼ 4.

The orbital velocity in the local model is

Equation (2)

This velocity, along with a constant density and zero magnetic field, is a steady-state solution to Equation (1). If the computational domain extends to |x|>(2/3)H = (2/3)cs/Ω, then the orbital speed is supersonic with respect to the grid.

The local model is studied numerically using the "shearing box" boundary conditions (e.g., Hawley et al. 1995). These boundary conditions isolate a rectangular region in the disk. The azimuthal (y) boundary conditions are periodic; the radial (x) boundary conditions are "nearly periodic," i.e., they connect the radial boundaries in a time-dependent way that enforces the mean shear flow. We use periodic boundary conditions in the vertical direction; this is the simplest possible version of the shearing box model.

2.1. Numerical Methods

Most of our models are evolved using ZEUS (Stone & Norman 1992). ZEUS is an operator-split, finite difference scheme on a staggered mesh. It uses artificial viscosity (not an anomalous viscosity!) to capture shocks. For the magnetic field evolution ZEUS uses the Method of Characteristics-Constrained Transport (MOC-CT) scheme, which is designed to accurately evolve Alfvén waves (MOC) and also to preserve the $\mbox{\boldmath $\nabla $}\cdot \mbox{\boldmath $B$}= 0$ constraint to machine precision (CT).

We have modified ZEUS to include "orbital advection" (Masset 2000; Gammie 2001; Johnson & Gammie 2005) with a magnetic field (Johnson et al. 2008). Advection by the orbital component of the velocity $\mbox{\boldmath $v$}_{{\rm orb}}$ (which may be supersonic with respect to the grid) is done using interpolation. With this modification the time step condition $\Delta t < {\it C} \Delta x/(|\delta \mbox{\boldmath $v$}| + c_{\mbox{max}})$ (cmax≡ maximum wave speed and C≡ Courant number) depends only on the perturbed velocity $\delta \mbox{\boldmath $v$}= \mbox{\boldmath $v$}- \mbox{\boldmath $v$}_{{\rm orb}}$ rather than $\mbox{\boldmath $v$}$. So when $|\mbox{\boldmath $v$}_{\mbox{orb}}| \gtrsim c_{\mbox{max}}$ (for shearing box models with v2A/c2s ≲ 1, when LH) the time step can be larger with orbital advection, and computational efficiency is improved.

Orbital advection also improves accuracy. ZEUS, like most Eulerian schemes, has a truncation error that increases as the speed of the fluid increases in the grid frame. In the shearing box without orbital advection the truncation error would then increase monotonically with |x|. Orbital advection reduces the amplitude of the truncation error and also makes it more nearly uniform in |x| (Johnson et al. 2008).

Do our results depend on the algorithm used to integrate the MHD equations? To find out, we have also evolved a subset of models using ATHENA, a second-order accurate Godunov scheme that solves the equations of ideal MHD in conservative form. The algorithm couples the dimensionally unsplit corner transport upwind (CTU) method of Colella (1990) with the third-order in space piecewise parabolic method (PPM) of Colella & Woodward (1984), and a constrained transport (CT) algorithm for preserving the $\mbox{\boldmath $\nabla $}\cdot \mbox{\boldmath $B$}= 0$ constraint. Details of the algorithm and test problems are described in Stone et al. (2008). The specific application of ATHENA to the shearing box (as used in this work) is described in Simon et al. (2009).5

2.2. Models with Zero Net Flux

We now consider a set of zero net field shearing box models to introduce our correlation function analysis. We use the same model parameters as FP07 so that these models also serve, by comparison with FP07, as a nonlinear code test.

The models in this section have size (Lx, Ly, Lz) = (1, π, 1)H. The initial magnetic field is Bz = Bz0 × sin(2πx/H), where Bz0 satisfies β ≡ 8πP0/B2z0 = 400. Noise is introduced in the initial velocity field to stimulate the growth of the unstable modes. The models are evolved to tf = 600 Ω−1.

We consider four resolutions: (Nx, Ny, Nz) = N(32, 50, 32), where N = 1, 2, 4, 8. The last three models correspond to runs std32, std64, and std128 in FP07, respectively. The evolution of 〈EB〉 ≡ 〈B2/(8πρoc2s)〉 is shown in Figure 1 (〈〉≡ volume average). The saturation 〈EB〉 decreases as the resolution increases.

Figure 1.

Figure 1. Evolution of magnetic energy in zero net field runs. From top to bottom radial resolution increases from 32/H, 64/H, 128/H to 256/H. The saturation level decreases in proportion to the grid scale.

Standard image High-resolution image

The dimensionless shear stress

Equation (3)

We measured the time average of α, denoted as $\overline{\alpha }$, (from tΩ = 250 to tΩ = 600) for each run, and recorded the results in Table 1.6 These averages are nearly identical to those obtained by FP07. This consistency enhances confidence in both sets of results.

Table 1. Shearing Box Runs with a Zero Net Vertical Field

Model Resolution $\overline{\alpha }$ λB,min λB,maj λB,z θB,tilt λv,min λv,maj λv,z θv,tilt λρ,min λρ,maj λρ,z θρ,tilt
z32 32 × 50 × 32 3.8 × 10−3 0.090 0.62 0.080 11 0.078 0.57 0.13 5.5 0.076 0.82 0.39 8.6
z64 64 × 100 × 64 4.2 × 10−3 0.059 0.38 0.050 13 0.074 0.45 0.18 7.1 0.056 0.60 0.33 6.6
z128 128 × 200 × 128 2.1 × 10−3 0.037 0.22 0.032 14 0.053 0.32 0.11 6.7 0.043 0.62 0.33 6.4
z256 256 × 400 × 256 1.1 × 10−3 0.024 0.17 0.024 14 0.035 0.24 0.10 5.9 0.019 0.34 0.18 5.8

Download table as:  ASCIITypeset image

For Nx ⩾ 64

Equation (4)

is a good fit to the numerical results. The magnetic field energy density, α, and the kinetic energy density are almost inversely proportional to Nx, and so, like FP07, we conclude that the zero net field models do not converge.

All shearing box models considered in this paper have $\overline{\alpha } \simeq \overline{\langle E_B\rangle }/2 = 1/(2\overline{\beta })$. This implies a characteristic orientation of the field, since (neglecting the Reynolds stress ρvxvy ∼ 0.25 × [ − BxBy]/[4π]) $\overline{\alpha } \approx - \overline{\langle B_x B_y\rangle }/(4\pi \langle \rho \rangle c_s^2) \equiv \overline{\langle B^2 \cos \theta _B \sin \theta _B \rangle }/(4\pi \langle \rho \rangle c_s^2)$, where θB is the angle between the field and the y-axis. Then $\overline{\alpha } = \overline{\langle E_B\rangle } 2 \cos \overline{\theta }_B \sin \overline{\theta }_B = \overline{\langle E_B\rangle }/2$. So $\overline{\theta }_B = \pi /12$ (15°) is the characteristic angle between the magnetic field and the y-axis.

Next, we turn to the structure of the zero net field turbulence. Consider the two-point correlation function for the density fluctuations

Equation (5)

(δρ ≡ ρ − 〈ρ〉), for the trace of the velocity fluctuation correlation tensor

Equation (6)

vivivi,orb − 〈vi〉), and there is an implied summation over i and for the trace of the magnetic field correlation tensor

Equation (7)

BiBi − 〈Bi〉). All correlation functions are calculated for fluctuating dynamical variables with zero mean. Figure 2 shows slices of correlation functions through ξ at z = 0 for run z128. The cores of the correlation functions are ellipsoidal, with three principal axes, and concentrated at $|\Delta \mbox{\boldmath $x$}| < H$. The correlations are localized.

Figure 2.

Figure 2. Two-point correlation function for density, velocity field, and magnetic field in Δx–Δy plane in run z128. The contours are set linearly from 0 to 0.009 for 20 levels; the heavy line is the 0 contour.

Standard image High-resolution image

We measure four features of the correlation functions: the angle θtilt between the correlation function major axis and the y-axis, and the correlation lengths along the major, minor, and z-axes (λmaj, λmin, λz), where the correlation length λi is defined 7 by

Equation (8)

l is the distance from $\Delta \mbox{\boldmath $x$}= 0$ along the principal axis defined by the unit vector $\hat{\mbox{\boldmath $x$}}_i$, and $\hat{\mbox{\boldmath $x$}}_{\rm maj} = \hat{\mbox{\boldmath $x$}} \sin \theta _{\rm tilt} - \hat{\mbox{\boldmath $y$}} \cos \theta _{\rm tilt}$, $\hat{\mbox{\boldmath $x$}}_{\rm min} = \hat{\mbox{\boldmath $x$}} \cos \theta _{\rm tilt} + \hat{\mbox{\boldmath $y$}} \sin \theta _{\rm tilt}$, and $\hat{\mbox{\boldmath $x$}}_z = \hat{\mbox{\boldmath $z$}}$.8,9 Figure 3 shows ξB along each of the principal axes in run z128. The dotted lines show ξ(0)exp(−li). The correlated regions in the magnetic field are narrow (λmin ≃ λz ≃ λmaj/6) filaments with a trailing spiral orientation.

Figure 3.

Figure 3. Magnetic field correlation function along the minor, major, and vertical principle axes in run z128. Solid lines: cut through the data; dotted lines: a simple model with exp(−λi), where λi is the measured correlation length along each principle axis. The correlation functions along the minor and z-axes are almost identical.

Standard image High-resolution image

What do the correlation functions mean? For the magnetic field, there is a characteristic orientation of the field $\overline{\theta }_B$ obtained through our measurement of the shear stress. The major axis of the correlation function is very nearly parallel to this, $\theta _{\rm tilt} \simeq \overline{\theta }_B$. It is reasonable to view the magnetic field correlation function, then, as tracing out a characteristic, filamentary structure in the magnetic field.

It is worth recalling briefly what we might expect for ξv in isotropic, homogeneous turbulence with an outer scale L0 = 2π/k0 and velocity dispersion σv. In the inertial range v2kk−11/3, so a reasonable functional form for the power spectrum is

Equation (9)

where N is a nondimensional normalization constant. The corresponding autocorrelation function is

Equation (10)

where K is the modified Bessel function and r ≡ |Δx|. For k0r ≪ 1,

Equation (11)

This is the usual r2/3 Kolmogorov dependence at small separation. For k0r ≫ 1,

Equation (12)

which yields the expected decorrelation for k0r ≫ 1. The correlation length defined in Equation (8) is 0.838/k0. Note that the power spectrum does not have zero power as k → 0; rather it asymptotes to Nσ2vk−30.

For MHD turbulence in disks, however, the correlation function is not isotropic, and its structure is not anticipated by any predictive theory. In the absence of such a theory it may be useful to have a convenient analytical representation of the numerical results. This can be obtained by stretching Equation (10) along each of the principal axes, that is, by replacing k0r by $u \equiv ((\Delta \mbox{\boldmath $x$}\cdot \hat{\mbox{\boldmath $x$}}_{\rm min}/\lambda _{\rm min})^2 + (\Delta \mbox{\boldmath $x$}\cdot \hat{\mbox{\boldmath $x$}}_{\rm maj}/\lambda _{\rm maj})^2 + (\Delta \mbox{\boldmath $x$}\cdot \hat{\mbox{\boldmath $x$}}_{\rm z}/\lambda _{\rm z})^2)^{1/2}$.

Do the correlation lengths converge? We find that the correlation lengths are resolution dependent, with

Equation (13)

λz and λmin are at most six zones. The scaling of λmin with zone size is clearly not linear, but the 2/3 power-law scaling is just a fit to the data and should not be taken too seriously. The nonlinear scaling does hint at the possibility that, as resolution is increased and λmin and λz are better resolved, there could be a transition in the outcome.

The major axis for ξB lies ∼15° from the y-axis. For the density and peculiar velocity field the tilt angle is ∼7°. The latter tilt is consistent with the measured Reynolds stress: 〈ρvxvy〉/〈ρv2〉 ∼ 1/8, so the average perturbed velocity is tilted at ∼7° to the y-axis.

Finally, note that there are low-amplitude features in ξv and ξρ at scales of a few correlation lengths (see particularly in Figure 4(a)). These features may be due to the excitation of rotationally modified sound waves by MHD turbulence. To test this hypothesis note that, for tightly wrapped (kykx), linear sound waves δρ2k20 ≃ δv2k/c2s. A field composed of these waves would then have ξρ20 = ξv/c2s. Taking the differential correlation function ξv/c2s − ξρ20 should therefore remove those pieces of ξv that are due to sound waves. Figure 4(b) shows ξv/c2s − ξρ20 at Δy = 0. Evidently much of the large-scale power is removed. Figure 5 shows another slice through ξv/c2s − ξρ20 at Δz = 0. Again the large-scale power is removed. This is consistent with the hypothesis that the largest scale features in the correlation functions are acoustic waves.

Figure 4.

Figure 4. Turbulent velocity correlation function and a differential correlation ξv/c2s − ξρ20 in the Δy = 0 plane for run z128. In ξv, apart from the compact core at small separations, there is a weak correlation at large scales that is likely due to the sound waves. The contours run linearly from 0 to 0.005 for 20 levels. The heavy line is the 0 contour.

Standard image High-resolution image
Figure 5.

Figure 5. Differential correlation function ξv/c2s − ξρ20 in the Δz = 0 plane in run z128. After removing the contribution due to the sound waves, the correlation ellipsoid is more compact and almost identical to that of the magnetic field. The contour levels are set the same as in Figure 2, linearly from 0 to 0.009.

Standard image High-resolution image

3. MODELS WITH A NET AZIMUTHAL FIELD

In this section, we study models with 〈By〉 ≠ 0. These models correspond more closely to what is observed in shearing box models with boundary conditions that allow the net field to evolve, what is seen in global MHD models of disks, and what is observed in galactic disks than the $\langle \mbox{\boldmath $B$}\rangle = 0$ models.

3.1. Convergence

To test convergence we use the same size models as the zero net field runs, (Lx, Ly, Lz) = (1, π, 1)H. 〈By〉 is set so that β = 400; in the initial conditions all other magnetic field components vanish. The models are evolved to t = 250 Ω−1. We use five different resolutions: (Nx, Ny, Nz) = N(32, 50, 32), where N = 1, 2, 4, 6, 8. In each run we average over the second half of the evolution to measure $\overline{\langle E_B\rangle }$ and $\overline{\alpha }$. We also measure the correlation lengths from ξρ, ξv, and ξB using an average of the correlation function calculated from eight data dumps in the second half of the run. Parameters and results for these runs are listed in Table 2.

Table 2. Shearing Box Runs with a Net Azimuthal Field

Model Algorithm Resolution $\overline{\alpha }$ $\overline{\langle E_B \rangle }/\rho _{0}c_s^2$ λB,min λB,maj λB,z θB,tilt λB,minx
y32a ZEUS 32 × 50 × 32   0.0094 0.019 0.10    0.40 0.084 12 3.2
y64a ZEUS 64 × 100 × 64 0.014 0.024 0.066 0.36 0.064 15 4.2
y128a ZEUS 128 × 200 × 128 0.015 0.028 0.053 0.28 0.049 15 6.8
y192a ZEUS 192 × 300 × 192 0.020 0.037 0.055 0.30 0.049 15 11
y256a ZEUS 256 × 400 × 256 0.021 0.041 0.049 0.27 0.045 15 13
y32b ATHENA 32 × 50 × 32 0.018 0.032 0.11    0.63 0.09 15 3.3
y64b ATHENA 64 × 100 × 64 0.015 0.027 0.070 0.38 0.060 16 4.5
y128b ATHENA 128 × 200 × 128 0.018 0.035 0.058 0.33 0.051 16 7.4
y192b ATHENA 192 × 300 × 192 0.025 0.050 0.052 0.30 0.047 17 10
y256b ATHENA 256 × 400 × 256 0.027 0.055 0.053 0.32 0.049 16 14

Download table as:  ASCIITypeset image

Figure 6 shows 〈EB〉(t) for various resolutions. In our two highest resolution runs, $\overline{\langle E_B\rangle } = 3.7\times 10^{-2} \rho _{0} c_s^2$ for run y192a and 4.1 × 10−2ρ0c2s for run y256a, respectively. A consistent fit is $\overline{\langle E_B\rangle } \simeq 0.03 (N_x/128)^{1/3}$ for 32 < Nx < 256. The saturation energy increases with resolution. It is unlikely that this trend continues indefinitely. As we will see below, the magnetic field correlation lengths are unresolved at Nx = 32 but resolved at Nx = 256. This suggests that the increase in $\overline{\langle E_B\rangle }$ is caused by resolution of magnetic structures near the correlation length. If there is little energy in structures much smaller than the correlation length, then $\overline{\langle E_B\rangle }$ should saturate at somewhat higher resolution.

Figure 6.

Figure 6. Evolution of the magnetic energy in the net azimuthal field run. From bottom to top the lines are: heavy dot–long dash: 32/H; gray short dash: 64/H; heavy dot–short dash: 128/H; gray dot: 192/H; heavy solid: 256/H. The saturation energy increases with resolution.

Standard image High-resolution image

To check the algorithm dependence of the results we ran an identical set of models using ATHENA. The results are listed in Table 2. Both sets of models show a weak upward trend in $\overline{\langle E_B\rangle }$ with resolution. If one corrects for the approximately 2 times higher effective resolution of ATHENA then the ATHENA and ZEUS results are quantitatively consistent with each other.

The correlation lengths for the zero net field runs do not converge. What about the net azimuthal field models? The magnetic field correlation lengths are listed in Table 2 (other correlation lengths are omitted for brevity, but they behave similarly). For N ⩾ 4 both ZEUS and ATHENA find

Equation (14)

For the N ⩾ 4 models λmin ∼ λz > 8Δx. In the highest resolution (N = 8) ATHENA model, λminx ≃ 14. The correlation lengths are both converged and resolved.

If MHD turbulence in disks has a forward energy cascade, as do three-dimensional hydrodynamic turbulence and the Goldreich–Sridhar model for strong MHD turbulence in a homogeneous medium, then it is natural to identify the correlation lengths with the outer, or energy injection, scale. Since λmin ≃ 15 grid zones at our highest resolution there is no resolved inertial range. We anticipate that future, higher-resolution numerical experiments with a mean azimuthal field will show the development of an inertial range.

3.2. Magnetic Energy Evolution

At what scale is magnetic energy generated in MRI-driven turbulence? To investigate this, we have studied the correlation function for each term driving the evolution of the volume-averaged magnetic energy:

Equation (15)

where D is the volume-averaged numerical dissipation rate. The terms on the right-hand side can be interpreted as describing the effects of advection, compression and expansion, field-line stretching, and numerical dissipation. On average the first term is small (it should vanish exactly for shearing box boundary conditions, but roundoff and truncation error make it nonzero), and the third term dominates the second by a factor of 20 in run y128b. In a time- and volume-averaged sense the right-hand side must be zero, so numerical dissipation must approximately balance energy injection by field-line stretching.

Previous studies (FP07; Simon et al. 2009) analyzed a version of this equation in the Fourier domain to study turbulent energy flow as a function of length scale in the shearing box simulations. Both studies found that the magnetic energy is generated on all scales by the background shear. Here, we perform a complementary analysis in the spatial domain.10

We have computed the autocorrelation function for the field-line stretching term from run y128b (without subtracting the mean). Figure 7 shows the autocorrelation function at Δz = 0. From this figure we conclude that: (1) the scale and shape of the correlation function is similar to that of the fundamental variables ($\mbox{\boldmath $B$}$, $\mbox{\boldmath $v$}$; recall that in y128b (λB,min, λB,maj, λB,z) = (0.058, 0.33, 0.051)H); (2) energy is injected at scales comparable to the correlation length of the fundamental variables; (3) a superposition of magnetic structures similar to ξB that are distributed with uniform probability in space would have a power spectrum that is flat (white noise) at low k. It is plausible that terms in the Fourier-transformed magnetic energy equation would also be flat at low k. This would not imply that energy is injected by dynamically meaningful structures at large scales; it would simply be the consequence of an uncorrelated superposition of small, localized features in the turbulence.

Figure 7.

Figure 7. Correlation function in the Δz = 0 plane for the "field-line stretching" term in the magnetic energy equation. The data are from the y128b ATHENA run and the contour levels run logarithmically from 10−5.1 to 10−3.6. The generation and dissipation of magnetic energy occurs in a local manner, consistent with the localization of the dynamical variables.

Standard image High-resolution image

3.3. Box Size

Does the saturation energy or correlation length depend on the size of the computational domain (box size)? To investigate, we fix the physical resolution at 64 zones/H and vary the model size: (Lx, Ly, Lz) = (1, π, 1)H, (2, π, 1)H, (1, 2π, 1)H, (2, 2π, 1)H, and (4, 4π, 1)H. The model parameters and outcomes are listed in Table 3.

Table 3. Shearing Box Runs with a Net Azimuthal Field: Effect of the Box Size

Model Size $\overline{\langle E_B \rangle }/\rho _{0}c_s^2$ ${\lambda _{B,{\rm maj}}} \over {H}$
y64 (1, π, 1)H 0.024 0.36
y64.x2 (2, π, 1)H 0.028 0.49
y64.y2 (1, 2π, 1)H 0.035 0.45
y64.x2y2 (2, 2π, 1)H 0.038 0.49
y64.x4y4 (4, 4π, 1)H 0.038 0.57

Download table as:  ASCIITypeset image

Evidently, there is a weak dependence of $\overline{\langle E_B\rangle }$ on box size; it increases from 0.024ρ0c2s for the smallest run to 0.038ρ0c2s for the largest run. The magnetic field correlation lengths also increase with box size, with λmaj = 0.36H for the smallest box (so Ly/(2λmaj) = 4.4) to λmaj = 0.57H for the largest box (so Ly/(2λmaj) = 11). This upward trend in correlation length is probably real, but it is sufficiently small that it is difficult to separate from noise in the correlation length measurements.

The correlation functions ξρ and ξv, unlike ξB, have low amplitude tails extending out to the box size. This can be seen in Figure 8, which shows ξρ and ξB for run y64.x4y4. The tails are likely due to sound waves, and their absence in the differential correlation function ξv/c2s − ξρ20, shown in the middle panel of Figure 8, is consistent with this.

Figure 8.

Figure 8. Density correlation function, a differential correlation ξv/c2s − ξρ20, and the magnetic correlation function in the Δz = 0 plane for run y64.x4y4. The contours are set linearly from −0.007 to 0.08 for 20 levels. The heavy line is the 0 contour.

Standard image High-resolution image

3.4. Field Strength

We now compare two models that differ only in their initial field strength. Both have a size Lx, Ly, Lz = (1, 2π, 1)H with resolution Nx, Ny, Nz = 64, 200, 64. One model has the same initial field strength as our other models, β0 = 400, while the other one starts with a stronger field, β0 = 100.

We found that the saturated magnetic energy for the β0 = 400 run is $\overline{\langle E_{B}\rangle } = 0.035\rho _{0}c_s^{2}$. For the β0 = 100 run $\overline{\langle E_{B}\rangle } = 0.079\rho _{0}c_s^{2}$, slightly more than twice the saturation level of the higher β0 run. Resolution may be playing a role here: the β0 = 100 run has twice the resolution per most unstable MRI wavelength, and we know that the saturation level depends on resolution when the field strength is constant.

Our results are consistent with the linear relation between $\overline{\langle E_B\rangle }$ and initial field strength for 〈By〉 ≠ 0 models reported in Hawley et al. (1995) (hereafter HGB95; although it may also be consistent with a wide range of exponents for this relation). Our results are inconsistent with HGB95's claim that $\overline{\langle E_B\rangle } \propto L_y$, at least if the box size is ≫λmaj (the good agreement found for HGB95's predictor may be a coincidence).

At the level we can determine from two data points, our results are consistent with $\overline{\langle E_B\rangle } \propto \rho _0 c_s V_{A,y0}$, where VA,y0 is the initial azimuthal Alfvén speed (i.e., here scale height cs/Ω replaces Ly in HGB95; there are no other length scales in the problem). This is interesting: it implies that α depends on the gas pressure, a result first reported by Sano et al. (2004) and thus compressibility plays a role in the saturation of the MRI!

Our results suggest that $\overline{\langle E_B\rangle }$ should scale differently in compressible and incompressible models. In the incompressible models the only length scale available is the size of the box, so in incompressible models we must have $\overline{\langle E_B\rangle } \sim \rho _0 (L\Omega)^a V_{A,y0}^b$, where L is some combination of Lx, Ly, and Lz, and the exponents a and b are not determined.

The correlation lengths for the magnetic field in the two runs are (λmin, λmaj, λz) ≃ (0.08, 0.45, 0.08)H for the β0 = 400 run and (λmin, λmaj, λz) ≃ (0.11, 0.58, 0.10)H for the β0 = 100 run. To sum up, the correlation length increases weakly as the initial field strength and the box size increases. It is not consistent with the scaling ∼By,0/(ρ0Ω) one would expect if the correlation length scaled with the most unstable wavelength of the background field, and it is not consistent with the scaling ∼〈B2y1/2/(ρ0Ω) ∼ B1/2y,0 one would expect if the correlation length is related to a characteristic MRI length scale for the (larger) fluctuating field.

4. SUMMARY

We have investigated the locality of MHD turbulence in an unstratified, Keplerian shearing box model using the two-point autocorrelation function.

We first considered models with zero net vertical field and the same parameters as FP07. Our slightly different orbital advection algorithm reproduces earlier results on the relation between the saturation level and resolution: zero net field models do not converge.

Consistent with this, we also find that as resolution increases the correlation lengths for the velocity, density, and magnetic field decrease. A fit to the results yields the following scaling for the magnetic field correlation lengths,

Equation (16)

i.e., the correlation length decreases as the resolution increases.

We then studied a set of models with net toroidal field and initial β = 400. These models are not completely converged in the sense that they show a trend of increasing $\overline{\alpha }$ with resolution. They are converged in that the correlation lengths are well resolved and constant near the highest resolution:

Equation (17)

But because λz and λmin are only just resolved (they are each at most 14 grid cells), we do not see an inertial range. We expect that future higher-resolution models will show the development of an inertial range.

We further examined the correlation function for the dominant (nonnumerical) field-line stretching term in the magnetic energy evolution equation. The correlation lengths are small compared to H, consistent with the correlation lengths of the dynamical variables. Evidently, energy is injected at scales comparable to or smaller than the correlation length.

We also explored the influence of the box size on the outcome in the net toroidal field models. We found a weak dependence on the box size but only for LxH. This suggests that in shearing box simulations the size of the box should be chosen to be at least a few scale heights so that the correlation lengths are not "squeezed" by the boundary conditions. We also varied the initial field strength, and consistent with earlier reports found that the saturation level ($\overline{\alpha }$ or $\overline{\langle E_B\rangle }$) scales linearly with the initial field strength. Correlation lengths also increase as the field strength increases, but not linearly.

So is disk turbulence really localized? Our answer is mixed. On the one hand, our net toroidal field models have almost all the correlation amplitudes contained within a region a few scale heights on a side and in this sense the turbulence is indeed local. On the other hand, we do see signs of radiation of compressive waves by the turbulence in the two-point correlation function. These signs are most impressively visible in Figure 4(a), which shows vertically extended tails on the density autocorrelation function in the Δy = 0 plane, and in Figure 8, which shows azimuthally extended tails on the density autocorrelation function in the Δz = 0 plane. These tails are matched by similar tails on the velocity autocorrelation function, consistent with our hypothesis that they are due to the excitation of rotationally modified sound waves.

The influence of compressive waves excited by MHD turbulence on disk evolution cannot, in the end, be assessed with the experiments and analysis in this paper. The key measurement needed is the radial damping length of compressive waves due to absorption and scattering of waves by turbulent eddies. This would be most easily measured in a separate experiment that studies the response of MHD turbulence to an imposed sound wave. Even this measurement would be incomplete because it neglects additional damping related to stratification (Lin et al. 1990; Lubow & Ogilvie 1998), but it would provide an upper limit on the damping length.

We are studying the locality of MHD turbulence in a highly idealized situation in which stratification and other aspects of the larger disk—such as the process that generates the imposed azimuthal magnetic field, perhaps a global dynamo—are absent. Our unstratified model is insensitive to some effects that could lead to the development of global (λ ∼ R) or mesoscale (R ≫ λ ≫ H) structures. Convection and rotation might reasonably be expected to lead to dynamo activity manifesting itself as large-scale structures in the magnetic field. Disk atmospheres might also develop large-scale, coronal structures that delocalize disk evolution by transmitting angular momentum and energy (Uzdensky & Goodman 2008).

Finite integration time and limited accuracy is also a concern. It is possible that large-scale structures emerge only over hundreds of rotation periods. If the disk is subject to a "viscous instability" of the sort mentioned in the introduction then the timescale for the growth of a feature on scale λ would be (αΩ)−1(λ/H)2; this timescale could be hundreds of rotation periods for the modest αs seen in our models if the instability is damped for λ< few ×H. Accurate, long duration, and expensive integrations will be required in future searches for viscous instability.

This work was supported by the National Science Foundation under grants AST 00-93091, PHY 02-05155, and AST 07-09246, and a Sony Faculty Fellowship, a University Scholar appointment, and a Richard and Margaret Romano Professorial Scholarship to C.F.G. Portions of this work were completed while C.F.G. was a member at the Institute for Advanced Study (2006–2007). The authors are grateful to Shane Davis, Peter Goldreich, Stu Shapiro, Fred Lamb, Friedrich Meyer, and John Hawley for discussions.

Footnotes

  • Simon et al. (2009) use P = (γ − 1) u in contrast to our P = c2sρ.

  • The combination of finite run time and fluctuations in α(t) introduces noise into $\overline{\alpha }$. To estimate the noise amplitude we divided the averaging interval into two and compared the two averages. In all the runs with Nx ⩾ 64 case they differed from the mean by ⩽10%.

  • Other definitions of λi are possible, e.g., the half-width at half-maximum (HWHM) of ξ can be used, with an exponential model for ξ, to find λi. For example, for ξB in run z128, the HWHM definition gives (λmaj, λmin, λz) = (0.031, 0.15, 0.023)H compared to (0.026, 0.15, 0.022)H from our definition. The differences are less than 20%.

  • In a periodic domain ∫d3Δxfξ = 0 if ∫d3xf = 0, but this does not imply that the line integral in Equation (8) vanishes.

  • The integral in Equation (8) is evaluated by linearly interpolating ξ in the Δx–Δy plane and summing over the interpolated values (trapezoidal rule) along the principal axis. We evaluate the line integral until ξ(l) = e−3ξ(0). The result is insensitive to the upper limit on the integral.

  • 10 

    Simon et al. (2009) study the k dependence of analogous terms on the right-hand side of an equation for $\dot{|B_k|^2}$ (their Equation (19)), which scale like B2. We directly autocorrelate the terms on the right-hand side of Equation (15), and this scales like B4.

Please wait… references are loading.
10.1088/0004-637X/694/2/1010